The area minimizing problem in conformal cones: II
Abstract.
In this paper we continue to study the connection among the area minimizing problem, certain area functional and the Dirichlet problem of minimal surface equations in a class of conformal cones with a similar motivation from [14]. These cones are certain generalizations of hyperbolic spaces. We describe the structure of area minimizing -integer multiplicity currents in bounded conformal cones with prescribed graphical boundary via a minimizing problem of these area functionals. As an application we solve the corresponding Dirichlet problem of minimal surface equations under a mean convex type assumption. We also extend the existence and uniqueness of a local area minimizing integer multiplicity current with star-shaped infinity boundary in hyperbolic spaces into a large class of complete conformal manifolds.
2010 Mathematics Subject Classification:
Primary 49Q20: Secondary 53C21. 53A10,. 35A01. 35J251. Introduction
In this paper we continue to study the area minimizing problem with prescribed boundary in a class of conformal cones similar to [14]. A conformal cone in this paper is defined as follows.
Definition 1.1.
Let be a -dimensional Riemannian manifold with a metric , let be the real line with the metric and be a positive function on . In this paper we call
(1.1) |
as a conformal product manifold. Let be a bounded domain in . We refer in as a conformal cone, denoted by .
Let be a function on and be its graph in . The area minimizing problem in a conformal cone is to find a -integer multiplicity current in , the closure of , to realize
(1.2) |
where is the mass of integer multiplicity currents in , and denotes the set of n-integer multiplicity currents with compact support in , i.e. for any , its support is contained in for some finite numbers . See subsection 4.1 for more details.
The main reason to study the conformal product manifold in Definition 1.1 is that hyperbolic space is a special case of (see remark 6.5). With this model in [22, Theorem 4.1], Hardt-Lin showed that there is a unique local area minimizing -integer multiplicity current for any prescribed infinity star-shaped boundary. Moreover it is a radial minimal graph over in hyperbolic spaces. On the other hand in [25, Section 4.1], Lin described the solution to the area minimizing problem (1.2) with graphical boundary in a bounded cylinder via a minimizing problem of an area functional of bounded variation (BV) functions. Motivated by Lin’s idea in [14] we studied the area minimizing problem (1.2) in a conformal product manifold . Note that there only depends on . Based on these preceding results, it is natural to consider the corresponding area minimizing problem (1.2) in conformal cones of . In particular it is desirable to know how many phenomena in hyperbolic sapce does not really rely on the hyperbolic structures. We refer readers to [2] for some history remarks on such kinds of area minimizing problems.
For any open set we denote the set of all bounded variation functions on by . A key concept for our study of the problem (1.2) is an area functional in defined as follows.
(1.3) |
where and are the volume form and the divergence of respectively, and denote the set of smooth functions and vector fields with compact support in respectively. Note that when , is the area of the graph of in .
Let be the domain in Definition 1.1. Suppose is a domain in such that . That is, the closure of is a compact set in . Suppose . The following minimizing problem
(1.4) |
also plays an important role to solve (1.2).
In fact the key idea to solve (1.2) is to establish the connection among the problem (1.2), the area functional minimizing problem (1.4) and the Dirichlet problem of minimal surface equations in . This can be easily seen when is a minimal graph of in over with boundary on . From Theorem 2.3 one has (1) should satisfy
(1.5) |
with on ;
(2) the area of is less than that of any compact surface with as the graph of containing in ; (3) realizes the minimum of (1.4) if requiring all .
In a similar connection between the problem (1.2) and the problem (1.4) is obtained in Theorem 4.1. It says that if is the solution to the problem (1.4), then solves the problem (1.2) in where is the subgraph of and is the corresponding integer multiplicity current. This generalizes Lin’s result [25] into conformal cones in . Its proof is based on the following three observations: (1) by Theorem 3.9, is just the perimeter of its subgraph ; (2) if solves (1.2) and for , then is a boundary of a Caccioppoli set (see Lemma 4.6); (3) by Theorem 3.10, for any Caccioppoli set in with , there is a such that is less than the perimeter of . Note that the observation (3) is a very general phenomenon in area-type functionals [29], [26], [30], [18],[15] and [7] (see remark 3.11). We refer it as the Miranda’s observation.
A direct application of Theorem 4.1 is the Dirichlet problem of minimal surface equations in . In Theorem 5.3 we show that if is -mean convex, then the Dirichlet problem (1.5) with continuous boundary data has a unique solution in . Here is -mean convex if the mean curvature of satisfies that
where is the outward normal vector of and . A more general form of the Dirichlet problem (1.5) was already considered by Casteras-Heinonen-Holopainen [8]. To obtain the esimate they relied on a lower bound of the Ricci curvature and the positive mean curvature of (see remark 5.4) . But out result in Theorem 5.3 is independent of the curvature of and also can not be obtained by the classical continuous method in [16, Chapter 11, 17.2] (see remark 5.5). A consequence of Theorem 5.3 is that if is -mean convex, the area minimizing integer multiplicity current in (1.2) is unique as the graph of a function to solve the Dirichlet problem (see Theorem 5.6).
At last we consider the existence and unqiueness of local area minimizing integer multiplicity current in with infinity boundary when can be written as which goes to as goes to zero in . Here is a compact Riemannian manifold with boundary and is the distance function in . Let be the set . In Theorem 6.4 we obtain that if there is such that for any is -mean convex, then for any and there is a unique local area minimizing integer multiplicity current with infinity boundary . Moreover is a minimal graph in over . This illustrates that Theorem 4.1 in [22] does not depend on the hyperbolic structure (see remark 6.5 ). Unlike the case in hyperbolic spaces our existence in Theorem 6.4 is from the Dirichlet problem of minimal surface equations in Theorem 5.3, not from the results in [5, 6] by Anderson via geometric measure theory.
For the idea of radial graphs in hyperbolic space we refer to [19, 9, 27, 30, 36, 31, 20, 21, 26] etc. For the Dirichlet problem of minimal surface equations in Riemannian manifolds, we refer to [8], [4], [35], [1], [3] and references therein. As for the variational method to study the Dirichlet problem of minimal surface equations, we refer to [18, 17, 7], [41] and references therein.
