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

Multiplicity one for min-max theory in compact manifolds with boundary and its applications

Ao Sun Department of Mathematics, University of Chicago, 5734 S. University Avenue, Chicago, IL, 60637 aosun@uchicago.edu Zhichao Wang Department of Mathematics, University of Toronto, Toronto, ON M5S2E4, Canada zhichao.wang@utoronto.ca  and  Xin Zhou Department of Mathematics, Cornell University, Ithaca, NY 14853, and Department of Mathematics, University of California Santa Barbara, Santa Barbara, CA 93106, USA xinzhou@cornell.edu
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 (Mn+1,M,g)(M^{n+1},\partial M,g) a compact Riemannian manifold with smooth boundary of dimension n+1n+1. In [Alm62], Almgren proved that the space of mod 2 relative cycles 𝒵n(M,M;2)\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}) is weakly homotopic to \mathbb{RP}^{\infty}; see also [LMN16]*§2.5. By performing a min-max procedure, Gromov [Gro88] defined the volume spectrum of MM, which is a sequence of non-decreasing positive numbers

0<ω1(M,g)ω2(M,g)ωk(M,g),0<\omega_{1}(M,g)\leq\omega_{2}(M,g)\leq\cdots\leq\omega_{k}(M,g)\to\infty,

depending only on MM and gg. Moreover, Gromov [Gro88] also showed that for each gg, ωk\omega_{k} grows like k1n+1k^{\frac{1}{n+1}}; see also Guth [Guth09]. Recently, Liokumovich-Marques-Neves [LMN16] proved that the volume spectrum obey a uniform Weyl Law.

When MM 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 3(n+1)73\leq(n+1)\leq 7,  each kk-width ωk\omega_{k} is realized by a disjoint collection of closed, embedded, minimal hypersurfaces with integer multiplicities, i.e. there exist integers {mjk}\{m_{j}^{k}\} and minimal hypersurfaces {Σjk}\{\Sigma_{j}^{k}\} so that

(1.1) ωk(M;g)=jlkmjkArea(Σjk) and  jlkindex(Σjk)k;\omega_{k}(M;g)=\sum_{j}^{l_{k}}m_{j}^{k}\cdot\mathrm{Area}(\Sigma_{j}^{k})\ \ \ \text{ and }\ \  \ \sum_{j}^{l_{k}}\operatorname{index}(\Sigma_{j}^{k})\leq k;

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]., mjk=1m_{j}^{k}=1 and Σjk\Sigma_{j}^{k} is two-sided for all kk\in\mathbb{N} and 1jlk1\leq j\leq l_{k}. 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 kk for each kk\in\mathbb{N} 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 kk, there exist integers {mjk}\{m_{j}^{k}\} and disjoint, embedded, free boundary minimal hypersurfaces {Σjk}\{\Sigma_{j}^{k}\} so that

(1.2) ωk(M;g)=jlkmjkArea(Σjk) and  jlkindex(Σjk)k.\omega_{k}(M;g)=\sum_{j}^{l_{k}}m_{j}^{k}\cdot\mathrm{Area}(\Sigma_{j}^{k})\ \ \ \text{ and }\ \  \ \sum_{j}^{l_{k}}\operatorname{index}(\Sigma_{j}^{k})\leq k.

Inspired by [Zhou19], it is natural to conjecture that mjk=1m_{j}^{k}=1 for generic metrics on compact manifolds with boundary. In this article we confirm this conjecture.

Theorem 1.1.

Let (Mn+1,M)(M^{n+1},\partial M) be a compact manifold with boundary of dimension 3(n+1)73\leq(n+1)\leq 7. For generic metrics gg on (M,M)(M,\partial M) and every integer k>0k>0, there exist a sequence of disjoint, two-sided, connected, embedded free boundary minimal hypersurfaces {Σjk}\{\Sigma_{j}^{k}\} satisfying

 ωk(M;g)=jArea(Σjk) and  jindex(Σjk)k. \omega_{k}(M;g)=\sum_{j}\mathrm{Area}(\Sigma_{j}^{k})\ \ \ \text{ and }\ \  \ \sum_{j}\operatorname{index}(\Sigma_{j}^{k})\leq k.
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 hh will be called an hh-hypersurface.

  1. (a)

    Generic properness for free boundary minimal hypersurfaces;

  2. (b)

    Existence and regularity for free boundary hh-hypersurfaces, including: the regularity for minimizers, compactness, and a generic min-max theory;

  3. (c)

    Genericity of good mean curvature prescribing functions hh for which the min-max theory for free boundary hh-hpersurfaces can be applied;

  4. (d)

    Generic countability of free boundary hh-hypersurface – an essential ingredient in the proof of Morse index estimates in the free boundary hh-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 (M,M)(M,\partial M), every embedded free boundary minimal hypersurface is both non-degenerate and proper.

Here a hypersurface Σ\Sigma is proper if it has empty touching set, i.e. Int(Σ)M=\mathrm{Int}(\Sigma)\cap\partial M=\emptyset.

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 𝔹n+1\mathbb{B}^{n+1}, and more generally in any strictly convex domain of n+1\mathbb{R}^{n+1}. 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 (Mn+1,M,g)(M^{n+1},\partial M,g) be a compact Riemannian manifold with non-negative Ricci curvature and strictly convex boundary of dimension 3(n+1)73\leq(n+1)\leq 7. Then there exist a sequence of two-sided, connected, embedded, free boundary minimal hypersurfaces {Σk}\{\Sigma_{k}\} satisfying

 Area(Σk)k1n+1 and index(Σk)k. \mathrm{Area}(\Sigma_{k})\sim k^{\frac{1}{n+1}}\ \ \ \ \text{ and }\ \ \ \ \operatorname{index}(\Sigma_{k})\leq k.

In particular, the unit ball 𝔹n+1\mathbb{B}^{n+1} with 3(n+1)73\leq(n+1)\leq 7 contains a sequence of free boundary minimal hypersurfaces with areas going to ++\infty. 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 n+1\mathbb{R}^{n+1} is called a self-shrinker if

H=x,ν2,H=\frac{\langle x,\nu\rangle}{2},

where HH is the mean curvature with respect to the normal vector field ν\nu. Equivalently, a self-shrinker is a minimal hypersurface in n+1\mathbb{R}^{n+1} under the Gaussian metric 𝒢=(4π)1e|x|22nδij\mathcal{G}=(4\pi)^{-1}e^{-\frac{|x|^{2}}{2n}}\delta_{ij}; 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 (n+1,𝒢)(\mathbb{R}^{n+1},\mathcal{G})) 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 (n+1,𝒢)(\mathbb{R}^{n+1},\mathcal{G}) 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 (3,𝒢)(\mathbb{R}^{3},\mathcal{G}) 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 3(n+1)73\leq(n+1)\leq 7, there exists a sequence of embedded self-shrinkers {Σk}\{\Sigma_{k}\} with entropy growth k1n+1k^{\frac{1}{n+1}} and index(Σk)k\mathrm{index}(\Sigma_{k})\leq k.

A direct corollary is the existence of self-shrinkers with arbitrary large entropy.

Corollary 1.7.

For any  n+1\mathbb{R}^{n+1} (n2n\geq 2) and Λ>0\Lambda>0, there exists an embedded self-shrinker with λ(Σ)(Λ,)\lambda(\Sigma)\in(\Lambda,\infty) and index<\mathrm{index}<\infty.

Proof of Corollary 1.7.

The corollary follows from Theorem 1.6 if (n+1)7(n+1)\leq 7. Now we assume that n7n\geq 7. Note that λ(Σ×)=λ(Σ)\lambda(\Sigma\times\mathbb{R})=\lambda(\Sigma) and Σ×\Sigma\times\mathbb{R} is a self-shrinker if and only if Σ\Sigma is a self-shrinker. Moreover, Σ\Sigma has finite index if and only if Σ×\Sigma\times\mathbb{R} has finite index by Theorem 5.3. ∎

There are several pioneer works which provided constructions of embedded self-shrinkers in 3\mathbb{R}^{3} 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 hh-hypersurfaces. Let (M,M,g)(M,\partial M,g) be a compact Riemannian manifold with boundary of dimension 3(n+1)73\leq(n+1)\leq 7. Moreover, we assume that every embedded free boundary minimal hypersurface in (M,M,g)(M,\partial M,g) is proper and does not have non-trivial Jacobi fields. Then the kk-the volume spectrum ωk(M;g)\omega_{k}(M;g) is realized by the min-max value of a homotopy class of kk-sweepouts. Note that the sweepouts are maps into 𝒵n(M,M;2)\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}). We can lift the kk-sweepouts to the space of Cappccioppoli set 𝒞(M)\mathcal{C}(M), and consider the associated relative homotopy class, denoted by Π~\widetilde{\Pi}. We will show that the relative min-max width is equal to ωk(M;g)\omega_{k}(M;g).

Take hC(M)h\in C^{\infty}(M) to be defined later. Let ϵj0\epsilon_{j}\rightarrow 0 and denote by Σj=Ωj\Sigma_{j}=\partial\Omega_{j} the free boundary ϵjh\epsilon_{j}h-hypersurface, produced by relative min-max theory (associated with Π~\widetilde{\Pi}). Letting jj\rightarrow\infty, and using compactness results, we get a free boundary minimal hypersurface Σ\Sigma with multiplicity mm. The aim is to prove m=1m=1 by taking suitable hh. Indeed, if m>1m>1, we can construct a non-trivial positive Jacobi field uu so that Lu=±2hLu=\pm 2h on Σ\Sigma and u𝜼=hM(ν,ν)u\frac{\partial u}{\partial\bm{\eta}}=h^{\partial M}(\nu,\nu)u, where LL is the Jacobi operator, ν\nu and 𝜼\bm{\eta} are respectively the unit normal and co-normal vector of Σ\Sigma, and hM(,)h^{\partial M}(\cdot,\cdot) is the second fundamental form of M\partial M. By choosing a suitable hh, we prove that this kind of uu 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 M~\widetilde{M} (which contains MM as a domain) so that each minimal hypersurface in M~\widetilde{M} with free boundary on M\partial M is non-degenerate and transverse to M\partial M. To obtain openness, we use the compactness to show that given a strongly bumpy metric gg and a constant CC, there is an open neighborhood of gg so that for each metric gg^{\prime} in that neighborhood, any free boundary minimal hypersurface in (M,M,g)(M,\partial M,g^{\prime}) with weak Morse index and area no more than CC is non-degenerate and proper. By taking intersections of such open and dense sets of metrics for a sequence CnC_{n}\rightarrow\infty, we obtain the desired genericity.

Second, to establish the min-max theory for free boundary hh-hypersurfaces, we need to generalize a number of regularity results. In Theorem 2.1, we prove the regularity for minimizing free boundary hh-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 hh-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 hh-hypersurfaces are much simpler since we only need to consider the case where the touching sets are contained in some (n1)(n-1)-dimensional subsets.

To obtain the aforementioned control on the touching sets, we need to restrict to “good” prescribing functions hh, where each free boundary hh-hypersurface has at most (n1)(n-1)-dimensional touching set (including self-touching and touching with M\partial M). To show the genericity of such good functions, we extend a given smooth function hh on MM to h~\widetilde{h} on M~\widetilde{M} and perturb it to h~1\widetilde{h}_{1} by [ZZ18]*Proposition 3.8 so that each h~1\widetilde{h}_{1}-hypersurface has small self-touching set. By genericity of Morse functions we can find another perturbed function h~2\widetilde{h}_{2} so that each h~2\widetilde{h}_{2}-hypersurface touches M\partial M in a small subset as desired. This implies h~2|M\widetilde{h}_{2}|_{M} is a good choice. Moreover, each function in a neighborhood of h~2|M\widetilde{h}_{2}|_{M} 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 hh-hypersurfaces for good pairs (see Subsection 2.4). Our proof of the countability follows from the following observation of the compactness of free boundary hh-hypersurfaces. Let Σj\Sigma_{j} be a sequence of free boundary hh-hypersurfaces with bounded index and area. Then it converges as varifolds to some limit hypersurface Σ\Sigma; (see Theorem 2.9). If the convergence is smooth, then Σ\Sigma has a non-trivial Jacobi field related to 𝒜h\mathcal{A}^{h}; on the contray, if the convergence is not smooth, Σ\Sigma has strictly less index than that of Σj\Sigma_{j} for sufficiently large jj. 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 n+1\mathbb{R}^{n+1}, 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 Σ×\Sigma\times\mathbb{R} whenever Σ\Sigma is a lower dimensional shrinker, and we prove in Theorem 5.3 that if Σ\Sigma has finite index, so does Σ×\Sigma\times\mathbb{R}. 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 𝒜h\mathcal{A}^{h}-minimizing hypersurfaces and the compactness of free boundary hh-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 hh-hypersurfaces. The regularity of hh-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 AD.

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 (Mn+1,Mn+1,g)(M^{n+1},\partial M^{n+1},g) be a compact, oriented, smooth Riemannian manifold with smooth boundary, of dimension 3n+173\leq n+1\leq 7. In this paper, we always isometrically embed (M,M,g)(M,\partial M,g) into a closed manifold (M~,g~)(\widetilde{M},\widetilde{g}) with dimension n+1n+1. Assume that (M~,g~)(\widetilde{M},\widetilde{g}) is embedded in some L,L\mathbb{R}^{L},L\in\mathbb{N}. {basedescript}\desclabelstyle\multilinelabel\desclabelwidth10em

the geodesic ball of (M~,g)(\widetilde{M},g);

the space of smooth vector fields on M~\widetilde{M};

the collection of X𝔛(M~)X\in\mathfrak{X}(\widetilde{M}) with X(p)Tp(M)X(p)\in T_{p}(\partial M) for pp in a neighborhood of Σ\partial\Sigma in M\partial M;

kk-dimensional Hausdorff measure;

the space of smooth Riemannian metrics on MM;

the space of integer rectifiable kk-currents with support in MM;

the space of mod 2 rectifiable kk-currents with support in MM;

the group \mathbb{Z} or 2\mathbb{Z}_{2};

the space of Tk(M;G)T\in\mathcal{R}_{k}(M;G) and spt(T)M\mathrm{spt}(\partial T)\subset\partial M;

the space of Tk(M;G)T\in\mathcal{R}_{k}(M;G) and Tk1(M;G)\partial T\in\mathcal{R}_{k-1}(M;G);

equivalent class of Zk(M,M;G)Z_{k}(M,\partial M;G), i.e. T1τT_{1}\in\tau if and only if TT1k(M;G)T-T_{1}\in\mathcal{R}_{k}(\partial M;G) for any TτT\in\tau;

the space of equivalent classes in Zk(M,M;G)Z_{k}(M,\partial M;G);

the integer rectifiable varifold associated with Tk(M;)T\in\mathcal{R}_{k}(M;\mathbb{Z});

the Radon measure associated with Tk(M;)T\in\mathcal{R}_{k}(M;\mathbb{Z});

the mass norm on k(M;G)\mathcal{R}_{k}(M;G);

𝐌(τ)=inf{𝐌(T):Tτ}\mathbf{M}(\tau)=\inf\{\mathbf{M}(T):T\in\tau\}.

the flat metric on k(M;G)\mathcal{R}_{k}(M;G);

(τ1,τ2)=inf{(T1,T2):T1τ1,T2τ2}\mathcal{F}(\tau_{1},\tau_{2})=\inf\{\mathcal{F}(T_{1},T_{2}):T_{1}\in\tau_{1},T_{2}\in\tau_{2}\} for τ1,τ2𝒵k(M,M;G)\tau_{1},\tau_{2}\in\mathcal{Z}_{k}(M,\partial M;G);

the space of sets ΩM\Omega\subset M with finite perimeter (Cappccioppoli set);

the space of smooth vector field in MM so that X(p)Tp(M)X(p)\in T_{p}(\partial M) for pMp\in\partial M;

the (reduced) boundary of Ω\llbracket\Omega\rrbracket restricted in the interior of MM, (thus, as an integer rectifiable current, ΩM=0\partial\Omega\llcorner\partial M=0);

the space of kk-dimensional integral varifolds;

the tangent space of VV at pp;

denotes the kk-dimensional Grassmannian bundle over MM;

𝐅(V,W)=sup{V(f)W(f):fCc(Gk(L)),|f|1,Lip(f)1},\mathbf{F}(V,W)=\sup\{V(f)-W(f):f\in C_{c}(G_{k}(\mathbb{R}^{L})),|f|\leq 1,\mathrm{Lip}(f)\leq 1\}, for V,W𝒱k(M)V,W\in\mathcal{V}_{k}(M);

𝐅(Ω1,Ω2)=(Ω1Ω2)+𝐅(|Ω1|,|Ω2|)\mathbf{F}(\Omega_{1},\Omega_{2})=\mathcal{F}(\Omega_{1}-\Omega_{2})+\mathbf{F}(|\partial\Omega_{1}|,|\partial\Omega_{2}|) for Ω1,Ω2𝒞(M)\Omega_{1},\Omega_{2}\in\mathcal{C}(M);

the integral current associated with MM;

the integral varifold associated with MM;

the collection of Ω𝒞(M)\Omega^{\prime}\in\mathcal{C}(M) so that 𝐅(Ω,Ω)ϵ\mathbf{F}(\Omega,\Omega^{\prime})\leq\epsilon, where Ω𝒞(M)\Omega\in\mathcal{C}(M).

For any τ𝒵k(M,M;G)\tau\in\mathcal{Z}_{k}(M,\partial M;G), there is a canonical representative TτT\in\tau where TZk(M,M;G)T\in Z_{k}(M,\partial M;G) and TM=0T\llcorner\partial M=0.

Given c>0c>0, a varifold V𝒱k(M)V\in\mathcal{V}_{k}(M) is said to have cc-bounded first variation in an open subset UMU\subset M, if

|δV(X)|cM|X|𝑑μV, for any X𝔛(U,UM),|\delta V(X)|\leq c\,\int_{M}|X|d\mu_{V},\text{ for any }X\in\mathfrak{X}(U,U\cap\partial M),

where 𝔛(U,UM)\mathfrak{X}(U,U\cap\partial M) is defined to be the subset of 𝔛(M,M)\mathfrak{X}(M,\partial M) compactly supported in UU.

Let Σ\Sigma be an nn-dimensional manifold with smooth boundary. Recall that an immersion is a map ϕ:(Σ,Σ)(M,M)\phi:(\Sigma,\partial\Sigma)\hookrightarrow(M,\partial M). In this paper, we often use Σ\Sigma to denote both the immersion and the hypersurface when there is no ambiguity.

We are interested in the following weighted area functional defined on 𝒞(M)\mathcal{C}(M). Given hC(M)h\in C^{\infty}(M), define the 𝒜h\mathcal{A}^{h}-functional on 𝒞(M)\mathcal{C}(M) by

(2.1) 𝒜h(Ω)=𝐌(Ω)Ωh𝑑n+1.\mathcal{A}^{h}(\Omega)=\mathbf{M}(\partial\Omega)-\int_{\Omega}h\,d\mathcal{H}^{n+1}.

The first variation formula for 𝒜h\mathcal{A}^{h} along X𝔛(M,M)X\in\mathfrak{X}(M,\partial M) is

δ𝒜h|Ω(X)=ΩdivΩX𝑑μΩΩhX,ν𝑑μΩ,\delta\mathcal{A}^{h}\big{|}_{\Omega}(X)=\int_{\partial\Omega}\mathrm{div}_{\partial\Omega}X\,d\mu_{\partial\Omega}-\int_{\partial\Omega}h\langle X,\nu\rangle\,d\mu_{\partial\Omega},

where ν\nu is the outward unit normal vector field of Ω\partial\Omega.

When the boundary Ω=Σ\partial\Omega=\Sigma has support on a smooth immersed hypersurface, by virtue of the divergence theorem, we have

(2.2) δ𝒜h|Ω(X)=Ω(Hh)X,ν𝑑μΩ+ΣX,𝜼𝑑μΣ,\delta\mathcal{A}^{h}\big{|}_{\Omega}(X)=\int_{\partial\Omega}(H-h)\langle X,\nu\rangle\,d\mu_{\partial\Omega}+\int_{\partial\Sigma}\langle X,\bm{\eta}\rangle\,d\mu_{\partial\Sigma},

where 𝜼\bm{\eta} is the unit outer co-normal vector field of Σ\partial\Sigma. If Ω\Omega is a critical point of 𝒜h\mathcal{A}^{h}, then (2.2) directly implies that Σ\Sigma must have mean curvature H=h|ΣH=h|_{\Sigma} and ννM\nu\perp\nu_{\partial M} along Σ\partial\Sigma, where νM\nu_{\partial M} is the outward unit normal vector field of M\partial M.

Recall that an immersed hypersurface is called an hh-hypersurface if its mean curvature H(x)=h(x)H(x)=h(x) everywhere.

2.1. Regularity for boundaries minimizing the 𝒜h\mathcal{A}^{h}-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 𝒜h\mathcal{A}^{h}-functional to prove the regularity of hh-hypersurfaces bounding a domain which minimizes the 𝒜h\mathcal{A}^{h}-functional; (see  [ZZ18]*Theorem 2.2). In this part, we provide the regularity for 𝒜h\mathcal{A}^{h}-minimizers with free boundaries.

Let BB be an (n+1)(n+1)-ball in n+1\mathbb{R}^{n+1}, and SBS\subset B a compact embedded nn-ball such that SB=\partial S\cap B=\emptyset. Denote by B+B^{+} and BB^{-} the two components of BSB\setminus S. Note that the 𝒜h\mathcal{A}^{h}-functional in (2.1) is well-defined for Ω𝒞(B+)\Omega\in\mathcal{C}(B^{+}).

Theorem 2.1.

Suppose that Ω𝒞(B+)\Omega\in\mathcal{C}(B^{+}) minimizes the 𝒜h\mathcal{A}^{h}-functional with free boundary on SS: that is, for any other Λ𝒞(B+)\Lambda\in\mathcal{C}(B^{+}) with sptΛΩB\operatorname{spt}\|\Lambda-\Omega\|\subset B, we have 𝒜h(Λ)𝒜h(Ω)\mathcal{A}^{h}(\Lambda)\geq\mathcal{A}^{h}(\Omega). Then except for a set of Hausdorff dimension at most (n7)(n-7), Ω\partial\Omega is smooth hypersurface embedded properly (under relative topology) in B+B^{+} and meets SB+S\cap B^{+} orthogonally.

We recall a few notations. Without loss of generality, we assume that SsptΩS\cap\operatorname{spt}\|\partial\Omega\|\neq\emptyset. For xBx\in B, let ξ(x)\xi(x) be the closest point to xx among SS. Clearly, ξ\xi is well-defined in a small neighborhood of SS. By possible rescaling, we assume that BB is contained in such a neighborhood. Then we define the reflection map σ\sigma across SS by

σ(x)=2ξ(x)x.\sigma(x)=2\xi(x)-x.

We have σ2=id\sigma^{2}=\mathrm{id} and define

Ω~=Ωσ#Ω.\widetilde{\Omega}=\Omega-\sigma_{\#}\Omega.

Thus Ω~𝒞(V)\widetilde{\Omega}\in\mathcal{C}(V), where V=B+(SB)σ(B+)V=B^{+}\cup(S\cap B)\cap\sigma(B^{+}). By possible rescaling, we may assume that B3RBB_{3R}\subset B. Set

Ω=Ω~BR.\Omega^{\prime}=\widetilde{\Omega}\llcorner B_{R}.

We also denote by B~r(p)=σ(Br(p))\widetilde{B}_{r}(p)=\sigma(B_{r}(p)). By shrinking rr if it is necesssary, we can assume that there exists κ>1\kappa>1 so that

(2.3) Lip(σ|B~2r(p))1+κr.\mathrm{Lip}(\sigma|_{\widetilde{B}_{2r}(p)})\leq 1+\kappa r.
Proof of Theorem 2.1.

Take pspt(Ω)p\in\operatorname{spt}(\partial\Omega^{\prime}). Then for r<(R|p|)/3r<(R-|p|)/3 and Λ𝒞(V)\Lambda\in\mathcal{C}(V) so that sptΛΩBr(p)\operatorname{spt}\|\Lambda-\Omega^{\prime}\|\subset B_{r}(p), we have

(2.4) μΩ(Br(p))\displaystyle\mu_{\partial\Omega^{\prime}}(B_{r}(p)) =μΩ(Br(p)BR+)+μΩ(Br(p)σ(BR+))\displaystyle=\mu_{\partial\Omega^{\prime}}(B_{r}(p)\cap B^{+}_{R})+\mu_{\partial\Omega^{\prime}}(B_{r}(p)\cap\sigma(B^{+}_{R}))
=μΩ(Br(p))+μσ#(Ω)(σ(B~r(p)BR+))\displaystyle=\mu_{\partial\Omega}(B_{r}(p))+\mu_{\sigma_{\#}(\partial\Omega)}\big{(}\sigma(\widetilde{B}_{r}(p)\cap B_{R}^{+})\big{)}
μΩ(Br(p))+(1+κr)nμΩ(B~r(p))\displaystyle\leq\mu_{\partial\Omega}(B_{r}(p))+(1+\kappa r)^{n}\mu_{\partial\Omega}(\widetilde{B}_{r}(p))
μΩ(Br(p))+μΩ(B~r(p))+β(n)κrμΩ(B~r(p))\displaystyle\leq\mu_{\partial\Omega}(B_{r}(p))+\mu_{\partial\Omega}(\widetilde{B}_{r}(p))+\beta(n)\kappa r\mu_{\partial\Omega}(\widetilde{B}_{r}(p))
μΩ(Br(p))+μΩ(B~r(p))+(1+κr)nβ(n)κrμσ#(Ω)(Br(p)).\displaystyle\leq\mu_{\partial\Omega}(B_{r}(p))+\mu_{\partial\Omega}(\widetilde{B}_{r}(p))+(1+\kappa r)^{n}\beta(n)\kappa r\mu_{\sigma_{\#}(\partial\Omega)}(B_{r}(p)).

Here (2.3) is used in the first and the last inequalities. The constant β(n)\beta(n) is allowed to  change from line to line. On the other hand,

(2.5) μΛ(Br(p))\displaystyle\mu_{\partial\Lambda}(B_{r}(p)) μΛ(Br(p)BR+)+μΛ(Br(p)σ(BR+))\displaystyle\geq\mu_{\partial\Lambda}(B_{r}(p)\cap B_{R}^{+})+\mu_{\partial\Lambda}(B_{r}(p)\cap\sigma(B_{R}^{+}))
μΛ(Br(p)BR+)+(1+κr)nμσ#(Λ)(B~r(p)BR+)\displaystyle\geq\mu_{\partial\Lambda}(B_{r}(p)\cap B_{R}^{+})+(1+\kappa r)^{-n}\mu_{\sigma_{\#}(\partial\Lambda)}\big{(}\widetilde{B}_{r}(p)\cap B_{R}^{+}\big{)}
μΛ(Br(p)BR+)+μσ#(Λ)(B~r(p)BR+)β(n)κrμσ#(Λ)(B~r(p)BR+)\displaystyle\geq\mu_{\partial\Lambda}(B_{r}(p)\cap B_{R}^{+})+\mu_{\sigma_{\#}(\partial\Lambda)}(\widetilde{B}_{r}(p)\cap B_{R}^{+})-\beta(n)\kappa r\mu_{\sigma_{\#}(\partial\Lambda)}(\widetilde{B}_{r}(p)\cap B_{R}^{+})
μΛ(Br(p)BR+)+μσ#(Λ)(B~r(p)BR+)β(n)κrμΛ(Br(p)).\displaystyle\geq\mu_{\partial\Lambda}(B_{r}(p)\cap B_{R}^{+})+\mu_{\sigma_{\#}(\partial\Lambda)}(\widetilde{B}_{r}(p)\cap B_{R}^{+})-\beta(n)\kappa r\mu_{\partial\Lambda}(B_{r}(p)).

Here the third inequality is from (2.3).

Recall that Ω\Omega is 𝒜h\mathcal{A}^{h}-minimizing. Assume that |h|c|h|\leq c. Hence we have

μΛ(Br(p)BR+)μΩ(Br(p))cn+1((ΛΩ)BR+)cn+1(ΛΩ),\displaystyle\mu_{\partial\Lambda}(B_{r}(p)\cap B_{R}^{+})-\mu_{\partial\Omega}(B_{r}(p))\geq-c\mathcal{H}^{n+1}((\Lambda\triangle\Omega)\cap B_{R}^{+})\geq-c\mathcal{H}^{n+1}(\Lambda\triangle\Omega^{\prime}),

and

μσ#(Λ)(B~r(p)BR+)μΩ(B~r(p))\displaystyle\mu_{\sigma_{\#}(\partial\Lambda)}(\widetilde{B}_{r}(p)\cap B_{R}^{+})-\mu_{\partial\Omega}(\widetilde{B}_{r}(p)) cn+1((σ#ΛΩ)B~r(p) BR+)\displaystyle\geq-c\mathcal{H}^{n+1}((\sigma_{\#}\Lambda\triangle\Omega)\cap\widetilde{B}_{r}(p)\cap B_{R}^{+})
cn+1(ΛΩ),\displaystyle\geq-c\mathcal{H}^{n+1}(\Lambda\triangle\Omega^{\prime}),

where \triangle 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) μΛ(Br(p))μΩ(Br(p))β(n,c)r[μΛ(Br(p))+μΩ(Br(p))].\mu_{\partial\Lambda}(B_{r}(p))-\mu_{\partial\Omega^{\prime}}(B_{r}(p))\geq-\beta^{\prime}(n,c)r[\mu_{\partial\Lambda}(B_{r}(p))+\mu_{\partial\Omega^{\prime}}(B_{r}(p))].

Here β(n,c)\beta^{\prime}(n,c) is a constant depends only on cc and nn.

Applying (2.6) to [Mor03]*Corollary 3.7, 3.8, ΩBr(p)\partial\Omega^{\prime}\llcorner B_{r}(p) is C1,1/2C^{1,1/2}. In particular, Ω\partial\Omega is a properly embedded free boundary C1,1/2C^{1,1/2} hypersurface. Since Ω\partial\Omega is an hh-hypersurface, the classical PDE argument implies the higher regularity. ∎

Note that such process works for any Riemannian manifold.  Let (Mn+1,M,g)(M^{n+1},\partial M,g) be compact Riemannian manifold with boundary which is isometrically embedded into a closed manifold (M~,g)(\widetilde{M},g). Recall that Br(p)B_{r}(p) is denoted by the geodesic ball of M~\widetilde{M} with radius rr and center at pp.  We then have the following regularity theorem:

Theorem 2.2.

Given Ω𝒞(M)\Omega\in\mathcal{C}(M), psptΩp\in\operatorname{spt}\|\partial\Omega\|, and some small s>0s>0, suppose that ΩBs(p)\Omega\llcorner B_{s}(p) minimizes the 𝒜h\mathcal{A}^{h}-functional: that is, for any other Λ𝒞(M)\Lambda\in\mathcal{C}(M) with sptΛΩBs(p)M\operatorname{spt}\|\Lambda-\Omega\|\subset B_{s}(p)\cap M, we have 𝒜h(Λ)𝒜h(Ω)\mathcal{A}^{h}(\Lambda)\geq\mathcal{A}^{h}(\Omega). Then except for a set of Hausdorff dimension at most (n7)(n-7), ΩBs(p)\partial\Omega\llcorner B_{s}(p) is a properly embedded hypersurface with free boundary on M\partial M, and is real analytic if the ambient metric on MM 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 M\partial M but do not cross.  A hypersurface is properly embedded if it is almost embedded and no sheets touch other sheets or M\partial M.

Let 𝒮=𝒮(g)\mathcal{S}=\mathcal{S}(g) be the collection of all Morse functions hh such that the zero set Σ0={h=0}\Sigma_{0}=\{h=0\} is a compact, smoothly embedded hypersurface so that

  • Σ0\Sigma_{0} is transverse to M\partial M and the mean curvature of Σ0\Sigma_{0} vanishes to at most finite order;

  • {xM:HM(x)=h(x) or HM(x)=h(x)}\{x\in\partial M:H_{\partial M}(x)=h(x)\text{ or }H_{\partial M}(x)=-h(x)\} is contained in an (n1)(n-1)-dimensional submanifold of M\partial M.

Lemma 2.3.

𝒮(g)\mathcal{S}(g) contains an open and dense subset in C(M)C^{\infty}(M).

Proof.

For any h0C(M)h_{0}\in C^{\infty}(M) and a neighborhood 𝒰0\mathcal{U}_{0} of h0h_{0} in C(M)C^{\infty}(M), we are going to find an open subset of 𝒰1𝒰0\mathcal{U}_{1}\subset\mathcal{U}_{0} so that 𝒰1𝒮(g)\mathcal{U}_{1}\subset\mathcal{S}(g). First, h0h_{0} can be extended to a function h~0C(M~)\widetilde{h}_{0}\in C^{\infty}(\widetilde{M}). We can also extend HM(x)H_{\partial M}(x) to a smooth function H~C(M~)\widetilde{H}\in C^{\infty}(\widetilde{M}).  Second, we take a neighborhood 𝒰~\widetilde{\mathcal{U}} of h~0\widetilde{h}_{0} in C(M~)C^{\infty}(\widetilde{M}) so that h~|M𝒰0\widetilde{h}|_{M}\in\mathcal{U}_{0} if h~𝒰~\widetilde{h}\in\widetilde{\mathcal{U}}.

