Multiplicity one for min-max theory in compact manifolds with boundary and its applications
Abstract.
We prove the multiplicity one theorem for min-max free boundary minimal hypersurfaces in compact manifolds with boundary of dimension between 3 and 7 for generic metrics. To approach this, we develop existence and regularity theory for free boundary hypersurface with prescribed mean curvature, which includes the regularity theory for minimizers, compactness theory, and a generic min-max theory with Morse index bounds. As applications, we construct new free boundary minimal hypersurfaces in the unit balls in Euclidean spaces and self-shrinkers of the mean curvature flows with arbitrarily large entropy.
1. Introduction
The main motivations of this article are two-folded. On one hand, we further advance the existence theory for minimal hypersurfaces with free boundary in compact manifolds with boundary, inspired by the recent exciting developments for closed minimal hypersurfaces. We have seen immense existence results concerning free boundary minimal hypersurfaces in recent years, c.f. [Fra00], [FrSch16], [MaNS17], [LZ16], [KL17], [Lin-Sun-Zhou20], [GLWZ19], [Wang20], [Pigati20], [CFS20]. In this article, we prove the free boundary version of the Multiplicity One Conjecture in min-max theory. This directly gives the existence of new free boundary minimal hypersurfaces even in the unit balls of the Euclidean spaces. On the other hand, we propose to use the results developed here and before by the authors to construct new minimal hypersurfaces in singular spaces. The idea is to use compact manifolds with boundary to approximate a given singular space. In fact, consider the compliments of tubular neighborhoods of the singular set, which are compact manifolds with nontrivial boundary; one can apply the min-max theory to obtain free boundary minimal hypersurfaces and take the limits. We apply this strategy to the singular Gaussian space, and construct connected minimal hypersurfaces therein with arbitrarily large Gaussian area. These minimal hypersurfaces are self-shrinkers of the mean curvature flow with arbitrarily large entropy, c.f. [CM12_1].
1.1. Multiplicity one theorem for the free boundary min-max theory
Denote by a compact Riemannian manifold with smooth boundary of dimension . In [Alm62], Almgren proved that the space of mod 2 relative cycles is weakly homotopic to ; see also [LMN16]*§2.5. By performing a min-max procedure, Gromov [Gro88] defined the volume spectrum of , which is a sequence of non-decreasing positive numbers
depending only on and . Moreover, Gromov [Gro88] also showed that for each , grows like ; see also Guth [Guth09]. Recently, Liokumovich-Marques-Neves [LMN16] proved that the volume spectrum obey a uniform Weyl Law.
When is closed with no boundary, by adapting the regularity theory given by Pitts [Pi] and Schoen-Simon [SS], and the index bounds given by Marques-Neves [MN16], it is known that, when , each -width is realized by a disjoint collection of closed, embedded, minimal hypersurfaces with integer multiplicities, i.e. there exist integers and minimal hypersurfaces so that
(1.1) |
see [IMN17]*Proposition 2.2. In [MN16, MN18], Marques-Neves raised the Multiplicity One Conjecture, which asserts that for bumpy metrics111Here a metric is said to be bumpy if each closed immersed minimal hypersurfaces has no Jacobi field, and the set of bumpy metrics in a closed manifold is generic in the Baire sense by White \citelist[Whi91][Whi17]., and is two-sided for all and . This conjecture was later confirmed by the last-named author [Zhou19]; (see also Chodosh-Mantoulidis [CM20]). This, together with [MN18], implies the existence of a closed, embedded minimal hypersurface of Morse index for each under a bumpy metric. This body of works established the Morse theory for the area functional. We also refer to [MN17, IMN17, MNS17, Song18, YY.Li19, Marques-Montezuma-Neves20, Song-Zhou20] for recent developments on the existence theory of closed minimal hypersurfaces, particularly on the resolution of Yau’s conjecture.
In this paper, we focus on compact manifolds with non-empty boundary. In [LZ16], Li-Zhou proved the general existence of free boundary minimal hypersurfaces by adapting the Almgren-Pitts theory. Inspired by \citelist[MN16][IMN17], the last two authors together with Q. Guang and M. Li generalized (1.1) to compact manifolds in [GLWZ19]. Precisely, for each positive integer , there exist integers and disjoint, embedded, free boundary minimal hypersurfaces so that
(1.2) |
Inspired by [Zhou19], it is natural to conjecture that for generic metrics on compact manifolds with boundary. In this article we confirm this conjecture.
Theorem 1.1.
Let be a compact manifold with boundary of dimension . For generic metrics on and every integer , there exist a sequence of disjoint, two-sided, connected, embedded free boundary minimal hypersurfaces satisfying
Remark 1.2.
Our proof follows generally the approach in [Zhou19], but there are several essential new challenges which stimulate novel ideas (we refer to Section 1.3 for elucidations). Recall that a hypersurface with mean curvature prescribed by an ambient function will be called an -hypersurface.
-
(a)
Generic properness for free boundary minimal hypersurfaces;
-
(b)
Existence and regularity for free boundary -hypersurfaces, including: the regularity for minimizers, compactness, and a generic min-max theory;
-
(c)
Genericity of good mean curvature prescribing functions for which the min-max theory for free boundary -hpersurfaces can be applied;
-
(d)
Generic countability of free boundary -hypersurface – an essential ingredient in the proof of Morse index estimates in the free boundary -hypersurface min-max theory.
To solve the above challenge (a), we prove the following strong bumpy metric theorem for compact manifolds with boundary.
Theorem 1.3.
For a generic metric on , every embedded free boundary minimal hypersurface is both non-degenerate and proper.
Here a hypersurface is proper if it has empty touching set, i.e. .
Remark 1.4.
In [ACS17]*Theorem 9, Ambrozio-Carlotto-Sharp proved a similar bumpy metric theorem, which says that for generic metrics, each properly embedded free boundary minimal hypersurface is non-degenerate. Our result is a strengthened version in the sense that each embedded free boundary minimal hypersurface is automatically proper in our generic metrics.
As a direct corollary of Theorem 1.1, we construct free boundary minimal hypersurfaces with arbitrarily large area in the unit ball , and more generally in any strictly convex domain of . Recall that a compact Riemannian manifold with non-negative Ricci curvature and strictly convex boundary does not contain any stable free boundary minimal hypersurfaces. Moreover, by Fraser-Li [FrLi]*Lemma 2.4, any two free boundary minimal hypersurfaces in this type of manifolds intersect with each other. Making use of the compactness \citelist[ACS17][GWZ18], we have the following result.
Corollary 1.5.
Let be a compact Riemannian manifold with non-negative Ricci curvature and strictly convex boundary of dimension . Then there exist a sequence of two-sided, connected, embedded, free boundary minimal hypersurfaces satisfying
In particular, the unit ball with contains a sequence of free boundary minimal hypersurfaces with areas going to . These are different from the hypersurfaces constructed by Fraser-Schoen [FrSch16], Kapouleas-Li [KL17] and Carlotto-Franz-Schulz [CFS20].
1.2. Application to Gaussian spaces
Besides the independent interest, the multiplicity one result for free boundary minimal hypersurfaces may also have many other applications. Here we present an application to construct new minimal hypersurfaces in certain non-compact and singular manifolds.
An embedded hypersurface in is called a self-shrinker if
where is the mean curvature with respect to the normal vector field . Equivalently, a self-shrinker is a minimal hypersurface in under the Gaussian metric ; see [CM12_1] for more details. The area of such a hypersurface under Gaussian metric is called the entropy, defined by Colding-Minicozzi in [CM12_1].
Self-shrinkers are models of the singularities of mean curvature flows, and the index of a self-shrinker (as a minimal hypersurface in ) characterizes the instability of the mean curvature flow at the singularity; see the discussions in [CM12_1]. Understanding self-shrinkers is very important to the study of singular behaviors of mean curvature flows.
The main challenge of using min-max method to construct self-shrinkers is that the Gaussian metric space is non-compact, and the space becomes singular at infinity – the curvature blows up at infinity, meanwhile the spheres with increasing radii have exponentially decreasing areas. Therefore, the classical min-max method can not be used directly. We note that Ketover-Zhou [Ketover-Zhou18] have established a version of the min-max theory in using a special flow near the singularity.
Our approximation method provides the following new constructions of embedded self-shrinkers. We believe that our method can also be applied to other non-compact singular spaces.
Theorem 1.6.
Given , there exists a sequence of embedded self-shrinkers with entropy growth and .
A direct corollary is the existence of self-shrinkers with arbitrary large entropy.
Corollary 1.7.
For any () and , there exists an embedded self-shrinker with and .
Proof of Corollary 1.7.
There are several pioneer works which provided constructions of embedded self-shrinkers in by other techniques, see \citelist[An92][Ng14][KKM18] [Mo11]. However, the entropy of all the self-shrinkers constructed in these works is bounded from above by a universal constant. In contrast, our min-max method constructs self-shrinkers with arbitrarily large entropy.
The entropy of a mean curvature flow is monotone non-increasing. As a consequence, any tangent flow of a mean curvature flow has entropy bounded from above by the entropy of the initial hypersurface. Based on this fact, there has been much research on low entropy mean curvature flow; cf. \citelist[Bernstein-Wang] [mramor2018low] [bernstein2020closed]. In contrast, our theorem implies that large entropy mean curvature flow may have very complicated tangent flows, which addresses the complication of mean curvature flow singularities.
1.3. Challenges and ideas
Inspired by the proof of the Multiplicity One Theorem in closed manifolds [Zhou19], we prove Theorem 1.1 by establishing the min-max theory for free boundary -hypersurfaces. Let be a compact Riemannian manifold with boundary of dimension . Moreover, we assume that every embedded free boundary minimal hypersurface in is proper and does not have non-trivial Jacobi fields. Then the -the volume spectrum is realized by the min-max value of a homotopy class of -sweepouts. Note that the sweepouts are maps into . We can lift the -sweepouts to the space of Cappccioppoli set , and consider the associated relative homotopy class, denoted by . We will show that the relative min-max width is equal to .