Our paper is organized as follows. In section 2 we show three properties of minimal graphs in and collect preliminary facts on BV functions. In section 3 we show the approximation theorem of (Theorem 3.8) and the Miranda’s observation (Theorem 3.9 and Theorem 3.10). In section 4 we consider the connection between the problem (1.2) and the problem (1.4) (Theorem 4.1). In section 5 we discuss the Dirichlet problem of minimal surface equations in on a -mean convex domain (Theorem 5.3). As an application we obtain the uniqueness of the problem (1.2) under the -mean convex assumption (Theorem 5.6). In section 6 we discuss the existence and uniqueness of local area minimizing -integer multiplicity current with infinity graphical boundary (Theorem 6.4). In Appendix A we give a proof of the interior estimate of minimal surface equations in (1.5) (Theorem A.3) following from an idea of Eichmair [11].
The first author is supported by the National Natural Science Foundation of China, No. 11771456. The second author is supported by the National Natural Science Foundation of China, No.
11801046 and by the Fundamental Research Funds for the Central Universities, China, No.2019CDXYST0015.
2. Preliminaries
Throughout this section we assume is a manifold with a metric , is a positive function defined on and is the conformal product manifold with the metric . We show three properties of minimal graphs in and collect some results on BV functions in Riemannian manifolds for later use.
2.1. Three properties
Definition 2.1.
Let be a orientable hypersurface in a Riemannian manifold with a normal vector . We call as the mean curvature of with respect to .
Let be a bounded domain in and . Let be the graph of . Then its upward normal vector in the product manifold is where denotes the gradient of and . Then we have
Lemma 2.2.
The mean curvature of with respect to its upward normal vector in is
(2.1) |
where , is the dimension of and is the divergence of .
Proof.
We write the metric of as where . The upward normal vector of in is . By [40, Lemma 3.1], the mean curvature of in with respect to is
(2.2) |
where is the mean curvature of with respect to in the product manifold and is the dimension of . With a straightforward computation is where is the divergence of . Putting and into (2.2), we obtain the conclusion. ∎
Now we summarize three properties of minimal graphs in as follows.
Theorem 2.3.
Suppose is a minimal graph of over with boundary . Then
-
(1)
satisfies that
(2.3) -
(2)
Let be a domain such that and . Further assuming on , then realizes
(2.4) where and is the volume form of .
-
(3)
The area of achieves the minimum of the area of all compact orientable hypersurface in with .
Proof.
Note that the property (1) is from Lemma 2.2.
For any graph over , its area in is
(2.5) |
where is the volume form on . Thus the property (2) follows from the property (3). It is sufficient to show the property (3).
Recall that . For any , define a map as . We define a new vector field in as
(2.6) |
where is the pushforword of . Thus is a smooth unit vector field in the tangent bundle of . By the definition of divergence, (see Page 423 in [23]) we define an -form as
(2.7) |
A key fact is
(2.8) |
where and is the divergence and volume form of . Note that is an isometry of for every . Because is minimal in , so is . Then on the whole .
Now let be any compact orientable hypersurface in satisfying . Without loss of generality we can assume and bounds a domain in . Suppose that the outward normal vector of on and are denoted by and , the volume form of on and are denoted by and respectively. It’s clear that the upward normal vector of in , , is . Applying the divergence theorem in the domain , we obtain
with equality if and only if . Thus we obtain the property (3). The proof is complete. ∎
2.2. BV functions
Now we collect some preliminary facts for BV functions in Riemannian manifolds. We refer readers to the books of Giusti[18], Evans-Gariepy[12] and the papers of McGonagle-Xiao [28], [41] etc. Recall that is a bounded domain in and is the volume form of . Let be the set of smooth vector fields with compact support in .
Now we define BV functions and Caccioppoli sets as follows:
Definition 2.4.
Let . Define
(2.9) |
where is the volume form of .
If , we say that has bounded variation in . The set of all functions with bounded variation is denoted by . If belongs to for any bounded domain , we say .
Let be a Borel set in and be the characteristic function of . If , then is called as a Caccioppoli set and is called as the perimeter of . In the reminder of this paper, we also write it as .
Remark 2.5.
In some settings, in order to emphasize the ambient manifold , we use the notation instead of .
For a Caccioppoli set all properties are unchanged if we make alterations of any Hausdorff measure zero set. Arguing exactly as Proposition 3.1 in [18], we can always choose a set differing by a Hausdorff measure zero set with and satisfying for any
(2.10) |
where is sufficiently small. From now on, we always assume that condition (2.10) holds for any Caccioppoli set .
Definition 2.6.
We say a sequence of measurable functions locally converges to in if for any open set we have
(2.11) |
Remark 2.7.
If converges to a.e, then this sequence locally converges to in .
By the definition in (2.9) it is easy to see that
Theorem 2.8 (Lower-semicontinuity).
Let be a sequence of functions in locally converging to a function in the . Then
(2.12) |
The following approximation result is not trivial when is not contained in a simply-connected domain of a Riemannian manifold. Because in this case no global symmetric mollifies exist for this domain as that in [18, page 15] for Euclidean spaces. For a complete proof, we refer to [41, section 3].
Theorem 2.9 (Theorem 3.6, [41]).
Suppose . Then there exists a sequence of functions in such that
(2.13) | |||
(2.14) |
As a conclusion, we can obtain the following well-known compactness result.
Theorem 2.10 (Compactness theorem).
Let be a bounded domain with Lipschitz boundary. Suppose there is a constant such that for
(2.15) |
Then there is a function such that there is a subsequence of converging to in .
We can also view a function as a Radon measure.
Theorem 2.11 (Theorem 2.6, [41]).
Let be a bounded domain. Suppose that .
-
(1)
There exists a Radon measure on such that
(2.16) for any open set and any nonnegative function .
-
(2)
There exists a vector field on satisfying
(2.17) where -a.e. for any .
3. The area functional
In this section we define a new area functional for BV functions to generalize the area of the graph of functions in (see (2.5)). Suppose is a fixed Riemannian manifold with metric and is a positive function on . Here is the set equipped with . We shall establish the approximation of and the Miranda’s observation mentioned in the introduction.
Recall that is a bounded domain in , and denote the set of all smooth vector fields and smooth functions with compact supports in respectively. Let denote the volume form of .