By [ZZ18]*Proposition 3.8, there exists a Morse function h~1𝒰~\widetilde{h}_{1}\in\widetilde{\mathcal{U}} so that {h~1=0}\{\widetilde{h}_{1}=0\} is a closed, embedded hypersurface with mean curvature vanishing at most finite order. By perturbing h~1\widetilde{h}_{1} slightly, according to the generic existence of Morse functions, we can find h~2𝒰~\widetilde{h}_{2}\in\widetilde{\mathcal{U}} so that

  • h~2\widetilde{h}_{2}, H~±h~2\widetilde{H}\pm\widetilde{h}_{2} are all Morse functions on M~\widetilde{M};

  • {h~2=0}\{\widetilde{h}_{2}=0\} is a closed, embedded hypersurface which is transverse to M\partial M and has mean curvature vanishing at most finite order;

  • {H~(x)=h~2(x)}\{\widetilde{H}(x)=\widetilde{h}_{2}(x)\} and {H~(x)=h~2(x)}\{\widetilde{H}(x)=-\widetilde{h}_{2}(x)\} are smooth, embedded hypersurfaces and transverse to M\partial M.

Denote by h=h~2|Mh=\widetilde{h}_{2}|_{M}. Then the last item implies that {xM:HM(x)=h(x) or h(x)}\{x\in\partial M:H_{\partial M}(x)=h(x)\text{ or }-h(x)\} is contained in an (n1)(n-1)-dimensional submanifold of M\partial M. Thus h𝒮(g)h\in\mathcal{S}(g). Moreover, by the choice of h~2\widetilde{h}_{2}, we have that hh and H~|M±h\widetilde{H}|_{M}\pm h are Morse functions on MM with 0 as regular value. Thus we conclude that for any hh^{\prime} in a CC^{\infty} neighborhood 𝒰1\mathcal{U}_{1} of hh in C(M)C^{\infty}(M), hh^{\prime} and H~|M±h\widetilde{H}|_{M}\pm h^{\prime} are still Morse functions with 0 as regular value. 𝒰1\mathcal{U}_{1} is the desired open set and Lemma 2.3 is proved. ∎

In the next, we prove that for each h𝒮(g)h\in\mathcal{S}(g), the touching set of an almost embedded free boundary hh-hypersurface has dimension less than or equal to (n1)(n-1).

Proposition 2.4.

Let h𝒮(g)h\in\mathcal{S}(g) and Σ1,Σ2\Sigma_{1},\Sigma_{2} are two different connected, embedded, free boundary hh-hypersurfaces in a connected open set UU. Then Σ1Σ2\Sigma_{1}\cap\Sigma_{2} and Σ1M\Sigma_{1}\cap\partial M are contained in a countable union of connected, smoothly embedded, (n1)(n-1)-dimensional submanifolds.

Proof.

Recall that the argument in [ZZ18]*Theorem 3.11 implies that Σ1Σ2Int(M)\Sigma_{1}\cap\Sigma_{2}\cap\mathrm{Int}(M) is contained in a countable union of connected, smoothly embedded, (n1)(n-1)-dimensional submanifolds. Note that Σ1Σ2(Σ1M)(Σ1Σ2Int(M))\Sigma_{1}\cap\Sigma_{2}\subset(\Sigma_{1}\cap\partial M)\cup(\Sigma_{1}\cap\Sigma_{2}\cap\mathrm{Int}(M)). Thus it suffices to prove that Σ1M\Sigma_{1}\cap\partial M is contained in the submanifolds as described.

To do this, we first take pIntΣ1Mp\in\mathrm{Int}\Sigma_{1}\cap\partial M so that h(p)±H(p)0h(p)\pm H(p)\neq 0. Then there exists a neighborhood U(p)MU(p)\subset M of pp so that Σ1U(p)\Sigma_{1}\cap U(p) can be written as a graph over MU(p)\partial M\cap U(p). Denote by u1u_{1} the graph function. By shrinking the neighborhood, we can also assume that

(2.7) HM(x)+h(x,u1(x))0  and  HM(x)h(x,u1(x))0H_{\partial M}(x)+h(x,u_{1}(x))\neq 0\  \text{ and }\  H_{\partial M}(x)-h(x,u_{1}(x))\neq 0

for all xMU(p)x\in\partial M\cap U(p). Note that such a function satisfies an inhomogeneous linear elliptic PDE of the form

Lu1=HM(x)+h(x,u1(x))  or  Lu1=HM(x)h(x,u1(x)).Lu_{1}=H_{\partial M}(x)+h(x,u_{1}(x))\  \text{ or }\  Lu_{1}=H_{\partial M}(x)-h(x,u_{1}(x)).

Together with (2.7), then the Hessian of u1u_{1} at pp has rank at least 1. The implicit function theorem then implies that, on a possibly smaller neighborhood of pp, the touching set {u1=u1=0}\{u_{1}=\nabla u_{1}=0\} is contained in an (n1)(n-1)-dimensional submanifold; see [ZZ17]*Lemma 2.8 for more details. Therefore, we conclude that

(2.8) {xMIntΣ1:HMh(x) and HMh(x)}\{x\in\partial M\cap\mathrm{Int}\Sigma_{1}:H_{\partial M}\neq h(x)\text{ and }H_{\partial M}\neq-h(x)\}

is contained in the submanifolds as described in the proposition. Recall that

(2.9) {xM:HM(x)=h(x) or HM(x)=h(x)}\{x\in\partial M:H_{\partial M}(x)=h(x)\text{ or }H_{\partial M}(x)=-h(x)\}

is contained in an (n1)(n-1)-dimensional submanifold of M\partial M. Then Proposition 2.4 follows from the fact that Σ1M\Sigma_{1}\cap\partial M is contained in the union of (2.8), (2.9) and Σ1\partial\Sigma_{1}. ∎

2.3. Compactness of free boundary hh-hypersurfaces

Recall that an almost embedded hypersurface Σ\Sigma is called a free boundary hh-hypersurface if HΣ=h|ΣH_{\Sigma}=h|_{\Sigma} and Σ\Sigma meets M\partial M orthogonally along Σ\partial\Sigma. Given h𝒮(g)h\in\mathcal{S}(g), denote by 𝒫h\mathcal{P}^{h} the collection of free boundary hh-hypersurfaces such that Σ=Ω\llbracket\Sigma\rrbracket=\partial\Omega for some open set ΩM\Omega\subset M.

Note that when h𝒮(g)h\in\mathcal{S}(g), the min-max free boundary hh-hypersurfaces produced in Theorem 3.10 satisfy the above requirements. Indeed, such Σ=Ω\llbracket\Sigma\rrbracket=\partial\Omega is a critical point of the weighted 𝒜h\mathcal{A}^{h} functional:

𝒜h(Ω)=Area(Σ)Ωh𝑑n+1.\mathcal{A}^{h}(\Omega)=\mathrm{Area}(\Sigma)-\int_{\Omega}h\,d\mathcal{H}^{n+1}.

The second variation formula for 𝒜h\mathcal{A}^{h} along normal vector field X=φνX=\varphi\nu is given by

(2.10) δ2𝒜h|Ω(X,X)=IIΣ(φ,φ)\displaystyle\delta^{2}\mathcal{A}^{h}|_{\Omega}(X,X)=\mathrm{II}_{\Sigma}(\varphi,\varphi)
=\displaystyle= Σ(|φ|2(Ric(ν,ν)+|A|2+νh)φ2)𝑑μΣΣhM(ν,ν)φ2𝑑μΣ.\displaystyle\int_{\Sigma}(|\nabla\varphi|^{2}-(\operatorname{Ric}(\nu,\nu)+|A|^{2}+\partial_{\nu}h)\varphi^{2})\,d\mu_{\Sigma}-\int_{\partial\Sigma}h^{\partial M}(\nu,\nu)\varphi^{2}\,d\mu_{\partial\Sigma}.

In the above formula, φ\nabla\varphi is the gradient of φ\varphi on Σ\Sigma; Ric\mathrm{Ric} is the Ricci curvature of MM; AA and hMh^{\partial M} are the second fundamental forms of Σ\Sigma and M\partial M with normal vector fields ν\nu and νM\nu_{\partial M}, respectively.

We remark that in (2.10), ΠΣ(,)\Pi_{\Sigma}(\cdot,\cdot) can also be defined for any immersed free boundary hh-hypersurfaces and it is a quadratic form on the space of CC^{\infty}-functions on Σ\Sigma (not ϕ(Σ)\phi(\Sigma)).

The Jacobi field of Σ\Sigma is defined to be a smooth function φ\varphi on Σ\Sigma (not ϕ(Σ)\phi(\Sigma)) so that

(2.11) {(Δ+|A|2+Ric(ν,ν)+νh)φ=0 on Σ,φ𝜼=hM(ν,ν)φ  on  Σ.\left\{\begin{aligned} &(\Delta+|A|^{2}+\operatorname{Ric}(\nu,\nu)+\partial_{\nu}h)\varphi=0\ \ \  \text{on }\Sigma,\\ &\frac{\partial\varphi}{\partial\bm{\eta}}=h^{\partial M}(\nu,\nu)\varphi\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \  \text{ on } \partial\Sigma.\end{aligned}\right.

The classical Morse index for Σ\Sigma is defined to be the number of negative eigenvalues of the above quadratic form. However, since Σ\Sigma may touch itself and the boundary of MM, a weaker version of the index is needed. Such a concept was introduced by Zhou [Zhou19, Definition 2.1, 2.3] for closed hh-hypersurfaces based on Marques-Neves [MN16]*Definition 4.1.

Definition 2.5.

Given Σ𝒫h\Sigma\in\mathcal{P}^{h} with Σ=Ω\Sigma=\partial\Omega, kk\in\mathbb{N} and ϵ0\epsilon\geq 0, we say that Σ\Sigma is kk-unstable in an ϵ\epsilon-neighborhood if there exist 0<c0<10<c_{0}<1 and a smooth family {Fv}vB¯kDiff(M)\{F_{v}\}_{v\in\overline{B}^{k}}\subset\mathrm{Diff}(M) with F0=Id,Fv=Fv1F_{0}=\mathrm{Id},F_{-v}=F_{v}^{-1} for all vB¯kv\in\overline{B}^{k} (the standard kk-dimensional unit ball in k\mathbb{R}^{k}) such that, for any Ω𝐁¯2ϵ𝐅(Ω)\Omega^{\prime}\in\overline{\mathbf{B}}^{\mathbf{F}}_{2\epsilon}(\Omega), the smooth function:

𝒜Ωh:B¯k[0,+),𝒜Ωh(v)=𝒜h(Fv(Ω))\mathcal{A}_{\Omega^{\prime}}^{h}:\overline{B}^{k}\rightarrow[0,+\infty),\ \ \ \ \mathcal{A}^{h}_{\Omega^{\prime}}(v)=\mathcal{A}^{h}(F_{v}(\Omega^{\prime}))

satisfies

  • 𝒜Ωh\mathcal{A}^{h}_{\Omega^{\prime}} has a unique maximum at m(Ω)Bc0/10k(0)m(\Omega^{\prime})\in B_{c_{0}/\sqrt{10}}^{k}(0);

  • 1c0IdD2𝒜Ωh(u)c0Id-\frac{1}{c_{0}}\mathrm{Id}\leq D^{2}\mathcal{A}^{h}_{\Omega^{\prime}}(u)\leq-c_{0}\mathrm{Id} for all uB¯ku\in\overline{B}^{k}.

Since Σ\Sigma is a critical point of 𝒜h\mathcal{A}^{h}, necessarily m(Ω)=0m(\Omega)=0.

When h0h\equiv 0, this reduces to the kk-unstable notion for free boundary minimal hypersurfaces defined in [GLWZ19, Definition 5.5].

Definition 2.6.

Assume that Σ𝒫h\Sigma\in\mathcal{P}^{h} or Σ\Sigma is a free boundary minimal hypersurface. Given kk\in\mathbb{N}, we say that the weak Morse index of Σ\Sigma is bounded (from above) by kk, denoted as

indexw(Σ)k,\mathrm{index}_{w}{(\Sigma)}\leq k,

if Σ\Sigma is not jj-unstable in 0-neighborhood for any jk+1j\geq k+1. Σ\Sigma is said to be weakly stable if indexw(Σ)=0\mathrm{index}_{w}(\Sigma)=0.

Remark 2.7.

We make several remarks:

  • If Σ𝒫h\Sigma\in\mathcal{P}^{h} is kk-unstable in a 0-neighborhood, then it is kk-unstable in an ϵ\epsilon-neighborhood for some ϵ>0\epsilon>0;

  • All the concepts can be localized to an open subset UMU\subset M by using Diff(U)\mathrm{Diff}(U) in place of Diff(M)\mathrm{Diff}(M);

  • If Σ𝒫h\Sigma\in\mathcal{P}^{h} is properly embedded, then Σ\Sigma is kk-unstable if and only if its classical Morse index is k\geq k.

We also have the following curvature estimates.

Theorem 2.8 (Curvature estimates for weakly stable free boundary hh-hypersurfaces).

Let 3(n+1)73\leq(n+1)\leq 7 and V,UMV,U\subset M be two relatively open subset so that V¯U\overline{V}\subset U. Let Σ𝒫h\Sigma\in\mathcal{P}^{h} be weakly stable in UU with Area(Σ)C\mathrm{Area}(\Sigma)\leq C, then there exists C1=C1(n,M,U,V,hC3,C)C_{1}=C_{1}(n,M,U,V,\|h\|_{C^{3}},C), such that

|A|(x)C1 for all xΣV.|A|(x)\leq C_{1}\text{\ for all }x\in\Sigma\cap V.

The proof is the same as the free boundary minimal cases [GWZ18]*Theorem 3.2.


Given h𝒮(g),Λ>0h\in\mathcal{S}(g),\Lambda>0 and II\in\mathbb{N}, let

(2.12) 𝒫h(Λ,I):={Σ𝒫h:Area(Σ)Λ,indexw(Σ)I}.\mathcal{P}^{h}(\Lambda,I):=\{\Sigma\in\mathcal{P}^{h}:\mathrm{Area}(\Sigma)\leq\Lambda,\mathrm{index}_{w}(\Sigma)\leq I\}.

The main purpose is to prove this theorem:

Theorem 2.9 (Compactness for free boundary hh-hypersurfaces).

Let (Mn+1,M,g)(M^{n+1},\partial M,g) be a compact Riemannian manifold with boundary of dimension 3(n+1)73\leq(n+1)\leq 7. Assume that {hk}k\{h_{k}\}_{k\in\mathbb{N}} is a sequence of smooth functions in 𝒮(g)\mathcal{S}(g) such that limkhk=h\lim_{k\rightarrow\infty}h_{k}=h_{\infty} in smooth topology, where h𝒮(g)h_{\infty}\in\mathcal{S}(g) or h=0h_{\infty}=0. Let {Σk}k\{\Sigma_{k}\}_{k\in\mathbb{N}} be a sequence of hypersurfaces such that Σk𝒫hk(Λ,I)\Sigma_{k}\in\mathcal{P}^{h_{k}}(\Lambda,I) for some fixed Λ>0\Lambda>0 and II\in\mathbb{N}. Then,

  1. (i)

    up to a subsequence, there exists a smooth, compact, almost embedded free boundary hh_{\infty}-hypersurface Σ\Sigma_{\infty} such that ΣkΣ\Sigma_{k}\rightarrow\Sigma_{\infty} (possibly with integer multiplicity) in the varifold sense, and hence also in the Hausdorff distance by monotonicity formula;

  2. (ii)

    there exists a finite set of points 𝒴Σ\mathcal{Y}\subset\Sigma_{\infty} with (𝒴)I\sharp(\mathcal{Y})\leq I, such that the convergence of ΣkΣ\Sigma_{k}\rightarrow\Sigma_{\infty} is locally smooth and graphical on Σ𝒴\Sigma_{\infty}\setminus\mathcal{Y};

  3. (iii)

    if h𝒮(g)h_{\infty}\in\mathcal{S}(g), then the multiplicity of Σ\Sigma_{\infty} is 1, and Σ𝒫h(Λ,I)\Sigma_{\infty}\in\mathcal{P}^{h_{\infty}}(\Lambda,I);

  4. (iv)

    assuming ΣkΣ\Sigma_{k}\neq\Sigma_{\infty} eventually and hk=h=h𝒮(g)h_{k}=h_{\infty}=h\in\mathcal{S}(g) for all kk and Σk\Sigma_{k} smoothly converges to Σ\Sigma_{\infty}, then 𝒴=\mathcal{Y}=\emptyset, and Σ\Sigma_{\infty} has a non-trivial Jacobi field;

  5. (v)

    if hk=h=h𝒮(g)h_{k}=h_{\infty}=h\in\mathcal{S}(g) and the convergence is not smooth, then 𝒴\mathcal{Y} is not empty and Σ\Sigma_{\infty} has strictly smaller weak Morse index than Σk\Sigma_{k} for all sufficiently large kk;

  6. (vi)

    if h0h_{\infty}\equiv 0 and Σ\Sigma_{\infty} is properly embedded, then the classical Morse index of Σ\Sigma_{\infty} satisfies index(Σ)I\mathrm{index}(\Sigma_{\infty})\leq I (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 Σk\Sigma_{k} converges locally smoothly and graphically to an almost embedded free boundary hh_{\infty}-hypersurface Σ\Sigma_{\infty} (possibly with integer multiplicity) away from at most II points, which we denote by 𝒴\mathcal{Y}. Now we prove that 𝒴\mathcal{Y} are all removable.

Case 1: Suppose pp is in the closure of Σ\Sigma and pMp\notin\partial M.

Then the argument is the same as that in [Zhou19]*Theorem 2.7, Part 1.

Case 2: Suppose that pMp\in\partial M.

Using the boundary removable singularity result Theorem A.1, we see that pp 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 δ2𝒜h\delta^{2}\mathcal{A}^{h} along Σ\Sigma_{\infty}. Recall that the Jacobi fields associated with δ2𝒜h\delta^{2}\mathcal{A}^{h} along a free boundary hh-hypersurface Σ\Sigma satisfy

(2.13) {LΣhφ=0, on Σ,φ𝜼=hM(ν,ν)φ, along Σ,\left\{\begin{aligned} &L^{h}_{\Sigma}\varphi=0,\text{\ \ on $\Sigma$},\\ &\frac{\partial\varphi}{\partial\bm{\eta}}=h^{\partial M}(\nu,\nu)\varphi,\text{\ \  along $\partial\Sigma,$}\end{aligned}\right.

where LΣh=ΔΣ+Ric(ν,ν)+|A|+νhL^{h}_{\Sigma}=\Delta_{\Sigma}+\operatorname{Ric}(\nu,\nu)+|A|+\partial_{\nu}h and 𝜼\bm{\eta} is the outward co-normal of Σ\Sigma.

Recall that (M,M,g)(M,\partial M,g) can be isometrically embedded into a closed Riemannian manifold (M~n+1,g~)(\widetilde{M}^{n+1},\widetilde{g}). Let 𝔛~(M,Σ)\mathfrak{\widetilde{X}}(M,\Sigma) be the space of vector fields X𝔛(M~)X\in\mathfrak{X}(\widetilde{M}) so that X(p)Tp(M)X(p)\in T_{p}(\partial M) for pp in a small neighborhood of Σ\partial\Sigma in M\partial M.

Denote by ϕ:(Σ,Σ)(M,M)\phi_{\infty}:(\Sigma_{\infty},\partial\Sigma_{\infty})\hookrightarrow(M,\partial M) the immersion. Now we fix a relative open set VV of Σ\Sigma_{\infty} so that ϕ|V\phi_{\infty}|_{V} is an embedding.

Now let X𝔛~(M,Σ)X\in\mathfrak{\widetilde{X}}(M,\Sigma_{\infty}) be an extension of the unit normal vector field of ϕ(V)\phi_{\infty}(V). Let Φ(x,t)\Phi(x,t) be the one-parameter family of diffeomorphisms of M~\widetilde{M} associated with XX, so that Φt(x,t)=X(Φ(x,t))\frac{\partial\Phi}{\partial t}(x,t)=X(\Phi(x,t)). Let VδV_{\delta} be the δ\delta-thickening of VV with respect to Φ\Phi so that

Vδ:={Φ(x,t):xV and |t|<δ}.V_{\delta}:=\{\Phi(x,t):x\in V\text{ and }|t|<\delta\}.

Then for sufficiently large kk, there exists φk(x)C(Σ)\varphi_{k}(x)\in C^{\infty}(\Sigma_{\infty}) so that

ϕk(Σk)Vδ={Φ(x,φk(x)):xϕ(V)},\displaystyle\phi_{k}(\Sigma_{k})\cap V_{\delta}=\{\Phi(x,\varphi_{k}(x)):x\in\phi_{\infty}(V)\},

where ϕk:(Σk,Σk)(M,M,g)\phi_{k}:(\Sigma_{k},\partial\Sigma_{k})\hookrightarrow(M,\partial M,g) is the immersion map. In the following, we often omit ϕk\phi_{k} and ϕ\phi_{\infty} for simplicity when there is no ambiguity.

Denote by Φk(x,t)=Φ(x,tφk(x))\Phi^{k}(x,t)=\Phi(x,t\varphi_{k}(x)) and

Vk(t)={Φk(x,t):xV}, Rk(t)={Φk(x,t):xVΣ}.V_{k}(t)=\{\Phi^{k}(x,t):x\in V\},\ \ \ \  R_{k}(t)=\{\Phi^{k}(x,t):x\in V\cap\partial\Sigma_{\infty}\}.

Then we have Vk(0)=VV_{k}(0)=V, Vk(1)=ΣkVδV_{k}(1)=\Sigma_{k}\cap V_{\delta}.

Now consider any vector field Z𝔛~(M,Σ)Z\in\mathfrak{\widetilde{X}}(M,\Sigma_{\infty}) such that |Z|+|Z|1|Z|+|\nabla Z|\leq 1 and Z|M~Vδ=0Z|_{\widetilde{M}\setminus V_{\delta}}=0. Let (Ψt)t(\Psi_{t})_{t} be the associated one-parameter family of diffeomorphisms of M~\widetilde{M}. From the fact that Σk\Sigma_{k} and Σ\Sigma_{\infty} are hh-hypersurfaces, we have

01ddt[Vk(t)divZhZ,νdμVk(t)]𝑑t=0.\int_{0}^{1}\frac{d}{dt}\Big{[}\int_{V_{k}(t)}\mathrm{div}Z-h\langle Z,\nu\rangle\,d\mu_{V_{k}(t)}\Big{]}\,dt=0.

The computation in Appendix B gives that

0=\displaystyle 0= 01ddt[Vk(t)divZhZ,νdμVk(t)]𝑑t\displaystyle\int_{0}^{1}\frac{d}{dt}\Big{[}\int_{V_{k}(t)}\mathrm{div}Z-h\langle Z,\nu\rangle\,d\mu_{V_{k}(t)}\Big{]}\,dt
=\displaystyle= 01[Vk(t)((X),(Z)Ric(X,Z)|A|2X,Z\displaystyle\int_{0}^{1}\Big{[}\int_{V_{k}(t)}\Big{(}\langle\nabla^{\perp}(X^{\perp}),\nabla^{\perp}(Z^{\perp})\rangle-\operatorname{Ric}(X^{\perp},Z^{\perp})-|A|^{2}\langle X^{\perp},Z^{\perp}\rangle
(νh)X,Z)dn+Rk(t)XZ,νMdn1+\displaystyle-(\partial_{\nu}h)\langle X^{\perp},Z^{\perp}\rangle\Big{)}\,d\mathcal{H}^{n}+\int_{R_{k}(t)}\langle\nabla_{X^{\perp}}Z^{\perp},\nu_{\partial M}\rangle\,d\mathcal{H}^{n-1}+
+Vk(t)Ξ~1(X,Z,𝐇)dn+Rk(t)Ξ~2(X,Z,𝐇,𝜼,νM)dn1]dt.\displaystyle+\int_{V_{k}(t)}\widetilde{\Xi}_{1}(X,Z,\mathbf{H})\,d\mathcal{H}^{n}+\int_{R_{k}(t)}\widetilde{\Xi}_{2}(X,Z,\mathbf{H},\bm{\eta},\nu_{\partial M})\,d\mathcal{H}^{n-1}\Big{]}\,dt.

Here we used the assumption of Z|M~Vδ=0Z|_{\widetilde{M}\setminus V_{\delta}}=0 in the last equality. By pulling everything back to Σ\Sigma_{\infty} and letting Z|Σ=ζνZ|_{\Sigma_{\infty}}=\zeta\nu, we obtain

0=\displaystyle 0= 01[V(φk,ζRic(ν,ν)φkζ|AΣ|2φkζ(νh)φkζ)dn\displaystyle\int_{0}^{1}\Big{[}\int_{V}\Big{(}\langle\nabla\varphi_{k},\nabla\zeta\rangle-\operatorname{Ric}(\nu,\nu)\varphi_{k}\zeta-|A^{\Sigma_{\infty}}|^{2}\varphi_{k}\zeta-(\partial_{\nu}h)\varphi_{k}\zeta\Big{)}\,d\mathcal{H}^{n}
VΣhM(ν,ν)φkζ𝑑n1+VW~k(t)(φk,ζ)𝑑n+\displaystyle-\int_{V\cap\partial\Sigma_{\infty}}h^{\partial M}(\nu,\nu)\varphi_{k}\zeta\,d\mathcal{H}^{n-1}+\int_{V}\widetilde{W}_{k}(t)(\varphi_{k},\zeta)\,d\mathcal{H}^{n}+
+VΣw~k(t)(φk,ζ)dn1]dt,\displaystyle+\int_{V\cap\partial\Sigma_{\infty}}\widetilde{w}_{k}(t)(\varphi_{k},\zeta)\,d\mathcal{H}^{n-1}\Big{]}\,dt,

where

W~k(t)(φk,ζ)ϵk(Σ,M~)|φk|,w~k(t)(φk,ζ)ϵk(Σ,M~)|φk|.\displaystyle\widetilde{W}_{k}(t)(\varphi_{k},\zeta)\leq\epsilon_{k}(\Sigma_{\infty},\widetilde{M})|\varphi_{k}|,\ \ \ \widetilde{w}_{k}(t)(\varphi_{k},\zeta)\leq\epsilon_{k}(\Sigma_{\infty},\widetilde{M})|\varphi_{k}|.

Here ϵk0\epsilon_{k}\rightarrow 0 uniformly as kk\rightarrow\infty and we used that |Z|+|Z|1|Z|+|\nabla Z|\leq 1. Now letting Wk=01W~k(t)𝑑tW_{k}=\int_{0}^{1}\widetilde{W}_{k}(t)dt and wk=01w~k(t)𝑑tw_{k}=\int_{0}^{1}\widetilde{w}_{k}(t)dt, by Fubini theorem, we have

(2.14) 0=\displaystyle 0= Σ(φk,ζRic(ν,ν)φkζ|AΣ|2φkζ(νh)φkζ)𝑑n\displaystyle\int_{\Sigma_{\infty}}\Big{(}\langle\nabla\varphi_{k},\nabla\zeta\rangle-\operatorname{Ric}(\nu,\nu)\varphi_{k}\zeta-|A^{\Sigma_{\infty}}|^{2}\varphi_{k}\zeta-(\partial_{\nu}h)\varphi_{k}\zeta\Big{)}\,d\mathcal{H}^{n}
ΣhM(ν,ν)φkζ𝑑n1+ΣWk(φk,ζ)𝑑n+Σwk(φk,ζ)𝑑n1.\displaystyle-\int_{\partial\Sigma_{\infty}}h^{\partial M}(\nu,\nu)\varphi_{k}\zeta\,d\mathcal{H}^{n-1}+\int_{\Sigma_{\infty}}W_{k}(\varphi_{k},\zeta)\,d\mathcal{H}^{n}+\int_{\partial\Sigma_{\infty}}w_{k}(\varphi_{k},\zeta)\,d\mathcal{H}^{n-1}.

Here we used the fact that ζ=0\zeta=0 on ΣV\Sigma_{\infty}\setminus V.

Now we take finitely many relatively open subsets {Vα}\{V_{\alpha}\} of Σ\Sigma_{\infty} so that ϕ|Vα\phi_{\infty}|_{V_{\alpha}} is an embedding for each jj and

αVα=Σ.\bigcup_{\alpha}V_{\alpha}=\Sigma_{\infty}.

Then we can also find finitely many non-negative cut-off functions ζαC(Σ)\zeta_{\alpha}\in C^{\infty}(\Sigma_{\infty}) so that ζα=0\zeta_{\alpha}=0 on ΣVα\Sigma_{\infty}\setminus V_{\alpha} and ζα=1\sum\zeta_{\alpha}=1. Then by (2.14), for each α\alpha and each uC(Σ)u\in C^{\infty}(\Sigma_{\infty}),

0=\displaystyle 0= Σ(φk,(ζαu)Ric(ν,ν)φkζαu|AΣ|2φkζαu(νh)φkζαu)𝑑n\displaystyle\int_{\Sigma_{\infty}}\Big{(}\langle\nabla\varphi_{k},\nabla(\zeta_{\alpha}u)\rangle-\operatorname{Ric}(\nu,\nu)\varphi_{k}\zeta_{\alpha}u-|A^{\Sigma_{\infty}}|^{2}\varphi_{k}\zeta_{\alpha}u-(\partial_{\nu}h)\varphi_{k}\zeta_{\alpha}u\Big{)}\,d\mathcal{H}^{n}
ΣhM(ν,ν)φkζαu𝑑n1+ΣWk(φk,ζαu)𝑑n+Σwk(φk,ζαu)𝑑n1.\displaystyle-\int_{\partial\Sigma_{\infty}}h^{\partial M}(\nu,\nu)\varphi_{k}\zeta_{\alpha}u\,d\mathcal{H}^{n-1}+\int_{\Sigma_{\infty}}W_{k}(\varphi_{k},\zeta_{\alpha}u)\,d\mathcal{H}^{n}+\int_{\partial\Sigma_{\infty}}w_{k}(\varphi_{k},\zeta_{\alpha}u)\,d\mathcal{H}^{n-1}.

Adding all of them together, we have

0=\displaystyle 0= Σ(φk,uRic(ν,ν)φku|AΣ|2φku(νh)φku)𝑑n\displaystyle\int_{\Sigma_{\infty}}\Big{(}\langle\nabla\varphi_{k},\nabla u\rangle-\operatorname{Ric}(\nu,\nu)\varphi_{k}u-|A^{\Sigma_{\infty}}|^{2}\varphi_{k}u-(\partial_{\nu}h)\varphi_{k}u\Big{)}\,d\mathcal{H}^{n}
ΣhM(ν,ν)φku𝑑n1+α[ΣWk(φk,ζαu)𝑑n+Σwk(φk,ζαu)𝑑n1].\displaystyle-\int_{\partial\Sigma_{\infty}}h^{\partial M}(\nu,\nu)\varphi_{k}u\,d\mathcal{H}^{n-1}+\sum_{\alpha}\Big{[}\int_{\Sigma_{\infty}}W_{k}(\varphi_{k},\zeta_{\alpha}u)\,d\mathcal{H}^{n}+\int_{\partial\Sigma_{\infty}}w_{k}(\varphi_{k},\zeta_{\alpha}u)\,d\mathcal{H}^{n-1}\Big{]}.

Let φ~k=φk/φkL2(Σ)\widetilde{\varphi}_{k}=\varphi_{k}/\|\varphi_{k}\|_{L^{2}(\Sigma_{\infty})}. Then the standard PDE theory implies that φ~k\widetilde{\varphi}_{k} converges smoothly to a nontrivial φC(Σ)\varphi\in C^{\infty}(\Sigma_{\infty}) 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 Σ\Sigma_{\infty} has weak index II. If Σk\Sigma_{k} does not smoothly converge to Σ\Sigma_{\infty}, then there exists a finite set 𝒴Σ\mathcal{Y}\subset\Sigma_{\infty} so that for any r>0r>0, Br(p)ΣkB_{r}(p)\cap\Sigma_{k} does not smoothly converge to Br(p)ΣB_{r}(p)\cap\Sigma_{\infty} and Σk\Sigma_{k} smoothly converges to Σ\Sigma_{\infty} outside Br(𝒴):=p𝒴Br(p)B_{r}(\mathcal{Y}):=\cup_{p\in\mathcal{Y}}B_{r}(p). We now prove that indexw(Σk)I+1\mathrm{index}_{w}(\Sigma_{k})\geq I+1 for large kk.

By the Definition 2.6 of weak index, there exists 0<c0<10<c_{0}<1 and smooth family {Fv}vB¯IDiff(M)\{F_{v}\}_{v\in\overline{B}^{I}}\subset\mathrm{Diff}(M) with F0=Id,Fv=Fv1F_{0}=\mathrm{Id},F_{-v}=F_{v}^{-1} for all vB¯Iv\in\overline{B}^{I} such that for any Ω𝐁¯2ϵ𝐅(Ω)\Omega^{\prime}\in\overline{\mathbf{B}}^{\mathbf{F}}_{2\epsilon}(\Omega_{\infty}) the smooth function:

𝒜Ωh:B¯I[0,+),𝒜Ωh(v)=𝒜h(Fv(Ω))\mathcal{A}_{\Omega^{\prime}}^{h}:\overline{B}^{I}\rightarrow[0,+\infty),\ \ \ \ \mathcal{A}^{h}_{\Omega^{\prime}}(v)=\mathcal{A}^{h}(F_{v}(\Omega^{\prime}))

satisfies

  • 𝒜Ωh\mathcal{A}^{h}_{\Omega^{\prime}} has a unique maximum at m(Ω)Bc0/10I(0)m(\Omega^{\prime})\in B_{c_{0}/\sqrt{10}}^{I}(0);

  • 1c0IdD2𝒜Ωh(u)c0Id-\frac{1}{c_{0}}\mathrm{Id}\leq D^{2}\mathcal{A}^{h}_{\Omega^{\prime}}(u)\leq-c_{0}\mathrm{Id} for all uB¯Iu\in\overline{B}^{I}.