Take to be defined later. Let and denote by the free boundary -hypersurface, produced by relative min-max theory (associated with ). Letting , and using compactness results, we get a free boundary minimal hypersurface with multiplicity . The aim is to prove by taking suitable . Indeed, if , we can construct a non-trivial positive Jacobi field so that on and , where is the Jacobi operator, and are respectively the unit normal and co-normal vector of , and is the second fundamental form of . By choosing a suitable , we prove that this kind of does not exist.
However, there are several new challenges in this process. First, we need to show the genericity of the metrics under which every embedded free boundary minimal hypersurface is proper and has no non-trivial Jacobi field. Such a metric is called strongly bumpy. The denseness of such metrics can be obtained by perturbing any given metric on a closed manifold (which contains as a domain) so that each minimal hypersurface in with free boundary on is non-degenerate and transverse to . To obtain openness, we use the compactness to show that given a strongly bumpy metric and a constant , there is an open neighborhood of so that for each metric in that neighborhood, any free boundary minimal hypersurface in with weak Morse index and area no more than is non-degenerate and proper. By taking intersections of such open and dense sets of metrics for a sequence , we obtain the desired genericity.
Second, to establish the min-max theory for free boundary -hypersurfaces, we need to generalize a number of regularity results. In Theorem 2.1, we prove the regularity for minimizing free boundary -hypersurfaces by using the reflecting method in [Gr87] and the regularity results in [Mor03]. The major step for this min-max theory is to establish the regularity for -almost minimizing varifolds. To finish the key gluing step therein, we apply the interior gluing procedure by by Zhou-Zhu [ZZ18], and then take the advantage of unique continuation to extend the gluing all the way to boundary. Comparing with the proof of the corresponding regularity results for free boundary minimal hypersurfaces in [LZ16], our arguments for free boundary -hypersurfaces are much simpler since we only need to consider the case where the touching sets are contained in some -dimensional subsets.
To obtain the aforementioned control on the touching sets, we need to restrict to “good” prescribing functions , where each free boundary -hypersurface has at most -dimensional touching set (including self-touching and touching with ). To show the genericity of such good functions, we extend a given smooth function on to on and perturb it to by [ZZ18]*Proposition 3.8 so that each -hypersurface has small self-touching set. By genericity of Morse functions we can find another perturbed function so that each -hypersurface touches in a small subset as desired. This implies is a good choice. Moreover, each function in a neighborhood of is also a good choice. This gives the genericity of good prescribing functions.
Finally, we have to prove the index upper bounds in the min-max procedure, and a crucial ingredient is the countability of free boundary -hypersurfaces for good pairs (see Subsection 2.4). Our proof of the countability follows from the following observation of the compactness of free boundary -hypersurfaces. Let be a sequence of free boundary -hypersurfaces with bounded index and area. Then it converges as varifolds to some limit hypersurface ; (see Theorem 2.9). If the convergence is smooth, then has a non-trivial Jacobi field related to ; on the contray, if the convergence is not smooth, has strictly less index than that of for sufficiently large . We refer to Lemma 2.12 for more details.
To construct self-shrinkers of large entropy, we use approximation methods. More precisely, we focus on balls with larger and larger radius in , perturb the Gaussian metric on the balls slightly to be generic, and use the theory developed in this article to construct free boundary minimal hypersurfaces of large area. By passing to a limit, these free boundary minimal hypersurfaces will converge to minimal hypersurface in the Gaussian space, or equivalently self-shrinkers. When passing to limits, there may be mass drop, and the limit shrinkers would not have large entropy. To overcome this issue, we prove a new monotonicity formula of minimal surfaces in almost Gaussian spaces, which is more close related to the monotonicity formula of minimal hypersurfaces in the Euclidean spaces compared with that for shrinkers. We also study the spectrum of the Jacobi operator for shrinkers of the form whenever is a lower dimensional shrinker, and we prove in Theorem 5.3 that if has finite index, so does . This allows us to construct smooth self-shrinkers with finite index and arbitrarily large entropy in high dimensional spaces.
Outline
This paper is organized as follows. In Section 2, we first collect some notations and then provide some technical results including the regularity for -minimizing hypersurfaces and the compactness of free boundary -hypersurfaces. We mention that Theorem 1.3 is proven in Subsection 2.5. Then in Section 3, we prove the relative min-max theory for free boundary -hypersurfaces. The regularity of -almost minimizing varifold with free boundary is proved in Theorem 3.10. Section 4 is devoted to the proof of Theorem 1.1. Finally, by applying our main result, we prove the existence of self-shrinkers with arbitrarily large entropy in Section 5. We also provide some well-known results and technical computations in Appendix A–D.
Acknowledgments
We would like to thank Professor Bill Minicozzi for his interest. Part of this work was carried out when Z.W. was a postdoc at Max-Planck Institute for Mathematics in Bonn and he thanks the institute for its hospitality and financial support. X. Z. is partially supported by NSF grant DMS-1811293, DMS-1945178 and an Alfred P. Sloan Research Fellowship.
2. Preliminaries
Notations
We collect some notions here.
Let be a compact, oriented, smooth Riemannian manifold with smooth boundary, of dimension . In this paper, we always isometrically embed into a closed manifold with dimension . Assume that is embedded in some . {basedescript}\desclabelstyle\multilinelabel\desclabelwidth10em
the geodesic ball of ;
the space of smooth vector fields on ;
the collection of with for in a neighborhood of in ;
-dimensional Hausdorff measure;
the space of smooth Riemannian metrics on ;
the space of integer rectifiable -currents with support in ;
the space of mod 2 rectifiable -currents with support in ;
the group or ;
the space of and ;
the space of and ;
equivalent class of , i.e. if and only if for any ;
the space of equivalent classes in ;
the integer rectifiable varifold associated with ;
the Radon measure associated with ;
the mass norm on ;
.
the flat metric on ;
for ;
the space of sets with finite perimeter (Cappccioppoli set);
the space of smooth vector field in so that for ;
the (reduced) boundary of restricted in the interior of , (thus, as an integer rectifiable current, );
the space of -dimensional integral varifolds;
the tangent space of at ;
denotes the -dimensional Grassmannian bundle over ;
for ;
for ;
the integral current associated with ;
the integral varifold associated with ;
the collection of so that , where .
For any , there is a canonical representative where and .
Given , a varifold is said to have -bounded first variation in an open subset , if
where is defined to be the subset of compactly supported in .
Let be an -dimensional manifold with smooth boundary. Recall that an immersion is a map . In this paper, we often use to denote both the immersion and the hypersurface when there is no ambiguity.
We are interested in the following weighted area functional defined on . Given , define the -functional on by
(2.1) |
The first variation formula for along is
where is the outward unit normal vector field of .
When the boundary has support on a smooth immersed hypersurface, by virtue of the divergence theorem, we have
(2.2) |
where is the unit outer co-normal vector field of . If is a critical point of , then (2.2) directly implies that must have mean curvature and along , where is the outward unit normal vector field of .
Recall that an immersed hypersurface is called an -hypersurface if its mean curvature everywhere.
2.1. Regularity for boundaries minimizing the -functional
In [Mor03], F. Morgan gives a general way to prove the regularity of isoperimetric hypersurfaces in Riemannian manifolds. His methods can be applied to -functional to prove the regularity of -hypersurfaces bounding a domain which minimizes the -functional; (see [ZZ18]*Theorem 2.2). In this part, we provide the regularity for -minimizers with free boundaries.
Let be an -ball in , and a compact embedded -ball such that . Denote by and the two components of . Note that the -functional in (2.1) is well-defined for .
Theorem 2.1.
Suppose that minimizes the -functional with free boundary on : that is, for any other with , we have . Then except for a set of Hausdorff dimension at most , is smooth hypersurface embedded properly (under relative topology) in and meets orthogonally.
We recall a few notations. Without loss of generality, we assume that . For , let be the closest point to among . Clearly, is well-defined in a small neighborhood of . By possible rescaling, we assume that is contained in such a neighborhood. Then we define the reflection map across by
We have and define
Thus , where . By possible rescaling, we may assume that . Set
We also denote by . By shrinking if it is necesssary, we can assume that there exists so that
(2.3) |
Proof of Theorem 2.1.
Take . Then for and so that , we have
(2.4) | ||||
Here (2.3) is used in the first and the last inequalities. The constant is allowed to change from line to line. On the other hand,
(2.5) | ||||
Here the third inequality is from (2.3).
Recall that is -minimizing. Assume that . Hence we have
and
where is the symmetric difference of two Cappccioppoli sets. The above two inequalities, together with (2.4) and (2.5) and the isoperimetric inequalities, imply that
(2.6) |
Here is a constant depends only on and .
Applying (2.6) to [Mor03]*Corollary 3.7, 3.8, is . In particular, is a properly embedded free boundary hypersurface. Since is an -hypersurface, the classical PDE argument implies the higher regularity. ∎
Note that such process works for any Riemannian manifold. Let be compact Riemannian manifold with boundary which is isometrically embedded into a closed manifold . Recall that is denoted by the geodesic ball of with radius and center at . We then have the following regularity theorem:
Theorem 2.2.
Given , , and some small , suppose that minimizes the -functional: that is, for any other with , we have . Then except for a set of Hausdorff dimension at most , is a properly embedded hypersurface with free boundary on , and is real analytic if the ambient metric on is real analytic.
2.2. Estimates of touching sets
An immersed hypersurface is almost embedded if it locally decomposes into smooth embedded sheets that touch other sheets or but do not cross. A hypersurface is properly embedded if it is almost embedded and no sheets touch other sheets or .
Let be the collection of all Morse functions such that the zero set is a compact, smoothly embedded hypersurface so that
-
•
is transverse to and the mean curvature of vanishes to at most finite order;
-
•
is contained in an -dimensional submanifold of .