Definition 3.1.
Let be a measurable function in . Define
(3.1) |
Remark 3.2.
If , is the area of the graph of in the conformal product manifold .
3.1. Properties of the area functional
By the definition we have
Lemma 3.3 (Lower-semicontinuity).
Let be a sequence converging to in . Then
(3.2) |
The following lemma establishes a relation between and .
Lemma 3.4.
There is a positive constant such that for any
(3.3) |
Proof.
The area functional induces a Radon measure on as follows.
Theorem 3.5.
Suppose . Then there is a Radon measure on such that for any open set in ,
(3.6) |
Proof.
For any nonnegative function , define
(3.7) |
It is easy to see that for any positive constant and . Thus is a positive linear functional on . By [34, remark 4.3] (see also [41, Theorem 2.5]), then there is a Radon measure on such that
(3.8) |
for any open set . Here denotes the closure of . From the definition it is clear that . The proof is complete. ∎
Fix any . Let be the exponential map at . Let be the finite number given by
(3.9) |
For any , let be the Euclidean ball centering at with radius . Thus is a diffeomorphism. Via this exponential map we identify with equipped with the metric . Then is called a normal ball. A vector field along can be represented by
(3.10) |
Let be a symmetric mollifier in , i.e. a function satisfying , and . Here denotes . If , let denote the convolution of where namely
(3.11) |
The convolution of a vector field and on is defined by
(3.12) |
Note that . Suppose . By a direct computation we have
(3.13) |
(3.14) |
To prove the approximation property of the area functional , we need the following two techinique lemmas from [41]. Both of their proofs follow from those of [41] with minor modifications. So we skip them here.
Lemma 3.6 (Theorem 2.12, [41]).
Let be a normal ball in with a metric . Let be a nonnegative continuous function and be a compact subset. Then for any , there exists a such that for all , any continuous function and any vector field satisfying
(3.15) |
we have
(3.16) |
where is defined by
(3.17) |
Lemma 3.7 (Lemma 2.13, [41]).
Let be a normal ball in with a metric . Suppose and be a smooth function with compact support in . Then for any there is a such that for all and any smooth vector field on satisfying .
(3.18) |
where is the covariant derivative of , and assume outside .
The approximation property of is stated as follows.
Theorem 3.8.
Let be a bounded domain in and . Then there is a sequence such that
(3.19) | |||
(3.20) |
Proof.
By Lemma 3.4 and that is bounded imply that is finite.
By Theorem 3.5, there is a Radon measure on such that for any open set . Fix any and . By [41, Theorem A.4], there exists a countable open cover of and a positive integer such that
-
(1)
Each is a normal ball such that and .
-
(2)
is a pairwise disjoint subcollection with
(3.21) -
(3)
The subcollection satisfies
(3.22) where is a positive integer only depending on and .
Take a partition of unity subordinate to the open cover , that is
(3.23) |
and for each . And for any in there is a compact neighborhood of such that on the summation above is finite.
Now fix any and any satisfying . For each by Lemma 3.6 and Lemma 3.7 we can choose so small independent of and such that ,
(3.24) | |||
(3.25) |
(3.26) |
Here . Now we define by
(3.27) |
Thus . By (3.26) we obtain
The last line we use the fact that . By definition we have
(3.28) |
Combining this with (3.21), (3.22) together we obtain
(3.29) |
Since are chosen arbitrarily satisfying and , we conclude that
(3.30) |
As a result we can choose such that . On the other hand by (3.24) converges to in . On the other hand by Lemma 3.3 . Thus we obtain the conclusion. The proof is complete. ∎
3.2. The Miranda’s observation
In this subsection we show the Miranda’s observation for mentioned in the introduction.
Let be a measurable function on . The set is called as the subgraph of , written as . For any Borel set we denote by its characterization function.
Let denote the perimeter of Caccioppoli sets in . Its connection with the area functional can be summarized as follows.
Theorem 3.9.
Suppose , is its subgraph and is a bounded domain. Then
(3.31) |
Proof.
By Lemma 3.4 is finite. If is a smooth function, the equality is obvious since both sides equal to the area of the graph of in .
Suppose . By Theorem 3.8 there is a sequence of smooth functions such that converges to in and converges to as . Let be the subgraph of . Note that converges to in a.e. By Lemma 2.8 we have
(3.32) |
To prove the converse first suppose that for some positive constant . Observe that we can add a constant to without changing and . Thus without loss of generality we assume that . Let and satisfying . Let be a smooth nonnegative function with compact support in such that in and on . Define
(3.33) |
Then and in . Observe that for each , we have
(3.34) |
where is a constant depending only on . We denote the product manifold with metric by . Let and be the divergence on and respectively. Note that (For example see [41, (2.11)]. Then by definition of perimeter, we have
Taking supremum over both sides for all with yields that . With (3.32) this yields (3.31) in the case of .
As for the general case we consider an approximation procedure based on the finite case. Let be the truncation of by , i.e. . Let be the subgraph of . Thus converges to a.e. in . By (3.32) and the lower semicontinuity in Lemma 3.3, is finite. Note that
(3.35) |
Combining this with Lemma 3.3 we obtain
(3.36) |
As a result we conclude
(3.37) |
Combining this with (3.32) yields the conclusion. The proof is complete. ∎
As an application we obtain a decreasing property of certain Caccioppoli sets in .
Theorem 3.10.
Let be a Caccioppoli set in with the following assumption: for almost every , there exists a such that for all and for all . Then the function
(3.38) |
is well-defined and
(3.39) |
Remark 3.11.
This result should be firstly observed by Miranda [29] in the case of and (see also [2]). It is a generalization of [18, Theorem 14.8] and the symmetrization of hyperbolic spaces in [25, remark 2.3] by Lin. It is also similar to the arrangement of constant mean curvature functional in [9, Theorem 3.1] by De Silva-Spruck, the decreasing perimeter property of singular area functional in [7, Lemma 9] by Bemelmans-Dierkes, [15, Lemma 3.3] by Gerhardt and [41, Lemma 5.8].
Proof.
By the assumption on it is not hard to see that is well defined. Moreover we assume that has finite perimeter in . Otherwise nothing needs to prove.