Denote by Xj(x)=ddt|t=0Ftej(x)X_{j}(x)=\frac{d}{dt}|_{t=0}F_{te_{j}}(x). Then Xj𝔛(M,M)X_{j}\in\mathfrak{X}(M,\partial M).

We can shrink rr so that {Xj|ΣBr(𝒴)}j=1I\{X_{j}|_{\Sigma_{\infty}\setminus B_{r}(\mathcal{Y})}\}_{j=1}^{I} is still linearly independent. Let ξr\xi_{r} be a cut-off function satisfying 0ξr10\leq\xi_{r}\leq 1 and ξ|Br(𝒴)=0\xi|_{B_{r}(\mathcal{Y})}=0 and Σ|ξr|20\int_{\Sigma_{\infty}}|\nabla\xi_{r}|^{2}\rightarrow 0 and ξr1\xi_{r}\rightarrow 1 as r0r\rightarrow 0. By Appendix C, we can shrink rr so that there exists ϵ>0\epsilon>0 satisfying

(2.15) (2/c0)jaj2<IIΣ(jajξrXj,jajξrXj)<(c0/2)jaj2, 1jI,-(2/c_{0})\sum_{j}a_{j}^{2}<\mathrm{II}_{\Sigma_{\infty}}(\sum_{j}a_{j}\xi_{r}X_{j}^{\perp},\sum_{j}a_{j}\xi_{r}X_{j}^{\perp})<-(c_{0}/2)\sum_{j}a_{j}^{2},\ \ 1\leq j\leq I,

for jaj20\sum_{j}a_{j}^{2}\neq 0. Recall that ΣkBr(𝒴)\Sigma_{k}\setminus B_{r}(\mathcal{Y}) smoothly converges to ΣBr(𝒴)\Sigma_{\infty}\setminus B_{r}(\mathcal{Y}). Hence for sufficiently large kk,

(2/c0)jaj2<IIΣk(jajξrXj,jajξrXj)<(c0/2)jaj2, 1jI.-(2/c_{0})\sum_{j}a_{j}^{2}<\mathrm{II}_{\Sigma_{k}}(\sum_{j}a_{j}\xi_{r}X_{j}^{\perp},\sum_{j}a_{j}\xi_{r}X_{j}^{\perp})<-(c_{0}/2)\sum_{j}a_{j}^{2},\ \ 1\leq j\leq I.

Here we used the fact that ξr=0\xi_{r}=0 on Br(𝒴)B_{r}(\mathcal{Y}).

By assumptions, Br(p)ΣkB_{r}(p)\cap\Sigma_{k} does not smoothly converge to Br(p)ΣB_{r}(p)\cap\Sigma_{\infty} for p𝒴p\in\mathcal{Y}. Hence Σk\Sigma_{k} is not weak stable in Br(p)B_{r}(p). This implies that for each kk, there exists {Fv}v[1,1]Diff(M)\{F^{\prime}_{v}\}_{v\in[-1,1]}\subset\mathrm{Diff}(M) such that

Fv|MBr(p)=Id, F^{\prime}_{v}|_{M\setminus B_{r}(p)}=\mathrm{Id}, 

and the smooth function 𝒜h(Ω)\mathcal{A}^{h}(\Omega^{\prime}) satisfying, for some c0>0c_{0}^{\prime}>0,

(2.16) 1c0IdD2𝒜Ωh(u)c0Id, -\frac{1}{c_{0}^{\prime}}\mathrm{Id}\leq D^{2}\mathcal{A}^{h}_{\Omega^{\prime}}(u)\leq-c_{0}^{\prime}\mathrm{Id}, 

for all u(1,1)u\in(-1,1) and Ω𝐁¯2ϵ𝐅(Ωk)\Omega^{\prime}\in\overline{\mathbf{B}}^{\mathbf{F}}_{2\epsilon}(\Omega_{k}).

For any X𝔛(M,M)X\in\mathfrak{X}(M,\partial M), denote by ΦtX\Phi^{X}_{t} the flow of XX. Now we define F~:B¯I+1Diff(M)\widetilde{F}:\overline{B}^{I+1}\rightarrow\mathrm{Diff}(M) by

 F~v=Fv0Φ1j=1IvjξrXj. \widetilde{F}_{v}=F_{v_{0}}^{\prime}\circ\Phi_{1}^{\sum_{j=1}^{I}v_{j}\xi_{r}X_{j}}.

Here v=(v0,,vI)v=(v_{0},\cdots,v_{I}). Then for any vB¯I+1v\in\overline{B}^{I+1} we have

d2dt2|t=0𝒜h(F~tv(Ωk))\displaystyle\frac{d^{2}}{dt^{2}}\Big{|}_{t=0}\mathcal{A}^{h}(\widetilde{F}_{tv}(\Omega_{k})) =d2dt2|t=0𝒜h(Ftv0(Ωk))+d2dt2|t=0𝒜h(F~t(0,v1,,vI)(Ωk))\displaystyle=\frac{d^{2}}{dt^{2}}\Big{|}_{t=0}\mathcal{A}^{h}(F^{\prime}_{tv_{0}}(\Omega_{k}))+\frac{d^{2}}{dt^{2}}\Big{|}_{t=0}\mathcal{A}^{h}(\widetilde{F}_{t(0,v_{1},\cdots,v_{I})}(\Omega_{k}))
=d2dt2|t=0𝒜h(Ftv0(Ωk))+IIΣk(j=1IvjξrXj,j=1IvjξrXj).\displaystyle=\frac{d^{2}}{dt^{2}}\Big{|}_{t=0}\mathcal{A}^{h}(F^{\prime}_{tv_{0}}(\Omega_{k}))+\mathrm{II}_{\Sigma_{k}}(\sum_{j=1}^{I}v_{j}\xi_{r}X_{j}^{\perp},\sum_{j=1}^{I}v_{j}\xi_{r}X_{j}^{\perp}).

Together with (2.15) and (2.16), we conclude that index(Σk)I+1\mathrm{index}(\Sigma_{k})\geq I+1. This finishes the proof of (v).

There is also a theorem analogous to the above one in the setting of changing ambient metrics on (M,M)(M,\partial M). The proof proceeds the same way when one realizes that the constant C1C_{1} in Theorem 2.5 depends only on the gC4\|g\|_{C^{4}} when gg is allowed to change.

Theorem 2.10.

Let (Mn+1,M)(M^{n+1},\partial M) be a closed manifold of dimension 3(n+1)73\leq(n+1)\leq 7, and {gk}k\{g_{k}\}_{k\in\mathbb{N}} be a sequence of metrics on (M,M)(M,\partial M) that converges smoothly to some limit metric gg. Let {hk}k\{h_{k}\}_{k\in\mathbb{N}} be a sequence of smooth functions with hk𝒮(gk)h_{k}\in\mathcal{S}(g_{k}) that converges smoothly to some limit hC(M)h_{\infty}\in C^{\infty}(M), where h𝒮(g)h_{\infty}\in\mathcal{S}(g) or h=0h_{\infty}=0. Let {Σk}k\{\Sigma_{k}\}_{k\in\mathbb{N}} be a sequence of hypersurfaces with Σk𝒫hk(Λ,I;gk)\Sigma_{k}\in\mathcal{P}^{h_{k}}(\Lambda,I;g_{k}) for some fixed λ>0\lambda>0 and II\in\mathbb{N}. Then there exists a smooth, compact, almost embedded free boundary hh_{\infty}-hypersurface Σ\Sigma_{\infty}, such that Theorem 2.9(i)(ii)(iii) are satisfied.

2.4. Generic existence of good pairs

Let (M~n+1,g)(\widetilde{M}^{n+1},g) be a closed Riemannian manifold of dimension 3(n+1)73\leq(n+1)\leq 7. Let MM be a compact domain of M~\widetilde{M} with smooth boundary. A pair (g~,h~)(\widetilde{g},\widetilde{h}) consisting of a Riemannian metric g~\widetilde{g} and smooth function h~C(M~)\widetilde{h}\in C^{\infty}(\widetilde{M}) is called a good pair related to MM, if

  1. (1)

    h~\widetilde{h} is a Morse function;

  2. (2)

    the zero set {h~=0}\{\widetilde{h}=0\} is a smoothly embedded closed hypersurface in M~\widetilde{M}, which is transverse to M\partial M and has mean curvature H~\widetilde{H} vanishing to at most finite order;

  3. (3)

    {xM:h~|M=±HM}\{x\in\partial M:\widetilde{h}|_{\partial M}=\pm H_{\partial M}\} has dimension less than or equal to (n1)(n-1);

  4. (4)

    g~\widetilde{g} is bumpy for 𝒜h\mathcal{A}^{h}, i.e., every almost embedded prescribed mean curvature hypersurface in M~\widetilde{M} with free boundary on M\partial M is non-degenerate.

Clearly, if (g~,h~)(\widetilde{g},\widetilde{h}) is a good pair related to MM, then h~|M𝒮(g~|M)\widetilde{h}|_{M}\in\mathcal{S}(\widetilde{g}|_{M}); (see Section 2.2). In this subsection, we are going to prove the generic existence of good pairs related to MM.

Denote by 𝒮~0\widetilde{\mathcal{S}}_{0} the set of smooth functions h~C(M~)\widetilde{h}\in C^{\infty}(\widetilde{M}) such that

  • h~\widetilde{h} is a Morse function;

  • {h~=0}\{\widetilde{h}=0\} is an embedded closed hypersurface in M~\widetilde{M}, which is transverse to M\partial M.

𝒮~0\widetilde{\mathcal{S}}_{0} is open and dense in C(M~)C^{\infty}(\widetilde{M}), 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 h~𝒮~0\widetilde{h}\in\widetilde{\mathcal{S}}_{0}, the set of Riemannian metrics g~\widetilde{g} on M~\widetilde{M} with (g~,h~)(\widetilde{g},\widetilde{h}) being a good pair related to MM is generic in the Baire sense.

Proof.

Firstly, we prove that (3) is generic. Let 𝚪~0\widetilde{\bm{\Gamma}}_{0} be the set of metrics g~\widetilde{g} so that {xM:h~|M=±H~M}\{x\in\partial M:\widetilde{h}|_{\partial M}=\pm\widetilde{H}_{\partial M}\} has dimension less than or equal to (n1)(n-1). Denote by 𝚪~1\widetilde{\bm{\Gamma}}_{1} the set of metrics on M~\widetilde{M} so that

{xM:HM(x)±h(x)=0}{ M(HM±h|M)0}.\{x\in\partial M:H_{\partial M}(x)\pm h(x)=0\}\subset\{ \nabla^{\partial M}(H_{\partial M}\pm h|_{\partial M})\neq 0\}.

Clearly, 𝚪~1𝚪~0\widetilde{\bm{\Gamma}}_{1}\subset\widetilde{\bm{\Gamma}}_{0} and 𝚪~1\widetilde{\bm{\Gamma}}_{1} is open. We claim that 𝚪~1\widetilde{\bm{\Gamma}}_{1} is also dense, which would imply that (3) is generic.

Indeed, given any g~\widetilde{g} and an open neighborhood 𝒰\mathcal{U} of g~\widetilde{g} in the space of metrics on M~\widetilde{M}, we can construct g~𝒰𝚪~1\widetilde{g}^{\prime}\in\mathcal{U}\cap\widetilde{\bm{\Gamma}}_{1} as follows. Let rr be the signed distance function to M\partial M under the metric g~\widetilde{g}. Let δ>0\delta>0 be small enough so that M\partial M has a tubular neighborhood WW and the exponential map from M×(δ,δ)\partial M\times(-\delta,\delta) to WW is a diffeomorphism. Fix a cut-off function η:(1,1)[0,1]\eta:(-1,1)\rightarrow[0,1] satisfying

η(z)=1 for z[δ/2,δ/2] and η(z)=0 for |z|δ.\text{$\eta(z)=1$ for $z\in[-\delta/2,\delta/2]$ and $\eta(z)=0$ for $|z|\geq\delta$}.

Now let φC(M)\varphi\in C^{\infty}(\partial M) be small enough so that

(2.17) {xM:nφ(x)+HM(x)±h(x)=0}{ M(nφ+HM±h|M)0}.\{x\in\partial M:n\varphi(x)+H_{\partial M}(x)\pm h(x)=0\}\subset\{ \nabla^{\partial M}(n\varphi+H_{\partial M}\pm h|_{\partial M})\neq 0\}.

This can be satisfied because the set of smooth Morse functions uu on M\partial M with empty singular set {u=u=0:xM}=\{u=\nabla u=0:x\in\partial M\}=\emptyset is open and dense.

Define ϕ(x)=r(x)η(r(x))φ(π(x))\phi(x)=-r(x)\eta(r(x))\varphi(\pi(x)), where π\pi is the projection of WW to its closest point in M\partial M. Let g~=e2ϕg~\widetilde{g}^{\prime}=e^{2\phi}\widetilde{g}. By taking φ\varphi in a small enough neighborhood of 0, we have that g~𝒰\widetilde{g}^{\prime}\in\mathcal{U}. Then the second fundamental form of M\partial M with respect to g~\widetilde{g}^{\prime} is given by (see \citelist[Bes87]*Section 1.163[IMN17]*Proposition 2.3)

AM,g~=AM,g~+g~(ϕ). A_{\partial M,\widetilde{g}^{\prime}}=A_{\partial M,\widetilde{g}}+\widetilde{g}\cdot(\nabla\phi)^{\perp}. 

Recall that g~|M=g~|M\widetilde{g}|_{\partial M}=\widetilde{g}^{\prime}|_{\partial M}. Hence the mean curvature with respect to g~\widetilde{g}^{\prime} is

HM,g~= HM,g~+nφ.H_{\partial M,\widetilde{g}^{\prime}}= H_{\partial M,\widetilde{g}}+n\varphi.

By (2.17), g~𝚪~1\widetilde{g}^{\prime}\in\widetilde{\bm{\Gamma}}_{1}. Thus 𝚪~1\widetilde{\bm{\Gamma}}_{1} is dense.

In the next, we prove that the set of metrics g~\widetilde{g} under which {h~=0}\{\widetilde{h}=0\} 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 𝒰\mathcal{U} be an open neighborhood of g~\widetilde{g}. Then from the above argument, for any φC({h~=0})\varphi\in C^{\infty}(\{\widetilde{h}=0\}) small enough, we can find a metric g~𝒰\widetilde{g}^{\prime}\in\mathcal{U} so that g~|{h~=0}=g~|{h~=0}\widetilde{g}^{\prime}|_{\{\widetilde{h}=0\}}=\widetilde{g}|_{\{\widetilde{h}=0\}} and  {h~=0}\{\widetilde{h}=0\} has mean curvature H0+φH_{0}+\varphi, where H0H_{0} is the mean curvature of {h~=0}\{\widetilde{h}=0\} under g~\widetilde{g}. We can choose φ\varphi so that H0+φH_{0}+\varphi is a Morse function on {h~=0}\{\widetilde{h}=0\}, hence H0+φH_{0}+\varphi 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 hh-hypersurfaces is countable for a good pair related to MM.

Given h𝒮(g),Λ>0h\in\mathcal{S}(g),\Lambda>0 and II\in\mathbb{N}, recall that 𝒫h(Λ,I)\mathcal{P}^{h}(\Lambda,I) the set of Σ𝒫h\Sigma\in\mathcal{P}^{h} satisfying Area(Σ)Λ\mathrm{Area}(\Sigma)\leq\Lambda and its weak Morse index is bounded by II from above.

Lemma 2.12.

Let (g~,h~)(\widetilde{g},\widetilde{h}) be a good pair related to MM. Denote by h=h~|Mh=\widetilde{h}|_{M} and g=g~|Mg=\widetilde{g}|_{M}. Then 𝒫h(Λ,I)\mathcal{P}^{h}(\Lambda,I) is countable, and hence I0𝒫h(Λ,I)\bigcup_{I\geq 0}\mathcal{P}^{h}(\Lambda,I) is countable.

Proof.

We prove it by an inductive method. By Theorem 2.9(v)(iii), there are only finitely many elements in 𝒫h(Λ,0)\mathcal{P}^{h}(\Lambda,0). Hence 𝒫h(Λ,0)\mathcal{P}^{h}(\Lambda,0) contains only finitely many hypersurfaces.

Assuming that 𝒫h(Λ,I)\mathcal{P}^{h}(\Lambda,I) is countable for some I0I\geq 0, we are going to prove that 𝒫h(Λ,I+1)\mathcal{P}^{h}(\Lambda,I+1) is also countable. Using Theorem 2.9(v)(iii) again, we know that for any r>0r>0, 𝒫h(Λ,I+1)Σ𝒫h(Λ,I)𝐁¯r𝐅(Σ)\mathcal{P}^{h}(\Lambda,I+1)\setminus\bigcup_{\Sigma\in\mathcal{P}^{h}(\Lambda,I)}\overline{\mathbf{B}}_{r}^{\mathbf{F}}(\Sigma) contains only finitely many elements. Therefore, 𝒫h(Λ,I+1)\mathcal{P}^{h}(\Lambda,I+1) is countable.

By induction, we have proved that 𝒫h(Λ,I)\mathcal{P}^{h}(\Lambda,I) is countable for all I0I\geq 0. ∎

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 M~\widetilde{M} and a two-sided, closed, embedded hypersurface NM~N\subset\widetilde{M}. (Later we will embed MM into M~\widetilde{M} and let N=MN=\partial M.) Let 𝚪~\widetilde{\mathbf{\Gamma}} be the set of smooth Riemannian metrics on M~\widetilde{M}. Let ~\mathcal{\widetilde{M}} be the space of all pairs (γ~,Σ~)(\widetilde{\gamma},\widetilde{\Sigma}) such that γ~𝚪~\widetilde{\gamma}\in\widetilde{\mathbf{\Gamma}} and (Σ~,Σ~)(\widetilde{\Sigma},\partial\widetilde{\Sigma}) is an embedded minimal hypersurface in (M~,γ~)(\widetilde{M},\widetilde{\gamma}) with (possibly empty) free boundary Σ~N\partial\widetilde{\Sigma}\subset N. Recall that the projection Π~:~𝚪~\widetilde{\Pi}:\mathcal{\widetilde{M}}\rightarrow\widetilde{\mathbf{\Gamma}} is defined as

Π~(γ~,Σ~)=γ~.\widetilde{\Pi}(\widetilde{\gamma},\widetilde{\Sigma})=\widetilde{\gamma}.

Denote by ~reg\mathcal{\widetilde{M}}_{reg} the set of (γ~,Σ~)~(\widetilde{\gamma},\widetilde{\Sigma})\in\mathcal{\widetilde{M}} such that Σ~\widetilde{\Sigma} 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 ~\mathcal{\widetilde{M}} is second countable, we know that ~reg\mathcal{\widetilde{M}}_{reg} is a countable union of open sets UjU_{j} such that Π~\widetilde{\Pi} maps each UjU_{j} homeomorphically onto an open subset of 𝚪~\widetilde{\mathbf{\Gamma}}.

Denote by ~0\mathcal{\widetilde{M}}_{0} the set of (γ~,Σ~)~reg(\widetilde{\gamma},\widetilde{\Sigma})\in\mathcal{\widetilde{M}}_{reg} such that Σ~\widetilde{\Sigma} is transverse to NN.

Lemma 2.13.

Π~(~~0)\widetilde{\Pi}(\mathcal{\widetilde{M}}\setminus\mathcal{\widetilde{M}}_{0}) is meager in 𝚪~\widetilde{\mathbf{\Gamma}}. As a corollary, for generic metrics on M~\widetilde{M}, each embedded minimal hypersurface with free boundary on NN is transverse to NN.

Proof.

The proof here is similar to [Whi19]*Corollary 5. Recall that ~reg\mathcal{\widetilde{M}}_{reg} is a countable union of open sets UjU_{j} so that Π~|Uj\widetilde{\Pi}|_{U_{j}} is a homeomorphism. Now

Π~(~~0)j=1Π((~~0) Uj)Π~(~~reg).\widetilde{\Pi}(\mathcal{\widetilde{M}}\setminus\mathcal{\widetilde{M}}_{0})\subset\bigcup_{j=1}^{\infty}\Pi((\mathcal{\widetilde{M}}\setminus\mathcal{\widetilde{M}}_{0})\cap U_{j})\cup\widetilde{\Pi}(\widetilde{\mathcal{M}}\setminus\widetilde{\mathcal{M}}_{reg}).

Now for any (γ~,Σ~)(~~0)Uj(\widetilde{\gamma},\widetilde{\Sigma})\in(\widetilde{\mathcal{M}}\setminus\widetilde{\mathcal{M}}_{0})\cap U_{j}, we can always take a smooth vector field X𝔛(M~)X\in\mathfrak{X}(\widetilde{M}) so that X|N=f𝐧X|_{N}=f\mathbf{n} satisfying, (here 𝐧\mathbf{n} is the unit normal vector field of NN),

  • X=0X=0 in a neighborhood UΣ~U_{\partial\widetilde{\Sigma}} of Σ~\partial\widetilde{\Sigma} in M~\widetilde{M};

  • For any xInt(Σ~)Nx\in\mathrm{Int}(\widetilde{\Sigma})\cap N, |f|=1|f|=1 and f=0\nabla f=0; (e.g. letting f1f\equiv 1 in a neighborhood of Int(Σ~)N\mathrm{Int}(\widetilde{\Sigma})\cap N).

Denote by (Ft)1<t<1(F_{t})_{-1<t<1} the 1-parameter family of diffeomorphisms of M~\widetilde{M} associated with XX.

Claim 1.

There exists δ>0\delta>0 so that for all 0<t<δ0<t<\delta, Σ~\widetilde{\Sigma} is transverse to Ft(N)F_{t}(N).

Proof.

By the choice of ff, we can take an open set VNV\subset N containing Int(Σ~)N\mathrm{Int}(\widetilde{\Sigma})\cap N such that X(x)0X(x)\neq 0 for any xVx\in V. By taking δ0>0\delta_{0}>0 small, denote by UU the set of

 {Ft(x):xV,t(δ0,δ0)}, \{F_{t}(x):x\in V,t\in(-\delta_{0},\delta_{0})\},

which is diffeomorphic to V×(δ0,δ0)V\times(-\delta_{0},\delta_{0}) by the map F(x,t):=Ft(x)F(x,t):=F_{t}(x). Define a continuous function Z:UZ:U\rightarrow\mathbb{R} by

Z(Ft(x))=d2(Ft(x),Σ~)+1𝐧Ft(x),d2.Z(F_{t}(x))=d^{2}(F_{t}(x),\widetilde{\Sigma})+1-\langle\mathbf{n}_{F_{t}(x)},\nabla d\rangle^{2}.

Here 𝐧Ft(x)\mathbf{n}_{F_{t}(x)} is the unit normal vector field of Ft(N)F_{t}(N), and d()d(\cdot) is the signed distance function to Σ~\widetilde{\Sigma}. It follows that Z(Ft(x))0Z(F_{t}(x))\geq 0. Moreover, if Z(F0(x0))=0Z(F_{0}(x_{0}))=0, then ZZ is smooth in a small neighborhood of (x0,0)(x_{0},0) and 2Z|F0(x0)0\nabla^{2}Z\big{|}_{F_{0}(x_{0})}\geq 0. Let {xj}j=1n\{x^{j}\}_{j=1}^{n} be a normal coordinate system of NN at F0(x0)F_{0}(x_{0}), and write ej=/xje_{j}=\partial/\partial x^{j}. By a direct computation, we have

 Z|F0(x0)=0, 2Z|F0(x0)(ei,t)=0, \nabla Z\big{|}_{F_{0}(x_{0})}=0,\ \ \  \nabla^{2}Z\big{|}_{F_{0}(x_{0})}(e_{i},\partial_{t})=0,

where t=dF(t)\partial_{t}=dF(\frac{\partial}{\partial t}). Moreover,

2Z(ei,ej)|F0(x0)\displaystyle\nabla^{2}Z(e_{i},e_{j})\Big{|}_{F_{0}(x_{0})} =2eiej𝐧Ft(x),d|F0(x0)\displaystyle=-2\nabla_{e_{i}}\nabla_{e_{j}}\langle\mathbf{n}_{F_{t}(x)},\nabla d\rangle\Big{|}_{F_{0}(x_{0})}
=2ei𝐧eid,ej𝐧ejd,\displaystyle=2\langle\nabla_{e_{i}}\mathbf{n}-\nabla_{e_{i}}\nabla d,\nabla_{e_{j}}\mathbf{n}-\nabla_{e_{j}}\nabla d\rangle,

and

d2dt2|F0(x0)Z(x,t)\displaystyle\ \frac{d^{2}}{dt^{2}}\Big{|}_{F_{0}(x_{0})}Z(x,t)
=\displaystyle= 2[ddt|(x,t)=(x0,0)d(Ft(x),Σ~)]22𝐧F0(x0),dd2dt2|(x,t)=(x0,0)𝐧Ft(x),d\displaystyle\ 2\Big{[}\frac{d}{dt}\Big{|}_{(x,t)=(x_{0},0)}d(F_{t}(x),\widetilde{\Sigma})\Big{]}^{2}-2\langle\mathbf{n}_{F_{0}(x_{0})},\nabla d\rangle\cdot\frac{d^{2}}{dt^{2}}\Big{|}_{(x,t)=(x_{0},0)}\langle\mathbf{n}_{F_{t}(x)},\nabla d\rangle
=\displaystyle= 2f2=2. UNKNOWN\displaystyle\ 2f^{2}=2.  

Therefore, we conclude that in a small neighborhood of (x0,0)(x_{0},0),

(2.18) Z(x,t)t2/2.Z(x,t)\geq t^{2}/2.

Now we assume on the contrary that there exists a sequence of positive ti0t_{i}\rightarrow 0 so that Σ~\widetilde{\Sigma} is not transverse to Ft(N)F_{t}(N). Then there exists (xi,ti)(x_{i},t_{i}) so that Fti(xi)Σ~F_{t_{i}}(x_{i})\in\widetilde{\Sigma} and Fti(N)F_{t_{i}}(N) is tangent to Σ~\widetilde{\Sigma}. By the definition of ZZ, we have Z(xi,ti)=0Z(x_{i},t_{i})=0. Assume that limixi=x0\lim_{i\rightarrow\infty}x_{i}=x_{0}, then Z(x0,0)=0Z(x_{0},0)=0. However, by (2.18), for sufficiently large ii,

 Z(xi,ti)ti2/2. Z(x_{i},t_{i})\geq t_{i}^{2}/2.

This leads to a contradiction and hence Claim 1 is proved. ∎

Claim 1 gives that for t>0t>0 small enough, Ft1(Σ~)F_{t}^{-1}(\widetilde{\Sigma}) is transverse to NN, hence (Ftγ~,Ft1Σ~)~0(F_{t}^{*}\widetilde{\gamma},F_{t}^{-1}\widetilde{\Sigma})\in\widetilde{\mathcal{M}}_{0}. Note that ~0\widetilde{\mathcal{M}}_{0} is open. This is to say that (~~0)Uj(\widetilde{\mathcal{M}}\setminus\widetilde{\mathcal{M}}_{0})\cap U_{j} is nowhere dense. Because Π~|Uj\widetilde{\Pi}|_{U_{j}} is a homeomorphism, Π~(~~0)Uj\widetilde{\Pi}(\widetilde{\mathcal{M}}\setminus\widetilde{\mathcal{M}}_{0})\cap U_{j} is nowhere dense. Hence Π~(~~0)\widetilde{\Pi}(\mathcal{\widetilde{M}}\setminus\mathcal{\widetilde{M}}_{0}) is meager in 𝚪~\widetilde{\mathbf{\Gamma}}. ∎

Now we are ready to prove Theorem 1.3. Let (Mn+1,M)(M^{n+1},\partial M) be a smooth compact manifold with boundary and 3(n+1)73\leq(n+1)\leq 7. Denote by 𝚪\mathbf{\Gamma} the space of smooth Riemannian metrics on (M,M)(M,\partial M). Let \mathcal{M} be the space of all pairs (γ,Σ)(\gamma,\Sigma) such that γ𝚪\gamma\in\mathbf{\Gamma} and (Σ,Σ)(\Sigma,\partial\Sigma) is an almost properly embedded, free boundary minimal hypersurface in (M,M,γ)(M,\partial M,\gamma). Recall that the projection Π:𝚪\Pi:\mathcal{M}\rightarrow\mathbf{\Gamma} is defined as

Π(γ,Σ)=γ.\Pi(\gamma,\Sigma)=\gamma.

Denote by 0\mathcal{M}_{0} the set of (γ,Σ)(\gamma,\Sigma)\in\mathcal{M} such that Σ\Sigma is non-degenerate and proper.

Proof of Theorem 1.3.

We first prove that 𝚪Π(0)\mathbf{\Gamma}\setminus\Pi(\mathcal{M}\setminus\mathcal{M}_{0}) is dense in 𝚪\mathbf{\Gamma}. Indeed, for any γ𝚪\gamma\in\mathbf{\Gamma}, (Mn+1,M,γ)(M^{n+1},\partial M,\gamma) can be isometrically embedded into a closed manifold (M~n+1,γ~)(\widetilde{M}^{n+1},\widetilde{\gamma}). By Lemma 2.13, there exists a sequence of γ~jγ~\widetilde{\gamma}_{j}\rightarrow\widetilde{\gamma} such that every embedded minimal hypersurface in (M~,γ~j)(\widetilde{M},\widetilde{\gamma}_{j}) with free boundary on M\partial M is non-degenerate and transverse to M\partial M. When restricting to (M,M)(M,\partial M), γ~j|Mγ\widetilde{\gamma}_{j}|_{M}\rightarrow\gamma and γ~j|M𝚪Π(0)\widetilde{\gamma}_{j}|_{M}\in\mathbf{\Gamma}\setminus\Pi(\mathcal{M}\setminus\mathcal{M}_{0}) for each jj.

Now for any C>0C>0, denote by 𝚪(C)\mathbf{\Gamma}(C) the set of γ𝚪\gamma\in\mathbf{\Gamma} so that every free boundary minimal hypersurface in MM with AreaC\mathrm{Area}\leq C and indexC\mathrm{index}\leq C is non-degenerate and proper. Then the above argument gives that 𝚪(C)\mathbf{\Gamma}(C) contains a dense set 𝚪Π(0)\mathbf{\Gamma}\setminus\Pi(\mathcal{M}\setminus\mathcal{M}_{0}).

Claim 2.

For any C>0C>0, 𝚪(C)\mathbf{\Gamma}(C) is open.

Proof of Claim 2.

Let γ𝚪(C)\gamma\in\mathbf{\Gamma}(C). Suppose for the contrary that {γj}\{\gamma_{j}\} is a sequence of metrics on (M,M)(M,\partial M) and {Σj}\{\Sigma_{j}\} a sequence of free boundary minimal hypersurfaces in (M,M,γj)(M,\partial M,\gamma_{j}) with γjγ\gamma_{j}\rightarrow\gamma, Area(Σj)C\mathrm{Area}(\Sigma_{j})\leq C and Index(Σj)C\mathrm{Index}(\Sigma_{j})\leq C, but Σj\Sigma_{j} is either degenerate or improper. Then by the compactness theorems in \citelist[ACS17][GWZ18], up to a subsequence, Σj\Sigma_{j} locally smoothly converges to a free boundary minimal hypersurface Σ\Sigma away from a finite set in (M,M,γ)(M,\partial M,\gamma) with AreaC\mathrm{Area}\leq C and IndexC\mathrm{Index}\leq C. Since γ𝚪(C)\gamma\in\mathbf{\Gamma}(C), Σ\Sigma is non-degenerate and proper. This implies that the convergence is smooth, and hence Σj\Sigma_{j} is also non-degenerate and proper for large jj. This is a contradiction and hence Claim 2 is proved. ∎

We thus have that 𝚪(C)\mathbf{\Gamma}(C) contains an open dense set of 𝚪\mathbf{\Gamma}. Recall that

𝚪Π(0)=k=1𝚪(k).\mathbf{\Gamma}\setminus\Pi(\mathcal{M}\setminus\mathcal{M}_{0})=\bigcap_{k=1}^{\infty}\mathbf{\Gamma}(k).

Obviously, this is a generic subset of 𝚪\mathbf{\Gamma}. ∎

3. Relative Min-max theory for free boundary hh-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 hh-hypersurfaces associated with multiple-parameter families in 𝒞(M)\mathcal{C}(M). 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 I=[0,1]I=[0,1] with 0-cells {[k3j]}\{[k\cdot 3^{-j}]\} and 1-cells {[k3j,(k+1)3j]}\{[k\cdot 3^{-j},(k+1)\cdot 3^{-j}]\};

the cubical complex I(1,j)I(1,j)I(1,j)I(1,j)\otimes I(1,j)\otimes\cdots\otimes I(1,j) (mm times);

the map from I(k)0I(k)_{0} to I(j)0I(j)_{0} defined as the unique element such that
d(x,𝐧(k,j)(x))=inf{d(x,y):yI(j)0}d(x,\mathbf{n}(k,j)(x))=\inf\{d(x,y):y\in I(j)_{0}\}, where kjk\geq j;

a cubical subcomplex of dimension kk\in\mathbb{N} in some I(m,j)I(m,j);

the union of all cells of I(m,l)I(m,l) whose support is contained in some cell of XX;

the set of all qq-cells in X(l)X(l).

