A Constrained Mean Curvature Flow On Capillary Hypersurface Supported on totally geodesic plane
Abstract.
We prove a new Minkowski type formula for capillary hypersurfaces supported on totally geodesic hyperplanes in hyperbolic space. It leads to a volume-preserving flow starting from a star-shaped initial hypersurface. We prove the long-time existence of the flow and its uniform convergence to a -totally umbilical cap. Additionally, we establish that a -totally umbilical cap is an energy minimizer for a given enclosed volume.
1. Introduction
Mean curvature flow has a rich history, dating back to significant works such as Huisken [13]. Huisken showed that a convex and closed hypersurface will flow to a sphere under the properly rescaled mean curvature flow. A constrained curvature flow is a flow which preserves some geometric quantities. In , Gage [5] used a constrained curve shortening flow to prove an isoperimetric inequality. In higher dimensions, a constrained mean curvature flow was applied to prove the isoperimetric inequality by Huisken [14]. This constrained flow preserves the enclosed volume while decreasing the area of the hypersurface.
An alternative approach to create a flow that preserves the enclosed volume is by employing the Minkowski formula on the hypersurface. This has been explored in [6], which investigated such flows in space forms. Furthermore, this approach has been extended to warped product spaces in [7], where they considered the flow satisfying
Under appropriate assumptions on the metric , the flow is expected to converge to a level set of . See for example [1], [2], [3], [9] and [10] for various types of fully nonlinear curvature flow and anisotropic curvature flow in different ambient spaces.
In recent years, there has been considerable interest in geometric flows of capillary hypersurfaces, for instance, inverse mean curvature flow with free boundary in the Euclidean unit ball [15].
Diving into the main topic of this paper, constrained curvature flow on capillary hypersurfaces has yielded significant results. These include: constrained inverse mean curvature type [19], [23] and mean curvature type flow [12], [20] for capillary hypersurfaces in the Euclidean unit ball; curvature flows [11], [16] and [21] in capillary hypersurfaces in Euclidean half space; mean curvature type flow [17] in geodesic ball in space forms.
In this paper, we consider a new constrained mean curvature type flow for capillary hypersurfaces, which are supported on totally geodesic hyperplanes in hyperbolic space . We use the well-known Poincar half space model , where is the -th coordinate, is the Euclidean metric and . Let be a totally geodesic hyperplane. Denote by the position vector in and the coordinate basis of .
Throughout the paper, we consider as an immersion of hypersurface in . If satisfies that
-
i.
,
-
ii.
,
-
iii.
and contacts at a constant angle on ,
we call a -capillary hypersurface supported on , and the supporting hypersurface.
On a -capillary hypersurface supported on totally geodesic hyperplane , we introduce a novel condition called star-shapedness with respect to .
Definition 1.
Let be a -capillary hypersurface supported on . We say is star-shaped with respect to if it satisfies that
We consider a flow, defined as a family of embeddings with , such that
(1) |
where the normal velocity is given by
We denote by the principal curvature of an umbilical hypersurface . Umbilical hypersurfaces in hyperbolic space can be classified into three types depending on as depicted in the figure. In the case of , it is a totally geodesic hyperplane; in the case of , it is a geodesic sphere, and for , it is an equidistant hypersurface, and if , it is a horosphere. In the Poincaré half space model, can be represented as a plane or sphere with respect to the Euclidean metric. It is easy to see that compact umbilical -capillary hypersurface, can be part of a geodesic sphere, a horosphere and an equidistant hypersurface.

To state our main theorem, we need the following definition.
Definition 2.
We define the -umbilical cap as follows
(2) |
where is a constant vector perpendicular to both and .
For a -umbilical cap, we define a constant by
Note that if and only if the is compact. Combining with Remark 2, we know that its principal curvature .
Now we are ready to state our main theorem.
Theorem 1.
Let be an embedding of a compact capillary hypersurface , supported on the totally geodesic plane with constant contact angle . Suppose there exist constants c, R such that , and is contained in the cap and star-shaped with respect to . We assume that in addition, satisfies that
(3) |
Then the flow (1) exists globally with uniform -estimates. Moreover, uniformly converges to an umbilical cap in topology as , with the same volume of its enclosed domain as .
Remark 1.
The following remark is crucial, which shows that some specific umbilical caps are static along the flow (1).