First we assume that there is a such that . Let be a compactly supported smooth function on such that on and . For each , define
(3.40) |
Choose and satisfying . Define
(3.41) |
Thus . Then we have
(3.42) |
where and denote the volume form and the divergence of respectively, is the divergence of the product manifold . By the assumption on and , for almost every we have
(3.43) | |||
(3.44) |
where is a constant depending only on and .
Let , we get
(3.45) |
Because and are arbitrarily choosen satisfying , we obtain (3.39) in the case of .
As for a general Caccioppoli set with finite perimeter, we set . Note that and . By the lower semicontinuity in Lemma 3.3, we obtain
(3.46) |
Define
By the assumption of , converges to a.e. as . Again by Lemma 3.3
(3.47) |
The proof is complete. ∎
Theorem 3.10 and the following result together are referred as the Miranda’s observation.
Theorem 3.12.
Let with . Suppose locally minimizes , i.e. if where are the subgraphs of respectively. Then locally minimizes the perimeter in , i.e.
for any Caccioppoli set satisfying .
4. The area minimizing problem
In this section we consider the area minimizing problem in the conformal product manifold (see Definition 1.1) similar to that in [14].
Throughout this section suppose and are two bounded domains in satisfying . Let be the set in and be its closure .
Suppose . We denote by the dimension of . Let denote the set of -integer multiplicity currents with compact support in , i.e. for any , its support is contained in for some finite numbers . Let be the graph of on . The area minimizing problem in this section is to find a solution to attain the value
(4.1) |
where is the mass of in (see subsection 4.1).
The main result of this section is stated as follows.
Theorem 4.1.
Remark 4.2.
In the case of , the above theorem is obtained by Lin in [25, section 4.1]. But we do not understand the proof of its uniqueness just assuming is bounded Lipschitz.
Remark 4.3.
We show that for in the property (2) . For we denote the set by . When is sufficiently small, is a domain. We define . It is obvious that is a graph near . Thus . And converges to and converges to in the sense of current as . Moreover converges to in the sense as . By the definition of the current . This is the reason that we assume .
4.1. Integer Multiplicity Currents
Here we collect some necessary facts on integer multiplicity currents. Our main references are [34, section 27] by Simon and [24] by Lin-Yang.
Let be an open domain in a Riemannian manifold with dimension and denotes the -dimensional Hausdorff measure in for any . Suppose is an integer in . Let be the set of all -smooth forms with compact support in . A -current in is a linear functional in .
Definition 4.4.
A set is said to be countable -rectifiable if
where and is a Lipschitz map for each .
Now we can define a -integer multiplicity rectifiable current.
Definition 4.5.
Let be a -current in . We say that is a -integer multiplicity current if
(4.3) |
where is a countable -rectifiable subset of , is a positive locally -integrable function which is integer-valued, and is a -form oriented the tangent space of a.e. . is also written as
.
The mass of in is
(4.4) |
where denotes the usual pairing of -form. The boundary of is defined by for any .
Remark 4.6.
For any -submanifold is a k-integer multiplicity current just choosing as its orientation which is equal to
If the dimension of an integer multiplicity current , written as , is equal to the dimension of we always choose as the volume form of . In this case is written as .
A good property of integer multiplicity currents is their compactness obtained by Federer and Fleming in [13] (see also [34, Theorem 27.3]).
Theorem 4.7.
Suppose is a sequence of -integer multiplicity currents with
for any open set . Then there is a -integer multiplicity current such that converges weakly to and
A useful way to construct integer multiplicity currents is the pushforward of local Lipschitz maps.
Definition 4.8.
Let be two open sets in (different) Riemannian manifolds. Suppose is local Lipschitz, is a -integer multiplicity current and is proper. We define by
(4.5) |
Theorem 4.9.
Let be the conformal cone in (see Definition 1.1) and . Let be a positive integer. Let be a -integer multiplicity current in satisfying . Then there is a -integer multiplicity current in such that . Here may be noncompact in .
Remark 4.10.
A similar proof also appears in [34, section 26.26].
Proof.
The proof is exactly the same as that of [14, Theorem 3.9]. Note that even with the metric the map as is still proper and local Lipschitz. Thus is well-defined. Moreover
The conclusion follows from that . ∎
Fix two numbers such that . Now consider an auxiliary problem of (4.1) as follows.
(4.6) |
By Theorem 4.7 there is a n-integer multiplicity current contained in with satisfying . We denote by an open domain in satisfying . Similar to [14, Lemma 4.6] we have
Lemma 4.11.
We can choose such that there is a Caccioppoli set in such that and is the subgraph of outside .
Remark 4.12.
Because is contained in one sees that for each it holds that for and for . If not the case we can replace , the complement of , with without any change of the perimeter.
Proof.
Since we can extend as a function on satisfying on and its subgraph is a Caccioppoli set in . Define . Then by remark 4.3 .
From Theorem 4.9 there is a -integer multiplicity current in such that
. Then we have
(4.7) |
Observe that can be represented as where is some integer value measurable function on . Since , outside i.e.
(4.8) |
Now define a function if and if . Let denote the function of .
Set . It is not hard to see that and
(4.9) |
As a result one has the following decomposition
(4.10) |
where is the integer multiplicity current . By (4.8) and the definition of , outside . As a result .
Now define for any integer . By the definition of , we have for any . Note that . Applying the decomposition theorem of integer multiplicity currents [34, corollary 27.8] we obtain
This implies that
(4.11) |
For applying the decomposition theorem [34, corollary 27.8] again gives that
(4.12) |
By (4.11) . As for the decomposition theorem gives that . Since , with (4.7) and (4.10), [14, Lemma 3.17] implies that
(4.13) |
and from (4.11). Thus . From (4.10),
(4.14) |
and . Since solves (4.1), (4.13) implies that we can choose as . By (4.10) and (4.9) coincides with the subgraph of outside . The proof is complete. ∎
4.2. Regularity of almost minimal boundary
In this subsection we recall some results on the regularity of the almost minimal boundary from [14, section 3.2]. All of their proofs are skipped here. We refer the reader to [18], [37] and [10] for more details. Throughout this subsection let be a -dimensional Riemannian manifold.