Let XkX^{k} be a cubical complex of dimension kk\in\mathbb{N} in some I(m,j)I(m,j) and ZXZ\subset X be a cubical subcomplex.

Let Φ0:X(𝒞(M),𝐅)\Phi_{0}:X\rightarrow(\mathcal{C}(M),\mathbf{F}) be a continuous map (under the 𝐅\mathbf{F}-topology on 𝒞(M)\mathcal{C}(M)). Let Π\Pi be the collection of all sequences of continuous maps {Φi:X𝒞(M)}i\{\Phi_{i}:X\rightarrow\mathcal{C}(M)\}_{i\in\mathbb{N}} such that

  1. (1)

    each Φi\Phi_{i} is homotopic to Φ0\Phi_{0} in the flat topology on 𝒞(M)\mathcal{C}(M);

  2. (2)

    there exist homotopy maps {Ψi:[0,1]×X𝒞(M)}i\{\Psi_{i}:[0,1]\times X\rightarrow\mathcal{C}(M)\}_{i\in\mathbb{N}} which are continuous in the flat topology, Ψi(0,)=Φi\Psi_{i}(0,\cdot)=\Phi_{i}, Ψi(1,)=Φ0\Psi_{i}(1,\cdot)=\Phi_{0}, and satisfy

    lim supisup{𝐅(Ψi(t,x),Φ0(x)):t[0,1],xZ}=0.\limsup_{i\rightarrow\infty}\sup\{\mathbf{F}(\Psi_{i}(t,x),\Phi_{0}(x)):t\in[0,1],x\in Z\}=0.

Given a pair (X,Z)(X,Z) and Φ0\Phi_{0} as above, {Φi}i\{\Phi_{i}\}_{i\in\mathbb{N}} is called a (X,Z)(X,Z)-homotopy sequence of mappings into 𝒞(M)\mathcal{C}(M), and Π\Pi is called the (X,Z)(X,Z)-homotopy class of Φ0\Phi_{0}. Then we define the hh-width by

𝐋h=𝐋h(Π):=inf{Φi}Πlim supisupxX{𝒜h(Φi(x))}.\mathbf{L}^{h}=\mathbf{L}^{h}(\Pi):=\inf_{\{\Phi_{i}\}\in\Pi}\limsup_{i\rightarrow\infty}\sup_{x\in X}\{\mathcal{A}^{h}(\Phi_{i}(x))\}.

A sequence {Φi}iΠ\{\Phi_{i}\}_{i\in\mathbb{N}}\in\Pi is called a minimizing sequence if 𝐋h({Φi})=𝐋h(Π)\mathbf{L}^{h}(\{\Phi_{i}\})=\mathbf{L}^{h}(\Pi), where

𝐋h({Φi}):=lim supisupxX{𝒜h(Φi(x))}.\mathbf{L}^{h}(\{\Phi_{i}\}):=\limsup_{i\rightarrow\infty}\sup_{x\in X}\{\mathcal{A}^{h}(\Phi_{i}(x))\}.

Given Φ0\Phi_{0} and Π\Pi, by the same argument as [Zhou19]*Lemma 1.5, there exists a minimizing sequence.

Definition 3.1.

If {Φi}i\{\Phi_{i}\}_{i\in\mathbb{N}} is a minimizing sequence in Π\Pi, the critical set of {Φi}\{\Phi_{i}\} is defined by

𝐂({Φi})={V=limj|Φij(xj)| as varifolds : with limj𝒜h(Φij(xj))=𝐋h(Π)}.\mathbf{C}(\{\Phi_{i}\})=\{V=\lim_{j\rightarrow\infty}|\partial\Phi_{i_{j}}(x_{j})|\text{\ as varifolds : with\ }\lim_{j\rightarrow\infty}\mathcal{A}^{h}(\Phi_{i_{j}}(x_{j}))=\mathbf{L}^{h}(\Pi)\}.

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 𝒵n(M,)\mathcal{Z}_{n}(M,\mathbb{Z}) or 𝒵n(M,2)\mathcal{Z}_{n}(M,\mathbb{Z}_{2}), they work well for sweepouts in 𝒞(M)\mathcal{C}(M) 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 X=[0,1]X=[0,1], there is no change for discrete sweepouts on XX.

Recall that given a map ϕ:X(k)0𝒞(M)\phi:X(k)_{0}\rightarrow\mathcal{C}(M), the fineness of ϕ\phi is defined as

𝐟(ϕ)=sup{(ϕ(x)ϕ(y))+𝐌(ϕ(x)ϕ(y)):x,y are adjacent vertices in X(k)0}.\mathbf{f}(\phi)=\sup\{\mathcal{F}(\phi(x)-\phi(y))+\mathbf{M}(\partial\phi(x)-\partial\phi(y)):x,y\text{ are adjacent vertices in }X(k)_{0}\}.

Here two vertices x,yX(k)0x,y\in X(k)_{0} are adjacent if they belong to a common cell in X(k)1X(k)_{1} .

Definition 3.2 ([MN17]*§3.7).

Given a continuous (in the flat topology) map Φ:X𝒞(M)\Phi:X\rightarrow\mathcal{C}(M), we say that Φ\Phi has no concentration of mass if

limr0sup{Φ(x)(Br(p)):pM,xX}=0.\lim_{r\rightarrow 0}\sup\{\|\partial\Phi(x)\|(B_{r}(p)):p\in M,x\in X\}=0.

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 Φ:X𝒞(M)\Phi:X\rightarrow\mathcal{C}(M) be a continuous map in the flat topology that has no concentration of mass, and supxX𝐌(Φ(x))<+\sup_{x\in X}\mathbf{M}(\partial\Phi(x))<+\infty. Assume that Φ|Z\Phi|_{Z} is continuous under the 𝐅\mathbf{F}-topology. Then there exist a sequence of maps

ϕi:X(ki)0𝒞(M),\phi_{i}:X(k_{i})_{0}\rightarrow\mathcal{C}(M),

and a sequence of homotopy maps:

ψi:I(ki)0×X(ki)0𝒞(M),\psi_{i}:I(k_{i})_{0}\times X(k_{i})_{0}\rightarrow\mathcal{C}(M),

with ki<ki+1,ψi(0,)=ϕi1𝐧(ki,ki1)k_{i}<k_{i+1},\psi_{i}(0,\cdot)=\phi_{i-1}\circ\mathbf{n}(k_{i},k_{i-1}), ψi(1,)=ϕi\psi_{i}(1,\cdot)=\phi_{i}, and a sequence of numbers {δi}i0\{\delta_{i}\}_{i\in\mathbb{N}}\rightarrow 0 such that

  1. (i)

    the fineness 𝐟(ψi)<δi\mathbf{f}(\psi_{i})<\delta_{i};

  2. (ii)
    sup{(ψi(t,x)Φ(x)):tI(ki)0,xX(ki)0}δi;\sup\{\mathcal{F}(\psi_{i}(t,x)-\Phi(x)):t\in I(k_{i})_{0},x\in X(k_{i})_{0}\}\leq\delta_{i};
  3. (iii)

    for some sequence lil_{i}\rightarrow\infty, with li<kil_{i}<k_{i},

    𝐌(ψi(t,x))sup{𝐌(Φ(y)):x,yα, for some αX(li)}+δi;\mathbf{M}(\partial\psi_{i}(t,x))\leq\sup\{\mathbf{M}(\partial\Phi(y)):x,y\in\alpha,\text{ for some }\alpha\in X(l_{i})\}+\delta_{i};

    and this directly implies that

    sup{𝐌(ϕi(x)):xX(k0)0}sup{𝐌(Φ(x)):xX}+δi.\sup\{\mathbf{M}(\partial\phi_{i}(x)):x\in X(k_{0})_{0}\}\leq\sup\{\mathbf{M}(\partial\Phi(x)):x\in X\}+\delta_{i}.

    As Φ|Z\Phi|_{Z} is continuous in the 𝐅\mathbf{F}-topology, we have from (iii) that for all tI(ki)0t\in I(k_{i})_{0} and xZ(ki)0x\in Z(k_{i})_{0},

    𝐌(ψi(t,x))𝐌(Φ(x))+ηi\mathbf{M}(\partial\psi_{i}(t,x))\leq\mathbf{M}(\partial\Phi(x))+\eta_{i}

    with ηi0\eta_{i}\rightarrow 0 as ii\rightarrow\infty. Applying [LZ16]*Lemma 3.13, we get by (ii) that

  4. (iv)
    sup{𝐅(ψi(t,x),Φ(x)):tI(ki)0,xZ(ki)0}0, as i.\sup\{\mathbf{F}(\psi_{i}(t,x),\Phi(x)):t\in I(k_{i})_{0},x\in Z(k_{i})_{0}\}\rightarrow 0,\text{ as }i\rightarrow\infty.

    Now given hC(M)h\in C^{\infty}(M), denoting c=supM|h|c=\sup_{M}|h|, then we have from (ii)(iii) that

  5. (v)
    𝒜h(ϕi(x))sup{𝒜h(Φ(y)):αX(li),x,yα}+(1+c)δi;\mathcal{A}^{h}(\phi_{i}(x))\leq\sup\{\mathcal{A}^{h}(\Phi(y)):\alpha\in X(l_{i}),x,y\in\alpha\}+(1+c)\delta_{i};

    and hence

    sup{𝒜h(ϕi(x)):xX(ki)0}sup{𝒜h(Φ(x)):xX}+(1+c)δi.\sup\{\mathcal{A}^{h}(\phi_{i}(x)):x\in X(k_{i})_{0}\}\leq\sup\{\mathcal{A}^{h}(\Phi(x)):x\in X\}+(1+c)\delta_{i}.
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 YY be a cubical subcomplex of I(m,k)I(m,k). Given two discrete maps ϕi:Y(li)0𝒞(M)\phi_{i}:Y(l_{i})_{0}\rightarrow\mathcal{C}(M), we say that ϕ1\phi_{1} is homotopic to ϕ2\phi_{2} with fineness less than η\eta, if there exist ll\in\mathbb{N}, l>l1,l2l>l_{1},l_{2} and a map

ψ:I(1,k+l)0×Y(l)0𝒞(M)\psi:I(1,k+l)_{0}\times Y(l)_{0}\rightarrow\mathcal{C}(M)

with fineness 𝐟(ψ)<η\mathbf{f}(\psi)<\eta and such that

ψ([i1],y)=ϕi(𝐧(k+l,k+li)(y)),i=1,2,yY(l)0.\psi([i-1],y)=\phi_{i}(\mathbf{n}(k+l,k+l_{i})(y)),i=1,2,y\in Y(l)_{0}.

The purpose of the next theorem is to construct a continuous map in the 𝐅\mathbf{F}-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 C0=C0(M,m)C_{0}=C_{0}(M,m) and δ0=δ0(M,m)\delta_{0}=\delta_{0}(M,m) so that if YY is a cubical subcomplex of I(m,k)I(m,k) and

ϕ:Y0𝒞(M)\phi:Y_{0}\rightarrow\mathcal{C}(M)

has 𝐟(ϕ)<δ0\mathbf{f}(\phi)<\delta_{0}, then there exists a map

Φ:Y𝒞(M)\Phi:Y\rightarrow\mathcal{C}(M)

continuous in the 𝐅\mathbf{F}-topology and satisfying

  1. (i)

    Φ(x)=ϕ(x)\Phi(x)=\phi(x) for all xY0x\in Y_{0};

  2. (ii)

    if α\alpha is some jj-cell in YY, then Φ\Phi restricted to α\alpha depends only on the values of ϕ\phi restricted on the vertices of α\alpha;

  3. (iii)
    sup{𝐅(Φ(x),Φ(y)):x,y lie in a common cell of Y}C0𝐟(ϕ).\sup\{\mathbf{F}(\Phi(x),\Phi(y)):x,y\text{ lie in a common cell of }Y\}\leq C_{0}\mathbf{f}(\phi).

Moreover, if ϕi:Y(li)0𝒞(M)\phi_{i}:Y(l_{i})_{0}\rightarrow\mathcal{C}(M) (i=1,2i=1,2) is homotopic to each other with fineness  η<δ0(M,m)\eta<\delta_{0}(M,m),

Φ1,Φ2:Y𝒞(M)\Phi_{1},\Phi_{2}:Y\rightarrow\mathcal{C}(M)

of ϕ1,ϕ2\phi_{1},\phi_{2}, respectively, are homotopic to each other in the 𝐅\mathbf{F}-topology.

Now for a continuous map Φ\Phi in Theorem 3.3, there exists a sequence of discretized maps {ϕi}\{\phi_{i}\}. Applying Theorem 3.4 to each ϕi\phi_{i}, we obtain Φi\Phi_{i} continuous in the 𝐅\mathbf{F}-topology. Then the next proposition says that Φi\Phi_{i} is homotopic to Φ\Phi.

Proposition 3.5 ([Zhou19]*Proposition 1.15).

Let {ϕi}i\{\phi_{i}\}_{i\in\mathbb{N}} and {ψi}i\{\psi_{i}\}_{i\in\mathbb{N}} be given by Theorem 3.3 applied to some Φ\Phi therein. Assume that Φ\Phi is continuous in the 𝐅\mathbf{F}-topology on XX. Then the Almgren extension Φi\Phi_{i} is homotopic to Φ\Phi in the 𝐅\mathbf{F}-topology for sufficiently large ii.

In particular, for ii large enough, there exist homotopy maps Ψi:[0,1]×X𝒞(M)\Psi_{i}:[0,1]\times X\rightarrow\mathcal{C}(M) continuous in the 𝐅\mathbf{F}-topology, Ψi(0,)=Φi\Psi_{i}(0,\cdot)=\Phi_{i}, Ψi(1,)=Φ\Psi_{i}(1,\cdot)=\Phi, and

lim supisupt[0,1],xX𝐅(Ψi(t,x),Φ(x))0.\limsup_{i\rightarrow\infty}\sup_{t\in[0,1],x\in X}\mathbf{F}(\Psi_{i}(t,x),\Phi(x))\rightarrow 0.

Therefore, for given hC(M)h\in C^{\infty}(M), we have

lim supisupxX𝒜h(Φi(x))supxX𝒜h(Φ(x)).\limsup_{i\rightarrow\infty}\sup_{x\in X}\mathcal{A}^{h}(\Phi_{i}(x))\leq\sup_{x\in X}\mathcal{A}^{h}(\Phi(x)).

3.3. Regularity of hh-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 “hh-almost minimizing with free boundary” concept.

Definition 3.6 (hh-almost minimizing varifolds with free boundary).

Let 𝝂\bm{\nu} be the \mathcal{F} or 𝐌\mathbf{M}-norm, or the 𝐅\mathbf{F}-metric. For any given ϵ,δ>0\epsilon,\delta>0 and a relative open subset UMU\subset M, we define 𝒜h(U;ϵ,δ;𝝂)\mathscr{A}^{h}(U;\epsilon,\delta;\bm{\nu}) to be the set of all Ω𝒞(M)\Omega\in\mathcal{C}(M) such that if Ω=Ω0,Ω1,Ω2,,Ωm𝒞(M)\Omega=\Omega_{0},\Omega_{1},\Omega_{2},...,\Omega_{m}\in\mathcal{C}(M) is a sequence with:

  1. (i)

    spt(ΩiΩ)U\operatorname{spt}(\Omega_{i}-\Omega)\subset U;

  2. (ii)

    𝝂(Ωi+1,Ωi)δ\bm{\nu}(\partial\Omega_{i+1},\partial\Omega_{i})\leq\delta;

  3. (iii)

    𝒜h(Ωi)𝒜h(Ω)+δ\mathcal{A}^{h}(\Omega_{i})\leq\mathcal{A}^{h}(\Omega)+\delta, for i=1,mi=1,...m, then

    𝒜h(Ωm)𝒜h(Ω)ϵ.\mathcal{A}^{h}(\Omega_{m})\geq\mathcal{A}^{h}(\Omega)-\epsilon.

We say that a varifold V𝒱n(M)V\in\mathcal{V}_{n}(M) is hh-almost minimizing in UU with free boundary if there exist sequences ϵi0\epsilon_{i}\rightarrow 0, δi0\delta_{i}\rightarrow 0, and Ωi𝒜h(U;ϵi,δi;)\Omega_{i}\in\mathscr{A}^{h}(U;\epsilon_{i},\delta_{i};\mathcal{F}) such that 𝐅(|Ωi|,V)ϵi.\mathbf{F}(|\partial\Omega_{i}|,V)\leq\epsilon_{i}.

For each pMp\in\partial M, as defined in [LZ16]*Definition A.4, the Fermi half-ball and half-sphere of radius rr centered at pp are

~r+(p)={qM:r~p(q)<r},𝒮~r+(p)={qM:r~p(q)=r},\widetilde{\mathcal{B}}^{+}_{r}(p)=\{q\in M:\widetilde{r}_{p}(q)<r\},\ \ \ \ \ \widetilde{\mathcal{S}}^{+}_{r}(p)=\{q\in M:\widetilde{r}_{p}(q)=r\},

where r~p()\widetilde{r}_{p}(\cdot) is the Fermi distance function to pp; (see [LZ16]*Definition A.1). For pMMp\in M\setminus\partial M, ~r+(p)\widetilde{\mathcal{B}}^{+}_{r}(p) and 𝒮~r+(p)\widetilde{\mathcal{S}}^{+}_{r}(p) are also used to denote the geodesic ball and sphere of radius rr at pp.

Definition 3.7.

A varifold V𝒱n(M)V\in\mathcal{V}_{n}(M) is said to be hh-almost minimizing in small annuli with free boundary if for each pMp\in M, there exists ram(p)>0r_{am}(p)>0 such that VV is hh-almost minimizing with free boundary in 𝒜s,r(p)\mathcal{A}_{s,r}(p) for all 0<s<rram(p)0<s<r\leq r_{am}(p), where 𝒜s,r(p)=~r+(p)~s+(p)\mathcal{A}_{s,r}(p)=\widetilde{\mathcal{B}}^{+}_{r}(p)\setminus\widetilde{\mathcal{B}}^{+}_{s}(p).

Before stating the regularity of hh-almost minimizing varifolds with free boundary, we provide the regularity of hh-replacements, which follows from Theorem 2.2.

Proposition 3.8 (Replacements [ZZ18]*Proposition 6.8).

Let V𝒱n(M)V\in\mathcal{V}_{n}(M) be hh-almost minimizing with free boundary in a relative open set UMU\subset M and KUK\subset U be a compact subset. Then there exists V𝒱n(M)V^{*}\in\mathcal{V}_{n}(M), called an hh-replacement of VV in KK such that, with c=sup|h|c=\sup|h|,

  1. (i)

    V(MK)=V(MK)V\llcorner(M\setminus K)=V^{*}\llcorner(M\setminus K);

  2. (ii)

    cVol(K)|V|(M)|V|(M)cVol(K)-c\cdot\mathrm{Vol}(K)\leq|V|(M)-|V^{*}|(M)\leq c\cdot\mathrm{Vol}(K);

  3. (iii)

    VV^{*} is hh-almost minimizing in UU with free boundary ;

  4. (iv)

    V=limi|Ωi|V^{*}=\lim_{i\rightarrow\infty}|\partial\Omega^{*}_{i}| as varifolds for some Ωi𝒞(M)\Omega^{*}_{i}\in\mathcal{C}(M) such that

    Ωi𝒜(U;ϵi,δi;) with ϵi,δi0;\Omega^{*}_{i}\in\mathscr{A}(U;\epsilon_{i},\delta_{i};\mathcal{F})\text{ with $\epsilon_{i},\delta_{i}\rightarrow 0$};

    furthermore Ωi\Omega^{*}_{i} locally minimizes 𝒜h\mathcal{A}^{h} in int(K)\mathrm{int}(K) (relative to M\partial M);

  5. (v)

    if VV has cc-bounded first variation in MM, then so does VV^{*}.

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 ϵi0\epsilon_{i}\rightarrow 0, δi0\delta_{i}\rightarrow 0 and Ωi𝒜h(U;ϵi,δi;)\Omega_{i}\in\mathscr{A}^{h}(U;\epsilon_{i},\delta_{i};\mathcal{F}) such that 𝐅(|Ωi|,V)ϵi\mathbf{F}(|\partial\Omega_{i}|,V)\leq\epsilon_{i}. Then for each Ωi\Omega_{i}, denote by Ωi\Omega_{i}^{*} the current by solving a constrained minimization problem; see [ZZ18]*Lemma 6.7. Then by Theorem 2.2, Ωi\partial\Omega^{*}_{i} is a properly embedded hh-hypersurface with free boundary. One can check that V:=limi|Ωi|V^{*}:=\lim_{i\rightarrow\infty}|\partial\Omega_{i}^{*}| (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 2n62\leq n\leq 6. Suppose that V𝒱n(M)V\in\mathcal{V}_{n}(M) has cc-bounded first variation in MM and is hh-almost minimizing in small annuli with free boundary. For any tangent varifold CVarTan(V,p)C\in\mathrm{VarTan}(V,p) with pspt|V|Mp\in\operatorname{spt}|V|\cap\partial M, we have either

  1. (i)

    C=Θn(|V|,p)|Tp(M)|C=\Theta^{n}(|V|,p)|T_{p}(\partial M)| where Θn(|V|,p)\Theta^{n}(|V|,p)\in\mathbb{N} or

  2. (ii)

    C=2Θn(|V|,p)|STpM|C=2\Theta^{n}(|V|,p)|S\cap T_{p}M| for some nn-plane S𝐆(L,n)S\in\mathbf{G}(L,n) such that STpM~S\subset T_{p}\widetilde{M} and STp(M)S\perp T_{p}(\partial M) and 2Θn(|V|,p)2\Theta^{n}(|V|,p)\in\mathbb{N}.

Moreover, for |V||V|-a.e. pspt|V|Mp\in\operatorname{spt}|V|\cap\partial M, the tangent varifold of VV at pp is unique, and the set of pMp\in\partial M in which case (ii) occurs as its unique tangent cones has |V||V|-measure 0; hence VV is rectifiable.

Proof.

The first step is to prove that CC is a stationary rectifiable cone in TpMT_{p}M 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.

Then applying Proposition 3.8, the proof of [LZ16]*Proposition 5.10 gives the desired classification. The only necessary modification is to replace the condition V(M)=V(M)\|V\|(M)=\|V^{*}\|(M) by Proposition 3.8 (ii) so that we can still get the volume growth for the blow up limit.  ∎

Now we are ready to prove the main regularity theorem for varifolds which is hh-almost minimizing with free boundary and has cc-bounded first variation.

Theorem 3.10 (Main regularity).

Let 2n62\leq n\leq 6, and (Mn+1,M,g)(M^{n+1},\partial M,g) be an (n+1)(n+1)-dimensional smooth, compact Riemannian manifold with boundary. Further let h𝒮(g)h\in\mathcal{S}(g) and set c=sup|h|c=\sup|h|. Suppose V𝒱n(M)V\in\mathcal{V}_{n}(M) is a varifold which

  1. (1)

    has cc-bounded first variation in MM, and

  2. (2)

    is hh-almost minimizing in small annuli with free boundary,

then VV is induced by Σ\Sigma, where Σ\Sigma is a compact, almost embedded hh-hypersurface with free boundary (possibly disconnected).

Proof.

We only need to prove the regularity of VV near an arbitrary point pspt|V|Mp\in\operatorname{spt}|V|\cap\partial M. Fix a pspt|V|Mp\in\operatorname{spt}|V|\cap\partial M, then there exists 0<r0<ram(p)0<r_{0}<r_{am}(p) such that for any r<r0r<r_{0}, the mean curvature HH of ~r+(p)M\partial\widetilde{\mathcal{B}}^{+}_{r}(p)\cap M in MM is greater than cc. Here ram(p)r_{am}(p) is as in Definition 3.7. In particular, if r<r0r<r_{0} and W𝒱n(M)W\in\mathcal{V}_{n}(M) has cc-bounded first variation in MM and W0W\neq 0 in ~r+(p)\mathcal{\widetilde{B}}_{r}^{+}(p), then

(3.1) sptW𝒮~r+(p)=Clos[spt|W|Clos(~r+(p))]𝒮~r+(p).\emptyset\neq\operatorname{spt}W\cap\mathcal{\widetilde{S}}_{r}^{+}(p)=\mathrm{Clos}[\operatorname{spt}|W|\setminus\mathrm{Clos}(\mathcal{\widetilde{B}}_{r}^{+}(p))]\cap\mathcal{\widetilde{S}}_{r}^{+}(p).

Here Clos()\mathrm{Clos}(\cdot) stands for the closure of some set.

We will show that V~r0+(p)V\llcorner\widetilde{\mathcal{B}}^{+}_{r_{0}}(p) is an almost embedded free boundary hh-hypersurface with density equal to 22 along its self-touching set.

The argument consists of six steps:

Step 1: Constructing successive hh-replacements VV^{*} and VV^{**} on two overlapping concentric annuli.

Step 2: Gluing the hh-replacements smoothly (as immersed hypersurfaces) on the overlap.

Step 3: Extending the hh-replacements to the point pp to get an ‘hh-replacement’ V~\widetilde{V} on the punctured ball.

Step 4: Showing that the singularity of V~\widetilde{V} at pp is removable, so that V~\widetilde{V} is regular.

Step 5: Showing that sptV~α+(p)\operatorname{spt}V\cap\widetilde{\mathcal{B}}^{+}_{\alpha}(p) is not contained in M\partial M for all α>0\alpha>0.

Step 6: Proving that VV coincides with the almost embedded hypersurface V~\widetilde{V}  on a small neighborhood of pp.

We now proceed to the proof.


Step 1. Fix any 0<s<t<r0<s<t<r. Since VV is hh-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 VV^{*} of VV on K=Clos(𝒜s,t(p))K=\mathrm{Clos}(\mathcal{A}_{s,t}(p)). By Theorem 2.2 and Proposition 3.8 (iv), if

Σ1:=spt|V|𝒜s,t(p),\Sigma_{1}:=\operatorname{spt}|V^{*}|\llcorner\mathcal{A}_{s,t}(p),

then (Σ1,Σ1)(M,M)(\Sigma_{1},\partial\Sigma_{1})\subset(M,\partial M) is an almost embedded stable free boundary hh-hypersurface with some unit normal ν1\nu_{1}; when the multiplicity is 11, Σ1\Sigma_{1} is locally a boundary so we can choose ν1\nu_{1} to be the outer normal.

Note that all the touching set 𝒮(Σ1)\mathcal{S}(\Sigma_{1}) is contained in a countable union of (n1)(n-1)-dimensional connected submanifolds S1(k)\bigcup S_{1}^{(k)}. Since a countable union of sets of measure zero still has measure zero, by Sard’s theorem we can choose s2(s,t)s_{2}\in(s,t) such that 𝒮~s2+(p)\widetilde{\mathcal{S}}_{s_{2}}^{+}(p) intersects Σ1\Sigma_{1} and all the S1(k)S^{(k)}_{1} transversally (even at Σ1\partial\Sigma_{1}). Then given any s1(0,s)s_{1}\in(0,s), following the argument in [ZZ17]*Theorem 6.1, Step 1, we can construct VV^{**}, which is a replacement of VV^{*} on Clos(𝒜s1,s2)\mathrm{Clos}(\mathcal{A}_{s_{1},s_{2}}). Denote by

Σ2:=spt|V|𝒜s1,s2(p).\Sigma_{2}:=\operatorname{spt}|V^{**}|\llcorner\mathcal{A}_{s_{1},s_{2}}(p).

Step 2.  We now show that Σ1\Sigma_{1} and Σ2\Sigma_{2} glue smoothly (as immersed hypersurfaces) across 𝒮~s2+(p)\widetilde{\mathcal{S}}_{s_{2}}^{+}(p). Indeed, define the intersection set

(3.2) Γ=Clos(Σ1)𝒮~s2+(p).\Gamma=\mathrm{Clos}(\Sigma_{1})\cap\widetilde{\mathcal{S}}_{s_{2}}^{+}(p).

Then by transversality, Γ\Gamma is an almost embedded hypersurface in 𝒮~s2+(p)\widetilde{\mathcal{S}}_{s_{2}}^{+}(p). Particularly, Γ\Gamma is not contained in M\partial M since 𝒮~s2+\widetilde{\mathcal{S}}_{s_{2}}^{+} intersect Σ1\partial\Sigma_{1} and 𝒮(Σ1)\mathcal{S}(\Sigma_{1}) transversally. For xΓMx\in\Gamma\setminus\partial M, following from the interior argument [ZZ18]*Theorem 7.1, Step 3, Σ1\Sigma_{1} coincides with Σ2\Sigma_{2} (with matching normal) in a neighborhood of xx. Using the unique continuation of hh-hypersurfaces, we conclude that

Σ1𝒜s,s2(p)=Σ2𝒜s,s2(p).\Sigma_{1}\llcorner\mathcal{A}_{s,s_{2}}(p)=\Sigma_{2}\llcorner\mathcal{A}_{s,s_{2}}(p).

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 pp. In fact, we use Vs1V_{s_{1}}^{**} to denote the second replacement that we constructed in Step 1 with inner radius s1s_{1}. Step 2 shows that this construction does not depend on s2s_{2}. Then we define V~\widetilde{V} to be the limit of Vs1V_{s_{1}}^{**} as s10s_{1}\to 0.

See [ZZ17]*Theorem 6.1, Step 3 for details.


Step 4. We now determine the regularity of V~\widetilde{V} at pp.

Firstly, observing that V~\widetilde{V} is still hh-almost minimizing in small annuli with free boundary and that V~\widetilde{V} is the varifold limit of a sequence Vs1V_{s_{1}}^{**}, which all have cc-bounded first variation, we know that V~\widetilde{V} also has cc-bounded first variation. This implies that the classification of tangent cones in Lemma 3.9 also holds true. Secondly, V~\widetilde{V}, when restricted to any small annulus 𝒜α,β(p)\mathcal{A}_{\alpha,\beta}(p) (0<α<β<s0<\alpha<\beta<s), already coincides with a smooth, almost embedded, weakly stable hh-boundary Σ\Sigma with free boundary. Using these two ingredients, by Theorem A.1, Σ\Sigma extends smoothly across pp as an almost embedded hypersurface in ~s+(p)\widetilde{\mathcal{B}}^{+}_{s}(p). Thus we complete Step 4.


Step 5. We argue by contradiction and assume that sptV~α0+(p)M\operatorname{spt}V\cap\widetilde{\mathcal{B}}^{+}_{\alpha_{0}}(p)\subset\partial M for some α0<s\alpha_{0}<s. Here we also use the chosen constants ss, tt, α\alpha, s2s_{2} in the previous steps. We first recall that by the Constancy Theorem [Si]*Theorem 41.1,

sptV~α0+(p)=M~α0+(p). \operatorname{spt}V\cap\widetilde{\mathcal{B}}^{+}_{\alpha_{0}}(p)=\partial M\cap\widetilde{\mathcal{B}}^{+}_{\alpha_{0}}(p). 

Now we take the first replacement VV^{*} of VV on K=Clos(𝒜s,t(p))K=\mathrm{Clos}(\mathcal{A}_{s,t}(p)). In the next paragraph, we are going to prove that for any α<α0\alpha<\alpha_{0}, ~α+(p)MsptV~\partial\widetilde{\mathcal{B}}^{+}_{\alpha}(p)\cap\partial M\subset\operatorname{spt}\widetilde{V}. As a result, for any α<α0\alpha<\alpha_{0}, ~α+(p)MsptV~\widetilde{\mathcal{B}}^{+}_{\alpha}(p)\cap\partial M\subset\operatorname{spt}\widetilde{V}, which leads to a contradiction to Proposition 2.4.

To conclude this step, we consider the second replacement VαV^{**}_{\alpha} of VV^{*} on Clos(𝒜α,s2(p))\mathrm{Clos}(\mathcal{A}_{\alpha,s_{2}}(p)). By the assumption ~α+(p)MsptV\widetilde{\mathcal{B}}^{+}_{\alpha}(p)\cap\partial M\subset\operatorname{spt}V and V~α+(p)=Vα~α+(p)V\llcorner\widetilde{\mathcal{B}}^{+}_{\alpha}(p)={V^{**}_{\alpha}}\llcorner\widetilde{\mathcal{B}}^{+}_{\alpha}(p), we have ~α+(p)MsptVα\partial\widetilde{\mathcal{B}}^{+}_{\alpha}(p)\cap\partial M\subset\operatorname{spt}V^{**}_{\alpha}. On the other hand, Vα𝒜α,s2(p)=V~𝒜α,s2(p)V^{**}_{\alpha}\llcorner\mathcal{A}_{\alpha,s_{2}}(p)=\widetilde{V}\llcorner\mathcal{A}_{\alpha,s_{2}}(p). Together with the classification of tangent cones in Lemma 3.9, we conclude that ~α+(p)MsptV~\partial\widetilde{\mathcal{B}}^{+}_{\alpha}(p)\cap\partial M\subset\operatorname{spt}\widetilde{V}. This concludes Step 5.


Step 6. It remains to show that VV coincides with V~\widetilde{V} in ~s+(p)\widetilde{\mathcal{B}}^{+}_{s}(p). Recall that by [ZZ18]*Theorem 7.1, VInt(M)V\llcorner\mathrm{Int}(M) is an almost embedded hh-hypersurface. Denote by 𝒮(V)\mathcal{S}(V) the self-touching set. Then by Proposition 2.4, 𝒮(V)\mathcal{S}(V) is contained in a countable union of smoothly embedded (n1)(n-1)-dimensional submanifolds. Hence we can take s<ss^{\prime}<s so that sptVIntM~s+(p)\operatorname{spt}V\cap\mathrm{Int}M\cap\partial\widetilde{\mathcal{B}}^{+}_{s^{\prime}}(p) is not contained in 𝒮(V)\mathcal{S}(V). Recall that VsV^{**}_{s^{\prime}} is the second replacement of VV^{*} in 𝒜s,s2\mathcal{A}_{s^{\prime},s_{2}}. Take

zsptVIntM~s+(p)𝒮(V).z\in\operatorname{spt}V\cap\mathrm{Int}M\cap\partial\widetilde{\mathcal{B}}^{+}_{s^{\prime}}(p)\setminus\mathcal{S}(V).