Remark 2.
For any , is identically zero on umbilical cap . It is well-known that the mean curvature by conformality can be written as
where and we use the fact that the mean curvature of with respect to the metric is . Hence, we get
Therefore umbilical caps are static along the flow (1).
Note that the principal curvature of an umbilical cap depends not only on the radius but also on the last coordinate of its center (in the Euclidean metric sense).
The paper is organized as follows:
In Section 2, we introduce basic notations and definitions of hypersurfaces. In Section 3, we prove a new Minkowski formula on capillary hypersurface supported on a totally geodesic hyperplane, comparing to the one in [4]. In Sections 4, 5 and 6, we follow the method in [20] and [16] to study the scalar equation of the flow (1), in particular, we prove the and estimates. In Section 7, we prove the uniform convergence of the flow.
Acknowledgments .
We would like to thank Juncheol Pyo for helpful comments and hospitality. X. Chai has been partially supported by National Research Foundation of Korea grant No. 2022R1C1C1013511. Y. Chen has been supported by National Research Foundation of Korea grant No. RS-2023-00247299 and partially supported by National Research Foundation of Korea grant No. NRF-2020R1A01005698.
2. Preliminaries
Throughout this paper, we assume that is an embedding of capillary hypersurface along . We denote the second fundamental form of the embedding by . Since is the outer normal field on , . Let be the eigenvalues of , i.e., the principal curvatures of . The -th mean curvature is defined by
and the normalized mean curvature is defined by .
The following so-called Newton transformation defined on the tangent bundle is essential to our formula:
The following properties are well-known,
Let denote the outer normal field of ( is the interior side), denote the outer normal field of the immersed hypersurface , denote the outer normal field of and denote the outer conormal field along . Without loss of generality, assuming as the angle between and , we have the following relation
(4) |
Then the following lemma is widely recognized, and we refer to [22] for its proof.
Lemma 1.
Let be an isometric immersion of a capillary hypersurface supported on . Then is a principal direction of , that is,
(5) |
This property of capillary hypersurfaces is essential in the proof of the Minkowski type formula in the next section.
3. Minkowski type formula
In this section, we introduce a new Minkowski type formula, which is based on the the following properties (see [8]).
Proposition 1.
It satisfies that
(6) |
(7) |
(8) |
(9) |
and
(10) |
These can be directly calculated by the relation between Levi-Civita connections of the metrics conformal to each other, we refer the proof to [8]. From (6), (8), and (9), we can easily see that , , and are conformal Killing vector fields, that is,
(11) |
(12) |
and
(13) |
for any , where denotes the metric .
Restricting the equation (13) to , we have
(14) |
(15) |
and
(16) |
Let be an immersion of -capillary hypersurface supported on the hyperplane . By using the facts above, we can prove the following Minkowski type formula on .
Proposition 2.
For and , it satisfies that
(17) |
where is the -th mean curvature of .
Proof.
Let acts on the both side of and integrate. Using divergence theorem, we have
Let , applying Proposition 1, we have
Let act on both sides, we get
Considering the vector field , from Proposition 1 we have
Similarly, we have
Combining all the equations above, we have
Therefore, we obtain the Minkowski type formula (17). ∎
Let , the Minkowski formula (17) becomes
(18) |
which holds for any capillary hypersurfaces supported on totally geodesic hyperplane . Here is the mean curvature of .
Remark 3.
The second author and Juncheol Pyo [4] gave another version of Minkowski type formula on capillary hypersurfaces supported on a totally geodesic plane, which is presented in Poincaré ball model as follows ( is an Euclidean unit ball and is the Euclidean metric in ).
(19) |
where is a -capillary hypersurface supported on a totally geodesic hypersurface (in Poincaré ball model), , is the position vector and . But unfortunately, we cannot find any umbilical -capillary hypersurfaces where the integrand in (19) is identically zero.
Let denote the domain enclosed by and , and be the domain enclosed by on . The energy functional defined by
is well-known since the critical hypersurface of this functional under any volume preserving variation is a -capillary hypersurface with constant mean curvature, see [18] and [22]. Under a flow with with the given normal velocity and capillary boundary condition as in (1), the following variation formula is well-known:
and
From the Minkowski formula , it is evident that the flow described in (1) is a volume preserving flow.
4. Scalar equation of the flow
In this section, we express the flow (1) by a scalar equation of the radius function.