Suppose is a -dimensional integer multiplicity current in , represented as where is a measurable set of . The relationship between the mass and BV functions can be summarized as follows:
(4.15) |
for any open set . In particular if is a Caccioppoli set in and we have
(4.16) |
where is the perimeter in . For a derivation see [34, subsection 27.7].
For a point and we denote by the open ball centered at with radius . Now we define an almost minimal set in an open set and a closed set respectively.
Definition 4.13.
Let be a domain. Fix and . Suppose is a Caccioppoli set in .
-
(1)
We say that is an -almost minimal set in if there is an and a constant with the property that for any , with ,
(4.17) where is any Caccioppoli set satisfying . In particular if , we say is a minimal set in .
-
(2)
We say that is an -almost minimal set in if there is an and a constant with the property that for any ,
(4.18) where is any Caccioppoli set satisfying . In particular if , we say is a minimal set in .
-
(3)
The regular set of is the set is a graph in a ball containing . The singular set of is the complement of the regular set in .
Remark 4.14.
By [41, Lemma 7.6] all bounded domains are almost minimal sets in an open neighborhood of their boundaries.
A good property of almost minimal sets in a domain is their boundary regularity.
Theorem 4.15 (Theorem 1 in [37], Theorem 5.6 in [10]).
Suppose a Caccioppoli set is a -almost minimal set in a domain . Let be the singular set of in . Then
-
(1)
if , ;
-
(2)
if , consists of isolated points;
-
(3)
if , for any . Here denotes the Hausdorff measure in .
where .
Remark 4.16.
Note that in general the boundary of almost minimal sets in closed sets does not have such good regularity.
The following result is very important.
Theorem 4.17 (Theorem 3.19 in [14]).
Let and be two domains in . Define . Fix a point in . Suppose is an - almost minimal set in (the closure of ) and passes through , then is a graph in an open ball containing for some .
Remark 4.18.
In general the boundary of is not .
4.3. The proof of Theorem 4.1
Proof.
Set and . Fix such that .
By Lemma 4.11 there is a Caccioppoli set in such that realizes the minimum in (4.6) and is the subgraph of outside .
We claim that has to be contained in the closed set . If not the case, assume
(4.19) |
Let be the point on achieving this maximum. Since , there is an embedded ball such that .
The first case is that for some . Note that is an minimal set in By Theorem 4.17 is still a graph near for some . Moreover is tangent to and at . But and is orthogonally transverse at . This gives a contradiction and the first case is impossible.
The second case is for some . By remark 4.12 . Thus is a minimal set in away from . Note that be the intersection of two domain. Again by Theorem 4.17 is a graph in a neighborhood of contained in . But we should observe that is also minimal in by Theorem 2.3. With respect to the upward normal vector, near in the Lipschitz sense. By the maximum principle in [14, Theorem A.1] coincides with near . From the connectedness of we obtain that . This is impossible since is contained in . Thus the second case is also impossible. This means . Arguing a similar derivation we will obtain that
(4.20) |
Thus the above claim is true.
For each with , is contained in and is uniformly bounded and . By Theorem 4.7 as goes to the sequence will converge to a such that and
Moreover from the compactness of BV functions in Theorem 2.10 there is a Caccioppoli set such that which is contained in and is the subgraph of outside . By Theorem 3.10 there is a BV function such that outside and
(4.21) |
where is the subgraph of and
(4.22) |
Moreover by remark 4.3. Since coincides with outside , then
(4.23) |
Thus also solves the area minimizing problem (4.1). This shows the existence of with the property (2).
Now fix any satisfying outside . Let be the subgraph of . By remark 4.3 . By the property (2)
(4.24) |
Because outside , by Theorem 3.9 and (4.16) we obtain
(4.25) |
Thus realizes the minimum of
(4.26) |
We conclude that statisfies the property (1).
At last we show that has the property (3). By the property (2) is a minimal set in . By Theorem 4.15, except a closed set with , is connected and for some . Since is minimal in and is , the regularity of minimal surface equation (2.3) implies that it is for any . Let be the normal vector of in pointing to the positive infinity. Define . Thus on in which is connected and . By the Harnack principle in corollary A.2, we have or on . Since is contained in and , only can happen in .
Let be the orthonormal projection of in . Therefore and . Moreover
(4.27) |
This implies that . By the property (1) in Theorem 4.1 is also the critical point of the functional
(4.28) |
From the removable singularity result of Simon [33, Theorem 1] is regular at every in . Thus satisfies (2.3) in the Lipschitz sense. Since , by the classical regularity of uniformly elliptic equations is for any . The proof is complete. ∎
5. the Dirichlet problem
In this section we apply Theorem 4.1 to solve the Dirichlet problem of minimal surface equations in .
Definition 5.1.
We say that is -mean convex if the mean curvature of satisfies . Here denotes the covariant derivative of and for the outward normal vector on .
Remark 5.2.
5.1. The Dirichlet problem on -mean convex domain
The first result of this section is given as follows.
Theorem 5.3.
Let be a bounded -mean convex domain in and let be a positive function in . For any the Dirichlet problem
(5.1) |
has a unique solution in . Here .
Remark 5.4.
Suppose is for some . Casteras-Heinonen-Holopainen [8, Theorem 2] showed that if there is a positive constant such that
-
(1)
, satisfying
-
(2)
the Ricci curvature of satisfies and
Then the Dirichlet problem
(5.2) |
with on for any has a solution in .
In the case of Theorem 5.3 removes the curvature assumption on in [8, Theorem 2]. Our mean curvature assumption on should be optimal. For example when this is confirmed by Serrin [32] in Euclidean spaces.
Remark 5.5.
The above result can not be obtained by the continuous method in [16, section 11.2, 11.3, Chapter 18]. The reason is that the boundary assumption may not be preserved by these methods.
Proof.
The uniqueness of the Dirichlet problem (5.1) is obvious by the maximum principle. We only need to show the existence.
First we assume that . Without loss generality we can assume for some strictly containing . Let be the set and be its closure. By Theorem 4.1 there is a with outside such that realizes the area minimizing problem (4.1). Moreover . As a result is a minimal graph in . By the conclusion (1) in Theorem 2.3 satisfies on . It only suffices to show that
Fix any . Now define
(5.3) |
where is the distance function in . By Theorem 4.1 for any . Thus is a finite number. Suppose . There is a sequence in such that and . Since is continuous at , there is a neighborhood of the point such that is disjoint with , the graph of in .