Such a set is non-empty by Step 5. Then VV coincides with VsV^{**}_{s^{\prime}} in a small neighborhood of zz by the construction of the second replacement. Then the unique continuation principle gives that V=V~V=\widetilde{V} in ~s+(p)\widetilde{\mathcal{B}}^{+}_{s}(p). This completes the proof of Theorem 3.10. ∎

3.4. Relative Min-max theory for free boundary hh-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 hh-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 hh-hypersurface setting with no change.

Recall that a minimizing sequence {Φi}iΠ\{\Phi_{i}\}_{i\in\mathbb{N}}\in\Pi such that every element of 𝐂({Φi})\mathbf{C}(\{\Phi_{i}\}) (see Definition 3.1) has cc-bounded variation or belongs to |Φ0|(Z)|\partial\Phi_{0}|(Z) is called a pull-tight.

The purpose of this part is to establish min-max theory for free boundary hh-hypersurfaces. Recall that the Morse index of an almost embedded free boundary hh-hypersurface Σ\Sigma is given in Definition 2.6.

Theorem 3.11.

Let (Mn+1,M,g)(M^{n+1},\partial M,g) be a compact Riemannian manifold of dimension 3(n+1)73\leq(n+1)\leq 7, and h𝒮(g)h\in\mathcal{S}(g) which satisfies Mh0\int_{M}h\geq 0. Given a kk-dimensional cubical complex XX and a subcomplex ZXZ\subset X, let Φ0:X𝒞(M)\Phi_{0}:X\rightarrow\mathcal{C}(M) be a map continuous in the 𝐅\mathbf{F}-topology, and Π\Pi be the associated (X,Z)(X,Z)-homotopy class of Φ0\Phi_{0}. Suppose

(3.3) 𝐋h(Π)>max{maxxZ𝒜h(Φ0(x)),0}.\mathbf{L}^{h}(\Pi)>\max\big{\{}\max_{x\in Z}\mathcal{A}^{h}(\Phi_{0}(x)),0\big{\}}.

Then there exists a nontrivial, smooth, compact, almost embedded hypersurface with free boundary (Σn,Σ)(M,M)(\Sigma^{n},\partial\Sigma)\subset(M,\partial M) , such that

  • Σ=Ω\llbracket\Sigma\rrbracket=\partial\Omega for some Ω𝒞(M)\Omega\in\mathcal{C}(M), where the mean curvature of Σ\Sigma with respect to the unit outer normal of Ω\Omega is hh, i.e.

    H|Σ=h|Σ;H|_{\Sigma}=h|_{\Sigma};
  • 𝒜h(Ω)=𝐋h(Π)\mathcal{A}^{h}(\Omega)=\mathbf{L}^{h}(\Pi);

  • indexw(Σ) k\mathrm{index}_{w}(\Sigma)\leq k.

Proof of Theorem 3.11.

The proof can be divided into five steps. In the first four steps, we always assume that (M,M,g)(M,\partial M,g) is isometrically embedded into a closed manifold (M~n+1,g~)(\widetilde{M}^{n+1},\widetilde{g}) and (g~,h~)(\widetilde{g},\widetilde{h}) is a good pair (e.g. Section 2.4) related to MM so that h~|M=h\widetilde{h}|_{M}=h.

Step A: We construct a pulled-tight minimizing sequence {Φi}iΠ\{\Phi_{i}\}_{i\in\mathbb{N}}\in\Pi so that every element of 𝐂({Φi})\mathbf{C}(\{\Phi_{i}\}) either has cc-bounded first variation, or belongs to |Φ0|(Z)|\partial\Phi_{0}|(Z).

Step B: There exists V𝐂({Φi})V\in\mathbf{C}(\{\Phi_{i}\}) so that VV is hh-almost minimizing in small annuli with free boundary.

Step C: VV has cc-bounded first variation, and hence VV is supported on an almost embedded free boundary hh-hypersurface Σ\Sigma satisfying Σ=Ω\llbracket\Sigma\rrbracket=\partial\Omega and 𝒜h(Ω)=𝐋h(Π)\mathcal{A}^{h}(\Omega)=\mathbf{L}^{h}(\Pi).

Step D: The {Φi}\{\Phi_{i}\} and VV in Step B can be chosen so that the support of VV has weak Morse index less than or equal to kk.

Step E: We provide the proof for general h𝒮(g)h\in\mathcal{S}(g).

Proof of Step A.

Let c=supM|h|c=\sup_{M}|h| and Lc=2𝐋h+cVol(M)L^{c}=2\mathbf{L}^{h}+c\mathrm{Vol}(M). Set

Ac={V𝒱n(M):V(M)Lc,V has c-bounded first variation}|Φ0|(Z),A^{c}_{\infty}=\{V\in\mathcal{V}_{n}(M):\|V\|(M)\leq L^{c},V\text{ has $c$-bounded first variation}\}\cup|\partial\Phi_{0}|(Z),

where |Φ0|(Z)={|Φ0(x)|:xZ}|\partial\Phi_{0}|(Z)=\{|\partial\Phi_{0}(x)|:x\in Z\}. Then we can follow [LZ16]*Proposition 4.17 to construct a continuous map:

H:[0,1]×(𝒞(M),𝐅){𝐌(Ω)Lc}(𝒞(M),𝐅){𝐌(Ω)Lc}H:[0,1]\times(\mathcal{C}(M),\mathbf{F})\cap\{\mathbf{M}(\partial\Omega)\leq L^{c}\}\rightarrow(\mathcal{C}(M),\mathbf{F})\cap\{\mathbf{M}(\partial\Omega)\leq L^{c}\}

such that:

  1. (i)

    H(0,Ω)=ΩH(0,\Omega)=\Omega for all Ω𝒞(M)\Omega\in\mathcal{C}(M);

  2. (ii)

    H(t,Ω)=ΩH(t,\Omega)=\Omega if |Ω|Ac|\partial\Omega|\in A^{c}_{\infty};

  3. (iii)

    if |Ω|Ac|\partial\Omega|\notin A^{c}_{\infty},

    𝒜h(H(1,Ω))𝒜h(Ω)L(𝐅(|Ω|,Ac))<0,\mathcal{A}^{h}(H(1,\Omega))-\mathcal{A}^{h}(\Omega)\leq-L(\mathbf{F}(|\partial\Omega|,A^{c}_{\infty}))<0,

    where L:[0,+)[0,+)L:[0,+\infty)\rightarrow[0,+\infty) is a continuous function with L(0)=0L(0)=0, L(t)>0L(t)>0 when t>0t>0;

  4. (iv)

    for every ϵ>0\epsilon>0, there exists δ>0\delta>0 such that

    xZ,𝐅(Ω,Φ0(x))<δ𝐅(H(t,Ω),Φ0(x))<ϵ, for all t[0,1];x\in Z,\mathbf{F}(\Omega,\Phi_{0}(x))<\delta\Rightarrow\mathbf{F}(H(t,\Omega),\Phi_{0}(x))<\epsilon,\text{ for all }t\in[0,1];

    this is a direct consequence of (ii) since |Φ0|(Z)Ac|\partial\Phi_{0}|(Z)\subset A^{c}_{\infty}.

Given a minimizing sequence {Φi}iΠ\{\Phi_{i}^{*}\}_{i\in\mathbb{N}}\in\Pi, we define Φi(x)=H(1,Φi(x))\Phi_{i}(x)=H(1,\Phi^{*}_{i}(x)) for every xXx\in X. Then {Φi}i\{\Phi_{i}\}_{i\in\mathbb{N}} is also a minimizing sequence in Π\Pi. Moreover, 𝐂({Φi})𝐂({Φi})\mathbf{C}(\{\Phi_{i}\})\subset\mathbf{C}(\{\Phi_{i}^{*}\}) and every element of 𝐂({Φi})\mathbf{C}(\{\Phi_{i}\}) either has cc-bounded first variation, or belongs to |Φ0|(Z)|\partial\Phi_{0}|(Z). 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 {Φi}iΠ\{\Phi_{i}\}_{i\in\mathbb{N}}\in\Pi be a pulled-tight minimizing sequence. For each Φi\Phi_{i}, Theorem 3.3 gives a sequence of maps:

ϕij:X(kij)0𝒞(M)\phi_{i}^{j}:X(k_{i}^{j})_{0}\rightarrow\mathcal{C}(M)

with kij<kij+1k_{i}^{j}<k_{i}^{j+1} and a sequence of positive δij0\delta_{i}^{j}\rightarrow 0 (as jj\rightarrow\infty), satisfying Theorem 3.3. Then for each ii, take sufficiently large j(i)j(i) and let φi=ϕij(i)\varphi_{i}=\phi_{i}^{j(i)}. Denote by S={φi}S=\{\varphi_{i}\}. Then we have 𝐋h(S)=𝐋h({Φi})\mathbf{L}^{h}(S)=\mathbf{L}^{h}(\{\Phi_{i}\}) and 𝐂(S)=𝐂({Φi})\mathbf{C}(S)=\mathbf{C}(\{\Phi_{i}\}), where

𝐋h(S)=lim supisup{𝒜h(φi(y)):y(Xij(i))0};\displaystyle\mathbf{L}^{h}(S)=\limsup_{i\rightarrow\infty}\sup\{\mathcal{A}^{h}(\varphi_{i}(y)):y\in(X_{i}^{j(i)})_{0}\};
𝐂(S)={V=limj|φij(yj)| as varifolds: with limj𝒜h(φij(yj))=𝐋h(S)}.\displaystyle\mathbf{C}(S)=\{V=\lim_{j\rightarrow\infty}|\partial\varphi_{i_{j}}(y_{j})|\text{ as varifolds: with }\lim_{j\rightarrow\infty}\mathcal{A}^{h}(\varphi_{i_{j}}(y_{j}))=\mathbf{L}^{h}(S)\}.

We now prove that there exists V𝐂(S)V\in\mathbf{C}(S) so that it is hh-almost minimizing in small annuli with free boundary. For a further reason, we need a stronger result:

Claim 3.

There exist a varifold VV satisfying the following: for any pMp\in M and any small enough annulus 𝒜r1,r2\mathcal{A}_{r_{1},r_{2}} centered at pp with radii 0<r1<r20<r_{1}<r_{2}, there exist two sequences of positive real numbers ϵj0,δj0\epsilon_{j}\rightarrow 0,\delta_{j}\rightarrow 0, a subsequence {ij}{i}\{i_{j}\}\subset\{i\} and yjdmnφijy_{j}\in\mathrm{dmn}\varphi_{i_{j}} (the domain of φij\varphi_{i_{j}}) so that

  • limj𝒜h(φij(yj))=𝐋h(S)\lim_{j\rightarrow\infty}\mathcal{A}^{h}(\varphi_{i_{j}}(y_{j}))=\mathbf{L}^{h}(S);

  • φij(yj)𝒜h(𝒜r1,r2;ϵj,δj;𝐌)\varphi_{i_{j}}(y_{j})\in\mathscr{A}^{h}(\mathcal{A}_{r_{1},r_{2}};\epsilon_{j},\delta_{j};\mathbf{M}); and

  • limj|φij(yj)|=V\lim_{j\rightarrow\infty}|\partial\varphi_{i_{j}}(y_{j})|=V.

If Claim 3 were not true, then using the argument in [Zhou19]*Theorem 1.16, we can find a sequence S~={φ~i}\widetilde{S}=\{\widetilde{\varphi}_{i}\} so that φ~i\widetilde{\varphi}_{i} is homotopic to φi\varphi_{i} with fineness converging to zero as ii\rightarrow\infty and 𝐋h(S~)<𝐋h(S)\mathbf{L}^{h}(\widetilde{S})<\mathbf{L}^{h}(S). The key point here is that 𝐂(S)\mathbf{C}(S) is compact in the sense of varifolds.

Then by Theorem 3.4, the Almgren extensions of φi,φ~i\varphi_{i},\widetilde{\varphi}_{i}:

 Φij(i),Φ~i:X𝒞(M), \Phi_{i}^{j(i)},\widetilde{\Phi}_{i}:X\rightarrow\mathcal{C}(M),

respectively, are homotopic to each other in the 𝐅\mathbf{F}-topology for large ii and

 lim supisup{𝒜h(Φ~i(x)):xX}𝐋h(S~)<𝐋h(S). \limsup_{i\rightarrow\infty}\sup\{\mathcal{A}^{h}(\widetilde{\Phi}_{i}(x)):x\in X\}\leq\mathbf{L}^{h}(\widetilde{S})<\mathbf{L}^{h}(S).

This leads to a contradiction and we have finished the proof of Claim 3. Obviously, such a varifold VV 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 VV has cc-bounded first variation. Indeed, from Step A, VV either has cc-bounded first variation or belongs to |Φ0|(Z)|\partial\Phi_{0}|(Z). Recall that being hh-almost minimizing in small annuli with free boundary always implies VV has cc-bounded first variation away from finitely many points. Then by a cut-off trick, we only need to prove that V\|V\| has at most rn12r^{n-\frac{1}{2}}-volume growth near these bad points . This essentially follows from [HL75]*Theorem 4.1 and we provide more details here.

Let pp be a bad point. Then we can take ϵ>0\epsilon>0 small enough so that ~ϵ+(p){p}\widetilde{\mathcal{B}}^{+}_{\epsilon}(p)\setminus\{p\} has no bad point. For any 0<t<ϵ<u<10<t<\epsilon<u<1, let ηu,ϕt:\eta_{u},\phi_{t}:\mathbb{R}\rightarrow\mathbb{R} be two cut-off functions so that

ηu0; ηu(x)=1 for xu; ηu(x)=0 for x1;\displaystyle\eta_{u}^{\prime}\leq 0;\  \ \eta_{u}(x)=1\text{ for }x\leq u;\  \ \eta_{u}(x)=0\text{ for }x\geq 1;
ϕt0; ϕt(x)=0 for xt/2; ϕt(x)=1 for xt.\displaystyle\phi_{t}^{\prime}\geq 0;\ \  \phi_{t}(x)=0\text{ for }x\leq t/2;\ \  \phi_{t}(x)=1\text{ for }x\geq t.

We only need to consider pMp\in\partial M. Denote by r~\widetilde{r} the Fermi distance function to pp in [LZ16]*Appendix A. Then  there exist CC and ϵ\epsilon so that for xMx\in M with r~(x)ϵ\widetilde{r}(x)\leq\epsilon,

(3.4) ||r~(x)|1|Cϵ  and |2r~2/2(x)g|Cϵ.\big{|}{|\nabla\widetilde{r}(x)|-1}\big{|}\leq C\epsilon\ \  \text{ and }\ \ |\nabla^{2}\widetilde{r}^{2}/2(x)-g|\leq C\epsilon.

Being cc-bounded first variation in {0<r~<ϵ}\{0<\widetilde{r}<\epsilon\} gives that for any ρ<ϵ\rho<\epsilon,

(3.5) divS(ηu(r~/ρ)ϕt(r~)r~2/2)𝑑V(x,S)\displaystyle\int\mathrm{div}_{S}\big{(}\eta_{u}(\widetilde{r}/\rho)\phi_{t}(\widetilde{r})\nabla\widetilde{r}^{2}/2\big{)}\,dV(x,S) cηu(r~/ρ)ϕt(r~)|r~2/2|𝑑V(x,S)\displaystyle\leq c\int\eta_{u}(\widetilde{r}/\rho)\phi_{t}(\widetilde{r})|\nabla\widetilde{r}^{2}/2|\,dV(x,S)
(1+Cϵ)cηu(r~/ρ)r~𝑑V.\displaystyle\leq(1+C\epsilon)c\int\eta_{u}(\widetilde{r}/\rho)\cdot\widetilde{r}\,dV.

By direct computation of the left hand side,

(3.6) divS(ηu(r~/ρ)ϕt(r~)r~2/2)𝑑V(x,S)\displaystyle\ \int\mathrm{div}_{S}\big{(}\eta_{u}(\widetilde{r}/\rho)\phi_{t}(\widetilde{r})\nabla\widetilde{r}^{2}/2\big{)}\,dV(x,S)
\displaystyle\geq ηu(r~/ρ)ρ1ϕt(r~)r~pS(r~),r~ +ηu(r~/ρ)ϕt(r~)divS(r~2/2)dV(x,S)\displaystyle\ \int\eta_{u}^{\prime}(\widetilde{r}/\rho)\cdot\rho^{-1}\cdot\phi_{t}(\widetilde{r})\widetilde{r}\langle p_{S}(\nabla\widetilde{r}),\nabla\widetilde{r}\rangle +\eta_{u}(\widetilde{r}/\rho)\phi_{t}(\widetilde{r})\mathrm{div}_{S}(\nabla\widetilde{r}^{2}/2)\,dV(x,S)
\displaystyle\geq ηu(r~/ρ)ρ1ϕt(r~)r~(1+Cϵ)+ηu(r~/ρ)ϕt(r~)(1Cϵ)ndV.\displaystyle\ \int\eta_{u}^{\prime}(\widetilde{r}/\rho)\cdot\rho^{-1}\cdot\phi_{t}(\widetilde{r})\widetilde{r}\cdot(1+C\epsilon)+\eta_{u}(\widetilde{r}/\rho)\phi_{t}(\widetilde{r})\cdot(1-C\epsilon)n\,dV.

Here pS()p_{S}(\cdot) is the projection to the hyperplane SS and(3.4) is used in the last inequality. Note that VV either has cc-bounded first variation or belongs to |Φ0|(Z)|\partial\Phi_{0}|(Z). Hence V({p})=0\|V\|(\{p\})=0. Combining (3.5) with (3.6), letting t0t\rightarrow 0, and by shrinking ϵ\epsilon if necessary, we have

(n1/2)I(ρ)ηu(r~/ρ)r~ρ1=ρddρηu(r~/ρ)𝑑V=ρI(ρ),(n-1/2)I(\rho)\leq-\int\eta_{u}^{\prime}(\widetilde{r}/\rho)\widetilde{r}\rho^{-1}=\rho\frac{d}{d\rho}\int\eta_{u}(\widetilde{r}/\rho)\,dV=\rho I^{\prime}(\rho),

where I(ρ)=ηu(r~/ρ)𝑑VI(\rho)=\int\eta_{u}(\widetilde{r}/\rho)\,dV. This implies the monotonicity of I(ρ)ρn+1/2I(\rho)\cdot\rho^{-n+1/2}. Hence for ρ<ϵ\rho<\epsilon, we have

I(ρ)I(ϵ)ϵn+1/2ρn1/2.I(\rho)\leq I(\epsilon)\epsilon^{-n+1/2}\rho^{n-1/2}.

Then we conclude that

V({xM:distM(x,p)<ρ/2})I(ρ)Cρn1/2. \|V\|(\{x\in M:\operatorname{dist}_{M}(x,p)<\rho/2\})\leq I(\rho)\leq C\rho^{n-1/2}. 

This proves that VV has cc-bounded first variation. Then by Theorem 3.10, VV is supported on an almost embedded hh-hypersurface Σ\Sigma with free boundary.

We now prove that Σ\Sigma is a free boundary hh-hypersurface satisfying Σ=Ω\llbracket\Sigma\rrbracket=\partial\Omega and 𝒜h(Ω)=𝐋h(Π)\mathcal{A}^{h}(\Omega)=\mathbf{L}^{h}(\Pi) for some Ω𝒞(M)\Omega\in\mathcal{C}(M). Recall that Claim 3 gives that V=limj|φij(yj)|V=\lim_{j\rightarrow\infty}|\partial\varphi_{i_{j}}(y_{j})|. Denote by Ωj=φij(yj)\Omega_{j}=\varphi_{i_{j}}(y_{j}). Then it suffices to prove that Ωj\partial\Omega_{j} subsequently converges to Σ\Sigma in the \mathcal{F} metric. Let Ω\Omega be a limit of Ωj\Omega_{j} in the \mathcal{F} topology. Then spt(Ω)Σ\operatorname{spt}(\partial\Omega)\subset\Sigma. Now for any pInt(Σ)Touch(Σ)p\in\mathrm{Int}(\Sigma)\setminus\mathrm{Touch}(\Sigma) and rp>0r_{p}>0 small enough, the Constancy Theorem [Si, Theorem 26.27] implies that ΩBrp(p)=ΣBrp(p)\partial\Omega\llcorner B_{r_{p}}(p)=\llbracket\Sigma\rrbracket\llcorner B_{r_{p}}(p) or 0. Here Touch(Σ)\mathrm{Touch}(\Sigma) is the touching set inside Σ\Sigma.

In the next, we prove that ΩBrp(p)=ΣBrp(p)\partial\Omega\llcorner B_{r_{p}}(p)=\llbracket\Sigma\rrbracket\llcorner B_{r_{p}}(p) for pInt(Σ)Touch(Σ)p\in\mathrm{Int}(\Sigma)\setminus\mathrm{Touch}(\Sigma). Recall that first hh-replacement VV^{*} in 𝒜r1,r2(p)\mathcal{A}_{r_{1},r_{2}}(p) coincides with VV in the Step 5 of Theorem 3.10. On the other hand, the argument in [ZZ18]*Proposition 7.3 gives that V𝒜r1,r2(p)=|Ω|𝒜r1,r2(p)V^{*}\llcorner\mathcal{A}_{r_{1},r_{2}}(p)=|\partial\Omega|\llcorner\mathcal{A}_{r_{1},r_{2}}(p). Hence ΩBrp(p)=ΣBrp(p)\partial\Omega\llcorner B_{r_{p}}(p)=\llbracket\Sigma\rrbracket\llcorner B_{r_{p}}(p).

Recall that by Proposition 2.4, the self-touching set and ΣM\Sigma\cap\partial M are contained in countable (n1)(n-1)-dimensional submanifolds. Therefore, we conclude that V=|Ω|V=|\partial\Omega|. Hence Step C is finished. ∎

Proof of Step D.

Recall that by Lemme 2.12, the set of almost embedded free boundary hh-hypersurface is countable. Then the proof here is parallel to [Zhou19]*Theorem 3.6, which is in fact a generalization of [MN16].

Denote by 𝒰\mathcal{U} the set of all (Σ,Ω)𝒫h×𝒞(M)(\Sigma,\Omega)\in\mathcal{P}^{h}\times\mathcal{C}(M) so that Σ=Ω\llbracket\Sigma\rrbracket=\partial\Omega. Set

𝒲={Ω𝒞(M):(Σ,Ω)𝒰,𝒜h(Ω)=𝐋h},\displaystyle\mathcal{W}=\{\Omega\in\mathcal{C}(M):(\Sigma,\Omega)\in\mathcal{U},\mathcal{A}^{h}(\Omega)=\mathbf{L}^{h}\},
𝒲(r)={Ω𝒲:(Σ,Ω)𝒰,𝐅([Σ],𝐂({Φi}i))r},\displaystyle\mathcal{W}(r)=\{\Omega\in\mathcal{W}:(\Sigma,\Omega)\in\mathcal{U},\mathbf{F}([\Sigma],\mathbf{C}(\{\Phi_{i}\}_{i\in\mathbb{N}}))\geq r\},
𝒲k+1={Ω𝒲:(Σ,Ω)𝒰,indexw(Σ)k+1}.\displaystyle\mathcal{W}^{k+1}=\{\Omega\in\mathcal{W}:(\Sigma,\Omega)\in\mathcal{U},\mathrm{index}_{w}(\Sigma)\geq k+1\}.

It suffices to show that for every r>0r>0, 𝒲(𝒲k+1𝒲(r))\mathcal{W}\setminus(\mathcal{W}^{k+1}\cup\mathcal{W}(r)) is not empty.

By the same proof with [Zhou19]*Lemma 3.7, there exist i0i_{0}\in\mathbb{N} and ϵ¯0>0\bar{\epsilon}_{0}>0 such that 𝐅(Φi(X),𝒲(r))>ϵ¯0\mathbf{F}(\Phi_{i}(X),\mathcal{W}(r))>\bar{\epsilon}_{0} for all i>i0i>i_{0}.

As (g~,h~)(\widetilde{g},\widetilde{h}) is a good pair related to MM, 𝒲k+1\mathcal{W}^{k+1} is countable by Lemma 2.12, and we can assume that

𝒲k+1𝐁¯ϵ¯0𝐅(𝒲(r))={Ω1,Ω2,},\mathcal{W}^{k+1}\setminus\overline{\mathbf{B}}_{\bar{\epsilon}_{0}}^{\mathbf{F}}(\mathcal{W}(r))=\{\Omega_{1},\Omega_{2},\cdots\},

and Ωj=Σj\partial\Omega_{j}=\llbracket\Sigma_{j}\rrbracket for j1j\geq 1. Then by taking ϵ1>0\epsilon_{1}>0 small enough, we can make sure 𝐁¯ϵ1𝐅(Ω1)𝐁¯ϵ¯0𝐅(𝒲(r))=\overline{\mathbf{B}}_{\epsilon_{1}}^{\mathbf{F}}(\Omega_{1})\cap\overline{\mathbf{B}}_{\bar{\epsilon}_{0}}^{\mathbf{F}}(\mathcal{W}(r))=\emptyset. Using the Deformation Theorem (\citelist[Zhou19]*Theorem 3.4[GLWZ19]*Theorem 5.8), by shrinking ϵ1>0\epsilon_{1}>0 if necessary, we can find i1i_{1}\in\mathbb{N}, and {Φi1}i\{\Phi_{i}^{1}\}_{i\in\mathbb{N}} so that

  • Φi1\Phi_{i}^{1} is homotopic to Φi\Phi_{i} in the 𝐅\mathbf{F}-topology for all ii\in\mathbb{N};

  • 𝐋h({Φi1}i)𝐋h(Π)\mathbf{L}^{h}(\{\Phi_{i}^{1}\}_{i})\leq\mathbf{L}^{h}(\Pi);

  • 𝐅(Φi1(X),𝐁¯ϵ1𝐅(Ω1)𝐁¯ϵ¯0𝐅(𝒲(r)))>0\mathbf{F}(\Phi_{i}^{1}(X),\overline{\mathbf{B}}_{\epsilon_{1}}^{\mathbf{F}}(\Omega_{1})\cup\overline{\mathbf{B}}_{\bar{\epsilon}_{0}}^{\mathbf{F}}(\mathcal{W}(r)))>0 for all ii1i\geq i_{1};

  • no Ωj\Omega_{j} belongs to 𝐁¯ϵ1𝐅(Ω1)\partial\overline{\mathbf{B}}_{\epsilon_{1}}^{\mathbf{F}}(\Omega_{1}); (this can be easily satisfied since {Ω1,Ω2,}\{\Omega_{1},\Omega_{2},\cdots\} is a countable set.)

Inductively, there are two possibilities. The first case is that we can find for all ll\in\mathbb{N}, there exist a sequence {Φil}i\{\Phi_{i}^{l}\}_{i\in\mathbb{N}}, ϵl\epsilon_{l}, ili_{l}\in\mathbb{N}, and Ωjl𝒲k+1 𝐁¯ϵ0𝐅(𝒲(r))\Omega_{j_{l}}\in\mathcal{W}^{k+1}\setminus \overline{\mathbf{B}}_{\epsilon_{0}}^{\mathbf{F}}(\mathcal{W}(r)) for some jlj_{l}\in\mathbb{N} so that

  • Φil\Phi_{i}^{l} is homotopic to Φi\Phi_{i} in the 𝐅\mathbf{F}-topology for all ii\in\mathbb{N};

  • 𝐋h({Φil}i)𝐋h(Π)\mathbf{L}^{h}(\{\Phi_{i}^{l}\}_{i})\leq\mathbf{L}^{h}(\Pi);

  • 𝐅(Φil(X),q=1l𝐁¯ϵq𝐅(Ωjq) 𝐁¯ϵ¯0𝐅(𝒲(r)))>0\mathbf{F}(\Phi_{i}^{l}(X),\cup_{q=1}^{l}\overline{\mathbf{B}}_{\epsilon_{q}}^{\mathbf{F}}(\Omega_{j_{q}})\cup \overline{\mathbf{B}}_{\bar{\epsilon}_{0}}^{\mathbf{F}}(\mathcal{W}(r)))>0 for all iili\geq i_{l};

  • {Ω1,,Ωl}q=1l𝐁¯ϵq𝐅(Ωjq)\{\Omega_{1},\cdots,\Omega_{l}\}\subset\cup_{q=1}^{l}\overline{\mathbf{B}}_{\epsilon_{q}}^{\mathbf{F}}(\Omega_{j_{q}});

  • no Ωj\Omega_{j} belongs to 𝐁¯ϵ1𝐅(Ω1) 𝐁¯ϵl𝐅(Ωjl)\partial\overline{\mathbf{B}}_{\epsilon_{1}}^{\mathbf{F}}(\Omega_{1})\cup\cdots\cup \partial\overline{\mathbf{B}}_{\epsilon_{l}}^{\mathbf{F}}(\Omega_{j_{l}}).

The second case is that the process stops in finitely many steps. This means that we can find some mm\in\mathbb{N}, a sequence {Φim}i\{\Phi_{i}^{m}\}_{i\in\mathbb{N}}, ϵ1,,ϵm>0\epsilon_{1},\cdots,\epsilon_{m}>0, imi_{m}\in\mathbb{N}, and Ωj1,,Ωjm𝒲k+1 𝐁¯ϵ¯0𝐅(𝒲(r))\Omega_{j_{1}},\cdots,\Omega_{j_{m}}\in\mathcal{W}^{k+1}\setminus \overline{\mathbf{B}}_{\bar{\epsilon}_{0}}^{\mathbf{F}}(\mathcal{W}(r)) so that

  • Φim\Phi_{i}^{m} is homotopic to Φi\Phi_{i} in the 𝐅\mathbf{F}-topology for all ii\in\mathbb{N};

  • 𝐋h({Φim}i)𝐋h(Π)\mathbf{L}^{h}(\{\Phi_{i}^{m}\}_{i})\leq\mathbf{L}^{h}(\Pi);

  • 𝐅(Φim(X),q=1m𝐁¯ϵq𝐅(Ωjq) 𝐁¯ϵ¯0𝐅(𝒲(r)))>0\mathbf{F}(\Phi_{i}^{m}(X),\cup_{q=1}^{m}\overline{\mathbf{B}}_{\epsilon_{q}}^{\mathbf{F}}(\Omega_{j_{q}})\cup \overline{\mathbf{B}}_{\bar{\epsilon}_{0}}^{\mathbf{F}}(\mathcal{W}(r)))>0 for all iimi\geq i_{m};

  • {Ωj:j1}q=1m𝐁¯ϵq𝐅(Ωjq)\{\Omega_{j}:j\geq 1\}\subset\cup_{q=1}^{m}\overline{\mathbf{B}}_{\epsilon_{q}}^{\mathbf{F}}(\Omega_{j_{q}}).

For the first case, we can choose a diagonal sequence {Φpll}l\{\Phi_{p_{l}}^{l}\}_{l\in\mathbb{N}} and set Ψl=Φpll\Psi_{l}=\Phi_{p_{l}}^{l}, where {pl}l\{p_{l}\}_{l\in\mathbb{N}} is an increasing sequence such that plilp_{l}\geq i_{l}  and

supxX𝒜h(Φpll(x))𝐋h(Π)+1l.\sup_{x\in X}\mathcal{A}^{h}(\Phi_{p_{l}}^{l}(x))\leq\mathbf{L}^{h}(\Pi)+\frac{1}{l}.

For the second case, we simply choose the last sequence {Φlm}l\{\Phi_{l}^{m}\}_{l}  and set pl=lp_{l}=l and Ψl=Φlm\Psi_{l}=\Phi_{l}^{m}. Now it is easy to see that for both cases, the new sequence {Ψl}l\{\Psi_{l}\}_{l\in\mathbb{N}} satisfies the following conditions:

  1. (1)

    Ψl\Psi_{l} is homotopic to Φpl\Phi_{p_{l}} in the 𝐅\mathbf{F}-topology for all ll\in\mathbb{N};

  2. (2)

    𝐋h({Ψl}l)𝐋h(Π)\mathbf{L}^{h}(\{\Psi_{l}\}_{l})\leq\mathbf{L}^{h}(\Pi);

  3. (3)

    given any subsequence {lj}{l}\{l_{j}\}\subset\{l\}, xjXx_{j}\in X, if limj𝒜h(Ψlj(xj))=𝐋h(Π)\lim_{j\rightarrow\infty}\mathcal{A}^{h}(\Psi_{l_{j}}(x_{j}))=\mathbf{L}^{h}(\Pi), then {Ψlj(xj)}j\{\Psi_{l_{j}}(x_{j})\}_{j\in\mathbb{N}} does not converge in the 𝐅\mathbf{F}-topology to any element in  𝒲k+1𝒲(r)\mathcal{W}^{k+1}\cup\mathcal{W}(r).

Then by Steps B and C, there exists V𝐂({Ψl})V\in\mathbf{C}(\{\Psi_{l}\}) so that its support has weak Morse index less than or equal to kk. This finishes the proof of Step D. ∎

Proof of Step E.

Assume that (Mn+1,M,g)(M^{n+1},\partial M,g) is isometrically embedded into (M~n+1,g~)(\widetilde{M}^{n+1},\widetilde{g}). Recall that 𝒮~0\widetilde{\mathcal{S}}_{0} is the set of smooth Morse functions so that the zero set is a properly embedded closed hypersurface in M~\widetilde{M}, and is transverse to M\partial M.