Let be the position vector defined on which is represented by
where defined on is the distance between and in the Euclidean metric. Since is smooth, star-shaped, the function is well-defined and smooth on . Let , then . Define , where is the gradient of with respect to the ordinary metric on . From the basic facts for radial function, it is well-known that
where .
Throughout the paper, we let the indices range from 1 to and we will apply the Einstein convention.
We use polar coordinates , where is the spherical coordinate on , and the star-shaped hypersurface can be written as
(20) |
where and . Then the Euclidean metric can be written as
and we have
Now we can represent , by the coordinate (20). Indeed, since
and
we can represent by the polar coordinate defined above,
(21) |
Similarly, since
and
we have
(22) |
Denoting and , from (21) and (22) we have
(23) |
and
(24) | ||||
By definition, we have
(25) |
Moreover, by the conformality of mean curvature,
(26) | ||||
where corresponds to the inverse of the metric on , is the -th component of (the gradient of on ) and is the -th component of the Hessian on . Then by calculation,
Writing the flow as a function of , from the flow (1), we see that
Therefore, we obtain the evolution equation of as follows,
(27) | ||||
As for the boundary condition in (1),
Therefore we have
(28) |
Combining (27) and (28), we know that the flow (1) can be written as the parabolic equation of the scalar function as follows,
(29) |
where and is the radial function with respect to of the initial hypersurface .
The short time existence of the flow can be guaranteed by applying the standard PDE theory to (29), due to the assumption on the star-shapedness of . In the following section, we will show the uniform and -estimates for the equation.
Before we start, we need the following calculation.
(30) |
(31) | ||||
(32) | ||||
(33) | ||||
(34) | ||||
where . Now we are ready to prove the and -estimates.
5. estimate
Wang and Weng [20] proved a -estimate for a similar constrained mean curvature type flow of capillary hypersurfaces in the Euclidean unit ball. More specifically, if the initial surface is bounded by two spherical caps that remains stationary under a specified constrained mean curvature flow, the flow will remain bounded by the same two spherical caps. Therefore, a certain pair of spherical caps can be used as barriers of their designed constrained mean curvature flow. A similar property holds in the case of geodesic ball in space forms, see [17].
According to the discussion in Remark 2, the umbilical caps defined by (2) can be regarded as barriers of the flow (1). Then we have a similar corresponding -estimate.
Proposition 3.
Proof.
Let be the defining logarithmic radial function of . Then, is a static solution of the equation of (29), we have
where
and
Since has no singular point if , is bounded and we can denote . We get
Let be the point where attains its nonnegative maximum value. By maximum principle, can only be located on the parabolic boundary, say . That is,
If by Hopf Lemma we have
where denotes the gradient of a function on , and is the normal derivative on . Then we have
From the boundary condition in (29),
which is a contradiction to by monotonicity of the function
Therefore we can only have , that is,
which gives an upper bound by the umbilical cap ’s defining function . Similarly, the desired lower bound can be obtained. We have finished the proof of Proposition 3. ∎
6. -estimate
In this section, we prove a -estimate of the flow (29). Inspired by [20], we use the similar auxiliary function. Let be a nonnegative smooth function on defined by
Indeed, the function is well defined on a neighborhood of but it can be extended to the whole and satisfies that
The following -estimate is essential for the flow (1).
Proposition 4.
If satisfies that , for any , we have
for some positive constant .
Proof.
Define an auxiliary function inspired by [20],
where is a constant to be determined later. Let be the point where attains its maximum. We will discuss case by case to prove the theorem.
Case 1: . At the point , we choose a local coordinate near , such that is an inner normal vector of , and be the geodesic coordinate near along , in the neighborhood of . Under this coordinate, on .
First of all, from the boundary condition of (29), we know that , denote , then from we have,
Applying the Gauss-Weingarten equation, we have
where we have used the fact that since is totally geodesic in . Then, from Hopf Lemma we have,
(35) | ||||
and for , we have
Differentiating the boundary condition (29), together with the fact above we have
therefore we know
(36) |
Then applying (36) to (35), we have
where is a universal constant which satisfies . Then can be chosen large enough so that it comes to a contradiction.
Case 2: If , then
Therefore for any .