Lemma 2.3.
contains an open and dense subset in .
Proof.
For any and a neighborhood of in , we are going to find an open subset of so that . First, can be extended to a function . We can also extend to a smooth function . Second, we take a neighborhood of in so that if .
By [ZZ18]*Proposition 3.8, there exists a Morse function so that is a closed, embedded hypersurface with mean curvature vanishing at most finite order. By perturbing slightly, according to the generic existence of Morse functions, we can find so that
-
•
, are all Morse functions on ;
-
•
is a closed, embedded hypersurface which is transverse to and has mean curvature vanishing at most finite order;
-
•
and are smooth, embedded hypersurfaces and transverse to .
Denote by . Then the last item implies that is contained in an -dimensional submanifold of . Thus . Moreover, by the choice of , we have that and are Morse functions on with as regular value. Thus we conclude that for any in a neighborhood of in , and are still Morse functions with as regular value. is the desired open set and Lemma 2.3 is proved. ∎
In the next, we prove that for each , the touching set of an almost embedded free boundary -hypersurface has dimension less than or equal to .
Proposition 2.4.
Let and are two different connected, embedded, free boundary -hypersurfaces in a connected open set . Then and are contained in a countable union of connected, smoothly embedded, -dimensional submanifolds.
Proof.
Recall that the argument in [ZZ18]*Theorem 3.11 implies that is contained in a countable union of connected, smoothly embedded, -dimensional submanifolds. Note that . Thus it suffices to prove that is contained in the submanifolds as described.
To do this, we first take so that . Then there exists a neighborhood of so that can be written as a graph over . Denote by the graph function. By shrinking the neighborhood, we can also assume that
(2.7) |
for all . Note that such a function satisfies an inhomogeneous linear elliptic PDE of the form
Together with (2.7), then the Hessian of at has rank at least 1. The implicit function theorem then implies that, on a possibly smaller neighborhood of , the touching set is contained in an -dimensional submanifold; see [ZZ17]*Lemma 2.8 for more details. Therefore, we conclude that
(2.8) |
is contained in the submanifolds as described in the proposition. Recall that
(2.9) |
is contained in an -dimensional submanifold of . Then Proposition 2.4 follows from the fact that is contained in the union of (2.8), (2.9) and . ∎
2.3. Compactness of free boundary -hypersurfaces
Recall that an almost embedded hypersurface is called a free boundary -hypersurface if and meets orthogonally along . Given , denote by the collection of free boundary -hypersurfaces such that for some open set .
Note that when , the min-max free boundary -hypersurfaces produced in Theorem 3.10 satisfy the above requirements. Indeed, such is a critical point of the weighted functional:
The second variation formula for along normal vector field is given by
(2.10) | ||||
In the above formula, is the gradient of on ; is the Ricci curvature of ; and are the second fundamental forms of and with normal vector fields and , respectively.
We remark that in (2.10), can also be defined for any immersed free boundary -hypersurfaces and it is a quadratic form on the space of -functions on (not ).
The Jacobi field of is defined to be a smooth function on (not ) so that
(2.11) |
The classical Morse index for is defined to be the number of negative eigenvalues of the above quadratic form. However, since may touch itself and the boundary of , a weaker version of the index is needed. Such a concept was introduced by Zhou [Zhou19, Definition 2.1, 2.3] for closed -hypersurfaces based on Marques-Neves [MN16]*Definition 4.1.
Definition 2.5.
Given with , and , we say that is -unstable in an -neighborhood if there exist and a smooth family with for all (the standard -dimensional unit ball in ) such that, for any , the smooth function:
satisfies
-
•
has a unique maximum at ;
-
•
for all .
Since is a critical point of , necessarily .
When , this reduces to the -unstable notion for free boundary minimal hypersurfaces defined in [GLWZ19, Definition 5.5].
Definition 2.6.
Assume that or is a free boundary minimal hypersurface. Given , we say that the weak Morse index of is bounded (from above) by , denoted as
if is not -unstable in 0-neighborhood for any . is said to be weakly stable if .
Remark 2.7.
We make several remarks:
-
•
If is -unstable in a 0-neighborhood, then it is -unstable in an -neighborhood for some ;
-
•
All the concepts can be localized to an open subset by using in place of ;
-
•
If is properly embedded, then is -unstable if and only if its classical Morse index is .
We also have the following curvature estimates.
Theorem 2.8 (Curvature estimates for weakly stable free boundary -hypersurfaces).
Let and be two relatively open subset so that . Let be weakly stable in with , then there exists , such that
The proof is the same as the free boundary minimal cases [GWZ18]*Theorem 3.2.
Given and , let
(2.12) |
The main purpose is to prove this theorem:
Theorem 2.9 (Compactness for free boundary -hypersurfaces).
Let be a compact Riemannian manifold with boundary of dimension . Assume that is a sequence of smooth functions in such that in smooth topology, where or . Let be a sequence of hypersurfaces such that for some fixed and . Then,
-
(i)
up to a subsequence, there exists a smooth, compact, almost embedded free boundary -hypersurface such that (possibly with integer multiplicity) in the varifold sense, and hence also in the Hausdorff distance by monotonicity formula;
-
(ii)
there exists a finite set of points with , such that the convergence of is locally smooth and graphical on ;
-
(iii)
if , then the multiplicity of is 1, and ;
-
(iv)
assuming eventually and for all and smoothly converges to , then , and has a non-trivial Jacobi field;
-
(v)
if and the convergence is not smooth, then is not empty and has strictly smaller weak Morse index than for all sufficiently large ;
-
(vi)
if and is properly embedded, then the classical Morse index of satisfies (without counting multiplicity)
Proof.
The proof follows essentially the same way as [Sharp17]*Theorem 2.3 and [GWZ18]*Theorem 4.1; we will only provide necessary modifications.
Part 1: The same argument in [Zhou19]*Theorem 2.6, by replacing [Zhou19]*Theorem 2.5 with Theorem 2.8, implies that converges locally smoothly and graphically to an almost embedded free boundary -hypersurface (possibly with integer multiplicity) away from at most points, which we denote by . Now we prove that are all removable.
Case 1: Suppose is in the closure of and .
Then the argument is the same as that in [Zhou19]*Theorem 2.7, Part 1.
Case 2: Suppose that .
Using the boundary removable singularity result Theorem A.1, we see that is also a removable singularity.
Up to here, we have finished proving (i) and (ii). The argument in [Zhou19]*Theorem 2.6(iii)(v) also works for (iii)(vi) here.
Part 2: We now prove (iv). It suffices to produce a Jacobi field for the second variation along . Recall that the Jacobi fields associated with along a free boundary -hypersurface satisfy
(2.13) |
where and is the outward co-normal of .
Recall that can be isometrically embedded into a closed Riemannian manifold . Let be the space of vector fields so that for in a small neighborhood of in .
Denote by the immersion. Now we fix a relative open set of so that is an embedding.
Now let be an extension of the unit normal vector field of . Let be the one-parameter family of diffeomorphisms of associated with , so that . Let be the -thickening of with respect to so that
Then for sufficiently large , there exists so that
where is the immersion map. In the following, we often omit and for simplicity when there is no ambiguity.
Denote by and
Then we have , .
Now consider any vector field such that and . Let be the associated one-parameter family of diffeomorphisms of . From the fact that and are -hypersurfaces, we have
The computation in Appendix B gives that
Here we used the assumption of in the last equality. By pulling everything back to and letting , we obtain
where
Here uniformly as and we used that . Now letting and , by Fubini theorem, we have
(2.14) | ||||
Here we used the fact that on .
Now we take finitely many relatively open subsets of so that is an embedding for each and
Then we can also find finitely many non-negative cut-off functions so that on and . Then by (2.14), for each and each ,
Adding all of them together, we have
Let . Then the standard PDE theory implies that converges smoothly to a nontrivial satisfying equation (2.13), so we finish proving (iv).
Part 3: In this part, we prove (v). The process is similar to [GLWZ19]*Proposition 5.2.
Assume has weak index . If does not smoothly converge to , then there exists a finite set so that for any , does not smoothly converge to and smoothly converges to outside . We now prove that for large .
By the Definition 2.6 of weak index, there exists and smooth family with for all such that for any the smooth function:
satisfies
-
•
has a unique maximum at ;
-
•
for all .
Denote by . Then .
We can shrink so that is still linearly independent. Let be a cut-off function satisfying and and and as . By Appendix C, we can shrink so that there exists satisfying
(2.15) |
for . Recall that smoothly converges to . Hence for sufficiently large ,
Here we used the fact that on .
By assumptions, does not smoothly converge to for . Hence is not weak stable in . This implies that for each , there exists such that
and the smooth function satisfying, for some ,
(2.16) |
for all and .
For any , denote by the flow of . Now we define by
Here . Then for any we have
Together with (2.15) and (2.16), we conclude that . This finishes the proof of (v).
∎
There is also a theorem analogous to the above one in the setting of changing ambient metrics on . The proof proceeds the same way when one realizes that the constant in Theorem 2.5 depends only on the when is allowed to change.
Theorem 2.10.
Let be a closed manifold of dimension , and be a sequence of metrics on that converges smoothly to some limit metric . Let be a sequence of smooth functions with that converges smoothly to some limit , where or . Let be a sequence of hypersurfaces with for some fixed and . Then there exists a smooth, compact, almost embedded free boundary -hypersurface , such that Theorem 2.9(i)(ii)(iii) are satisfied.
2.4. Generic existence of good pairs
Let be a closed Riemannian manifold of dimension . Let be a compact domain of with smooth boundary. A pair consisting of a Riemannian metric and smooth function is called a good pair related to , if
-
(1)
is a Morse function;
-
(2)
the zero set is a smoothly embedded closed hypersurface in , which is transverse to and has mean curvature vanishing to at most finite order;
-
(3)
has dimension less than or equal to ;
-
(4)
is bumpy for , i.e., every almost embedded prescribed mean curvature hypersurface in with free boundary on is non-degenerate.