Given h𝒮(g)h\in\mathcal{S}(g), we can take an extension h~\widetilde{h} of hh so that h~|M=h\widetilde{h}|_{M}=h and h~𝒮~0\widetilde{h}\in\widetilde{\mathcal{S}}_{0}. By Lemma 2.11, there exists a sequence of smooth metrics g~j\widetilde{g}_{j} on M~\widetilde{M} so that g~j|Mg\widetilde{g}_{j}|_{M}\rightarrow g smoothly and (g~j,h~)(\widetilde{g}_{j},\widetilde{h}) is a good pair related MM for each jj. Then by Steps A-D, for each jj, there exists an almost embedded free boundary hh-hypersurface Σ~j\widetilde{\Sigma}_{j} with Ω~j=Σ~j\partial{\widetilde{\Omega}_{j}}=\llbracket\widetilde{\Sigma}_{j}\rrbracket for some Ω~j𝒞(M)\widetilde{\Omega}_{j}\in\mathcal{C}(M) and Σ~j\widetilde{\Sigma}_{j} has weak Morse index less than or equal to kk (with respect to g~j|M\widetilde{g}_{j}|_{M}). Recall that h𝒮(g)h\in\mathcal{S}(g). Let Σ\Sigma_{\infty} (with Ω=Σ\partial\Omega_{\infty}=\llbracket\Sigma_{\infty}\rrbracket) be the limit of {Σ~j}j\{\widetilde{\Sigma}_{j}\}_{j\in\mathbb{N}} 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 𝒜h(Ω)=𝐋h(Π,g)\mathcal{A}^{h}(\Omega_{\infty})=\mathbf{L}^{h}(\Pi,g) and indexw(Σ)k\mathrm{index}_{w}(\Sigma_{\infty})\leq k. ∎

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 𝒜ϵh\mathcal{A}^{\epsilon h} for some prescribing function when ϵ0\epsilon\rightarrow 0. By the index estimates and the multiplicity one result for 𝒜ϵh\mathcal{A}^{\epsilon h}, we prove that the limit free boundary minimal hypersurfaces also have multiplicity one.

Recall that a Riemannian metric gg is said to be bumpy if every finite cover of a smooth immersed free boundary minimal hypersurface is non-degenerate. A bumpy metric gg 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 (Mn+1,M,g)(M^{n+1},\partial M,g) be a compact Riemannian manifold with boundary of dimension 3(n+1)73\leq(n+1)\leq 7. Let X be a kk-dimensional cubical subcomplex of I(m,j)I(m,j) and ZXZ\subset X be a subcomplex, and Φ0:X𝒞(M)\Phi_{0}:X\rightarrow\mathcal{C}(M) be a map continuous in the 𝐅\mathbf{F}-topology. Let Π\Pi be the associated (X,Z)(X,Z)-homotopy class of Φ0\Phi_{0}. Assume that

𝐋(Π)>max(maxxZ𝐌(Φ0(x)),0),\mathbf{L}(\Pi)>\max\big{(}\max_{x\in Z}\mathbf{M}(\partial\Phi_{0}(x)),0\big{)},

where we let h0h\equiv 0 in Section 3.1.

If gg is a strongly bumpy metric, then there exists a disjoint collection of smooth, connected, compact, two-sided, properly embedded, free boundary minimal hypersurfaces Σ=i=1NΣi\Sigma=\cup_{i=1}^{N}\Sigma_{i} so that

𝐋(Π)=i=1NArea(Σi), and index(Σ)=i=1Nindex(Σi)k.\mathbf{L}(\Pi)=\sum_{i=1}^{N}\mathrm{Area}(\Sigma_{i}),\text{ and }\mathrm{index}(\Sigma)=\sum_{i=1}^{N}\mathrm{index}(\Sigma_{i})\leq k.

In particular, each component of Σ\Sigma is two-sided and has exactly multiplicity one.

Proof.

Fix a sequence of positive numbers ϵk0\epsilon_{k}\rightarrow 0 as kk\rightarrow\infty. Recall that 𝒮(g)\mathcal{S}(g) is open and dense in C(M)C^{\infty}(M). Thus

j=1{f:ϵjf𝒮(g)}\bigcap_{j=1}^{\infty}\{f:\epsilon_{j}f\in\mathcal{S}(g)\}

is generic in the Baire sense. Pick an hh in this generic set with Mh0\int_{M}h\geq 0 (to be fixed at the end, in Part 5), and ϵ>0\epsilon>0 small enough so that

𝐋(Π)maxxZ𝐌(Φ0(x))>2ϵsupM|h|Vol(M).\mathbf{L}(\Pi)-\max_{x\in Z}\mathbf{M}(\partial\Phi_{0}(x))>2\epsilon\sup_{M}|h|\cdot\mathrm{Vol}(M).

Note that ϵjh𝒮(g)\epsilon_{j}\cdot h\in\mathcal{S}(g) for each jj.

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 ϵjh\epsilon_{j}\cdot h-hypersurface Σϵj=Ωϵj\Sigma_{\epsilon_{j}}=\partial\Omega_{\epsilon_{j}}; moreover, 𝒜h(Ωϵj)=𝐋ϵjh=𝐋ϵjh(Π)\mathcal{A}^{h}(\Omega_{\epsilon_{j}})=\mathbf{L}^{\epsilon_{j}h}=\mathbf{L}^{\epsilon_{j}h}(\Pi) and indexw(Σϵj)k\mathrm{index}_{w}(\Sigma_{\epsilon_{j}})\leq k. We also have 𝐋ϵjh𝐋(Π)\mathbf{L}^{\epsilon_{j}h}\rightarrow\mathbf{L}(\Pi) as jj\rightarrow\infty.

Now applying Theorem 2.9, by the strong bumpiness of gg, there exists a subsequence (still denoted by {ϵj}0\{\epsilon_{j}\}\rightarrow 0) such that Σj=Σϵj\Sigma_{j}=\Sigma_{\epsilon_{j}} converges to a smooth, compact, embedded, free boundary minimal hypersurface Σ\Sigma (with integer multiplicity). We denote by 𝒴\mathcal{Y} the set of points where the convergence fails to be smooth. By Theorem 2.9,

𝐌(Σ)=𝐋(Π), and index(Σ)k.\mathbf{M}(\Sigma)=\mathbf{L}(\Pi),\text{ and }\mathrm{index}(\Sigma)\leq k.

We now prove that every component of Σ\Sigma_{\infty} is two-sided and has multiplicity one. Without loss of generality, we may assume that Σ\Sigma_{\infty} has only one connected component with multiplicity mm\in\mathbb{N}. We will prove by contradiction.

Part 1:  We assume that Σ\Sigma_{\infty} is 2-sided; otherwise, we would consider the double cover of Σ\Sigma_{\infty} just like [Zhou19]*Proof of Theorem 4.1, Part 8. Let ν\nu be the global unit normal of Σ\Sigma and X𝔛~(M,Σ)X\in\widetilde{\mathfrak{X}}(M,\Sigma) be an extension of ν\nu. Suppose that ϕt\phi_{t} is a one-parameter family of diffeomorphisms generated by XX. For any domain UΣU\subset\Sigma and  small δ>0\delta>0, ϕt\phi_{t} produces a neighborhood UδU_{\delta} of UU with thickness δ\delta, i.e., Uδ={ϕt(x)|xU,|t|δ}U_{\delta}=\{\phi_{t}(x)\,|\,x\in U,|t|\leq\delta\}. We will use (t,x)(t,x) as coordinates on UδU_{\delta}. If UU is in the interior of Σ\Sigma, then  for δ\delta small UδU_{\delta} is the same as U×[δ,δ]U\times[-\delta,\delta] in the geodesic normal coordinates of Σ\Sigma. Now fix a domain UΣ𝒴U\subset\subset\Sigma\setminus\mathcal{Y}, by the convergence ΣjmΣ\Sigma_{j}\to m\Sigma, we know that for jj sufficiently large, ΣjUδ\Sigma_{j}\cap U_{\delta} can be decomposed to mm graphs over UU which can be ordered by height

uj1uj2ujm, and uji0, in smooth topology as j..u_{j}^{1}\leq u_{j}^{2}\leq\cdots\leq u_{j}^{m},\text{ and $u_{j}^{i}\rightarrow 0$, in smooth topology as $j\rightarrow\infty$.}.

Since Σj\Sigma_{j} is the boundary of some set Ωj\Omega_{j}, we know that the unit outer normal νj\nu_{j} of Ωj\Omega_{j} will alternate orientations along these graphs; see [Zhou19]*Proof of Theorem 4.1, Part 3.

Part 2: We first deal with an easier case: mm is an odd number. Hence m3m\geq 3. Let ff be any smooth function on Σ\Sigma and Z𝔛~(M,Σ)Z\in\mathfrak{\widetilde{X}}(M,\Sigma) be an extension of fνf\nu. Construct a family of hypersurfaces

Σj,t:={ϕ((1t)uj1+tujm,x)|xU},\Sigma_{j,t}:=\{\phi((1-t)u_{j}^{1}+tu_{j}^{m},x)|x\in U\},

where ϕ(t,x):=ϕt(x)\phi(t,x):=\phi_{t}(x). Then Σk,0\Sigma_{k,0} and Σk,1\Sigma_{k,1} are the bottom and top sheets of Σj\Sigma_{j}. Since Σj\Sigma_{j} is a free boundary ϵjh\epsilon_{j}h-hypersurface, we have

01[ddtΣj,t(divZϵjhZ,ν)𝑑n]𝑑t=0.\int_{0}^{1}\Big{[}\frac{d}{dt}\int_{\Sigma_{j,t}}(\mathrm{div}Z-\epsilon_{j}h\langle Z,\nu\rangle)\,d\mathcal{H}^{n}\Big{]}\,dt=0.

Then the computation in Appendix B gives that

0=\displaystyle 0= 01[Σj,t((X),(Z)Ric(X,Z)|A|2X,Z\displaystyle\int_{0}^{1}\Big{[}\int_{\Sigma_{j,t}}\Big{(}\langle\nabla^{\perp}(X^{\perp}),\nabla^{\perp}(Z^{\perp})\rangle-\operatorname{Ric}(X^{\perp},Z^{\perp})-|A|^{2}\langle X^{\perp},Z^{\perp}\rangle
ϵj(νh)X,Z)dn+Σj,tXZ,νMdn1+\displaystyle-\epsilon_{j}(\partial_{\nu}h)\langle X^{\perp},Z^{\perp}\rangle\Big{)}\,d\mathcal{H}^{n}+\int_{\partial\Sigma_{j,t}}\langle\nabla_{X^{\perp}}Z^{\perp},\nu_{\partial M}\rangle\,d\mathcal{H}^{n-1}+
+Σj,tΞ~1(X,Z,𝐇)dn+Σj,tΞ~2(X,Z,𝐇,𝜼,νM)dn1]dt,\displaystyle+\int_{\Sigma_{j,t}}\widetilde{\Xi}_{1}(X,Z,\mathbf{H})\,d\mathcal{H}^{n}+\int_{\partial\Sigma_{j,t}}\widetilde{\Xi}_{2}(X,Z,\mathbf{H},\bm{\eta},\nu_{\partial M})\,d\mathcal{H}^{n-1}\Big{]}\,dt,

where X|Σ=(ujmuj1)νX|_{\Sigma}=(u_{j}^{m}-u_{j}^{1})\nu, 𝜼\bm{\eta} is the co-normal of Σj,t\Sigma_{j,t}, and

|Ξ~1(X,Z,𝐇)|+|Ξ~2(X,Z,𝐇,𝜼,νM)| C(|X|(|𝐇+ϵjhν|+|𝜼νM|+|Z|)+|X|).|\widetilde{\Xi}_{1}(X,Z,\mathbf{H})|+|\widetilde{\Xi}_{2}(X,Z,\mathbf{H},\bm{\eta},\nu_{\partial M})|\leq C\big{(}|X|(|\mathbf{H}+\epsilon_{j}h\nu|+|\bm{\eta}-\nu_{\partial M}|+|Z^{\top}|)+|X^{\top}|\big{)}.

Here 𝐇\mathbf{H} is the mean curvature vector. Denote by φj=ujmuj1\varphi_{j}=u_{j}^{m}-u_{j}^{1}. By pulling everything back to UU, we obtain

0=\displaystyle 0= 01[U(φj,fRic(ν,ν)φjf|AΣ|2φjfϵj(νh)φjf)dn\displaystyle\int_{0}^{1}\Big{[}\int_{U}\Big{(}\langle\nabla\varphi_{j},\nabla f\rangle-\operatorname{Ric}(\nu,\nu)\varphi_{j}f-|A^{\Sigma}|^{2}\varphi_{j}f-\epsilon_{j}(\partial_{\nu}h)\varphi_{j}f\Big{)}\,d\mathcal{H}^{n}
ΣUhM(ν,ν)φjfdn1+UW~j(t)(φj,f)dn+ΣUw~j(t)(φj,f)dn1]dt,\displaystyle-\int_{\partial\Sigma\cap U}h^{\partial M}(\nu,\nu)\varphi_{j}f\,d\mathcal{H}^{n-1}+\int_{U}\widetilde{W}_{j}(t)(\varphi_{j},f)\,d\mathcal{H}^{n}+\int_{\partial\Sigma\cap U}\widetilde{w}_{j}(t)(\varphi_{j},f)\,d\mathcal{H}^{n-1}\Big{]}\,dt,

where

(4.1) W~j(t)(φj,f)ϵ~j(f,Σ,M)|φj|,w~j(t)(φj,f)ϵ~j(f,Σ,M)|φj|.\widetilde{W}_{j}(t)(\varphi_{j},f)\leq\widetilde{\epsilon}_{j}(f,\Sigma,M)|\varphi_{j}|,\ \widetilde{w}_{j}(t)(\varphi_{j},f)\leq\widetilde{\epsilon}_{j}(f,\Sigma,M)|\varphi_{j}|.

Here ϵ~j0\widetilde{\epsilon}_{j}\rightarrow 0 uniformly as jj\rightarrow\infty. Now letting Wj=01W~j(t)𝑑tW_{j}=\int_{0}^{1}\widetilde{W}_{j}(t)dt and wj=01w~j(t)𝑑tw_{j}=\int_{0}^{1}\widetilde{w}_{j}(t)dt, by Fubini theorem, we have

0=\displaystyle 0= U(φj,fRic(ν,ν)φjf|AΣ|2φjfϵj(νh)φjf)𝑑n\displaystyle\int_{U}\Big{(}\langle\nabla\varphi_{j},\nabla f\rangle-\operatorname{Ric}(\nu,\nu)\varphi_{j}f-|A^{\Sigma}|^{2}\varphi_{j}f-\epsilon_{j}(\partial_{\nu}h)\varphi_{j}f\Big{)}\,d\mathcal{H}^{n}
ΣUhM(ν,ν)φjf𝑑n1+UWj(φj,f)𝑑n+ΣUwj(φj,f)𝑑n1.\displaystyle-\int_{\partial\Sigma\cap U}h^{\partial M}(\nu,\nu)\varphi_{j}f\,d\mathcal{H}^{n-1}+\int_{U}W_{j}(\varphi_{j},f)\,d\mathcal{H}^{n}+\int_{\partial\Sigma\cap U}w_{j}(\varphi_{j},f)\,d\mathcal{H}^{n-1}.

Take qUq\in U. Let φ~j=φj/φj(q)\widetilde{\varphi}_{j}=\varphi_{j}/\varphi_{j}(q). Then the standard PDE theory implies that φ~j\widetilde{\varphi}_{j} converges smoothly to a positive φC(U)\varphi\in C^{\infty}(U) satisfying

(4.2) {LΣφ=0, on U,φ𝜼=hM(ν,ν)φ, along ΣU.\left\{\begin{aligned} &L_{\Sigma}\varphi=0,\text{\ \ on $U$},\\ &\frac{\partial\varphi}{\partial\bm{\eta}}=h^{\partial M}(\nu,\nu)\varphi,\text{\ \  along $\partial\Sigma\cap U$.}\end{aligned}\right.

Note that above argument works for any UΣ𝒴U\subset\subset\Sigma\setminus\mathcal{Y}. Taking an exhaustion of Σ𝒴\Sigma\setminus\mathcal{Y}, we can extend φ\varphi to Σ𝒴\Sigma\setminus\mathcal{Y} and such that

(4.3) {LΣφ=0, on Σ𝒴,φ𝜼=hM(ν,ν)φ, along Σ𝒴.\left\{\begin{aligned} &L_{\Sigma}\varphi=0,\text{\ \ on $\Sigma\setminus\mathcal{Y}$},\\ &\frac{\partial\varphi}{\partial\bm{\eta}}=h^{\partial M}(\nu,\nu)\varphi,\text{\ \  along $\partial\Sigma\setminus\mathcal{Y}$.}\end{aligned}\right.

Part 3: Next we use White’s local foliation argument to prove that φ\varphi extends smoothly across 𝒴\mathcal{Y}, and this will contradict the bumpy assumption of gg.

By the work for interior singularity in [Zhou19]*Proof of Theorem 4.1, Part 5, it suffices to show the uniform boundedness for p𝒴Mp\in\mathcal{Y}\cap\partial M. Since Σ\Sigma has empty touching set, then pΣp\in\partial\Sigma. Using the hh-hypersurface with free boundary (see Proposition D.1), we can also prove that φ\varphi is bounded. Then the classical PDE gives that φ\varphi is smooth across 𝒴\mathcal{Y}. Thus, we conclude that there is a positive Jacobi field on Σ\Sigma, which is a contradiction to the fact that gg is a strongly bumpy metric.

Part 4: We now take care the case when mm is even. Hence m2m\geq 2. Then without loss of generality we may assume that

H|Graph(ujm)(x)=ϵjh(x,ujm(x)), and H|Graph(uj1)(x)=ϵjh(x,uj1(x)), for x U.H|_{\mathrm{Graph}(u_{j}^{m})}(x)=-\epsilon_{j}h(x,u_{j}^{m}(x)),\text{ and }H|_{\mathrm{Graph}(u_{j}^{1})}(x)=\epsilon_{j}h(x,u_{j}^{1}(x)),\text{ for }x\in U.

Then by the argument in Part 2, we have

0=\displaystyle 0= U(φj,fRic(ν,ν)φjf|AΣ|2φjfϵj(νh)φjf+2ϵjhf)𝑑n\displaystyle\int_{U}\Big{(}\langle\nabla\varphi_{j},\nabla f\rangle-\operatorname{Ric}(\nu,\nu)\varphi_{j}f-|A^{\Sigma}|^{2}\varphi_{j}f-\epsilon_{j}(\partial_{\nu}h)\varphi_{j}f+2\epsilon_{j}hf\Big{)}\,d\mathcal{H}^{n}
ΣUhM(ν,ν)φjf𝑑n1+UWj(φj,f)𝑑n+ΣUwj(φj,f)𝑑n1.\displaystyle-\int_{\partial\Sigma\cap U}h^{\partial M}(\nu,\nu)\varphi_{j}f\,d\mathcal{H}^{n-1}+\int_{U}W_{j}(\varphi_{j},f)\,d\mathcal{H}^{n}+\int_{\partial\Sigma\cap U}w_{j}(\varphi_{j},f)\,d\mathcal{H}^{n-1}.

Here φj=ujmuj1\varphi_{j}=u^{m}_{j}-u^{1}_{j}; WW and ww are defined as the integral of W~\widetilde{W} and w~\widetilde{w} in (4.1). Fix a point qΣ𝒴q\in\Sigma\setminus\mathcal{Y}.

Case 1: lim supkφj(q)/ϵj=+\limsup_{k\rightarrow\infty}\varphi_{j}(q)/\epsilon_{j}=+\infty. Then the renormalizations φ~j=φj/φj(q)\widetilde{\varphi}_{j}=\varphi_{j}/\varphi_{j}(q) converges locally smoothly to a nontrivial function φ0\varphi\geq 0 on Σ𝒴\Sigma\setminus\mathcal{Y}, and by same reasoning as Part 2, we have

{LΣφ=0, on Σ𝒴,φ𝜼=hM(ν,ν)φ, along Σ𝒴.\left\{\begin{aligned} &L_{\Sigma}\varphi=0,\text{\ \ on $\Sigma\setminus\mathcal{Y}$},\\ &\frac{\partial\varphi}{\partial\bm{\eta}}=h^{\partial M}(\nu,\nu)\varphi,\text{\ \  along $\partial\Sigma\setminus\mathcal{Y}$.}\end{aligned}\right.

Case 2: lim supkφj(q)/ϵj<+\limsup_{k\rightarrow\infty}\varphi_{j}(q)/\epsilon_{j}<+\infty. Consider renormalizations φ~j=φj/ϵj\widetilde{\varphi}_{j}=\varphi_{j}/\epsilon_{j}. Then again by the same reasoning, φ~j\widetilde{\varphi}_{j} converges locally smoothly to a nonnegative φ0\varphi\geq 0 on Σ𝒴\Sigma\setminus\mathcal{Y}, and such that

(4.4) {LΣφ=2h|Σ, on Σ𝒴,φ𝜼=hM(ν,ν)φ, along Σ𝒴.\left\{\begin{aligned} &L_{\Sigma}\varphi=2h|_{\Sigma},\text{\ \ on $\Sigma\setminus\mathcal{Y}$},\\ &\frac{\partial\varphi}{\partial\bm{\eta}}=h^{\partial M}(\nu,\nu)\varphi,\text{\ \  along $\partial\Sigma\setminus\mathcal{Y}$.}\end{aligned}\right.

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 φ\varphi is smooth across 𝒴\mathcal{Y} 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 h𝒮(g)h\in\mathcal{S}(g), 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 gg is strongly bumpy. Recall that by \citelist[GWZ18][Wang19], there are only finitely many free boundary minimal hypersurfaces with Area𝐋(Π)\mathrm{Area}\leq\mathbf{L}(\Pi) and indexk\mathrm{index}\leq k, denoted by {Σ1,Σ2,,ΣN}\{\Sigma_{1},\Sigma_{2},\cdots,\Sigma_{N}\}. Then as in [Zhou19], we can take disjoint neighborhood Uj±ΣjU_{j}^{\pm}\subset\Sigma_{j} so that Uj±M=U_{j}^{\pm}\cap\partial M=\emptyset. Since all the small neighborhoods are disjoint, we can take a smooth function uu defined on Uj±\bigcup U_{j}^{\pm} whose supports are compact and

  • u|Uj+u|_{U_{j}^{+}} is nonnegative and is positive at some point;

  • u|Uju|_{U_{j}^{-}} is nonpositive and is negative at some point.

Then take h0C(M)h_{0}\in C^{\infty}(M) be an extension of Lu/2Lu/2 such that hh equals to 0 in a neighborhood of M\partial M. Recall that

j=1{f:ϵjf𝒮(g)}\bigcap_{j=1}^{\infty}\{f:\epsilon_{j}f\in\mathcal{S}(g)\}

is generic. Take hh in this generic set which is close to h0h_{0} as we wanted. Then the solution of (4.4) would be close to u/2u/2 on each Σj\Sigma_{j}, therefore it must change sign. Thus, such an hh 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 (Mn+1,M,g)(M^{n+1},\partial M,g) be a compact Riemannian manifold with boundary. Let XX be a cubical subcomplex of Im=[0,1]I^{m}=[0,1] for some mm\in\mathbb{N}. Given kk\in\mathbb{N}, a continuous map Φ:X𝒵n(M,M;2)\Phi:X\rightarrow\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}) is called a kk-sweepout if

0Φ(λ¯k)Hk(X,2),0\neq\Phi^{*}(\bar{\lambda}^{k})\in H^{k}(X,\mathbb{Z}_{2}),

where λ¯H1(𝒵n(M,M;2))=2\bar{\lambda}\in H^{1}(\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}))=\mathbb{Z}_{2} is the generator. A sweepout Φ\Phi is said to be admissible if it has no concentration of mass (see Definition 3.2). Denote by 𝒫k\mathcal{P}_{k} the set of all admissible kk-sweepouts. Then

Definition 4.3.

The kk-width of (M,M,g)(M,\partial M,g) is

ωk(M,g)=infΦ𝒫ksup{𝐌(Φ(x)):xdmn(Φ)},\omega_{k}(M,g)=\inf_{\Phi\in\mathcal{P}_{k}}\sup\{\mathbf{M}(\Phi(x)):x\in\mathrm{dmn}(\Phi)\},

where dmn(Φ)\mathrm{dmn}(\Phi) is the domain of Φ\Phi.

The kk-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 kk-width.

Theorem 4.4 ([LMN16]*§‘ 1.1).

There exists a constant a(n)>0a(n)>0 such that, for every compact Riemannian manifold (Mn+1,g)(M^{n+1},g) with (possibly empty) boundary, we have

(4.5) limpωp(M)p1n+1=a(n)vol(M)nn+1.\lim_{p\to\infty}\omega_{p}(M)p^{-\frac{1}{n+1}}=a(n)\mathrm{vol}(M)^{\frac{n}{n+1}}.

Assume from now on that 3(n+1)73\leq(n+1)\leq 7. It was later observed by Marques-Neves in [MN16] (see also [GLWZ19]*Section 4) that one can restrict to a subclass of 𝒫k\mathcal{P}_{k} in the definition of ωk(M,g)\omega_{k}(M,g). In particular, let 𝒫~k\widetilde{\mathcal{P}}_{k} denote those elements Φ𝒫k\Phi\in\mathcal{P}_{k} which is continuous under the 𝐅\mathbf{F}-topology, and whose domain X=dmn(Φ)X=\mathrm{dmn}(\Phi) has dimension kk (and is identical to its kk-skeleton). Then

ωk(M,g)=infΦ𝒫~ksup{𝐌(Φ(x)):xdmn(Φ)}.\omega_{k}(M,g)=\inf_{\Phi\in\widetilde{\mathcal{P}}_{k}}\sup\{\mathbf{M}(\Phi(x)):x\in\mathrm{dmn}(\Phi)\}.

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 kk\in\mathbb{N}, there exists a disjoint collection of smooth, connected, almost properly embedded, free boundary minimal hypersurfaces {Σik:i=1,,lk}\{\Sigma^{k}_{i}:i=1,\cdots,l_{k}\} with integer multiplicities {mik:i=1,,lk}\{m^{k}_{i}:i=1,\cdots,l_{k}\}\subset\mathbb{N}, such that

ωk(M,g)=i=1lkmikArea(Σik),i=1lkindex(Σik) k.\omega_{k}(M,g)=\sum_{i=1}^{l_{k}}m^{k}_{i}\cdot\mathrm{Area}(\Sigma^{k}_{i}),\ \ \sum_{i=1}^{l_{k}}\mathrm{index}(\Sigma^{k}_{i})\leq k.

Before stating the main theorem, we recall an observation by [MN17]*Corollary 3.4. Denote S1S^{1} by the unit circle.

Lemma 4.5 (\citelist[MN17]*Corollary 3.4[LMN16]*Proposition 2.12).

Let τ𝒵n(M,M;2)\tau\in\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}) so that its support is a properly embedded free boundary minimal hypersurface in (M,M,g)(M,\partial M,g). There exists ϵ¯\bar{\epsilon} sufficiently small, depending on τ\tau and MM so that every map Φ:S1𝒵n(M,M;2)\Phi:S^{1}\rightarrow\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}) with

Φ(S1)Bϵ¯(τ)={τ1𝒵n(M,M;2):(τ1,τ)<ϵ¯}.\Phi(S^{1})\subset B_{\bar{\epsilon}}^{\mathcal{F}}(\tau)=\{\tau_{1}\in\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}):\mathcal{F}(\tau_{1},\tau)<\bar{\epsilon}\}.

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 gg is bumpy, then for each kk\in\mathbb{N}, there exist a kk-dimensional cubical complex XX and a map Φ0,k:X𝒵n(M,M,𝐅;2)\Phi_{0,k}:X\rightarrow\mathcal{Z}_{n}(M,\partial M,\mathbf{F};\mathbb{Z}_{2}) continuous in the 𝐅\mathbf{F}-topology with Φ0,k𝒫~k\Phi_{0,k}\in\widetilde{\mathcal{P}}_{k}, such that

𝐋(Πk)=ωk(M,g),\mathbf{L}(\Pi_{k})=\omega_{k}(M,g),

where Πk=Π(Φ0,k)\Pi_{k}=\Pi(\Phi_{0,k}) is the class of all maps Φ:X𝒵n(M,M,𝐅;2)\Phi:X\rightarrow\mathcal{Z}_{n}(M,\partial M,\mathbf{F};\mathbb{Z}_{2}) continuous in the 𝐅\mathbf{F}-topology that are homotopic to Φ0,k\Phi_{0,k} in flat topology.

Moreover, there exists a pulled-tight (see [GLWZ19]*Theorem 5.8) minimizing sequence {Φi}i\{\Phi_{i}\}_{i\in\mathbb{N}} of Πk\Pi_{k} such that if Σ𝐂({Φi}i)\Sigma\in\mathbf{C}(\{\Phi_{i}\}_{i\in\mathbb{N}}) has support a compact, smooth, almost properly embedded, free boundary minimal hypersurface, then

Σ(M)=ωk(M,g), and indexw(support of Σ)k.\|\Sigma\|(M)=\omega_{k}(M,g),\ \text{ and }\ \mathrm{index}_{w}(\text{support of $\Sigma$})\leq k.

Now we are going to state and prove our main theorem.

Theorem 4.7 (same with Theorem 1.1).

If gg is a strongly bumpy metric and 3(n+1)73\leq(n+1)\leq 7, then for each kk\in\mathbb{N}, there exists a disjoint collection of compact, smooth, connected, properly embedded, two-sided, free boundary minimal hypersurfaces {Σik:i=1,,lk}\{\Sigma^{k}_{i}:i=1,\cdots,l_{k}\}, such that

ωk(M,g)=i=1lkArea(Σik),index(Σik)k.\omega_{k}(M,g)=\sum_{i=1}^{l_{k}}\mathrm{Area}(\Sigma^{k}_{i}),\ \ \mathrm{index}(\Sigma^{k}_{i})\leq k.

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 gg is bumpy, then there are only finitely many compact, almost properly embedded, free boundary minimal hypersurfaces with AreaΛ\mathrm{Area}\leq\Lambda and indexI\mathrm{index}\leq I for given Λ>0,I\Lambda>0,I\in\mathbb{N} by \citelist[GWZ18][Wang19]; see [ACS17] for strongly bumpy metrics.

Now we fix kk\in\mathbb{N} and omit the sub-index kk in the following. Take Π=[Φ0:X𝒵n(M,M,𝐅;2)]\Pi=[\Phi_{0}:X\rightarrow\mathcal{Z}_{n}(M,\partial M,\mathbf{F};\mathbb{Z}_{2})] with 𝐋(Π)=ωk\mathbf{L}(\Pi)=\omega_{k}.

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 {Φi}i\{\Phi_{i}\}_{i\in\mathbb{N}}, such that for ii sufficiently large, either |Φi(x)||\Phi_{i}(x)| is close to a regular min-max free boundary minimal hypersurface, or the mass 𝐌(Φi(x))\mathbf{M}(\Phi_{i}(x)) is strictly less than ωk(M,g)\omega_{k}(M,g).

We recall the following observation by [MN17]*Claim 6.2. Let 𝒮\mathcal{S} be the set of all stationary integral varifolds with Areaωk\mathrm{Area}\leq\omega_{k} whose support is a compact, smooth, almost properly embedded, free boundary minimal hypersurface with index(support)k\mathrm{index}(\text{support})\leq k. Consider the set 𝒯\mathcal{T} of all τ𝒵n(M,M;2)\tau\in\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}) with 𝐌(τ)ωk\mathbf{M}(\tau)\leq\omega_{k} and either τ=0\tau=0 or the support of τ\tau is a compact, smooth, properly embedded, free boundary minimal hypersurface with indexk\mathrm{index}\leq k. By the bumpy assumption, both sets 𝒮\mathcal{S} and 𝒯\mathcal{T} are finite. Moreover,

Claim 4 ([MN17]*Claim 6.2).

For every ϵ¯>0\bar{\epsilon}>0, there exists ϵ>0\epsilon>0 such that τ𝒵n(M,M;2)\tau\in\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}) with 𝐅(|τ|,𝒮) 3ϵ(τ,𝒯)<ϵ¯\mathbf{F}(|\tau|,\mathcal{S})\leq 3\epsilon\Rightarrow\mathcal{F}(\tau,\mathcal{T})<\bar{\epsilon}.

Proof of Claim 4.

Here since gg is strongly bumpy, then each element in 𝒮\mathcal{S} 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 {Φi}i\{\Phi_{i}\}_{i\in\mathbb{N}} be chosen as in Lemma 4.6. We choose ϵ¯\bar{\epsilon} as Lemma 4.5 so that every map Φ:S1𝒵n(M,M;2)\Phi:S^{1}\rightarrow\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}) with Φ(S1)Bϵ¯(T)\Phi(S^{1})\subset B_{\bar{\epsilon}}^{\mathcal{F}}(T) is homotopically trivial. According to such ϵ¯\bar{\epsilon}, we then choose ϵ>0\epsilon>0 by Claim 4. Take a sequence {ki}i\{k_{i}\}_{i\in\mathbb{N}}\rightarrow\infty so that

sup{𝐅(Φi(x),Φi(y)):αX(ki),x,yα}ϵ/2.\sup\{\mathbf{F}(\Phi_{i}(x),\Phi_{i}(y)):\alpha\in X(k_{i}),x,y\in\alpha\}\leq\epsilon/2.