Case 3: In the case of , we have
(37) | ||||
Choose a geodesic coordinate, up to a rotation of the one in Case 1, such that
Assume is large enough such that , and are equivalent. Otherwise, we could obtain the desire estimate for . Here we denote .
Firstly, from (37) by letting , we have
(38) |
and letting , we have
(39) |
We can see from the (38) that
(40) |
where , and it is clear that , if there exists some positive such that . Indeed, if not, we would have already proved the theorem.
(41) | ||||
Applying the condition of second derivative on , we have
(42) | ||||
Let us examine the term and at first. Differentiating the parabolic equation (29) with respect to , we have
(43) |
Ricci identity on gives that
(44) |
And by definition, we have
(45) | ||||
Applying (43), (44) and (45) to the terms and gives,
(46) | ||||
It is obvious that . And using (41) we have
(47) | ||||
Similarly, the term be written as follows,
(48) | ||||
And for the same reason, we have and .
For the term , we have
(49) | ||||
Then we compile the following three terms
(50) | ||||
To estimate , we need the following arguments. Let be a constant which satisfies , for some . Then we can assume that
(51) |
Otherwise,
would give an upper bound for , the required estimate is obtained.
Considering and using (51), we have
(52) | ||||
Combining (50) and (52), we have
For the last inequality, Proposition 3 shows that , and hence . Therefore
Let , from the -estimate in Proposition 3, there is an uniform lower bound on . Hence letting , we have
(53) |
Now we consider and . Using (41), (42) and (45), we obtain
and it is easy to see .
7. Convergence of the flow
The higher order a-priori estimates of follow from the uniform and estimates. The same argument as in [20] gives the following result.
Proposition 5.
Before we prove Theorem 1, we need the following lemma.
Lemma 2.
Let and defined by (2). If their enclosed volume are equal, that is, , it holds that
that is, the umbilical -caps with the same enclosed volume have the same principal curvature.
Proof.
Let be an isometry composed by two isometries in , where is a translation along the hyperbolic geodesic , such that,
and is a translation defined by
They are both isometric transformation. Hence
Since is monotonically decreasing as decreasing, so . ∎
Remark 4.
We can see from the proof that, the umbilical caps and on totally geodesic hyperplane with the same enclosed volume is equivalent to the same area and the same wetting area, which is the area of the domain enclosed by and .
Now we are ready to prove Theorem 1.
Proof of the Theorem 1.
Let in the Minkowski type formula (17), we have
Let be the domain enclosed by on . The first variation formula of the energy gives that
where and ’s are the principle curvatures at identified as a point on . Therefore, the energy is monotonically decreasing from Proposition 3, we know that the energy is bounded from above and below. Then integrating both sides of the equation above on , we have
From the uniform estimate in Proposition 5, we have
Therefore any convergent subsequence of must converge to an umbilical cap as . Hence from the boundary condition in (1), we know that they are given by
where is a constant vector perpendicular to both and .
It remains to show that the limit umbilical cap is unique. We follow the proof in [17] and [21]. Let be the unique number such that .
Denote by the radius of the unique umbilical cap
passing through the point . Let and there exists a point attaining by compactness. It then follows that is non-increasing since the is a barrier of by Proposition 3. We claim that
(54) |
We suppose otherwise, then there exist and large enough, such that
(55) |
from the definition of , we have
or equivalently
Taking the derivative of the equation above with respect to , we have
(56) |
Now we evaluate at the point , since is tangential to at the point, then the normal vector at is
(57) | ||||
From the calculation in Remark 2, we have
(58) |
then
(59) |
From (25), (58) and (59), we can see that for small enough, there exist a such that for any , the following
(60) | ||||
holds at .
On the other hand, since converge to and is uniquely determined, we get
Therefore, for , there exists a , such that for any ,
(61) |
Then from (55), (60) and (61), we get
(62) | ||||
In the second inequality we have used Proposition 5. On the other hand, since , we have
(63) |
then combining (57), (62) and (63), we have
This leads to a contradiction to that . Hence (54) holds. Similarly, we can show
where is defined by and is the point achieving . Therefore , and we obtain the uniqueness of the limit. ∎
Remark 5.
In Poincaré half space model of hyperbolic space, the volume of an umbilical cap is determined not only by its Euclidean radius but also by the location of its center, particularly the -th coordinate. As indicated by Remark 2, the location of its center also influences the principal curvatures of the cap. Therefore, we cannot identity the radius by the volume of the domain bounded by initial hypersurfaces .