By Theorem 4.1 is a minimal set in and passes through the point . Since is , Theorem 4.17 implies that is a graph near . Since is a minimal set in , then near with respect to the outward normal vector of in the Lipschitz sense. Since is -mean convex, by remark 5.2 is a mean convex domain in . Because and is tangent to . From the maximum principle in [14, Theorem A.1], coincides with in . This contradicts to the definition of . Thus
(5.4) |
With a similar derivation we also obtain
(5.5) |
Combining the above two facts together yields that is continous for any fixed , thus . As a result we obtain the existence of the Dirichlet problem (5.1) when .
Now we consider the general case of . Then there are two sequences in such that
(5.6) |
and both of them converge to in the . By the previous argument for any positive integer let be the solution of the Dirichlet problem (5.1) on with on . Applying the maximum principle on (5.1), we obtain
-
(1)
where ;
-
(2)
on ;
According to Theorem A.3, has a local uniformly bound of their norms in . Now we can choose a subsequence, still written as such that on and and satisfies on .
The only thing to left is to show that and on . By the conclusion (2) above applying the maximum principle on yields that
(5.7) |
for any positive integers . This implies that for any
(5.8) |
Letting go to the positive infinity we obtain for any . Thus we define on and . The proof is complete. ∎
5.2. The uniqueness of area minimizing currents
Theorem 5.6.
Suppose is a -mean convex domain and . Set . Let be the solution of the Dirichlet problem (5.1) with boundary data . Let be the subgraph of .
Remark 5.7.
This theorem is a finite version of [22, Theorem 4.1] in finite conformal cones in .
Proof.
By Theorem 5.3, let be a n-integer multiplicity current with compact support in to realize (4.1). For any , let be the subgraph of over . By (5.1), is a minimal graph over in and the boundary of is disjoint with if .
Now the following number
(5.9) |
is well-defined since is compact. Suppose . Then is tangent to at some point .
By Lemma 4.11, where is a minimal set in the set . From the definition of , . By Theorem 4.17 is near . In the following let denote the mean curvature. Thus near with respect to the upward normal vector in the Lipschitz sense. Note that near . Then the maximum principle in [14, Theorem A.1] implies that coincides with near . By the connectedness of , we have . Thus for some , the point is contained in .
Since , there is a neighborhood of disjoint with . Then is a minimal set in . By Remark 5.2, is mean convex in . Again by [14, Theorem A.1], coincides with near . With the connectedness, the set is contained in . This is impossible since is compact. Thus and . With a similar derivation, we can show that . Finally combining the above two facts yields that where
is the subgraph of . We obtain the conclusion (1).
Suppose there are two functions in to solve (4.2). Let be the subgraph of and for . By (2) in Theorem 4.1, should solve (4.1). By the conclusion (1) we just obtained, and on . We arrive the conclusion (2). The proof is complete.
∎
6. Infinity boundary cases
In this section we apply the results in previous sections to generalize the existence and the uniqueness of area minimizing graphs with the infinity star-shaped boundary in hyperbolic spaces.
Throughout this section, we assume that is a compact -dimensional Riemannian manifold with its metric and compact embedded boundary . We define
(6.1) |
where denotes , the distance between and . The following lemma is obvious.
Lemma 6.1.
Since is embedded and compact, there is a such that for any , there is a unique such that is equal to .
From now on fix . Thus is on . Suppose is a positive function on with
(6.2) |
where is a positive function. Then
(6.3) |
is a complete Riemannian manifold. A natural compactification is
(6.4) |
equipped with the product metric topology of .
Definition 6.2.
Let be a complete -integer multiplicity current in . Its infinity asymptotic boundary in is the set where is the closure of in the product metric topology.
Definition 6.3.
Let be the mass of currents in . We say that is a local area minimizing -integer multiplicity current in if is contained in for some finite interval , for any -integer multiplicity current satisfying outside a compact set in , then .
The main result of this section is given as follows.
Theorem 6.4.
Remark 6.5.
Recall that is the upper half space equipped with the metric
(6.5) |
Let be the upper half hemisphere in with the induced metric . We introduce the pole coordinate on (the pole at the north pole) . Thus . Note that for and . Letting we can represent Hyperbolic spaces as
(6.6) |
which is a special case of (1.1).
The above theorem is to generalize Hardt-Lin’s result in Hyperbolic spaces [22, Theorem 4.1]. We verify this claim as follows. Fixing any number we define a domain
(6.7) |
Now we are ready to show Theorem 6.4.
Proof.
First we show the existence of a minimal graph in with the infinity boundary . The local area minimizing property can be easily obtained by Theorem 4.1.
Step one: Existence. First assume . Since is compact, by Lemma 6.1 we extend into a function on such that
(6.9) |
Because for each is -mean convex, by Theorem 5.3 there is a in to solve the Dirichlet problem
(6.10) |
Let be a positive constant such that . By the maximum principle, for each ,
(6.11) |
For any fixed embedded open ball with , Theorem A.3 implies that
(6.12) |
where is a constant depending on , on and . Thus after choosing a subsequence from , denoted by , converges uniformly to in the sense on any compact set of disjoint with . Thus
(6.13) |
Next we show for any . To achieve this goal we construct the supersolution and subsolution of in (6.10) on for some .
Lemma 6.6.
Proof.
Let be a local coordinate in and be the corresponding frame. We write as and . For any smooth function , , denote the first and second covariant derivatives of and with the sum over .
Note that
(6.17) |
Here , and denote the Laplacian and the covariant derivative of respectively. Now assume are given in (6.14) where are determined later.
By (6.2) and (6.9) on we have and
(6.18) |
where and . With (6.18) we compute on as follows.
Since is -mean convex for each and , the first term above is nonnegative. By (6.14) . Now assuming on satisfies
(6.19) |
for some positive constant only depending on . Now take we obtain on . This shows (6.15) for and we have to determine and in a such way and . We have
(6.20) | |||
(6.21) |
All conditions are satisfied provided we take small enough and large satisfying
(6.22) |
This gives (6.16) for .