Consider ZiZ_{i} to be the cubical subcomplex of X(ki)X(k_{i}) consisting of all cells αX(ki)\alpha\in X(k_{i}) such that

𝐅(|Φi(x)|,𝒮)ϵ, for every vertex x in α.\mathbf{F}(|\Phi_{i}(x)|,\mathcal{S})\geq\epsilon,\text{ for every vertex $x$ in $\alpha$}.

Hence 𝐅(|Φi(x)|,𝒮)ϵ/2\mathbf{F}(|\Phi_{i}(x)|,\mathcal{S})\geq\epsilon/2 for all xZix\in Z_{i}. Consider this sub-coordinating sequence {Φi|Zi}i\{\Phi_{i}|_{Z_{i}}\}_{i\in\mathbb{N}}. 𝐋({Φi|Zi})\mathbf{L}(\{\Phi_{i}|_{Z_{i}}\}) and 𝐂({Φi|Zi})\mathbf{C}(\{\Phi_{i}|_{Z_{i}}\}) are defined in the same way as in Section 3.1 with 𝒜h\mathcal{A}^{h} replaced by 𝐌\mathbf{M}.

Let Yi=XZi¯Y_{i}=\overline{X\setminus Z_{i}}. It then follows that

𝐅(|Φi(x)|,𝒮)<32ϵ, for all xYi.\mathbf{F}(|\Phi_{i}(x)|,\mathcal{S})<\frac{3}{2}\epsilon,\text{ for all $x\in Y_{i}$}.

We also denote Bi=YiZiB_{i}=Y_{i}\cap Z_{i}. In fact, BiB_{i} is the topological boundary of YiY_{i} and ZiZ_{i}. For a later purpose, we also consider the set

𝐁i= the union of all cells αZi such that αBi.\mathbf{B}_{i}=\text{ the union of all cells $\alpha\in Z_{i}$ such that $\alpha\cap B_{i}\neq\emptyset$}.

𝐁i\mathbf{B}_{i} can be thought of the “thickening” of BiB_{i} inside ZiZ_{i}.

Let λ=(Φi)λ¯H1(X,2)\lambda=(\Phi_{i})^{*}{\bar{\lambda}}\in H^{1}(X,\mathbb{Z}_{2}). Let Yi=Yi𝐁iY_{i}^{\prime}=Y_{i}\cup\mathbf{B}_{i} and Zi=Zi𝐁i¯Z_{i}^{\prime}=\overline{Z_{i}\setminus\mathbf{B}_{i}}, and i1:YiXi_{1}^{\prime}:Y_{i}^{\prime}\rightarrow X and i2:ZiXi^{\prime}_{2}:Z_{i}^{\prime}\rightarrow X be the inclusion maps. Then by Lemma 4.5, we have

(i1)(λ)=0H1(Yi;2), and (i2)(λk1)0Hk1(Zi;2).(i^{\prime}_{1})^{*}({\lambda})=0\in H^{1}(Y_{i}^{\prime};\mathbb{Z}_{2}),\text{ and }(i^{\prime}_{2})^{*}({\lambda}^{k-1})\neq 0\in H^{k-1}(Z_{i}^{\prime};\mathbb{Z}_{2}).

Now consider B~i=YiZi\widetilde{B}_{i}=Y_{i}^{\prime}\cap Z_{i}^{\prime} and the set

𝐁~i= the union of all cells αZi such that αB~i.\widetilde{\mathbf{B}}_{i}=\text{ the union of all cells $\alpha\in Z_{i}^{\prime}$ such that $\alpha\cap\widetilde{B}_{i}\neq\emptyset$}.

Let Y~i=Yi𝐁~i\widetilde{Y}_{i}=Y_{i}^{\prime}\cup\widetilde{\mathbf{B}}_{i} and Z~i=Zi𝐁~i¯\widetilde{Z}_{i}=\overline{Z_{i}^{\prime}\setminus\widetilde{\mathbf{B}}_{i}} and i~1:Y~iX\widetilde{i}_{1}:\widetilde{Y}_{i}\rightarrow X and i~2:Z~iX\widetilde{i}_{2}:\widetilde{Z}_{i}\rightarrow X be the inclusion maps. Then by Lemma 4.5, we also have

(i~1)(λ)=0H1(Y~i;2), and (i~2)(λk1)0Hk1(Z~i;2).(\widetilde{i}_{1})^{*}({\lambda})=0\in H^{1}(\widetilde{Y}_{i};\mathbb{Z}_{2}),\text{ and }(\widetilde{i}_{2})^{*}({\lambda}^{k-1})\neq 0\in H^{k-1}(\widetilde{Z}_{i};\mathbb{Z}_{2}).
Claim 5.

{Φi}\{\Phi_{i}\} can be deformed so that

 𝐋({Φi|Zi}i)<𝐋(Π)=ωk. \mathbf{L}(\{\Phi_{i}|_{Z_{i}}\}_{i\in\mathbb{N}})<\mathbf{L}(\Pi)=\omega_{k}.
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 2\mathbb{Z}_{2}-coefficients), together with Lemma 4.6, we also have the following dichotomy:

  • no element V𝐂({Φi|Zi}i)V\in\mathbf{C}(\{\Phi_{i}|_{Z_{i}}\}_{i\in\mathbb{N}}) is 2\mathbb{Z}_{2}-almost minimizing in small annuli with free boundary (see [LZ16]*Definition 4.19),

  • or

    (4.6) 𝐋({Φi|Zi}i)<𝐋(Π)=ωk.\mathbf{L}(\{\Phi_{i}|_{Z_{i}}\}_{i\in\mathbb{N}})<\mathbf{L}(\Pi)=\omega_{k}.

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 Φi\Phi_{i}, we can use the Discretization Theorem [LZ16]*Theorem 4.12 (or [GLWZ19]*Theorem 4.5) to get a sequence of discrete maps ϕij:X(kij)0𝒵n(M,M;2)\phi_{i}^{j}:X(k_{i}^{j})_{0}\rightarrow\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}). Taking j(i)j(i) sufficiently large and by assumptions, we can deform ϕij(i)\phi_{i}^{j(i)} and then use the Almgren extension [LZ16]*Theorem 4.14 to get a sequence of maps Ψi:X𝒵n(M,M;2)\Psi_{i}:X\rightarrow\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}) which is still a minimizing sequence of Π\Pi and

𝐋({Ψi|Zi}i)<𝐋(Π)=ωk.\mathbf{L}(\{\Psi_{i}|_{Z_{i}}\}_{i\in\mathbb{N}})<\mathbf{L}(\Pi)=\omega_{k}.

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 𝐁~iZi\widetilde{\mathbf{B}}_{i}\subset Z_{i}^{\prime}, we have

𝐋({Φi|𝐁~i}i)<ωk.\mathbf{L}(\{\Phi_{i}|_{\widetilde{\mathbf{B}}_{i}}\}_{i\in\mathbb{N}})<\omega_{k}.

Step II: Now we want to produce sweepouts in 𝒞(M)\mathcal{C}(M) by lifting to the double cover :𝒞(M)𝒵n(M,M;Z2)\partial:\mathcal{C}(M)\rightarrow\mathcal{Z}_{n}(M,\partial M;Z_{2}) so as to produce sweepouts satisfying the assumption of Theorem 4.1.

Note that Φii~1:Y~i𝒵n(M,M;2)\Phi_{i}\circ\widetilde{i}_{1}:\widetilde{Y}_{i}\rightarrow\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2}) is homotopically trivial. Then by the lifting criterion [Hat]*Proposition 1.33, there exist lifting maps Φ~i:Y~i(𝒞(M),𝐅)\widetilde{\Phi}_{i}:\widetilde{Y}_{i}\rightarrow(\mathcal{C}(M),\mathbf{F}) so that Φ~i=Φii~1\partial\circ\widetilde{\Phi}_{i}=\Phi_{i}\circ\widetilde{i}_{1}.

Lemma 4.8.

For ii large enough, if Π~i\widetilde{\Pi}_{i} is the (Y~i,𝐁~i)(\widetilde{Y}_{i},\widetilde{\mathbf{B}}_{i})-homotopy class associated with Φ~i\widetilde{\Phi}_{i}, then we have

𝐋(Π~i)𝐋(Π)>maxx𝐁~i𝐌(Φ~i(x)).\mathbf{L}(\widetilde{\Pi}_{i})\geq\mathbf{L}(\Pi)>\max_{x\in\widetilde{\mathbf{B}}_{i}}\mathbf{M}(\partial\widetilde{\Phi}_{i}(x)).
Proof of Lemma 4.8.

Fix ii large, so that

supx𝐁~i𝐌(Φi(x))<𝐋(Π),\sup_{x\in\widetilde{\mathbf{B}}_{i}}\mathbf{M}(\Phi_{i}(x))<\mathbf{L}(\Pi),

and we will omit the sub-index ii in the following proof.

If the conclusion were not true, then we can find a sequence of maps {Ψ~j:Y~(𝒞(M),𝐅)}Π~\{\widetilde{\Psi}_{j}:\widetilde{Y}\rightarrow(\mathcal{C}(M),\mathbf{F})\}\subset\widetilde{\Pi}, such that

(4.7) lim supjsup{𝐌(Ψ~j(x)):xY~}<𝐋(Π),\limsup_{j\rightarrow\infty}\sup\{\mathbf{M}(\partial\widetilde{\Psi}_{j}(x)):x\in\widetilde{Y}\}<\mathbf{L}(\Pi),

and homotopy maps {H~j:[0,1]×Y~𝒞(M)}\{\widetilde{H}_{j}:[0,1]\times\widetilde{Y}\rightarrow\mathcal{C}(M)\}, which are continuous in the flat topology, H~j(0,)=Φ~\widetilde{H}_{j}(0,\cdot)=\widetilde{\Phi}, H~j(1,)=Ψ~j\widetilde{H}_{j}(1,\cdot)=\widetilde{\Psi}_{j}, and

(4.8) lim supjsup{𝐅(H~j(t,x),Φ~(x)):t[0,1],x𝐁~}=0.\limsup_{j\rightarrow\infty}\sup\{\mathbf{F}(\widetilde{H}_{j}(t,x),\widetilde{\Phi}(x)):t\in[0,1],x\in\widetilde{\mathbf{B}}\}=0.

We now construct new sequences of maps {Ψj:X𝒵n(M,M;2)}j\{\Psi_{j}:X\rightarrow\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2})\}_{j\in\mathbb{N}} and {Hj:[0,1]×X𝒵n(M,M;2)}\{H_{j}:[0,1]\times X\rightarrow\mathcal{Z}_{n}(M,\partial M;\mathbb{Z}_{2})\} defined as:

  • Hj(t,x)=Φ(x)H_{j}(t,x)=\Phi(x) if xXY~x\in X\setminus\widetilde{Y};

  • Hj(t,x)=H~j(t(1s(x)),x)H_{j}(t,x)=\partial\circ\widetilde{H}_{j}(t(1-s(x)),x) if xY~x\in\widetilde{Y}, where s(x)=min{1,dist(x,Y)}s(x)=\min\{1,\operatorname{dist}(x,Y^{\prime})\};

  • Ψj=H~j(1,)\Psi_{j}=\widetilde{H}_{j}(1,\cdot).

Then Ψ\Psi is continuous and homotopic to Φ\Phi in the flat topology. Moreover, (4.7) and (4.8) give that

lim supjsupxY~𝐌(Ψj(x))<𝐋(Π).\limsup_{j\rightarrow\infty}\sup_{x\in\widetilde{Y}}\mathbf{M}(\Psi_{j}(x))<\mathbf{L}(\Pi).

Recall that Ψ(x)=Φ(x)\Psi(x)=\Phi(x) for xXY~x\in X\setminus\widetilde{Y}. Then by Claim 5,

𝐋({Ψj|XY~}i)=𝐋({Φ|XY~}i)<𝐋(Π)=ωk.\mathbf{L}(\{\Psi_{j}|_{X\setminus\widetilde{Y}}\}_{i\in\mathbb{N}})=\mathbf{L}(\{\Phi|_{X\setminus\widetilde{Y}}\}_{i\in\mathbb{N}})<\mathbf{L}(\Pi)=\omega_{k}.

Thus we conclude that

(4.9) lim supjsupxX𝐌(Ψj(x))<𝐋(Π)=ωk.\limsup_{j\rightarrow\infty}\sup_{x\in X}\mathbf{M}(\Psi_{j}(x))<\mathbf{L}(\Pi)=\omega_{k}.

Note that (4.8) also gives that {Ψj}\{\Psi_{j}\} has no concentration of mass. This implies that Ψ\Psi is also a kk-sweepout, so (4.9) contradicts the definition of ωk\omega_{k}. ∎

Step III: Now we are ready to prove Theorem 4.7.

For ii large enough in Lemma 4.8, Theorem 4.1 applied to Π~i\widetilde{\Pi}_{i} gives a disjoint collection of compact, smooth, connected, properly embedded, 2-sided, free boundary minimal hypersurfaces Σi=j=1NiΣi,j\Sigma_{i}=\cup^{N_{i}}_{j=1}\Sigma_{i,j}, such that

𝐋(Π~i)=j=1NiArea(Σi,j),  and index(Σi)k.\mathbf{L}(\widetilde{\Pi}_{i})=\sum_{j=1}^{N_{i}}\mathrm{Area}(\Sigma_{i,j}),\  \text{ and }\ \mathrm{index}(\Sigma_{i})\leq k.

Recall that 𝐋(Π~i)𝐋(Φi)𝐋(Π)=ωk\mathbf{L}(\widetilde{\Pi}_{i})\leq\mathbf{L}(\Phi_{i})\rightarrow\mathbf{L}(\Pi)=\omega_{k} as ii\rightarrow\infty.Together with Lemma 4.8, 𝐋(Π~i)ωk\mathbf{L}(\widetilde{\Pi}_{i})\rightarrow\omega_{k} as ii\rightarrow\infty. Counting the fact that there are only finitely many compact, smooth, properly embedded, free boundary hypersurfaces with Areaωk+1\mathrm{Area}\leq\omega_{k}+1 and indexk\mathrm{index}\leq k, for ii sufficiently large we have

𝐋(Π~i)=𝐋(Π~i+1)==ωk.\mathbf{L}(\widetilde{\Pi}_{i})=\mathbf{L}(\widetilde{\Pi}_{i+1})=\cdots=\omega_{k}.

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 n+1\mathbb{R}^{n+1} with the Gaussian metric by virtue of Theorem 4.7. We will prove that each kk-width is realized by a connected, embedded self-shrinker with multiplicity one.

Recall that an embedded hypersurface Σn+1\Sigma\subset\mathbb{R}^{n+1} is called a self-shrinker if and only if

 H=x,ν2, H=\frac{\langle x,\nu\rangle}{2},

where HH is the mean curvature with respect to the unit normal vector field ν\nu. It is equivalent to say that Σ\Sigma is minimal under the Gaussian metric (n+1,(4π)1e|x|22nδij)(\mathbb{R}^{n+1},(4\pi)^{-1}e^{-\frac{|x|^{2}}{2n}}\delta_{ij}). 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 n+1\mathbb{R}^{n+1} with generic metrics near the Gaussian metric. Then by passing to limits we will obtain complete minimial hypersurfaces in (n+1,(4π)1e|x|22nδij)(\mathbb{R}^{n+1},(4\pi)^{-1}e^{-\frac{|x|^{2}}{2n}}\delta_{ij}).  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 𝒢=(4π)1e|x|22nδij\mathcal{G}=(4\pi)^{-1}e^{-\frac{|x|^{2}}{2n}}\delta_{ij}. We first give the definition of volume spectrum for (n+1;𝒢)(\mathbb{R}^{n+1};\mathcal{G}).

Definition 5.1.

Given a sequence of compact domain K1KjK_{1}\subset\cdots\subset K_{j}\subset\cdots exhausting n+1\mathbb{R}^{n+1}, then for any positive integer kk, we define

ωk(n+1;𝒢)=limjωk(Kj;𝒢).\omega_{k}(\mathbb{R}^{n+1};\mathcal{G})=\lim_{j\rightarrow\infty}\omega_{k}(K_{j};\mathcal{G}).

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 Σ\Sigma be an embedded self-shrinker in n+1\mathbb{R}^{n+1}. We say that Σ\Sigma has Indexk\mathrm{Index}\geq k if there exists a kk-dimensional subspace WW of C(Σ)C^{\infty}(\Sigma) such that each nonzero fWf\in W has compact support and

 Σ(|f|2|A|2f212f2)e|x|24𝑑n<0. \int_{\Sigma}(|\nabla f|^{2}-|A|^{2}f^{2}-\frac{1}{2}f^{2})e^{-\frac{|x|^{2}}{4}}d\mathcal{H}^{n}<0.

Σ\Sigma has index kk if and only if Σ\Sigma has indexk\operatorname{index}\geq k but does not have index(k+1)\operatorname{index}\geq(k+1).

Now we are going to prove the main theorem in this section.

Proof of Theorem 1.6.

In the following, BR(0)B_{R}(0) always denotes the ball in n+1\mathbb{R}^{n+1} with radius RR under the Euclidean metric. Let {Rj}\{R_{j}\} be a sequence of positive numbers with RjR_{j}\rightarrow\infty. Denote by Ωj=BRj(0)\Omega_{j}=B_{R_{j}}(0). For each jj, we now take a perturbed metric gjg_{j} on Ωj\Omega_{j} so that

  • for any compact domain Ωn+1\Omega\subset\mathbb{R}^{n+1}, gj|Ω𝒢g_{j}|_{\Omega}\rightarrow\mathcal{G} smoothly;

  • under the metric gjg_{j}, BR(0)B_{R}(0) has mean concave boundary for each 3<RRj3<R\leq R_{j};

  • gj|Ωjg_{j}|_{\Omega_{j}} is strongly bumpy;

  • ωk(Ωj;gj)ωk(n+1;𝒢)\omega_{k}(\Omega_{j};g_{j})\rightarrow\omega_{k}(\mathbb{R}^{n+1};\mathcal{G}).

The last item can be satisfied because ωk(Ω;g)\omega_{k}(\Omega;g) depends continuously on the metric gg by Irie-Marques-Neves [IMN17]*Lemma 2.1.  Hence without loss of generality,

ωk(Ωj;gj)<Λk:=ω(n+1;𝒢)+1.\omega_{k}(\Omega_{j};g_{j})<\Lambda_{k}:=\omega(\mathbb{R}^{n+1};\mathcal{G})+1.

Since gjg_{j} is strongly bumpy, by Theorem 4.7, there exists a free boundary minimal hypersurface

(Σj,Σj)(Ωj,Ωj;gj)(\Sigma_{j},\partial\Sigma_{j})\subset(\Omega_{j},\partial\Omega_{j};g_{j})

so that Area(Σj;gj)=ωk(Ωj;gj)\mathrm{Area}(\Sigma_{j};g_{j})=\omega_{k}(\Omega_{j};g_{j}). By the compactness of minimal surfaces with bounded area and index (see [Sharp17]), Σj\Sigma_{j} subsequently, locally smoothly converges to a smooth minimal hypersurface Σ\Sigma in (n+1;𝒢)(\mathbb{R}^{n+1};\mathcal{G}) with multiplicity m1m\geq 1 away from a finite set 𝒲\mathcal{W}. Then using the fact of non-existence of stable minimal hypersurfaces in (n+1;𝒢)(\mathbb{R}^{n+1};\mathcal{G}), mm can only be 11. Thus Σj\Sigma_{j} locally smoothly converges to Σ\Sigma. We refer to [CM12] for more details about this kind of convergence. Then by the Frankel property for self-shrinkers [CCMS]*Corollary C.4, Σ\Sigma is connected. To finish the proof, we prove that no mass is lost in the convergence.

Claim 6.

There exist constants C,L>0C,L>0 depending only on nn so that for all R>LR>L and sufficiently large jj,

Area(ΣjBR(0);gj)<CΛkR12neR24(112n).\mathrm{Area}(\Sigma_{j}\setminus B_{R}(0);g_{j})<C\Lambda_{k}R^{\frac{1}{2}-n}e^{-\frac{R^{2}}{4}(1-\frac{1}{2n})}.
Proof of Claim 6.

For s>0s>0, let r(s)=14πset2/(4n)𝑑tr(s)=\frac{1}{\sqrt{4\pi}}\int_{s}^{\infty}e^{-t^{2}/(4n)}dt. Clearly,

(5.1) lims4πses2/(4n)r(s)=2n.\lim_{s\rightarrow\infty}\sqrt{4\pi}\cdot s\cdot e^{s^{2}/(4n)}r(s)=2n.

Hence we can take L>4nL>4n large enough so that for all s>Ls>L,

(5.2) 2n+110>4πses2/8r(s)>2n110.2n+\frac{1}{10}>\sqrt{4\pi}\cdot s\cdot e^{s^{2}/8}r(s)>2n-\frac{1}{10}.

For x3x\in\mathbb{R}^{3}, we define r(x)=r(|x|)r(x)=r(|x|). Then r(x)r(x) is the distance to \infty under (3;𝒢)(\mathbb{R}^{3};\mathcal{G}). Denote by \nabla and (j)\nabla^{(j)} the Levi-Civita connection associated with 𝒢\mathcal{G} and gjg_{j}, respectively. Then |r|𝒢=1|\nabla r|_{\mathcal{G}}=1 and

(5.3) 𝒢(e1r2/2,e2)=𝒢(e1,r)𝒢(e2,r)+r𝒢(e1r,e2).\mathcal{G}(\nabla_{e_{1}}\nabla r^{2}/2,e_{2})=\mathcal{G}(e_{1},\nabla r)\mathcal{G}(e_{2},\nabla r)+r\mathcal{G}(\nabla_{e_{1}}\nabla r,e_{2}).

By direct computations, rr=0\nabla_{\nabla r}\nabla r=0, and for e1,e2Tx(B|x|(0))e_{1},e_{2}\in T_{x}(\partial B_{|x|}(0)),

 𝒢(e1r,e2)=4πe|x|24n(|x|2n1|x|)𝒢(e1,e2), \mathcal{G}(\nabla_{e_{1}}\nabla r,e_{2})=\sqrt{4\pi}\cdot e^{\frac{|x|^{2}}{4n}}\big{(}\frac{|x|}{2n}-\frac{1}{|x|}\big{)}\mathcal{G}(e_{1},e_{2}),

which implies that for |x|>L>4n|x|>L>4n and eTx(B|x|(0))e\in T_{x}(\partial B_{|x|}(0)),

𝒢(er2/2,e)4π re|x|2/8|x|2n(12n|x|2)𝒢(e,e)>(114n)𝒢(e,e).\mathcal{G}(\nabla_{e}\nabla r^{2}/2,e)\geq\sqrt{4\pi}\cdot r\cdot e^{|x|^{2}/8}\cdot\frac{|x|}{2n}(1-\frac{2n}{|x|^{2}})\cdot\mathcal{G}(e,e)>(1-\frac{1}{4n})\mathcal{G}(e,e).

Here we used (5.2) in the last inequality. Together with (5.3), we have that for |x|>L|x|>L and any eTx3e\in T_{x}\mathbb{R}^{3},

 𝒢(er2/2,e)>(114n)𝒢(e,e). \mathcal{G}(\nabla_{e}\nabla r^{2}/2,e)>(1-\frac{1}{4n})\mathcal{G}(e,e).

Then can assume that for any eTx3e\in T_{x}\mathbb{R}^{3} and |x|>L|x|>L (by taking gjg_{j} close to 𝒢\mathcal{G}),

(5.4) 1+1/j>gj(r,r)>11/j and gj(e(j)(j)r2/2,e)>(114n)gj(e,e).1+1/j>g_{j}(\nabla r,\nabla r)>1-1/j\ \text{ and }\ g_{j}(\nabla_{e}^{(j)}\nabla^{(j)}r^{2}/2,e)>(1-\frac{1}{4n})\cdot g_{j}(e,e).

Given 0<t<u<10<t<u<1, we define two sequences of cut-off functions ϕt,ηu:\phi_{t},\eta_{u}:\mathbb{R}\rightarrow\mathbb{R} so that

ηu0; ηu(x)=1 for xu; ηu(x)=0 for x1;\displaystyle\eta_{u}^{\prime}\leq 0;\  \ \eta_{u}(x)=1\text{ for }x\leq u;\  \ \eta_{u}(x)=0\text{ for }x\geq 1;
ϕt0; ϕt(x)=0 for xt/2; ϕt(x)=1 for xt.\displaystyle\phi_{t}^{\prime}\geq 0;\ \  \phi_{t}(x)=0\text{ for }x\leq t/2;\ \  \phi_{t}(x)=1\text{ for }x\geq t.

Since Σj\Sigma_{j} is a minimal hypersurface in (Ωj;gj)(\Omega_{j};g_{j}), then by the divergence theorem, for any ρ(r(Rj),r(L))\rho\in(r(R_{j}),r(L)),

0\displaystyle 0 =ΣjdivΣj(j)[ϕt(r(x)r(Rj))ηu(r/ρ)(j)r2/2]𝑑μgj\displaystyle=\int_{\Sigma_{j}}\mathrm{div}^{(j)}_{\Sigma_{j}}\big{[}\phi_{t}(r(x)-r(R_{j}))\cdot\eta_{u}(r/\rho)\cdot\nabla^{(j)}r^{2}/2\big{]}\,d\mu_{g_{j}}
Σjϕtηu(r/ρ)1ρgj((j)r,(j)r2/2)+ϕtηu(r/ρ)divΣj(j)((j)r2/2)dμgj.\displaystyle\geq\int_{\Sigma_{j}}\phi_{t}\cdot\eta^{\prime}_{u}(r/\rho)\cdot\frac{1}{\rho}\cdot g_{j}(\nabla^{(j)}r,\nabla^{(j)}r^{2}/2)+\phi_{t}\cdot\eta_{u}(r/\rho)\cdot\mathrm{div}^{(j)}_{\Sigma_{j}}(\nabla^{(j)}r^{2}/2)\,d\mu_{g_{j}}.
Σjϕtηu(r/ρ)1ρ(1+1/j)r+ϕtηu(r/ρ)(n14)dμgj.\displaystyle\geq\int_{\Sigma_{j}}\phi_{t}\cdot\eta_{u}^{\prime}(r/\rho)\cdot\frac{1}{\rho}\cdot(1+1/j)r+\phi_{t}\cdot\eta_{u}(r/\rho)\cdot(n-\frac{1}{4})\,d\mu_{g_{j}}.

Here (5.4) is used in the last inequality. Letting t0t\rightarrow 0, we have

0Σjρddρηu(r/ρ)+ηu(r/ρ)(n12)dμgj.\displaystyle 0\geq\int_{\Sigma_{j}}-\rho\frac{d}{d\rho}\eta_{u}(r/\rho)+\eta_{u}(r/\rho)\cdot(n-\frac{1}{2})\,d\mu_{g_{j}}.

Therefore, we conclude that for ρ[r(Rj),r(L)]\rho\in[r(R_{j}),r(L)] and sufficiently large jj,

(n12)Ij(ρ)ρddρIj(ρ),(n-\frac{1}{2})\cdot I_{j}(\rho)\leq\rho\frac{d}{d\rho}I_{j}(\rho),

where Ij(ρ)=Σjηu(r/ρ)I_{j}(\rho)=\int_{\Sigma_{j}}\eta_{u}(r/\rho). Such an inequality implies that Ij(ρ)ρ12nI_{j}(\rho)\rho^{\frac{1}{2}-n} is monotone increasing. Thus, for ρ[r(Rj),r(L)]\rho\in[r(R_{j}),r(L)],

(5.5) Ij(ρ)Ij(r(L))(r(L))12nρn12.I_{j}(\rho)\leq I_{j}(r(L))(r(L))^{\frac{1}{2}-n}\rho^{n-\frac{1}{2}}.

Now letting u1u\rightarrow 1 in ηu\eta_{u}, then

Ij(r(R))Area({xΣj:r(|x|)<r(R)};gj)=Area(ΣjBR(0);gj).I_{j}(r(R))\rightarrow\mathrm{Area}(\{x\in\Sigma_{j}:r(|x|)<r(R)\};g_{j})=\mathrm{Area}(\Sigma_{j}\setminus B_{R}(0);g_{j}).

Together with (5.1), the inequality in (5.5) becomes

Area(ΣjBR;gj)Λk (r(L))12nR12neR24(112n)\mathrm{Area}(\Sigma_{j}\setminus B_{R};g_{j})\leq\Lambda_{k}\cdot (r(L))^{\frac{1}{2}-n}R^{\frac{1}{2}-n}e^{-\frac{R^{2}}{4}(1-\frac{1}{2n})}

for all R(L,Rj)R\in(L,R_{j}). The proof of Claim 6 is finished by taking C=(r(L))12nC=(r(L))^{\frac{1}{2}-n}. ∎

To proceed the proof of theorem, it follows from Claim 6 that

Area(Σ;𝒢)=limjArea(Σj;gj)=limjωk(Ωj;gj)=ωk(3;𝒢).\mathrm{Area}(\Sigma;\mathcal{G})=\lim_{j\rightarrow\infty}\mathrm{Area}(\Sigma_{j};g_{j})=\lim_{j\rightarrow\infty}\omega_{k}(\Omega_{j};g_{j})=\omega_{k}(\mathbb{R}^{3};\mathcal{G}).

We conclude that Theorem 1.6 is finished. ∎

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 Σ\Sigma be an embedded self-shrinker in n+1\mathbb{R}^{n+1}. Throughout, LΣ=ΔΣ+|A|2+12x2L_{\Sigma}=\Delta_{\Sigma}+|A|^{2}+\frac{1}{2}-\frac{x}{2}\cdot\nabla will be the Jacobi operator from the second variation formula. Such an operator is associated with a bi-linear form:

 𝔅(u,v)=Σ(u,v|A|2uv12uv)e|x|24. \mathfrak{B}(u,v)=\int_{\Sigma}(\langle\nabla u,\nabla v\rangle-|A|^{2}uv-\frac{1}{2}uv)e^{-\frac{|x|^{2}}{4}}.

Then bottom of the spectrum μ1\mu_{1} of Σ\Sigma is defined as

 μ1:=inff𝔅(f,f)Σf2e|x|24, \mu_{1}:=\inf_{f}\frac{\mathfrak{B}(f,f)}{\int_{\Sigma}f^{2}e^{-\frac{|x|^{2}}{4}}},

where the infimum is taken over smooth functions ff with compact support. Since Σ\Sigma may be noncompact, we allow the possibility that μ1=\mu_{1}=-\infty.

In the following we will focus on the index of a self-shrinker of the form Σ×\Sigma\times\mathbb{R}, where Σ\Sigma is a lower dimensional self-shrinker. We will always use xx to denote a point on Σ\Sigma and use yy to denote a point in \mathbb{R}. We will also use LΣL_{\Sigma} and LΣ×L_{\Sigma\times\mathbb{R}} to denote the second variational operator on Σ\Sigma and Σ×\Sigma\times\mathbb{R} respectively. Note that

(5.6) LΣ×=LΣ+y212yy.L_{\Sigma\times\mathbb{R}}=L_{\Sigma}+\partial_{y}^{2}-\frac{1}{2}y\partial_{y}.

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 Σ\Sigma is a self-shrinker with finite index, then Σ×\Sigma\times\mathbb{R} is a self-shrinker with finite index.

We need the following correspondence between the eigenfunctions and eigenvalues on Σ\Sigma and Σ×\Sigma\times\mathbb{R} respectively. In the following the eigenvalues and eigenfunctions are under the Dirichlet boundary conditions.

Let ΩΣ\Omega\subset\Sigma be a bound domain with smooth boundary. Denote by 𝒲01,2(Ω)\mathcal{W}^{1,2}_{0}(\Omega) the closure of

 {fC(Σ):sptfIntΩ} \{f\in C^{\infty}(\Sigma):\operatorname{spt}f\subset\mathrm{Int}\ \Omega\}

in the topology of 𝒲1,2(Σ)\mathcal{W}^{1,2}(\Sigma). Then the self-adjoint operator LΣL_{\Sigma} has discrete Dirichlet eigenvalues

μ1(Ω)<μ2(Ω)μ3(Ω)+,\mu_{1}(\Omega)<\mu_{2}(\Omega)\leq\mu_{3}(\Omega)\leq\cdots\rightarrow+\infty,

and associated eigenfunctions {fj}𝒲01,2(Ω)\{f_{j}\}\subset\mathcal{W}^{1,2}_{0}(\Omega). Moreover, {fj}\{f_{j}\} forms a complete basis of the weighted L2(Ω)L^{2}(\Omega); see [Str08]*§11.3 Theorem 2.

Proposition 5.4.

Suppose ΩΣ\Omega\subset\Sigma is a bounded compact subset with smooth boundary. Suppose μ1(Ω)<μ2(Ω)\mu_{1}(\Omega)<\mu_{2}(\Omega)\leq\cdots are eigenvalues of LΣL_{\Sigma} on Ω\Omega, with associated eigenfunctions {ϕi}i=1\{\phi_{i}\}_{i=1}^{\infty}; suppose ν1T<ν2T\nu_{1}^{T}<\nu_{2}^{T}\leq\cdots are eigenvalues of y212yy\partial_{y}^{2}-\frac{1}{2}y\partial_{y} on [T,T][-T,T]\subset\mathbb{R}, with associated eigenfunctions {ψj}j=1\{\psi_{j}\}_{j=1}^{\infty}. Then the eigenvalues of LΣ×L_{\Sigma\times\mathbb{R}} on Ω×[T,T]\Omega\times[-T,T] are {μi(Ω)+νjT}i,j=1\{\mu_{i}(\Omega)+\nu_{j}^{T}\}_{i,j=1}^{\infty}, with associated eigenfunctions ϕiψj\phi_{i}\psi_{j}.