Clearly, if is a good pair related to , then ; (see Section 2.2). In this subsection, we are going to prove the generic existence of good pairs related to .
Denote by the set of smooth functions such that
-
•
is a Morse function;
-
•
is an embedded closed hypersurface in , which is transverse to .
is open and dense in , and is independent of the choice of a metric.
The following lemma is a generalization of [Zhou19]*Lemma 3.5 by the last author.
Lemma 2.11.
Given , the set of Riemannian metrics on with being a good pair related to is generic in the Baire sense.
Proof.
Firstly, we prove that (3) is generic. Let be the set of metrics so that has dimension less than or equal to . Denote by the set of metrics on so that
Clearly, and is open. We claim that is also dense, which would imply that (3) is generic.
Indeed, given any and an open neighborhood of in the space of metrics on , we can construct as follows. Let be the signed distance function to under the metric . Let be small enough so that has a tubular neighborhood and the exponential map from to is a diffeomorphism. Fix a cut-off function satisfying
Now let be small enough so that
(2.17) |
This can be satisfied because the set of smooth Morse functions on with empty singular set is open and dense.
Define , where is the projection of to its closest point in . Let . By taking in a small enough neighborhood of , we have that . Then the second fundamental form of with respect to is given by (see \citelist[Bes87]*Section 1.163[IMN17]*Proposition 2.3)
Recall that . Hence the mean curvature with respect to is
By (2.17), . Thus is dense.
In the next, we prove that the set of metrics under which has mean curvature vanishing to at most finite order is an open and sense subset. Clearly, it is open and it suffices to prove denseness. Let be an open neighborhood of . Then from the above argument, for any small enough, we can find a metric so that and has mean curvature , where is the mean curvature of under . We can choose so that is a Morse function on , hence vanishes to at most finite order. This gives the denseness and we conclude that (2) is generic.
It remains to prove that (4) is generic. The proof is the same with the Bumpy metric theorem [Whi91]*Theorem 2.2. See also [ACS17]*Theorem 9 and an alternative version [GWZ18]*Theorem 2.8. ∎
In the end of this subsection, we prove that the space of almost embedded free boundary -hypersurfaces is countable for a good pair related to .
Given and , recall that the set of satisfying and its weak Morse index is bounded by from above.
Lemma 2.12.
Let be a good pair related to . Denote by and . Then is countable, and hence is countable.
Proof.
We prove it by an inductive method. By Theorem 2.9(v)(iii), there are only finitely many elements in . Hence contains only finitely many hypersurfaces.
Assuming that is countable for some , we are going to prove that is also countable. Using Theorem 2.9(v)(iii) again, we know that for any , contains only finitely many elements. Therefore, is countable.
By induction, we have proved that is countable for all . ∎
2.5. Generic properness for free boundary minimal hypersurfaces
In this part, we review White’s Generically Transversality Theorem and prove an adapted version for free boundary minimal hypersurfaces, i.e. Theorem 1.3.
We focus on the co-dimension one and embedded case. We first consider a closed manifold and a two-sided, closed, embedded hypersurface . (Later we will embed into and let .) Let be the set of smooth Riemannian metrics on . Let be the space of all pairs such that and is an embedded minimal hypersurface in with (possibly empty) free boundary . Recall that the projection is defined as
Denote by the set of such that is non-degenerate (with no nontrivial Jacobi field). Then by the work of White \citelist[Whi91][Whi17] and Ambrozio-Carlotto-Sharp [ACS17], together with the fact that is second countable, we know that is a countable union of open sets such that maps each homeomorphically onto an open subset of .
Denote by the set of such that is transverse to .
Lemma 2.13.
is meager in . As a corollary, for generic metrics on , each embedded minimal hypersurface with free boundary on is transverse to .
Proof.
The proof here is similar to [Whi19]*Corollary 5. Recall that is a countable union of open sets so that is a homeomorphism. Now
Now for any , we can always take a smooth vector field so that satisfying, (here is the unit normal vector field of ),
-
•
in a neighborhood of in ;
-
•
For any , and ; (e.g. letting in a neighborhood of ).
Denote by the 1-parameter family of diffeomorphisms of associated with .
Claim 1.
There exists so that for all , is transverse to .
Proof.
By the choice of , we can take an open set containing such that for any . By taking small, denote by the set of
which is diffeomorphic to by the map . Define a continuous function by
Here is the unit normal vector field of , and is the signed distance function to . It follows that . Moreover, if , then is smooth in a small neighborhood of and . Let be a normal coordinate system of at , and write . By a direct computation, we have
where . Moreover,
and
Therefore, we conclude that in a small neighborhood of ,
(2.18) |
Now we assume on the contrary that there exists a sequence of positive so that is not transverse to . Then there exists so that and is tangent to . By the definition of , we have . Assume that , then . However, by (2.18), for sufficiently large ,
This leads to a contradiction and hence Claim 1 is proved. ∎
Claim 1 gives that for small enough, is transverse to , hence . Note that is open. This is to say that is nowhere dense. Because is a homeomorphism, is nowhere dense. Hence is meager in . ∎
Now we are ready to prove Theorem 1.3. Let be a smooth compact manifold with boundary and . Denote by the space of smooth Riemannian metrics on . Let be the space of all pairs such that and is an almost properly embedded, free boundary minimal hypersurface in . Recall that the projection is defined as
Denote by the set of such that is non-degenerate and proper.
Proof of Theorem 1.3.
We first prove that is dense in . Indeed, for any , can be isometrically embedded into a closed manifold . By Lemma 2.13, there exists a sequence of such that every embedded minimal hypersurface in with free boundary on is non-degenerate and transverse to . When restricting to , and for each .
Now for any , denote by the set of so that every free boundary minimal hypersurface in with and is non-degenerate and proper. Then the above argument gives that contains a dense set .
Claim 2.
For any , is open.
Proof of Claim 2.
Let . Suppose for the contrary that is a sequence of metrics on and a sequence of free boundary minimal hypersurfaces in with , and , but is either degenerate or improper. Then by the compactness theorems in \citelist[ACS17][GWZ18], up to a subsequence, locally smoothly converges to a free boundary minimal hypersurface away from a finite set in with and . Since , is non-degenerate and proper. This implies that the convergence is smooth, and hence is also non-degenerate and proper for large . This is a contradiction and hence Claim 2 is proved. ∎
We thus have that contains an open dense set of . Recall that
Obviously, this is a generic subset of . ∎
3. Relative Min-max theory for free boundary -hypersurfaces
Here we generalize multi-parameter min-max theory for prescribed mean curvature hypersurface to compact manifolds with boundary.
3.1. Notations for Min-max construction
In this part, we describe the setup for min-max theory for free boundary -hypersurfaces associated with multiple-parameter families in . All the setups here are the same as those in [Zhou19]*§1.1. First, we list some notations for cubical complex. {basedescript}\desclabelstyle\multilinelabel\desclabelwidth6em
the cubical complex on with 0-cells and 1-cells ;
the cubical complex ( times);
the map from to defined as the unique element such that
, where ;
a cubical subcomplex of dimension in some ;
the union of all cells of whose support is contained in some cell of ;
the set of all -cells in .
Let be a cubical complex of dimension in some and be a cubical subcomplex.
Let be a continuous map (under the -topology on ). Let be the collection of all sequences of continuous maps such that
-
(1)
each is homotopic to in the flat topology on ;
-
(2)
there exist homotopy maps which are continuous in the flat topology, , , and satisfy
Given a pair and as above, is called a -homotopy sequence of mappings into , and is called the -homotopy class of . Then we define the -width by
A sequence is called a minimizing sequence if , where
Given and , by the same argument as [Zhou19]*Lemma 1.5, there exists a minimizing sequence.
Definition 3.1.
If is a minimizing sequence in , the critical set of is defined by
3.2. Discretization and Interpolation
We record several discretization and interpolation results developed by Marques-Neves \citelist[MN14][MN17] (for closed manifolds), and later by Li and the last author [LZ16] (for compact manifolds with boundary). Though these results were proven for sweepouts in or , they work well for sweepouts in like in [Zhou19]*§1.3. We will point out necessary modifications.
We refer to [ZZ18]*Section 4 for the notion of discrete sweepouts. Though all definitions therein were made when , there is no change for discrete sweepouts on .
Recall that given a map , the fineness of is defined as
Here two vertices are adjacent if they belong to a common cell in .
Definition 3.2 ([MN17]*§3.7).
Given a continuous (in the flat topology) map , we say that has no concentration of mass if
The purpose of the next theorem is to construct discrete maps out of a continuous map in the flat topology.
Theorem 3.3 (Discretization, [Zhou19]*Theorem 1.11).
Let be a continuous map in the flat topology that has no concentration of mass, and . Assume that is continuous under the -topology. Then there exist a sequence of maps
and a sequence of homotopy maps:
with , , and a sequence of numbers such that
-
(i)
the fineness ;
-
(ii)
-
(iii)
for some sequence , with ,
and this directly implies that
-
(iv)
-
(v)
and hence
Proof.
The last named author \citelist[Zhou17]*Theorem 5.1[Zhou19]*Theorem 1.11 proved this for closed manifolds. The argument works well here by using the isoperimetric lemmas given in [LZ16]*§3.2.
∎
Before stating the next result, we first recall the notion of homotopic equivalence between discrete sweepouts. Let be a cubical subcomplex of . Given two discrete maps , we say that is homotopic to with fineness less than , if there exist , and a map
with fineness and such that
The purpose of the next theorem is to construct a continuous map in the -topology from a discrete map with small fineness, which is called an Almgren extension. Moreover, the Almgren extensions from two homotopic maps are also homotopic to each other.
Theorem 3.4 (Interpolation, [Zhou19]*Theorem 1.12 and Proposition 1.14).