A similar derivation yields the conclusions for . The proof is complete.
∎
Now we continue to show Theorem 6.4. Now for any , with the maximum principle, (6.10) and (6.11) Lemma 6.6 implies that on . Let go to we obtain that
(6.23) |
Since both and are continuous on and equal to on , thus
(6.24) |
Let be the subgraph of in . Let be the corresponding integer multiplicity current. Thus with respect to the product topology of , .
Now we obtain the existence of a local integer multiplicity current in with the desirable infinity boundary for any . Moreover it is a minimal graph over in .
Step Three: General Case As for , we can construct two monotone sequences and in such that the former one converges increasingly to and the latter one converges decreasingly to in the sense. Let be the solutions of on with asymptotic value for . From the maximum principle and the translating invariant of minimal graphs, we have
is an increasing sequence on and for any . Note that both sequences are uniformly bounded. By the interior estimate of in , locally converges to a function in satisfying .
Thus for any , on we obtain
(6.25) |
Thus (6.24) still holds for . As a result for any we show there is a minimal graph in with the infinity boundary . It is also a local area minimizing integer multiplicity current according to Theorem 4.1.
Step Four: Uniqueness. Suppose is a local area minimizing -integer multiplicity current with infinity boundary . For any , we define for any and . By (1.1), is an isometry of . Thus is also a local area minimizing -integer multiplicity current in with infinity boundary . Recall that . We claim that for any . Otherwise there is a such that the regular part of and intersects transversely. Since for any . By the area minimzing property, we can replace a piece of by the piece of that is cut off by and still have a local area minimizing property. But then the singular set of the resulting integer multiplicity current would include the intersection of and with dimension. This contradicts to the dimensional singular set of local area minimizing integer multiplicity currents. Thus is a graph over . Arguing similar as in the proof of the property (3) in Theorem 4.1, is a minimal graph in with the infinity boundary . Since is compact, applying the maximum principle into yields that the uniqueness of the minimal graph.
The proof is complete.
∎
Appendix A Interior estimate of mean curvature equations
Throughout this section let be a complete Riemannian manifold with a metric and be a function in . Let be the product manifold equipped with the metric . Let be the covariant derivative of . Let be an orientable hypersurface in with the normal vector for any . We have the following theorem about its angle function . Comparing to [8, section 2.3] we use a moving frame method from [11, section 2] to obtain the interior estimate of mean curvature equations.
Lemma A.1.
Let . Suppose the mean curvature of with respect satisfies that
(A.1) |
where is the gradient of on . Then in the Lipschitz sense it holds that
(A.2) |
where is the second fundamental form of , is the covariant derivative of and is the Ricci curvature of the product manifold , is the Hessian of in .
Proof.
Notice that is a hypersurface. We denote by a local orthonormal frame of . Note that the derivation in Lemma 2.2, [41] is true for any dimension. That is by Lemma 2.2 in [41] we have
(A.3) |
where is the mean curvature of . Let is the corvariant derivative of . Since ,
(A.4) |
where . Note that . Thus
(A.5) |
Here we use the fact . Since the metric of is the product metric, for any tangent vector field . Thus . Combining this with (A.4), (A.5) we obtain
(A.6) |
Putting this into (A.3) we obtain the conclusion in the case .
In the case of , is a hypersurface for any in . The result comes from the classical approximating method. The proof is complete.
∎
As a corollary we obtain the following Harnack type result.
Corollary A.2.
Let be a bounded domain in and be a function in . Suppose is a connected orientable hypersurface satisfying (A.1) with its normal vector . If on , or on the whole .
Proof.
We have the following interior estimate of mean curvature equations.
Theorem A.3.
Let . Fix and let be an embedded ball centered at with radius . Suppose satisfying
(A.8) |
where is the gradient of , and is the divergence of . If for some positive constant , then
(A.9) |
where is a constant only depending on the Ricci curvature, the norm of , and .
Proof.
Our proof follows from the idea of Lemma 2.1 in [11] by Eichmair.
First we assume is a function in . By the classical Schauder estimate is for any . We denote by the norm of on . i.e.
(A.10) |
Let be the graph of with its upward normal vector in . By (A.8), the mean curvature of with respect to , , in satisfies that . Let which is a function. By Lemma A.1, we have
(A.11) |
Thus satisfies that
(A.12) |
where is a constant only depending on the Ricci curvature and .
Let be the distance function between and . Now we define
(A.13) |
Define . Thus . Set where is a positive constant determined later. Thus the maximum of is obtained at a point in , for example . At this point, and
(A.14) |
Observe that
(A.15) |
Since , from (A.14) and (A.15) we obtain
(A.16) |
Note that for any function , where is the orthonormal frame on . Thus
As a result,
(A.17) |
where is a constant depending on , the Ricci curvature on . Let be the covariant derivative on . Note that
(A.18) |
Here is a fixed constant depending on . Now combining (A.16) with (A.17), (A.18) together we obtain
(A.19) |
When taking sufficiently large, we obtain at , . Note that only depends on , and , the Ricci curvature of on . Thus
(A.20) |
on . Note that for any , by (A.13) . Thus . Here is a constant only depending on , and .
The proof is complete.
∎
References
- [1] A. Aiolfi, J. Ripoll, and M. Soret. The Dirichlet problem for the minimal hypersurface equation on arbitrary domains of a Riemannian manifold. Manuscripta Math., 149(1-2):71–81, 2016.
- [2] F. Almgren. Review: Enrico Giusti, Minimal surfaces and functions of bounded variation. Bull. Amer. Math. Soc. (N.S.), 16(1):167–171, 01 1987.
- [3] Y. N. Alvarez. PDE and hypersurfaces with prescribed mean curvature, Nov. 2019. arXiv:1911.12679.
- [4] Y. N. Alvarez and R. Sa Earp. Sharp solvability criteria for Dirichlet problems of mean curvature type in Riemannian manifolds: non-existence results. Calc. Var. Partial Differential Equations, 58(6):Paper No. 197, 12, 2019.
- [5] M. T. Anderson. Complete minimal varieties in hyperbolic space. Invent. Math., 69(3):477–494, 1982.