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 ϕiψj\phi_{i}\psi_{j} are eigenfunctions of LΣ×L_{\Sigma\times\mathbb{R}}, with eigenvalue μi(Ω)+νjT\mu_{i}(\Omega)+\nu_{j}^{T}.

Recall that {ϕi}i=1\{\phi_{i}\}_{i=1}^{\infty} is a basis in the weighted L2(Ω)L^{2}(\Omega) sense. Similarly, {ψi}i=1\{\psi_{i}\}_{i=1}^{\infty} is a basis in the weighted L2([T,T])L^{2}([-T,T]) sense. Therefore, {ϕiψj}i,j=1\{\phi_{i}\psi_{j}\}_{i,j=1}^{\infty} is a basis in the weighted L2(Ω×[T,T])L^{2}(\Omega\times[-T,T]) sense. This shows that {ϕiψj}i,j=1\{\phi_{i}\psi_{j}\}_{i,j=1}^{\infty} 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 Σ\Sigma has index k<k<\infty and we are going to prove that μ1(Σ)>\mu_{1}(\Sigma)>-\infty.

By Definition 5.2, there exists a kk-dimensional subspace WC(Σ)W\subset C^{\infty}(\Sigma) so that each non-zero fWf\in W has compact support and

 𝔅(f,f)<0. \mathfrak{B}(f,f)<0.

Without loss of generality, we may assume that sptfBR(0)Σ\operatorname{spt}f\subset B_{R}(0)\cap\Sigma for all fWf\in W. Denote by κ0=supΣB2R(0)|A(x)|+1\kappa_{0}=\sup_{\Sigma\cap B_{2R}(0)}|A(x)|+1.

We pause to take a cut-off function η:[0,1]\eta:\mathbb{R}\rightarrow\mathbb{[}0,1] and a universal constant C0>1C_{0}>1 so that

C0>η(t)0 and |η′′(t)|<C0 for t;η(t)=0 for t1; η(t)=1 for t2.C_{0}>\eta^{\prime}(t)\geq 0\text{ and }|\eta^{\prime\prime}(t)|<C_{0}\text{ for }t\in\mathbb{R};\ \ \ \eta(t)=0\text{ for }t\leq 1;\ \ \  \eta(t)=1\text{ for }t\geq 2.

Let κ=2nC0κ02\kappa=2nC_{0}\kappa^{2}_{0}. We now assume on the contrary that μ1(Σ)=\mu_{1}(\Sigma)=-\infty. Then there exist fC(Σ)f\in C^{\infty}(\Sigma) has compact support and satisfies

(5.7) Σ(|f|2|A|2f212f2)e|x|24<4κΣf2e|x|24.\int_{\Sigma}(|\nabla f|^{2}-|A|^{2}f^{2}-\frac{1}{2}f^{2})e^{-\frac{|x|^{2}}{4}}<-4\kappa\int_{\Sigma}f^{2}e^{-\frac{|x|^{2}}{4}}.

Define ϕ(x)=η(|x|/R)\phi(x)=\eta(|x|/R). Then ϕ=1Rη(|x|)x|x|\nabla\phi=\frac{1}{R}\cdot\eta^{\prime}(|x|)\cdot\frac{x^{\top}}{|x|} and

(5.8) Δϕ\displaystyle\Delta\phi =divΣ[η(|x|)x|x|1R]\displaystyle=\mathrm{div}_{\Sigma}\big{[}\eta^{\prime}(|x|)\cdot\frac{x^{\top}}{|x|}\cdot\frac{1}{R}\big{]}
=η′′(|x|)|x|2|x|21R2η(|x|/R)|x|2|x|31R+1R(n2H2)η(|x|)/|x|\displaystyle=\eta^{\prime\prime}(|x|)\frac{|x^{\top}|^{2}}{|x|^{2}}\cdot\frac{1}{R^{2}}-\eta^{\prime}(|x|/R)\frac{|x^{\top}|^{2}}{|x|^{3}}\cdot\frac{1}{R}+\frac{1}{R}(n-2H^{2})\eta^{\prime}(|x|)/|x|
C0C0/|x|nC0κ02/|x|κ.\displaystyle\geq-C_{0}-C_{0}/|x|-nC_{0}\kappa_{0}^{2}/|x|\geq-\kappa.

Here ()(\cdot)^{\top} is the projection to TxΣT_{x}\Sigma. The second equality used the fact that divΣ(x)=n2H2\mathrm{div}_{\Sigma}(x^{\top})=n-2H^{2}. The first inequality follows from |H|2n|A|2nκ02|H|^{2}\leq n|A|^{2}\leq n\kappa_{0}^{2} for |x|2R|x|\leq 2R. For |x|>2R|x|>2R, Δϕ=0\Delta\phi=0.

Then by a direct computation, we have

(5.9) Σ|(ϕf)|2e|x|24\displaystyle\int_{\Sigma}|\nabla(\phi f)|^{2}e^{-\frac{|x|^{2}}{4}} =Σ(|f|2ϕ2+|ϕ|2f2+12ϕ2,f2)e|x|24\displaystyle=\int_{\Sigma}(|\nabla f|^{2}\phi^{2}+|\nabla\phi|^{2}f^{2}+\frac{1}{2}\langle\nabla\phi^{2},\nabla f^{2}\rangle)e^{-\frac{|x|^{2}}{4}}
=Σ(|f|2+|ϕ|2f212f2Δϕ2+14x,ϕ2)e|x|24\displaystyle=\int_{\Sigma}(|\nabla f|^{2}+|\nabla\phi|^{2}f^{2}-\frac{1}{2}f^{2}\Delta\phi^{2}+\frac{1}{4}\langle x,\nabla\phi^{2}\rangle)e^{-\frac{|x|^{2}}{4}}
Σ[|f|2+κf2+12Rϕη(|x|)|x|2|x|]e|x|24\displaystyle\leq\int_{\Sigma}\Big{[}|\nabla f|^{2}+\kappa f^{2}+\frac{1}{2R}\cdot\phi\eta^{\prime}(|x|)\cdot\frac{|x^{\top}|^{2}}{|x|}\Big{]}e^{-\frac{|x|^{2}}{4}}
Σ(|f|2+2κf2)e|x|24.\displaystyle\leq\int_{\Sigma}(|\nabla f|^{2}+2\kappa f^{2})e^{-\frac{|x|^{2}}{4}}.

Here the first inequality is from the divergence theorem and ηC0\eta^{\prime}\leq C_{0}. (5.8) is used in the second inequality. In the last inequality, we note that η=0\eta^{\prime}=0 for |x|2R|x|\geq 2R.

Now we prove that ϕf\phi f has contribution for index. Indeed,

𝔅(ϕf,ϕf)\displaystyle\mathfrak{B}(\phi f,\phi f) =Σ(|(ϕf)|2|A|2f2ϕ212f2ϕ2)e|x|24\displaystyle=\int_{\Sigma}(|\nabla(\phi f)|^{2}-|A|^{2}f^{2}\phi^{2}-\frac{1}{2}f^{2}\phi^{2})e^{-\frac{|x|^{2}}{4}}
Σ(|f|2+2κf2|A|2f2)e|x|24+ΣB2R(0)|A|2f2e|x|24\displaystyle\leq\int_{\Sigma}(|\nabla f|^{2}+2\kappa f^{2}-|A|^{2}f^{2})e^{-\frac{|x|^{2}}{4}}+\int_{\Sigma\cap B_{2R}(0)}|A|^{2}f^{2}e^{-\frac{|x|^{2}}{4}}
Σ(2κ+1)f2e|x|24+ΣB2R(0)κ02f2e|x|24\displaystyle\leq\int_{\Sigma}(-2\kappa+1)f^{2}e^{-\frac{|x|^{2}}{4}}+\int_{\Sigma\cap B_{2R}(0)}\kappa_{0}^{2}f^{2}e^{-\frac{|x|^{2}}{4}}
κΣf2ϕ2e|x|24.\displaystyle\leq-\kappa\int_{\Sigma}f^{2}\phi^{2}e^{-\frac{|x|^{2}}{4}}.

Here we used (5.9) in the first inequality. The second one is from (5.7) and the definition of κ0\kappa_{0}.

Note that ϕf\phi f and ψW\psi\in W have disjoint support set. This implies that Σ\Sigma has indexk+1\operatorname{index}\geq k+1, which lead to a contradiction. This finishes Part I.

Part II: We now prove Σ×\Sigma\times\mathbb{R} has finite index.

It suffices to prove that for any compact domain Ω×[T,T]Σ×\Omega\times[-T,T]\subset\Sigma\times\mathbb{R}, the Dirichlet eigenproblem of LΣ×L_{\Sigma\times\mathbb{R}} has index bounded from above uniformly. Recall that μ1\mu_{1} is the first eigenvalue of LΣL_{\Sigma}. Suppose μ1(Ω)<μ2(Ω)\mu_{1}(\Omega)<\mu_{2}(\Omega)\leq\cdots are eigenvalues of LΣL_{\Sigma} on Ω\Omega, which correspond to the eigenfunctions {ϕi}i=1\{\phi_{i}\}_{i=1}^{\infty}; suppose ν1T<ν2T\nu_{1}^{T}<\nu_{2}^{T}\leq\cdots are Dirichlet eigenvalues of y212yy\partial_{y}^{2}-\frac{1}{2}y\partial_{y} (=:L0=:L_{0}) on [T,T][-T,T]\subset\mathbb{R}, which correspond to the eigenfunctions {ψj}j=1\{\psi_{j}\}_{j=1}^{\infty}. Recall that the νjT\nu_{j}^{T} converges to the eigenvalues of L0L_{0} on \mathbb{R} as TT\rightarrow\infty; see [CM2020]*Lemma 6.1.

We first note that by definition, μ1μ1(Ω)\mu_{1}\leq\mu_{1}(\Omega) for any compact domain Ω\Omega. Secondly, note that the eigenvalues of L0L_{0} on \mathbb{R} are given by non-negative half-integers with multiplicity one, i.e. νjTj12\nu^{T}_{j}\rightarrow\frac{j-1}{2} as TT\rightarrow\infty. Thus we can take T0>0T_{0}>0 so that

(5.10) νjT>μ1\nu_{j}^{T}>-\mu_{1} for all j>12μ1j>1-2\mu_{1} and T>T0T>T_{0}.

To proceed the proof, we take T>T0T>T_{0}. By Proposition 5.4, the Dirichlet eigenvalues of LΣ×L_{\Sigma\times\mathbb{R}} on Ω×[T,T]\Omega\times[-T,T] are {μi(Ω)+νjT}i,j=1\{\mu_{i}(\Omega)+\nu_{j}^{T}\}_{i,j=1}^{\infty}, which correspond to the eigenfunctions ϕiψj\phi_{i}\psi_{j}.

Since Σ\Sigma has finite index kk, the index of Ω\Omega is bounded from above by kk uniformly. From Part I, μ1\mu_{1}\neq-\infty. This implies that, μi(Ω)+νjT\mu_{i}(\Omega)+\nu_{j}^{T} must be non-negative if j>12μ1j>1-2\mu_{1} . So the index of Ω×[T,T]\Omega\times[-T,T] is bounded from above by an uniform constant k(12μ1)k(1-2\mu_{1}). This completes the proof. ∎

Appendix A Removing singularity for weakly stable free boundary hh-hypersurfaces

Theorem A.1 (cf. [ACS17]*Theorem 27).

Let (Mn+1,M,g)(M^{n+1},\partial M,g) be a compact Riemannian manifold with boundary of dimension 3(n+1)73\leq(n+1)\leq 7. Given hC(M)h\in C^{\infty}(M) and ΣBϵ(p){p}\Sigma\subset B_{\epsilon}(p)\setminus\{p\} an almost embedded free boundary hh-hypersurface with ΣIntMBϵ(p){p}=\partial\Sigma\cap\mathrm{Int}M\cap B_{\epsilon}(p)\setminus\{p\}=\emptyset, assume that Σ\Sigma is weakly stable in Bϵ(p){p}B_{\epsilon}(p)\setminus\{p\} as in Remark 2.7. If [Σ][\Sigma] represents a varifold of bounded first variation in Bϵ(p)B_{\epsilon}(p), then Σ\Sigma extends smoothly across pp as an almost embedded hypersurface in Bϵ(p)B_{\epsilon}(p).

Proof.

Given any sequence of positive λi0\lambda_{i}\rightarrow 0, consider the blowups {𝝁p,λi(Σ)𝝁p,λi(M)}\{\bm{\mu}_{p,\lambda_{i}}(\Sigma)\subset\bm{\mu}_{p,\lambda_{i}}(M)\}, where 𝝁p,λi(x)=xpλi\bm{\mu}_{p,\lambda_{i}}(x)=\frac{x-p}{\lambda_{i}}.  Since Σ\Sigma has bounded first variation, 𝝁p,λi(Σ)\bm{\mu}_{p,\lambda_{i}}(\Sigma) converges (up to a subsequence) to a stationary integral rectifiable cone CC in a half Euclidean space +n+1=TpM\mathbb{R}^{n+1}_{+}=T_{p}M with free boundary on +n+1=Tp(M)\partial\mathbb{R}^{n+1}_{+}=T_{p}(\partial M). By weak stability and Theorem 2.9, the convergence is locally smooth and graphical away from the origin, so CC is an integer multiple of some embedded minimal hypercone with free boundary. Hence the reflection of CC across +n+1\partial\mathbb{R}^{n+1}_{+} is a stable minimal cone in n+1{0}\mathbb{R}^{n+1}\setminus\{0\}, and hence a plane with integer multiplicities. Therefore, we conclude that C=mTp(M)C=m\cdot T_{p}(\partial M) or 2mP2m\cdot P for some hyperplane P+n+1P\subset\mathbb{R}^{n+1}_{+} with P+n+1P\perp\partial\mathbb{R}^{n+1}_{+}. Here m=Θn(Σ,p)m=\Theta^{n}(\Sigma,p). Note that PP may depend on the choice of {λi}\{\lambda_{i}\}.

If C=mTp(M)C=m\cdot T_{p}(\partial M), then the argument in [Zhou19]*Theorem B.1 implies the removability of pp.

If C=2mPC=2m\cdot P for some half-hyperplane PP, then by the locally smooth and graphical convergence, there exists σ0>0\sigma_{0}>0 small enough, such that for any 0<σσ00<\sigma\leq\sigma_{0}, Σ\Sigma has an mm-sheeted, ordered, graphical decomposition in the annulus Aσ/2,σ(p)=MBσ(p)Bσ/2(p)A_{\sigma/2,\sigma}(p)=M\cap B_{\sigma}(p)\setminus B_{\sigma/2}(p):

ΣAσ/2,σ(p)=i=1mΣi(σ).\Sigma\cap A_{\sigma/2,\sigma}(p)=\bigcup_{i=1}^{m}\Sigma_{i}(\sigma).

Here each Σi(σ)\Sigma_{i}(\sigma) is a graph over Aσ/2,σ(p)PA_{\sigma/2,\sigma}(p)\cap P. We can continue each Σi(σ)\Sigma_{i}(\sigma) all the way to Bσ0(p){p}B_{\sigma_{0}}(p)\setminus\{p\}, and denote the continuation by Σi\Sigma_{i}. Then each Σi\Sigma_{i} is a free boundary hh-hypersurface in MBΣ0(p){p}M\cap B_{\Sigma_{0}}(p)\setminus\{p\} and Θn(Σi,p)=1/2\Theta^{n}(\Sigma_{i},p)=1/2.  By the Allard type regularity theorem for rectifiable varifolds with free boundary and bounded mean curvature [GJ86]*Theorem 4.13, Σi\Sigma_{i} extends as a C1,αC^{1,\alpha} free boundary hypersurface across pp for some α(0,1)\alpha\in(0,1). Higher regularity of Σi\Sigma_{i} follows from the prescribing mean curvature equation and elliptic regularity. ∎

Appendix B The second variation of 𝒜h\mathcal{A}^{h} for smooth hypersurfaces

Our goal is to derive the second variation of 𝒜h\mathcal{A}^{h} for arbitrary hypersurfaces with boundary in (Mn+1,M,g)(M^{n+1},\partial M,g). Note that we always assume (Mn+1,M,g)(M^{n+1},\partial M,g) is isometrically embedded in some closed (n+1)(n+1)-dimensional Riemannian manifold (M~,g~)L(\widetilde{M},\widetilde{g})\subset\mathbb{R}^{L}. Let (Σn,Σ)(M,M)(\Sigma^{n},\partial\Sigma)\subset(M,\partial M) be an embedded hypersurface in MM with boundary on M\partial M. We consider a two-parameter family of ambient variations Σ(t,s)\Sigma(t,s) of Σ=Σ(0,0)\Sigma=\Sigma(0,0) defined by

Σ(t,s)=Ft,s(Σ)=ΨsΦt(Σ),\Sigma(t,s)=F_{t,s}(\Sigma)=\Psi_{s}\circ\Phi_{t}(\Sigma),

where Φt\Phi_{t} and Ψs\Psi_{s} are the flows generated by compactly supported vector fields XX and ZZ in L\mathbb{R}^{L}, respectively. We assume both X,Z𝔛~(M,Σ)X,Z\in\widetilde{\mathfrak{X}}(M,\Sigma). Therefore each Σ(t,s)\Sigma(t,s) is an embedded hypersurface in M~\widetilde{M} with boundary lying on M\partial M.

We have

Ft,st(x)|t=s=0=X(x), Ft,ss(x)|t=s=0=Z(x),and  tsFt,s(x)|t=s=0=DXZ(x).\frac{\partial F_{t,s}}{\partial t}(x)\Big{|}_{t=s=0}=X(x),\  \frac{\partial F_{t,s}}{\partial s}(x)\Big{|}_{t=s=0}=Z(x),\ \text{and } \frac{\partial}{\partial t}\frac{\partial}{\partial s}F_{t,s}(x)\Big{|}_{t=s=0}=D_{X}Z(x).

The computation in [ACS17]*Appendix A gives that

(B.1) ts|t=s=0n(Σ(t,s))=t|t=0Σ(t,0)divZ𝑑n\displaystyle\frac{\partial}{\partial t}\frac{\partial}{\partial s}\Big{|}_{t=s=0}\mathcal{H}^{n}(\Sigma(t,s))=\frac{\partial}{\partial t}\Big{|}_{t=0}\int_{\Sigma(t,0)}\mathrm{div}Z\,d\mathcal{H}^{n}
=\displaystyle= Σ((X),(Z)Ric(X,Z)|A|2X,Z)𝑑n+ΣXZ,νM𝑑n1+\displaystyle\int_{\Sigma}\Big{(}\langle\nabla^{\perp}(X^{\perp}),\nabla^{\perp}(Z^{\perp})\rangle-\operatorname{Ric}(X^{\perp},Z^{\perp})-|A|^{2}\langle X^{\perp},Z^{\perp}\rangle\Big{)}\,d\mathcal{H}^{n}+\int_{\partial\Sigma}\langle\nabla_{X^{\perp}}Z^{\perp},\nu_{\partial M}\rangle\,d\mathcal{H}^{n-1}+
+ΣΞ1(X,Z,𝐇)𝑑n+ΣΞ2(X,Z,𝐇,𝜼,νM)𝑑n1,\displaystyle+\int_{\Sigma}\Xi_{1}(X,Z,\mathbf{H})\,d\mathcal{H}^{n}+\int_{\partial\Sigma}\Xi_{2}(X,Z,\mathbf{H},\bm{\eta},\nu_{\partial M})\,d\mathcal{H}^{n-1},

where

Ξ1(X,Z,𝐇)=X,𝐇Z,𝐇X(Z),𝐇[X,Z],𝐇,\displaystyle\Xi_{1}(X,Z,\mathbf{H})=\langle X^{\perp},\mathbf{H}\rangle\langle Z^{\perp},\mathbf{H}\rangle-\langle\nabla_{X^{\perp}}(Z^{\perp}),\mathbf{H}\rangle-\langle[X^{\perp},Z^{\top}],\mathbf{H}\rangle,

and

Ξ2(X,Z,𝐇,𝜼,νM)=\displaystyle\Xi_{2}(X,Z,\mathbf{H},\bm{\eta},\nu_{\partial M})= XZ,𝜼νM+[X,Z],𝜼+divΣ(Z)X,𝜼\displaystyle\langle\nabla_{X^{\perp}}Z^{\perp},\bm{\eta}-\nu_{\partial M}\rangle+\langle[X^{\perp},Z^{\top}],\bm{\eta}\rangle+\mathrm{div}_{\Sigma}(Z^{\top})\langle X^{\top},\bm{\eta}\rangle
Z,𝐇X,𝜼X,𝐇Z,𝜼.\displaystyle-\langle Z^{\perp},\mathbf{H}\rangle\langle X^{\top},\bm{\eta}\rangle-\langle X^{\perp},\mathbf{H}\rangle\langle Z^{\top},\bm{\eta}\rangle.

Here 𝐇=Hν\mathbf{H}=-H\nu is the mean curvature vector; 𝜼\bm{\eta} is the unit co-normal vector field of Σ\Sigma; \nabla is the connection on MM; \perp and \top are the normal and tangential parts on Σ(t,s)\Sigma(t,s), respectively.

Let hh be a smooth function on M~\widetilde{M}. Then a direct computation gives that

(B.2) t|t=0Σ(t,0)hZ,ν𝑑n+1\displaystyle\frac{\partial}{\partial t}\Big{|}_{t=0}\int_{\Sigma(t,0)}h\langle Z,\nu\rangle\,d\mathcal{H}^{n+1}
=\displaystyle= ΣZ,hνdivX+XZ,hνdn\displaystyle\int_{\Sigma}\langle Z,h\nu\rangle\mathrm{div}X+\nabla_{X}\langle Z,h\nu\rangle\,d\mathcal{H}^{n}
=\displaystyle= Σ(Z,hνdivXZ,hν𝐇,X+XZ,hν+(Xh)Z,ν+h(XZ,ν))𝑑n\displaystyle\int_{\Sigma}\Big{(}\langle Z,h\nu\rangle\mathrm{div}X^{\top}-\langle Z,h\nu\rangle\langle\mathbf{H},X^{\perp}\rangle+\nabla_{X^{\top}}\langle Z,h\nu\rangle+(\nabla_{X^{\perp}}h)\langle Z,\nu\rangle+h(\nabla_{X^{\perp}}\langle Z,\nu\rangle)\Big{)}\,d\mathcal{H}^{n}
=\displaystyle= Σ((νh)X,ZX,𝐇Z,hν+X(Z),hν)𝑑n+ΣZ,hνX,𝜼𝑑n1.\displaystyle\int_{\Sigma}\Big{(}(\partial_{\nu}h)\langle X^{\perp},Z^{\perp}\rangle-\langle X,\mathbf{H}\rangle\langle Z,h\nu\rangle+\langle\nabla_{X^{\perp}}(Z^{\perp}),h\nu\rangle\Big{)}\,d\mathcal{H}^{n}+\int_{\partial\Sigma}\langle Z^{\perp},h\nu\rangle\langle X^{\top},\bm{\eta}\rangle\,d\mathcal{H}^{n-1}.

By (B.1) and (B.2), we have

t|t=0Σ(t,0)divZhZ,νdn\displaystyle\frac{\partial}{\partial t}\Big{|}_{t=0}\int_{\Sigma(t,0)}\mathrm{div}Z-h\langle Z,\nu\rangle\,d\mathcal{H}^{n}
=\displaystyle= Σ((X),(Z)Ric(X,Z)|A|2X,Z(νh)X,Z)𝑑n\displaystyle\int_{\Sigma}\Big{(}\langle\nabla^{\perp}(X^{\perp}),\nabla^{\perp}(Z^{\perp})\rangle-\operatorname{Ric}(X^{\perp},Z^{\perp})-|A|^{2}\langle X^{\perp},Z^{\perp}\rangle-(\partial_{\nu}h)\langle X^{\perp},Z^{\perp}\rangle\Big{)}\,d\mathcal{H}^{n}
+\displaystyle+ ΣXZ,νM𝑑n1+ΣΞ~1(X,Z,𝐇)𝑑n+ΣΞ~2(X,Z,𝐇,𝜼,νM)𝑑n1,\displaystyle\int_{\partial\Sigma}\langle\nabla_{X^{\perp}}Z^{\perp},\nu_{\partial M}\rangle\,d\mathcal{H}^{n-1}+\int_{\Sigma}\widetilde{\Xi}_{1}(X,Z,\mathbf{H})\,d\mathcal{H}^{n}+\int_{\partial\Sigma}\widetilde{\Xi}_{2}(X,Z,\mathbf{H},\bm{\eta},\nu_{\partial M})\,d\mathcal{H}^{n-1},

where

Ξ~1(X,Z,𝐇)\displaystyle\widetilde{\Xi}_{1}(X,Z,\mathbf{H}) =Ξ1(X,Z,𝐇)+X,𝐇Z,hνX(Z),hν\displaystyle=\Xi_{1}(X,Z,\mathbf{H})+\langle X^{\perp},\mathbf{H}\rangle\langle Z^{\perp},h\nu\rangle-\langle\nabla_{X^{\perp}}(Z^{\perp}),h\nu\rangle
=X,𝐇Z,𝐇+hνX(Z),𝐇+hν[X,Z],𝐇\displaystyle=\langle X^{\perp},\mathbf{H}\rangle\langle Z^{\perp},\mathbf{H}+h\nu\rangle-\langle\nabla_{X^{\perp}}(Z^{\perp}),\mathbf{H}+h\nu\rangle-\langle[X^{\perp},Z^{\top}],\mathbf{H}\rangle
=X,𝐇Z,𝐇+hνX(Z),𝐇+hνA(X,Z)ν,𝐇,\displaystyle=\langle X^{\perp},\mathbf{H}\rangle\langle Z^{\perp},\mathbf{H}+h\nu\rangle-\langle\nabla_{X^{\perp}}(Z^{\perp}),\mathbf{H}+h\nu\rangle-A(X^{\top},Z^{\top})\langle\nu,\mathbf{H}\rangle,

and

Ξ~2(X,Z,𝐇,𝜼,νM)\displaystyle\widetilde{\Xi}_{2}(X,Z,\mathbf{H},\bm{\eta},\nu_{\partial M}) =Ξ2(X,Z,𝐇,𝜼,νM)Z,hνX,𝜼\displaystyle=\Xi_{2}(X,Z,\mathbf{H},\bm{\eta},\nu_{\partial M})-\langle Z^{\perp},h\nu\rangle\langle X^{\top},\bm{\eta}\rangle
=XZ,𝜼νM+[X,Z],𝜼+divΣ(Z)X,𝜼\displaystyle=\langle\nabla_{X^{\perp}}Z^{\perp},\bm{\eta}-\nu_{\partial M}\rangle+\langle[X^{\perp},Z^{\top}],\bm{\eta}\rangle+\mathrm{div}_{\Sigma}(Z^{\top})\langle X^{\top},\bm{\eta}\rangle
Z,𝐇+hνX,𝜼X,𝐇Z,𝜼.\displaystyle\ \ \ \ -\langle Z^{\perp},\mathbf{H}+h\nu\rangle\langle X^{\top},\bm{\eta}\rangle-\langle X^{\perp},\mathbf{H}\rangle\langle Z^{\top},\bm{\eta}\rangle.

In the equality of Ξ~1(X,Z,𝐇)\widetilde{\Xi}_{1}(X,Z,\mathbf{H}), we used the following identity:

[X,Z],ν\displaystyle\langle[X^{\perp},Z^{\top}],\nu\rangle =X(Z)Z(X),ν\displaystyle=\langle\nabla_{X^{\perp}}(Z^{\top})-\nabla_{Z^{\top}}(X^{\perp}),\nu\rangle
=X(Z)X(Z)Z(X),ν\displaystyle=\langle\nabla_{X}(Z^{\top})-\nabla_{X^{\top}}(Z^{\top})-\nabla_{Z^{\top}}(X^{\perp}),\nu\rangle
=Z,XνZ(X),ν+A(X,Z)\displaystyle=-\langle Z^{\top},\nabla_{X}\nu\rangle-\langle\nabla_{Z^{\top}}(X^{\perp}),\nu\rangle+A(X^{\top},Z^{\top})
=Z,X,νZ(X),ν+A(X,Z)\displaystyle=\langle Z^{\top},\nabla\langle X,\nu\rangle\rangle-\langle\nabla_{Z^{\top}}(X^{\perp}),\nu\rangle+A(X^{\top},Z^{\top})
=A(X,Z).\displaystyle=A(X^{\top},Z^{\top}).

We remark that

|Ξ~1(X,Z,𝐇)|+|Ξ~2(X,Z,𝐇,𝜼,νM)| C(|X|(|𝐇+hν|+|𝜼νM|+|Z|)+|X|).\displaystyle|\widetilde{\Xi}_{1}(X,Z,\mathbf{H})|+|\widetilde{\Xi}_{2}(X,Z,\mathbf{H},\bm{\eta},\nu_{\partial M})|\leq C\big{(}|X|(|\mathbf{H}+h\nu|+|\bm{\eta}-\nu_{\partial M}|+|Z^{\top}|)+|X^{\top}|\big{)}.

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 (Mn+1,M,g)(M^{n+1},\partial M,g) be a compact manifold with boundary of dimension (n+1)3(n+1)\geq 3. Let (Σ,Σ)(M,M)(\Sigma,\partial\Sigma)\subset(M,\partial M) be an almost embedded free boundary hh-hypersurface and φ\varphi be a smooth function on Σ\Sigma. Then for any pΣp\in\Sigma, there exists a family of cut-off functions (ξr)0<r<ϵ(\xi_{r})_{0<r<\epsilon} for some ϵ>0\epsilon>0 so that ξr(p)=0\xi_{r}(p)=0

IIΣ(φ,φ)=limr0IIΣ(ξrφ,ξrφ).\mathrm{II}_{\Sigma}(\varphi,\varphi)=\lim_{r\rightarrow 0}\mathrm{II}_{\Sigma}(\xi_{r}\varphi,\xi_{r}\varphi).
Proof.

Set

(C.1) ξr(x)={0,|x|<r22log|x|logr,r2|x|r1,|x|>r,\xi_{r}(x)=\left\{\begin{aligned} &0,&|x|<r^{2}\\ &2-\frac{\log|x|}{\log r},&r^{2}\leq|x|\leq r\\ &1,&|x|>r\end{aligned}\right.,

where |x|=distM(x,p)|x|=\operatorname{dist}_{M}(x,p). Then we have Σ|ξr|2<C(n)/|logr|0\int_{\Sigma}|\nabla\xi_{r}|^{2}<C(n)/|\log r|\rightarrow 0 and ξr1\xi_{r}\rightarrow 1 as r0r\rightarrow 0. Then it suffices to prove Σ|(ξrφ)|2Σ|φ|2\int_{\Sigma}|\nabla(\xi_{r}\varphi)|^{2}\rightarrow\int_{\Sigma}|\nabla\varphi|^{2} as r0r\rightarrow 0. This follows from Σ|ξr|20\int_{\Sigma}|\nabla\xi_{r}|^{2}\rightarrow 0

Appendix D Local hh-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 (Mn+1,M,g)(M^{n+1},\partial M,g) be a compact Riemannian manifold with boundary, and let (Σ,Σ)(M,M)(\Sigma,\partial\Sigma)\subset(M,\partial M) be an embedded, free boundary minimal hypersurface. Given a point pΣp\in\partial\Sigma, there exist ϵ>0\epsilon>0 and a neighborhood UMU\subset M of pp such that if h:Uh:U\rightarrow\mathbb{R} is a smooth function with hC2,α<ϵ\|h\|_{C^{2,\alpha}}<\epsilon and

w:ΣU satisfies wC2,α<ϵ,w:\Sigma\cap U\rightarrow\mathbb{R}\text{ satisfies }\|w\|_{C^{2,\alpha}}<\epsilon,

then for any t(ϵ,ϵ)t\in(-\epsilon,\epsilon), there exists a C2,αC^{2,\alpha}-function vt:UΣv_{t}:U\cap\Sigma\rightarrow\mathbb{R}, whose graph GtG_{t} meets M\partial M orthogonally along UΣU\cap\partial\Sigma and satisfies:

HGt=h|Gt,H_{G_{t}}=h|_{G_{t}},

(where HGtH_{G_{t}} is evaluated with respect to the upward pointing normal of GtG_{t}), and

vt(x)=w(x)+t, if x(UΣ)IntM.v_{t}(x)=w(x)+t,\text{ if }x\in\partial(U\cap\Sigma)\cap\mathrm{Int}M.

Furthermore, vtv_{t} depends on t,h,wt,h,w in C1C^{1} and the graphs {Gt:t[ϵ,ϵ]}\{G_{t}:t\in[-\epsilon,\epsilon]\} 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 Φ\Phi in [ACS17]*Section 3:

Ψ:×X×Y×Y×YZ1×Z2×Z3.\Psi:\mathbb{R}\times X\times Y\times Y\times Y\rightarrow Z_{1}\times Z_{2}\times Z_{3}.

The map Ψ\Psi is defined by

Ψ(t,g,h,w,u)=(Hg(t+w+u)h,g(Ng(t+w+u),νg(t+w+u)),u|Γ2);\Psi(t,g,h,w,u)=(H_{g(t+w+u)}-h,g(N_{g}(t+w+u),\nu_{g}(t+w+u)),u|_{\Gamma_{2}});

here all the notions are the same as [ACS17]*Section 3. ∎

References