There exist some positive constants and so that if is a cubical subcomplex of and
has , then there exists a map
continuous in the -topology and satisfying
-
(i)
for all ;
-
(ii)
if is some -cell in , then restricted to depends only on the values of restricted on the vertices of ;
-
(iii)
Moreover, if () is homotopic to each other with fineness ,
of , respectively, are homotopic to each other in the -topology.
Now for a continuous map in Theorem 3.3, there exists a sequence of discretized maps . Applying Theorem 3.4 to each , we obtain continuous in the -topology. Then the next proposition says that is homotopic to .
Proposition 3.5 ([Zhou19]*Proposition 1.15).
Let and be given by Theorem 3.3 applied to some therein. Assume that is continuous in the -topology on . Then the Almgren extension is homotopic to in the -topology for sufficiently large .
In particular, for large enough, there exist homotopy maps continuous in the -topology, , , and
Therefore, for given , we have
3.3. Regularity of -almost minimizing varifolds with free boundary
One key ingredient in the Almgren-Pitts theory to prove regularity of min-max varifold is to introduce the “-almost minimizing with free boundary” concept.
Definition 3.6 (-almost minimizing varifolds with free boundary).
Let be the or -norm, or the -metric. For any given and a relative open subset , we define to be the set of all such that if is a sequence with:
-
(i)
;
-
(ii)
;
-
(iii)
, for , then
We say that a varifold is -almost minimizing in with free boundary if there exist sequences , , and such that
For each , as defined in [LZ16]*Definition A.4, the Fermi half-ball and half-sphere of radius centered at are
where is the Fermi distance function to ; (see [LZ16]*Definition A.1). For , and are also used to denote the geodesic ball and sphere of radius at .
Definition 3.7.
A varifold is said to be -almost minimizing in small annuli with free boundary if for each , there exists such that is -almost minimizing with free boundary in for all , where .
Before stating the regularity of -almost minimizing varifolds with free boundary, we provide the regularity of -replacements, which follows from Theorem 2.2.
Proposition 3.8 (Replacements [ZZ18]*Proposition 6.8).
Let be -almost minimizing with free boundary in a relative open set and be a compact subset. Then there exists , called an -replacement of in such that, with ,
-
(i)
;
-
(ii)
;
-
(iii)
is -almost minimizing in with free boundary ;
-
(iv)
as varifolds for some such that
furthermore locally minimizes in (relative to );
-
(v)
if has -bounded first variation in , then so does .
Proof.
The proof here is same as [ZZ18]*Proposition 6.8. We only sketch the steps and point out the difference here.
By Definition 3.7, there exist sequences , and such that . Then for each , denote by the current by solving a constrained minimization problem; see [ZZ18]*Lemma 6.7. Then by Theorem 2.2, is a properly embedded -hypersurface with free boundary. One can check that (as varifolds) is the desired replacement; (see [ZZ17]*Proposition 5.8 for details). ∎
Making use of Theorem 3.8 and the monotonicity formula for varifolds with bounded first variation, we can classify the tangent varifolds.
Lemma 3.9 ([LZ16]*Proposition 5.10).
Let . Suppose that has -bounded first variation in and is -almost minimizing in small annuli with free boundary. For any tangent varifold with , we have either
-
(i)
where or
-
(ii)
for some -plane such that and and .
Moreover, for -a.e. , the tangent varifold of at is unique, and the set of in which case (ii) occurs as its unique tangent cones has -measure 0; hence is rectifiable.
Proof.
The first step is to prove that is a stationary rectifiable cone in with free boundary. Such a result follows from the monotonicity formula, which also holds true for varifolds with bounded first variation (see [GLZ16]) together with the argument in [LZ16]*Lemma 5.8.
Now we are ready to prove the main regularity theorem for varifolds which is -almost minimizing with free boundary and has -bounded first variation.
Theorem 3.10 (Main regularity).
Let , and be an -dimensional smooth, compact Riemannian manifold with boundary. Further let and set . Suppose is a varifold which
-
(1)
has -bounded first variation in , and
-
(2)
is -almost minimizing in small annuli with free boundary,
then is induced by , where is a compact, almost embedded -hypersurface with free boundary (possibly disconnected).
Proof.
We only need to prove the regularity of near an arbitrary point . Fix a , then there exists such that for any , the mean curvature of in is greater than . Here is as in Definition 3.7. In particular, if and has -bounded first variation in and in , then
(3.1) |
Here stands for the closure of some set.
We will show that is an almost embedded free boundary -hypersurface with density equal to along its self-touching set.
The argument consists of six steps:
Step 1: Constructing successive -replacements and on two overlapping concentric annuli.
Step 2: Gluing the -replacements smoothly (as immersed hypersurfaces) on the overlap.
Step 3: Extending the -replacements to the point to get an ‘-replacement’ on the punctured ball.
Step 4: Showing that the singularity of at is removable, so that is regular.
Step 5: Showing that is not contained in for all .
Step 6: Proving that coincides with the almost embedded hypersurface on a small neighborhood of .
We now proceed to the proof.
Step 1. Fix any . Since is -almost minimizing on small annuli with free boundary, we can apply [ZZ18]*Lemma 6.7 by replacing [ZZ18]*Theorem 2.2 with Theorem 2.2 to obtain a first replacement of on . By Theorem 2.2 and Proposition 3.8 (iv), if
then is an almost embedded stable free boundary -hypersurface with some unit normal ; when the multiplicity is , is locally a boundary so we can choose to be the outer normal.
Note that all the touching set is contained in a countable union of -dimensional connected submanifolds . Since a countable union of sets of measure zero still has measure zero, by Sard’s theorem we can choose such that intersects and all the transversally (even at ). Then given any , following the argument in [ZZ17]*Theorem 6.1, Step 1, we can construct , which is a replacement of on . Denote by
Step 2. We now show that and glue smoothly (as immersed hypersurfaces) across . Indeed, define the intersection set
(3.2) |
Then by transversality, is an almost embedded hypersurface in . Particularly, is not contained in since intersect and transversally. For , following from the interior argument [ZZ18]*Theorem 7.1, Step 3, coincides with (with matching normal) in a neighborhood of . Using the unique continuation of -hypersurfaces, we conclude that
Then we finish the proof of Step 2.
Step 3. Then we extend the replacements, via the unique continuation from Step 2, all the way to . In fact, we use to denote the second replacement that we constructed in Step 1 with inner radius . Step 2 shows that this construction does not depend on . Then we define to be the limit of as .
See [ZZ17]*Theorem 6.1, Step 3 for details.
Step 4. We now determine the regularity of at .
Firstly, observing that is still -almost minimizing in small annuli with free boundary and that is the varifold limit of a sequence , which all have -bounded first variation, we know that also has -bounded first variation. This implies that the classification of tangent cones in Lemma 3.9 also holds true. Secondly, , when restricted to any small annulus (), already coincides with a smooth, almost embedded, weakly stable -boundary with free boundary. Using these two ingredients, by Theorem A.1, extends smoothly across as an almost embedded hypersurface in . Thus we complete Step 4.
Step 5. We argue by contradiction and assume that for some . Here we also use the chosen constants , , , in the previous steps. We first recall that by the Constancy Theorem [Si]*Theorem 41.1,
Now we take the first replacement of on . In the next paragraph, we are going to prove that for any , . As a result, for any , , which leads to a contradiction to Proposition 2.4.
To conclude this step, we consider the second replacement of on . By the assumption and , we have . On the other hand, . Together with the classification of tangent cones in Lemma 3.9, we conclude that . This concludes Step 5.
Step 6. It remains to show that coincides with in . Recall that by [ZZ18]*Theorem 7.1, is an almost embedded -hypersurface. Denote by the self-touching set. Then by Proposition 2.4, is contained in a countable union of smoothly embedded -dimensional submanifolds. Hence we can take so that is not contained in . Recall that is the second replacement of in . Take
Such a set is non-empty by Step 5. Then coincides with in a small neighborhood of by the construction of the second replacement. Then the unique continuation principle gives that in . This completes the proof of Theorem 3.10. ∎
3.4. Relative Min-max theory for free boundary -hypersurfaces
The existence of almost minimizing varifolds follows from a combinatorial argument of Pitts [Pi]*page 165-page 174 inspired by early work of Almgren [Alm65]. Pitts’s argument works well in the construction of min-max -hypersurfaces; see [ZZ18]*Theorem 6.4. Marques-Neves has generalized Pitts’s combinatorial argument to a more general form in [MN17]*§2.12, and we can adapt their result to the free boundary -hypersurface setting with no change.
Recall that a minimizing sequence such that every element of (see Definition 3.1) has -bounded variation or belongs to is called a pull-tight.
The purpose of this part is to establish min-max theory for free boundary -hypersurfaces. Recall that the Morse index of an almost embedded free boundary -hypersurface is given in Definition 2.6.
Theorem 3.11.
Let be a compact Riemannian manifold of dimension , and which satisfies . Given a -dimensional cubical complex and a subcomplex , let be a map continuous in the -topology, and be the associated -homotopy class of . Suppose
(3.3) |
Then there exists a nontrivial, smooth, compact, almost embedded hypersurface with free boundary , such that
-
•
for some , where the mean curvature of with respect to the unit outer normal of is , i.e.
-
•
;
-
•
.
Proof of Theorem 3.11.
The proof can be divided into five steps. In the first four steps, we always assume that is isometrically embedded into a closed manifold and is a good pair (e.g. Section 2.4) related to so that .
Step A: We construct a pulled-tight minimizing sequence so that every element of either has -bounded first variation, or belongs to .
Step B: There exists so that is -almost minimizing in small annuli with free boundary.
Step C: has -bounded first variation, and hence is supported on an almost embedded free boundary -hypersurface satisfying and .
Step D: The and in Step B can be chosen so that the support of has weak Morse index less than or equal to .
Step E: We provide the proof for general .
Proof of Step A.