Note that all the isometries appearing in the proof of Lemma 2 on Poincar half-space model will keep the ratio of an umbilical cap .
From the monotonicity of the energy , we have the following corollary.
Corollary 1.
Let be a -capillary hypersurface supported on the totally geodesic hyperplane . If
-
(1)
is contained in an umbilical cap which satisfies that .
-
(2)
the contacting angle satisfies
then -umbilical caps with the same enclosed volume with are the only minimizers of the energy .
References
- [1] Ben Andrews “Volume-preserving anisotropic mean curvature flow” In Indiana University mathematics journal JSTOR, 2001, pp. 783–827
- [2] Ben Andrews and Yong Wei “Quermassintegral preserving curvature flow in hyperbolic space” In Geometric and Functional Analysis 28.5 Springer, 2018, pp. 1183–1208
- [3] Ben Andrews and Yong Wei “Volume preserving flow by powers of the th mean curvature” In Journal of Differential Geometry 117.2 Lehigh University, 2021, pp. 193–222
- [4] Yimin Chen and Juncheol Pyo “Some rigidity results on compact hypersurfaces with capillary boundary in Hyperbolic space” In arXiv preprint arXiv:2206.09062, 2022
- [5] Michael Gage “On an area-preserving evolution equation for plane curves” In Nonlinear problems in geometry 51 American Mathematical Society, 1986, pp. 51–62
- [6] Pengfei Guan and Junfang Li “A mean curvature type flow in space forms” In International Mathematics Research Notices 2015.13 Oxford University Press, 2015, pp. 4716–4740
- [7] Pengfei Guan, Junfang Li and Mu-Tao Wang “A volume preserving flow and the isoperimetric problem in warped product spaces” In Transactions of the American Mathematical Society 372.4, 2019, pp. 2777–2798
- [8] 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
- [9] Yingxiang Hu and Haizhong Li “Geometric inequalities for static convex domains in hyperbolic space” In Transactions of the American Mathematical Society 375.08, 2022, pp. 5587–5615
- [10] Yingxiang Hu, Haizhong Li and Yong Wei “Locally constrained curvature flows and geometric inequalities in hyperbolic space” In Mathematische Annalen 382.3 Springer, 2022, pp. 1425–1474
- [11] 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
- [12] Yingxiang Hu, Yong Wei, Bo Yang and Tailong Zhou “On the mean curvature type flow for convex capillary hypersurfaces in the ball” In Calculus of Variations and Partial Differential Equations 62.7 Springer, 2023, pp. 209
- [13] Gerhard Huisken “Flow by mean curvature of convex surfaces into spheres” In Journal of Differential Geometry 20.1 Lehigh University, 1984, pp. 237–266
- [14] Gerhard Huisken “The volume preserving mean curvature flow.” In Journal für die reine und angewandte Mathematik 1987.382, 1987, pp. 35–48 DOI: doi:10.1515/crll.1987.382.35
- [15] Ben Lambert and Julian Scheuer “The inverse mean curvature flow perpendicular to the sphere” In Mathematische Annalen 364 Springer, 2016, pp. 1069–1093
- [16] Xinqun Mei, Guofang Wang and Liangjun Weng “A Constrained Mean Curvature Flow and Alexandrov–Fenchel Inequalities” In International Mathematics Research Notices 2024.1 Oxford University Press, 2024, pp. 152–174
- [17] Xinqun Mei and Liangjun Weng “A constrained mean curvature type flow for capillary boundary hypersurfaces in space forms” In The Journal of Geometric Analysis 33.6 Springer, 2023, pp. 195
- [18] Antonio Ros and Rabah Souam “On stability of capillary surfaces in a ball” In pacific journal of mathematics 178.2 Mathematical Sciences Publishers, 1997, pp. 345–361
- [19] 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
- [20] Guofang Wang and Liangjun Weng “A mean curvature type flow with capillary boundary in a unit ball” In Calculus of Variations and Partial Differential Equations 59.5 Springer, 2020, pp. 149
- [21] 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
- [22] Guofang Wang and Chao Xia “Uniqueness of stable capillary hypersurfaces in a ball” In Math. Ann. 374.3-4, 2019, pp. 1845–1882
- [23] 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