- [6] M. T. Anderson. Complete minimal hypersurfaces in hyperbolic -manifolds. Comment. Math. Helv., 58(2):264–290, 1983.
- [7] J. Bemelmans and U. Dierkes. On a singular variational integral with linear growth. I. Existence and regularity of minimizers. Arch. Rational Mech. Anal., 100(1):83–103, 1987.
- [8] J.-B. Casteras, E. Heinonen, and I. Holopainen. Dirichlet problem for -minimal graphs. J. Anal. Math., 138(2):917–950, 2019.
- [9] D. De Silva and J. Spruck. Rearrangements and radial graphs of constant mean curvature in hyperbolic space. Calc. Var. Partial Differential Equations, 34(1):73–95, 2009.
- [10] F. Duzaar and K. Steffen. minimizing currents. Manuscripta Math., 80(4):403–447, 1993.
- [11] M. Eichmair. The Plateau problem for marginally outer trapped surfaces. J. Differential Geom., 83(3):551–583, 2009.
- [12] L. C. Evans and R. F. Gariepy. Measure theory and fine properties of functions. Textbooks in Mathematics. CRC Press, Boca Raton, FL, revised edition, 2015.
- [13] H. Federer and W. H. Fleming. Normal and integral currents. Ann. of Math. (2), 72:458–520, 1960.
- [14] Q. Gao and H. Zhou. The area minimizing problem in conformal cones, 2020. arXiv:2001.06207.
- [15] C. Gerhardt. Closed hypersurfaces of prescribed mean curvature in locally conformally flat Riemannian manifolds. J. Differential Geom., 48(3):587–613, 1998.
- [16] D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
- [17] E. Giusti. Generalized solutions for the mean curvature equation. Pacific J. Math., 88(2):297–321, 1980.
- [18] E. Giusti. Minimal surfaces and functions of bounded variation, volume 80 of Monographs in Mathematics. Birkhäuser Verlag, Basel, 1984.
- [19] B. Guan and J. Spruck. Hypersurfaces of constant mean curvature in hyperbolic space with prescribed asymptotic boundary at infinity. Amer. J. Math., 122(5):1039–1060, 2000.
- [20] E. M. Guio and R. Sa Earp. Existence and non-existence for a mean curvature equation in hyperbolic space. Commun. Pure Appl. Anal., 4(3):549–568, 2005.
- [21] E. M. Guio and R. Sa Earp. Errata: “Existence and non-existence for a mean curvature equation in hyperbolic space” [Commun. Pure Appl. Anal. 4 (2005), no. 3, 549–568; mr2167187]. Commun. Pure Appl. Anal., 7(2):465, 2008.
- [22] R. Hardt and F.-H. Lin. Regularity at infinity for area-minimizing hypersurfaces in hyperbolic space. Invent. Math., 88(1):217–224, 1987.
- [23] J. M. Lee. Introduction to smooth manifolds, volume 218 of Graduate Texts in Mathematics. Springer, New York, second edition, 2013.
- [24] F. Lin and X. Yang. Geometric measure theory—an introduction, volume 1 of Advanced Mathematics (Beijing/Boston). Science Press Beijing, Beijing; International Press, Boston, MA, 2002.
- [25] F.-H. Lin. Regularity for a class of parametric obstacle problems integrand integral current prescribed mean curvature minimal surface system. ProQuest LLC, Ann Arbor, MI, 1985. Thesis (Ph.D.)–University of Minnesota.
- [26] F.-H. Lin. On the Dirichlet problem for minimal graphs in hyperbolic space. Invent. Math., 96(3):593–612, 1989.
- [27] L. Lin and L. Xiao. Modified mean curvature flow of star-shaped hypersurfaces in hyperbolic space. Comm. Anal. Geom., 20(5):1061–1096, 2012.
- [28] M. McGonagle and L. Xiao. Minimal Graphs and Graphical Mean Curvature Flow in . J. Geom. Anal., 30(2):2189–2224, 2020.
- [29] M. Miranda. Superfici cartesiane generalizzate ed insiemi di perimetro localmente finito sui prodotti cartesiani. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3), 18:515–542, 1964.
- [30] B. Nelli and J. Spruck. On the existence and uniqueness of constant mean curvature hypersurfaces in hyperbolic space. In Geometric analysis and the calculus of variations, pages 253–266. Int. Press, Cambridge, MA, 1996.
- [31] J. Ripoll and M. Telichevesky. On the asymptotic plateau problem for CMC hypersurfaces in hyperbolic space. Bull. Braz. Math. Soc. (N.S.), 50(2):575–585, 2019.
- [32] J. Serrin. The problem of Dirichlet for quasilinear elliptic differential equations with many independent variables. Philos. Trans. Roy. Soc. London Ser. A, 264:413–496, 1969.
- [33] L. Simon. On a theorem of de Giorgi and Stampacchia. Math. Z., 155(2):199–204, 1977.
- [34] L. Simon. Lectures on geometric measure theory, volume 3 of Proceedings of the Centre for Mathematical Analysis, Australian National University. Australian National University, Centre for Mathematical Analysis, Canberra, 1983.
- [35] L. Simon. The minimal surface equation. In Geometry, V, volume 90 of Encyclopaedia Math. Sci., pages 239–272. Springer, Berlin, 1997.
- [36] J. Spruck. Interior gradient estimates and existence theorems for constant mean curvature graphs in . Pure Appl. Math. Q., 3(3, Special Issue: In honor of Leon Simon. Part 2):785–800, 2007.
- [37] I. Tamanini. Boundaries of Caccioppoli sets with Hölder-continuous normal vector. J. Reine Angew. Math., 334:27–39, 1982.
- [38] N. S. Trudinger. On Harnack type inequalities and their application to quasilinear elliptic equations. Comm. Pure Appl. Math., 20:721–747, 1967.
- [39] H. Zhou. Inverse mean curvature flows in warped product manifolds. J. Geom. Anal., 28(2):1749–1772, 2018.
- [40] H. Zhou. The boundary behavior of domains with complete translating, minimal and CMC graphs in . Sci. China Math., 62(3):585–596, 2019.
- [41] H. Zhou. Generalized solutions to the Dirichlet problem of translating mean curvature equations., 2019. arXiv:1902.01512.