Let and . Set
where . Then we can follow [LZ16]*Proposition 4.17 to construct a continuous map:
such that:
-
(i)
for all ;
-
(ii)
if ;
-
(iii)
if ,
where is a continuous function with , when ;
- (iv)
Given a minimizing sequence , we define for every . Then is also a minimizing sequence in . Moreover, and every element of either has -bounded first variation, or belongs to . We refer to [Zhou19]*Lemma 1.8 for the details of verification. This finishes proving Step A. ∎
Proof of Step B.
The proof here is parallel to [Zhou19]*Theorem 1.7 and we just sketch the idea for completeness.
Let be a pulled-tight minimizing sequence. For each , Theorem 3.3 gives a sequence of maps:
with and a sequence of positive (as ), satisfying Theorem 3.3. Then for each , take sufficiently large and let . Denote by . Then we have and , where
We now prove that there exists so that it is -almost minimizing in small annuli with free boundary. For a further reason, we need a stronger result:
Claim 3.
There exist a varifold satisfying the following: for any and any small enough annulus centered at with radii , there exist two sequences of positive real numbers , a subsequence and (the domain of ) so that
-
•
;
-
•
; and
-
•
.
If Claim 3 were not true, then using the argument in [Zhou19]*Theorem 1.16, we can find a sequence so that is homotopic to with fineness converging to zero as and . The key point here is that is compact in the sense of varifolds.
Then by Theorem 3.4, the Almgren extensions of :
respectively, are homotopic to each other in the -topology for large and
This leads to a contradiction and we have finished the proof of Claim 3. Obviously, such a varifold is almost minimizing with free boundary in small annuli by Definition 3.7. Therefore, Step B is also completed. ∎
Proof of Step C.
We first prove that has -bounded first variation. Indeed, from Step A, either has -bounded first variation or belongs to . Recall that being -almost minimizing in small annuli with free boundary always implies has -bounded first variation away from finitely many points. Then by a cut-off trick, we only need to prove that has at most -volume growth near these bad points . This essentially follows from [HL75]*Theorem 4.1 and we provide more details here.
Let be a bad point. Then we can take small enough so that has no bad point. For any , let be two cut-off functions so that
We only need to consider . Denote by the Fermi distance function to in [LZ16]*Appendix A. Then there exist and so that for with ,
(3.4) |
Being -bounded first variation in gives that for any ,
(3.5) | ||||
By direct computation of the left hand side,
(3.6) | ||||
Here is the projection to the hyperplane and(3.4) is used in the last inequality. Note that either has -bounded first variation or belongs to . Hence . Combining (3.5) with (3.6), letting , and by shrinking if necessary, we have
where . This implies the monotonicity of . Hence for , we have
Then we conclude that
This proves that has -bounded first variation. Then by Theorem 3.10, is supported on an almost embedded -hypersurface with free boundary.
We now prove that is a free boundary -hypersurface satisfying and for some . Recall that Claim 3 gives that . Denote by . Then it suffices to prove that subsequently converges to in the metric. Let be a limit of in the topology. Then . Now for any and small enough, the Constancy Theorem [Si, Theorem 26.27] implies that or . Here is the touching set inside .
In the next, we prove that for . Recall that first -replacement in coincides with in the Step 5 of Theorem 3.10. On the other hand, the argument in [ZZ18]*Proposition 7.3 gives that . Hence .
Recall that by Proposition 2.4, the self-touching set and are contained in countable -dimensional submanifolds. Therefore, we conclude that . Hence Step C is finished. ∎
Proof of Step D.
Recall that by Lemme 2.12, the set of almost embedded free boundary -hypersurface is countable. Then the proof here is parallel to [Zhou19]*Theorem 3.6, which is in fact a generalization of [MN16].
Denote by the set of all so that . Set
It suffices to show that for every , is not empty.
By the same proof with [Zhou19]*Lemma 3.7, there exist and such that for all .
As is a good pair related to , is countable by Lemma 2.12, and we can assume that
and for . Then by taking small enough, we can make sure . Using the Deformation Theorem (\citelist[Zhou19]*Theorem 3.4[GLWZ19]*Theorem 5.8), by shrinking if necessary, we can find , and so that
-
•
is homotopic to in the -topology for all ;
-
•
;
-
•
for all ;
-
•
no belongs to ; (this can be easily satisfied since is a countable set.)
Inductively, there are two possibilities. The first case is that we can find for all , there exist a sequence , , , and for some so that
-
•
is homotopic to in the -topology for all ;
-
•
;
-
•
for all ;
-
•
;
-
•
no belongs to .
The second case is that the process stops in finitely many steps. This means that we can find some , a sequence , , , and so that
-
•
is homotopic to in the -topology for all ;
-
•
;
-
•
for all ;
-
•
.
For the first case, we can choose a diagonal sequence and set , where is an increasing sequence such that and
For the second case, we simply choose the last sequence and set and . Now it is easy to see that for both cases, the new sequence satisfies the following conditions:
-
(1)
is homotopic to in the -topology for all ;
-
(2)
;
-
(3)
given any subsequence , , if , then does not converge in the -topology to any element in .
Then by Steps B and C, there exists so that its support has weak Morse index less than or equal to . This finishes the proof of Step D. ∎
Proof of Step E.
Assume that is isometrically embedded into . Recall that is the set of smooth Morse functions so that the zero set is a properly embedded closed hypersurface in , and is transverse to .
Given , we can take an extension of so that and . By Lemma 2.11, there exists a sequence of smooth metrics on so that smoothly and is a good pair related for each . Then by Steps A-D, for each , there exists an almost embedded free boundary -hypersurface with for some and has weak Morse index less than or equal to (with respect to ). Recall that . Let (with ) be the limit of given in Theorem 2.10, then the multiplicity one (see Theorem 2.9 (iii)) and locally smoothly convergence away from a finite set imply that and . ∎
By putting all above together, Theorem 3.11 is finished. ∎
4. Multiplicity one for free boundary minimal hypersurfaces
4.1. Multiplicity one for relative sweepouts
In this subsection, we approximate the area functional by the weighted functionals for some prescribing function when . By the index estimates and the multiplicity one result for , we prove that the limit free boundary minimal hypersurfaces also have multiplicity one.
Recall that a Riemannian metric is said to be bumpy if every finite cover of a smooth immersed free boundary minimal hypersurface is non-degenerate. A bumpy metric is said to be strongly bumpy if the touching set of every immersed free boundary minimal hypersurface is empty. By Theorem 1.3, the set of strongly bumpy metrics is generic in the Baire sense.
Theorem 4.1.
Let be a compact Riemannian manifold with boundary of dimension . Let X be a -dimensional cubical subcomplex of and be a subcomplex, and be a map continuous in the -topology. Let be the associated -homotopy class of . Assume that
where we let in Section 3.1.
If is a strongly bumpy metric, then there exists a disjoint collection of smooth, connected, compact, two-sided, properly embedded, free boundary minimal hypersurfaces so that
In particular, each component of is two-sided and has exactly multiplicity one.
Proof.
Fix a sequence of positive numbers as . Recall that is open and dense in . Thus
is generic in the Baire sense. Pick an in this generic set with (to be fixed at the end, in Part 5), and small enough so that
Note that for each .
Then we can follow the argument in [Zhou19]*Theorem 4.1, by replacing [Zhou19]*3.1 with Theorem 3.11, to produce a non-trivial, smooth, compact, almost embedded, free boundary -hypersurface ; moreover, and . We also have as .
Now applying Theorem 2.9, by the strong bumpiness of , there exists a subsequence (still denoted by ) such that converges to a smooth, compact, embedded, free boundary minimal hypersurface (with integer multiplicity). We denote by the set of points where the convergence fails to be smooth. By Theorem 2.9,
We now prove that every component of is two-sided and has multiplicity one. Without loss of generality, we may assume that has only one connected component with multiplicity . We will prove by contradiction.
Part 1: We assume that is 2-sided; otherwise, we would consider the double cover of just like [Zhou19]*Proof of Theorem 4.1, Part 8. Let be the global unit normal of and be an extension of . Suppose that is a one-parameter family of diffeomorphisms generated by . For any domain and small , produces a neighborhood of with thickness , i.e., . We will use as coordinates on . If is in the interior of , then for small is the same as in the geodesic normal coordinates of . Now fix a domain , by the convergence , we know that for sufficiently large, can be decomposed to graphs over which can be ordered by height
Since is the boundary of some set , we know that the unit outer normal of will alternate orientations along these graphs; see [Zhou19]*Proof of Theorem 4.1, Part 3.
Part 2: We first deal with an easier case: is an odd number. Hence . Let be any smooth function on and be an extension of . Construct a family of hypersurfaces
where . Then and are the bottom and top sheets of . Since is a free boundary -hypersurface, we have
Then the computation in Appendix B gives that
where , is the co-normal of , and
Here is the mean curvature vector. Denote by . By pulling everything back to , we obtain
where
(4.1) |
Here uniformly as . Now letting and , by Fubini theorem, we have
Take . Let . Then the standard PDE theory implies that converges smoothly to a positive satisfying
(4.2) |
Note that above argument works for any . Taking an exhaustion of , we can extend to and such that
(4.3) |
Part 3: Next we use White’s local foliation argument to prove that extends smoothly across , and this will contradict the bumpy assumption of .
By the work for interior singularity in [Zhou19]*Proof of Theorem 4.1, Part 5, it suffices to show the uniform boundedness for . Since has empty touching set, then . Using the -hypersurface with free boundary (see Proposition D.1), we can also prove that is bounded. Then the classical PDE gives that is smooth across . Thus, we conclude that there is a positive Jacobi field on , which is a contradiction to the fact that is a strongly bumpy metric.
Part 4: We now take care the case when is even. Hence . Then without loss of generality we may assume that
Then by the argument in Part 2, we have
Here ; and are defined as the integral of and in (4.1). Fix a point .
Case 1: . Then the renormalizations converges locally smoothly to a nontrivial function on , and by same reasoning as Part 2, we have
Case 2: . Consider renormalizations . Then again by the same reasoning, converges locally smoothly to a nonnegative on , and such that
(4.4) |
Then by the argument in [Zhou19]*Proof of Theorem 4.1, Part 7 together with using the foliations by Proposition D.1, we can also prove that is smooth across in both cases.
Part 5: Following the step in [Zhou19]*Proof of Theorem 4.1, Part 9, one can also show that for a nicely chosen , the (unique) solutions to (4.4) must change sign. Thus there is no 1-sided component, and the multiplicity for 2-sided component must be one.
We only sketch the proof for two-sided case here. Note that every almost properly embedded free boundary minimal hypersurface has empty touching set since is strongly bumpy. Recall that by \citelist[GWZ18][Wang19], there are only finitely many free boundary minimal hypersurfaces with and , denoted by . Then as in [Zhou19], we can take disjoint neighborhood so that . Since all the small neighborhoods are disjoint, we can take a smooth function defined on whose supports are compact and
-
•
is nonnegative and is positive at some point;
-
•
is nonpositive and is negative at some point.
Then take be an extension of such that equals to in a neighborhood of . Recall that
is generic. Take in this generic set which is close to as we wanted. Then the solution of (4.4) would be close to on each , therefore it must change sign. Thus, such an is a desired function. ∎
Remark 4.2.
We remark here the theorem is stated only for stongly bumpy metrics. However, we also believe the multiplicity one holds true for metrics which are only bumpy. Such a result may need a highly nontrivial argument for constructing Jacobi fields; see [Wang19].
4.2. Application to volume spectrum
In this part, we will apply the result in Section 4.1 to study volume spectrum introduced by Gromov [Gro88], Guth [Guth09], and Marques-Neves [MN17]. In particular, we will prove that for a strongly bumpy metric, the volume spectrum can be realized by the area of min-max minimal hypersurfaces with free boundary produced by Theorem 4.1.
We first recall the definition of volume spectrum following [MN17]*Section 4. Let be a compact Riemannian manifold with boundary. Let be a cubical subcomplex of for some . Given , a continuous map is called a -sweepout if
where is the generator. A sweepout is said to be admissible if it has no concentration of mass (see Definition 3.2). Denote by the set of all admissible -sweepouts. Then
Definition 4.3.
The -width of is
where is the domain of .
The -width satisfies a Weyl’s asymptotic law. This asymptotic behaviour was first conjectured by Gromov in [Gro88] and studied by Guth in [Guth09]. Finally, Liokumovich-Marques-Neves proved the following Weyl law for -width.
Theorem 4.4 ([LMN16]*§‘ 1.1).
There exists a constant such that, for every compact Riemannian manifold with (possibly empty) boundary, we have
(4.5) |
Assume from now on that . It was later observed by Marques-Neves in [MN16] (see also [GLWZ19]*Section 4) that one can restrict to a subclass of in the definition of . In particular, let denote those elements which is continuous under the -topology, and whose domain has dimension (and is identical to its -skeleton). Then
Following the idea of Marques-Neves [MN16], the last two authors together with Q.Guang and M. Li also proved in [GLWZ19] that for each , there exists a disjoint collection of smooth, connected, almost properly embedded, free boundary minimal hypersurfaces with integer multiplicities , such that
Before stating the main theorem, we recall an observation by [MN17]*Corollary 3.4. Denote by the unit circle.
Lemma 4.5 (\citelist[MN17]*Corollary 3.4[LMN16]*Proposition 2.12).
Let so that its support is a properly embedded free boundary minimal hypersurface in . There exists sufficiently small, depending on and so that every map with
is homotopically trivial.
We will also use the following Lemma proved in [GLWZ19]. Such a result follows from the Morse index upper bound estimates for the free boundary min-max theory in [GLWZ19].
Lemma 4.6 ([GLWZ19], Theorem 2.1).
Suppose is bumpy, then for each , there exist a -dimensional cubical complex and a map continuous in the -topology with , such that
where is the class of all maps continuous in the -topology that are homotopic to in flat topology.
Moreover, there exists a pulled-tight (see [GLWZ19]*Theorem 5.8) minimizing sequence of such that if has support a compact, smooth, almost properly embedded, free boundary minimal hypersurface, then
Now we are going to state and prove our main theorem.
Theorem 4.7 (same with Theorem 1.1).
If is a strongly bumpy metric and , then for each , there exists a disjoint collection of compact, smooth, connected, properly embedded, two-sided, free boundary minimal hypersurfaces , such that
That is to say, the min-max minimal hypersurfaces with free boundary are all two-sided and have multiplicity one for generic metrics.
Proof of Theorem 4.7.
Since is bumpy, then there are only finitely many compact, almost properly embedded, free boundary minimal hypersurfaces with and for given by \citelist[GWZ18][Wang19]; see [ACS17] for strongly bumpy metrics.
Now we fix and omit the sub-index in the following. Take with .
We proceed the proof by the following three steps.
Step I: In this step, we show how to find another minimizing sequence, still denoted by , such that for sufficiently large, either is close to a regular min-max free boundary minimal hypersurface, or the mass is strictly less than .
We recall the following observation by [MN17]*Claim 6.2. Let be the set of all stationary integral varifolds with whose support is a compact, smooth, almost properly embedded, free boundary minimal hypersurface with . Consider the set of all with and either or the support of is a compact, smooth, properly embedded, free boundary minimal hypersurface with . By the bumpy assumption, both sets and are finite. Moreover,
Claim 4 ([MN17]*Claim 6.2).
For every , there exists such that with .
Proof of Claim 4.
Here since is strongly bumpy, then each element in is supported on a properly embedded hypersurface. This implies that Constancy Theorem [Si, §26.27] can be applied. Then the proof is the same with [MN17]*Claim 6.2. ∎
Let be chosen as in Lemma 4.6. We choose as Lemma 4.5 so that every map with is homotopically trivial. According to such , we then choose by Claim 4. Take a sequence so that
Consider to be the cubical subcomplex of consisting of all cells such that
Hence for all . Consider this sub-coordinating sequence . and are defined in the same way as in Section 3.1 with replaced by .
Let . It then follows that
We also denote . In fact, is the topological boundary of and . For a later purpose, we also consider the set
can be thought of the “thickening” of inside .
Let . Let and , and and be the inclusion maps. Then by Lemma 4.5, we have
Now consider and the set
Let and and and be the inclusion maps. Then by Lemma 4.5, we also have
Claim 5.
can be deformed so that
Proof of Claim 5.
By the work of the last author and M. Li [LZ16] (see also [GLWZ19]*Theorem 4.5 for the adaptions to -coefficients), together with Lemma 4.6, we also have the following dichotomy:
-
•
no element is -almost minimizing in small annuli with free boundary (see [LZ16]*Definition 4.19),
-
•
or
(4.6)
For the latter case, we are done.
We now assume the first case happens. The argument has been well-known and we only sketch the ideas here. For each , we can use the Discretization Theorem [LZ16]*Theorem 4.12 (or [GLWZ19]*Theorem 4.5) to get a sequence of discrete maps . Taking sufficiently large and by assumptions, we can deform and then use the Almgren extension [LZ16]*Theorem 4.14 to get a sequence of maps which is still a minimizing sequence of and
This finishes the proof of Claim 5. We refer to [Zhou19]*Proof of Theorem 5.2, Step 2 for more details. ∎
We remark that since , we have
Step II: Now we want to produce sweepouts in by lifting to the double cover so as to produce sweepouts satisfying the assumption of Theorem 4.1.
Note that is homotopically trivial. Then by the lifting criterion [Hat]*Proposition 1.33, there exist lifting maps so that .
Lemma 4.8.
For large enough, if is the -homotopy class associated with , then we have
Proof of Lemma 4.8.
Fix large, so that
and we will omit the sub-index in the following proof.
If the conclusion were not true, then we can find a sequence of maps , such that
(4.7) |
and homotopy maps , which are continuous in the flat topology, , , and
(4.8) |
We now construct new sequences of maps and defined as:
-
•
if ;
-
•
if , where ;
-
•
.
Then is continuous and homotopic to in the flat topology. Moreover, (4.7) and (4.8) give that
Recall that for . Then by Claim 5,
Thus we conclude that
(4.9) |
Note that (4.8) also gives that has no concentration of mass. This implies that is also a -sweepout, so (4.9) contradicts the definition of . ∎
Step III: Now we are ready to prove Theorem 4.7.
For large enough in Lemma 4.8, Theorem 4.1 applied to gives a disjoint collection of compact, smooth, connected, properly embedded, 2-sided, free boundary minimal hypersurfaces , such that
Recall that as .Together with Lemma 4.8, as . Counting the fact that there are only finitely many compact, smooth, properly embedded, free boundary hypersurfaces with and , for sufficiently large we have
Hence we finish the proof of Theorem 4.7. ∎
5. Existence of self-shrinkers with arbitrarily large entropy
In this section, we consider the min-max theory in with the Gaussian metric by virtue of Theorem 4.7. We will prove that each -width is realized by a connected, embedded self-shrinker with multiplicity one.
Recall that an embedded hypersurface is called a self-shrinker if and only if
where is the mean curvature with respect to the unit normal vector field . It is equivalent to say that is minimal under the Gaussian metric . We refer to [CM12_1] for more results concerning about self-shrinkers.
Before we go to details, let us first sketch the idea of the proof. We will construct min-max free boundary minimal hypersurfaces in larger and larger balls in with generic metrics near the Gaussian metric. Then by passing to limits we will obtain complete minimial hypersurfaces in . The limits are non-trivial because the areas of the sequence of the free boundary minimal hypersurfaces outside a large ball are uniformly small (see Claim 6). The Weyl law ensures that the limits have arbitrarily large area. Finally an index estimate (Theorem 5.3) allows us to say something about the indices of higher dimensional self-shrinkers constructing by multiplying a low dimensional self-shrinker with linear spaces.
5.1. Min-max theory in Gaussian metric spaces
Denote by . We first give the definition of volume spectrum for .
Definition 5.1.
Given a sequence of compact domain exhausting , then for any positive integer , we define
We remark that such a definition does not depend on the choice of the sequences.
The following definition of index is well-known:
Definition 5.2.
Let be an embedded self-shrinker in . We say that has if there exists a -dimensional subspace of such that each nonzero has compact support and
has index if and only if has but does not have .
Now we are going to prove the main theorem in this section.
Proof of Theorem 1.6.
In the following, always denotes the ball in with radius under the Euclidean metric. Let be a sequence of positive numbers with . Denote by . For each , we now take a perturbed metric on so that
-
•
for any compact domain , smoothly;
-
•
under the metric , has mean concave boundary for each ;
-
•
is strongly bumpy;
-
•
.
The last item can be satisfied because depends continuously on the metric by Irie-Marques-Neves [IMN17]*Lemma 2.1. Hence without loss of generality,
Since is strongly bumpy, by Theorem 4.7, there exists a free boundary minimal hypersurface
so that . By the compactness of minimal surfaces with bounded area and index (see [Sharp17]), subsequently, locally smoothly converges to a smooth minimal hypersurface in with multiplicity away from a finite set . Then using the fact of non-existence of stable minimal hypersurfaces in , can only be . Thus locally smoothly converges to . We refer to [CM12] for more details about this kind of convergence. Then by the Frankel property for self-shrinkers [CCMS]*Corollary C.4, is connected. To finish the proof, we prove that no mass is lost in the convergence.
Claim 6.
There exist constants depending only on so that for all and sufficiently large ,
Proof of Claim 6.
For , let . Clearly,
(5.1) |
Hence we can take large enough so that for all ,
(5.2) |
For , we define . Then is the distance to under . Denote by and the Levi-Civita connection associated with and , respectively. Then and
(5.3) |
By direct computations, , and for ,
which implies that for and ,
Here we used (5.2) in the last inequality. Together with (5.3), we have that for and any ,
Then can assume that for any and (by taking close to ),
(5.4) |
Given , we define two sequences of cut-off functions so that
Since is a minimal hypersurface in , then by the divergence theorem, for any ,
Here (5.4) is used in the last inequality. Letting , we have
Therefore, we conclude that for and sufficiently large ,
where . Such an inequality implies that is monotone increasing. Thus, for ,
(5.5) |
Now letting in , then
Together with (5.1), the inequality in (5.5) becomes
for all . The proof of Claim 6 is finished by taking . ∎
5.2. Index estimates
In order to prove Corollary 1.7, we provide some equivalent conditions for a self-shrinker to have finite index.
Let be an embedded self-shrinker in . Throughout, will be the Jacobi operator from the second variation formula. Such an operator is associated with a bi-linear form:
Then bottom of the spectrum of is defined as
where the infimum is taken over smooth functions with compact support. Since may be noncompact, we allow the possibility that .
In the following we will focus on the index of a self-shrinker of the form , where is a lower dimensional self-shrinker. We will always use to denote a point on and use to denote a point in . We will also use and to denote the second variational operator on and respectively. Note that
(5.6) |
The main theorem in this section is the following index estimate, which would help us to obtain the finiteness of index in Corollary 1.7.
Theorem 5.3.
Supposing is a self-shrinker with finite index, then is a self-shrinker with finite index.
We need the following correspondence between the eigenfunctions and eigenvalues on and respectively. In the following the eigenvalues and eigenfunctions are under the Dirichlet boundary conditions.
Let be a bound domain with smooth boundary. Denote by the closure of
in the topology of . Then the self-adjoint operator has discrete Dirichlet eigenvalues
and associated eigenfunctions . Moreover, forms a complete basis of the weighted ; see [Str08]*§11.3 Theorem 2.
Proposition 5.4.
Suppose is a bounded compact subset with smooth boundary. Suppose are eigenvalues of on , with associated eigenfunctions ; suppose are eigenvalues of on , with associated eigenfunctions . Then the eigenvalues of on are , with associated eigenfunctions .
Proof.
This theorem is just the classical result on the spectrum of the elliptic operators on product manifolds. Here we sketch the proof for completeness. Direct calculation shows that are eigenfunctions of , with eigenvalue .
Recall that is a basis in the weighted sense. Similarly, is a basis in the weighted sense. Therefore, is a basis in the weighted sense. This shows that are all possible eigenfunctions up to rescaling. ∎
Proof of Theorem 5.3.
The proof is divided into two parts.
Part I: We first assume that has index and we are going to prove that .
By Definition 5.2, there exists a -dimensional subspace so that each non-zero has compact support and
Without loss of generality, we may assume that for all . Denote by .
We pause to take a cut-off function and a universal constant so that
Let . We now assume on the contrary that . Then there exist has compact support and satisfies
(5.7) |
Define . Then and
(5.8) | ||||
Here is the projection to . The second equality used the fact that . The first inequality follows from for . For , .
Then by a direct computation, we have
(5.9) | ||||
Here the first inequality is from the divergence theorem and . (5.8) is used in the second inequality. In the last inequality, we note that for .
Now we prove that has contribution for index. Indeed,
Here we used (5.9) in the first inequality. The second one is from (5.7) and the definition of .
Note that and have disjoint support set. This implies that has , which lead to a contradiction. This finishes Part I.
Part II: We now prove has finite index.
It suffices to prove that for any compact domain , the Dirichlet eigenproblem of has index bounded from above uniformly. Recall that is the first eigenvalue of . Suppose are eigenvalues of on , which correspond to the eigenfunctions ; suppose are Dirichlet eigenvalues of () on , which correspond to the eigenfunctions . Recall that the converges to the eigenvalues of on as ; see [CM2020]*Lemma 6.1.
We first note that by definition, for any compact domain . Secondly, note that the eigenvalues of on are given by non-negative half-integers with multiplicity one, i.e. as . Thus we can take so that
(5.10) | for all and . |
To proceed the proof, we take . By Proposition 5.4, the Dirichlet eigenvalues of on are , which correspond to the eigenfunctions .
Since has finite index , the index of is bounded from above by uniformly. From Part I, . This implies that, must be non-negative if . So the index of is bounded from above by an uniform constant . This completes the proof. ∎
Appendix A Removing singularity for weakly stable free boundary -hypersurfaces
Theorem A.1 (cf. [ACS17]*Theorem 27).
Let be a compact Riemannian manifold with boundary of dimension . Given and an almost embedded free boundary -hypersurface with , assume that is weakly stable in as in Remark 2.7. If represents a varifold of bounded first variation in , then extends smoothly across as an almost embedded hypersurface in .
Proof.
Given any sequence of positive , consider the blowups , where . Since has bounded first variation, converges (up to a subsequence) to a stationary integral rectifiable cone in a half Euclidean space with free boundary on . By weak stability and Theorem 2.9, the convergence is locally smooth and graphical away from the origin, so is an integer multiple of some embedded minimal hypercone with free boundary. Hence the reflection of across is a stable minimal cone in , and hence a plane with integer multiplicities. Therefore, we conclude that or for some hyperplane with . Here . Note that may depend on the choice of .
If , then the argument in [Zhou19]*Theorem B.1 implies the removability of .
If for some half-hyperplane , then by the locally smooth and graphical convergence, there exists small enough, such that for any , has an -sheeted, ordered, graphical decomposition in the annulus :
Here each is a graph over . We can continue each all the way to , and denote the continuation by . Then each is a free boundary -hypersurface in and . By the Allard type regularity theorem for rectifiable varifolds with free boundary and bounded mean curvature [GJ86]*Theorem 4.13, extends as a free boundary hypersurface across for some . Higher regularity of follows from the prescribing mean curvature equation and elliptic regularity. ∎
Appendix B The second variation of for smooth hypersurfaces
Our goal is to derive the second variation of for arbitrary hypersurfaces with boundary in . Note that we always assume is isometrically embedded in some closed -dimensional Riemannian manifold . Let be an embedded hypersurface in with boundary on . We consider a two-parameter family of ambient variations of defined by
where and are the flows generated by compactly supported vector fields and in , respectively. We assume both . Therefore each is an embedded hypersurface in with boundary lying on .
We have
The computation in [ACS17]*Appendix A gives that
(B.1) | ||||
where
and
Here is the mean curvature vector; is the unit co-normal vector field of ; is the connection on ; and are the normal and tangential parts on , respectively.
Let be a smooth function on . Then a direct computation gives that
(B.2) | ||||
In the equality of , we used the following identity:
We remark that
Appendix C Cut-off trick
In this section, we provide a lemma which has been used in Part 3 of the proof of Theorem 2.9(v). Such a result has also been used in [GLWZ19].
Lemma C.1.
Let be a compact manifold with boundary of dimension . Let be an almost embedded free boundary -hypersurface and be a smooth function on . Then for any , there exists a family of cut-off functions for some so that
Proof.
Set
(C.1) |
where . Then we have and as . Then it suffices to prove as . This follows from
∎
Appendix D Local -foliation with free boundary
The following proposition is a generalization of minimal foliation given by B. White [Whi87]. This description has already been stated in [Wang20]*Proposition A.2.
Proposition D.1.
Let be a compact Riemannian manifold with boundary, and let be an embedded, free boundary minimal hypersurface. Given a point , there exist and a neighborhood of such that if is a smooth function with and
then for any , there exists a -function , whose graph meets orthogonally along and satisfies:
(where is evaluated with respect to the upward pointing normal of ), and
Furthermore, depends on in and the graphs forms a foliation.
Proof.
The proof follows from [Whi87]*Appendix together with the free boundary version [ACS17]*Section 3. The only modification is that we need to use the following map to replace in [ACS17]*Section 3:
The map is defined by
here all the notions are the same as [ACS17]*Section 3. ∎