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

Mean Convex Smoothing of Mean Convex Cones

Zhihan Wang Department of Mathematics, Princeton University, Fine Hall, 304 Washington Road, Princeton, NJ 08540, USA [email protected]
Abstract.

We show that any minimizing hypercone can be perturbed into one side to a properly embedded smooth minimizing hypersurface in the Euclidean space, and every viscosity mean convex cone admits a properly embedded smooth mean convex self-expander asymptotic to it near infinity. These two together confirm a conjecture of Lawson [Bro86, Problem 5.7].

1. Introduction

Regularity theory for minimal hypersurfaces has long been studied in the history. By the landmark work of [DG61, FF60, Sim68, Fed70], an area-minimizing hypersurface in the Euclidean space or in a Riemannian manifold is always smooth away from some closed singular set of Hausdorff codimension 7\geq 7; Near each singularity, minimizing hypersurfaces are modeled on minimizing hypercones, i.e. minimizing hypersurfaces in the Euclidean space invariant under rescalings. This regularity result was generalized to stable minimal hypersurfaces without codimension 11 edge-type singularities by [SS81, Wic14]. Later, [Sim93, CN13, NV20] carefully studied the structure of singular set, proving that they are in fact codimension 77-rectifiable. On the other hand, [Sim21b] constructed examples of stable minimal hypersurfaces in m+7+1\mathbb{R}^{m+7+1} (m1m\geq 1) under CC^{\infty} Riemannian metrics arbitrarily close to the Euclidean metric, whose singular set is any prescribed closed subset in m×{𝟎}\mathbb{R}^{m}\times\{\mathbf{0}\}. This suggests that the singular set can be bad in general.

On the other hand, those singularities of minimizing hypersurfaces seem to be unstable under perturbation. In [HS85], for a regular minimizing hypercone Cn+1C\subset\mathbb{R}^{n+1} (regular means CC has only isolated singularity at the origin), it was shown that there exists a unique (up to rescaling) minimizing hypersurface SS lying on one side of CC which has no singularity. Moreover, such SS is a radial graph over certain domain on 𝕊n\mathbb{S}^{n}. Based on this, [Sma93] shown that for a closed oriented 88-manifold MM with nontrivial 77-th homology group and a CC^{\infty}-generic Riemannian metric gg, every (homologically) minimizing hypersurface in (M,g)(M,g) has no singularity. Later, this was generalized by [CLS20, LW20] to generic metric on an arbitrary closed 88-manifold to construct at least one entirely smooth minimal hypersurface (not necessarily area-minimizing).

In this paper, we generalize the existence result by Hardt and Simon to arbitrary minimizing hypercones, not necessarily with only isolated singularity.

Theorem 1.1 (c.f. Theorem 5.2).

Let En+1E\subset\mathbb{R}^{n+1} be a closed subset such that C:=EC:=\partial E is a minimizing hypercone in n+1\mathbb{R}^{n+1}. Then there exists SInt(E)S\subset\text{Int}(E) a properly embedded smooth minimizing hypersurface in n+1\mathbb{R}^{n+1} with distant 11 to the origin and CC to be the tangent cone at infinity. Moreover, SS is a radial graph over Int(E)𝕊n\text{Int}(E)\cap\mathbb{S}^{n}.

We might expect this theorem to be a first step towards the full generic regularity problem for minimizing hypersurfaces in general dimension. However, we assert that the uniqueness (up to rescaling) of such one-sided smooth perturbation is still open, even in the splitting case, i.e. when C=C0×kC=C_{0}\times\mathbb{R}^{k} for some regular minimizing hypercone C0C_{0}. In [Sim21a], uniqueness in the splitting case is proved under further strict stability assumption. We also mention that [Loh18] claimed the existence and uniqueness of such one-sided minimizing perturbation, with a seemingly incomplete proof. On the other hand, even if the uniqueness (up to rescaling) is established, it’s still unclear how to use it to smoothing a general minimizing hypersurface in a Riemannian manifold.

When CC is not minimizing in EE, by a barrier argument, such minimal hypersurface lying in the interior of EE should not exist. [Lin87] shown that for n+18n+1\leq 8, for a regular minimal hypercone C=En+1C=\partial E\subset\mathbb{R}^{n+1}, there exists a smooth mean convex hypersurface asymptotic to CC near infinity. [Din20] shown that more generally, if the cone CC is C3,αC^{3,\alpha}-regular and mean convex of arbitrary dimension, but not minimizing in EE, then there exists a unique smooth strictly mean convex self-expander sitting inside Int(E)\text{Int}(E) and properly embedded in n+1\mathbb{R}^{n+1}. Here a self-expander is a hypersurface satisfying the equation,

HX2=0,\vec{H}-\frac{X^{\perp}}{2}=0,

where H\vec{H} is the mean curvature vector and XX^{\perp} is the projection of position vector XX onto the normal direction. Self-expander was introduced in [EH89] as model of long time behavior for mean curvature flow, and also studied by [Ilm98] as model for mean curvature flow coming out of a singularity. See Section 2.2 for more on geometry of self-expanders.

We also generalize Ding’s result to arbitrary mean convex hypercone, not necessarily regular.

Theorem 1.2 (c.f. Theorem 5.3).

Let En+1E\subset\mathbb{R}^{n+1} be a closed subset such that C:=EC:=\partial E is a mean convex hypercone in n+1\mathbb{R}^{n+1} and not minimizing in EE. Then there exists SInt(E)S\subset\text{Int}(E) a smooth strictly mean convex self-expander properly embedded in n+1\mathbb{R}^{n+1} with CC to be the tangent cone at infinity. In particular, SS is a radial graph over Int(E)𝕊n\text{Int}(E)\cap\mathbb{S}^{n}.

Here, mean convexity assumption for EE is in viscosity sense, see Definition 2.2. In particular, Theorem 1.2 applies when E\partial E is a stationary minimal hypercone, possibly with singularities, which is not minimizing in EE. On the other hand, uniqueness is again unclear in this non-smooth case.

Combining Theorem 1.1 and 1.2, we confirm a conjecture of Lawson [Bro86, Problem 5.7]:

Theorem 1.3.

Given a stable minimal hypercone Cn+1C\subset\mathbb{R}^{n+1} and every ϵ>0\epsilon>0, there exists a properly embedded smooth hypersurface with positive mean curvature in 𝔹1n+1(𝟎)\mathbb{B}_{1}^{n+1}(\mathbf{0}) within Hausdorff distant ϵ\leq\epsilon from C𝔹1n+1(𝟎)C\cap\mathbb{B}_{1}^{n+1}(\mathbf{0}).

A key ingredient of the proof of both Theorem 1.1 and 1.2 is the following divergence of positive super-solution to Jacobi field equation near singularities.

Theorem 1.4 (c.f. Corollary 4.3).

Let (M,g)(M,g) be a smooth Riemannian manifold (not necessarily complete), Λ>0\Lambda>0. Let ΣM\Sigma\subset M be a two-sided stable minimal hypersurface with singular set of codimension 7\geq 7. Let uCloc2(Σ)u\in C^{2}_{loc}(\Sigma) be a positive function such that

ΔΣu+|AΣ|2uΛu0,\Delta_{\Sigma}u+|A_{\Sigma}|^{2}u-\Lambda u\leq 0,

on Σ\Sigma. Then for every xSing(Σ)x\in\mathrm{Sing}(\Sigma),

lim infyxu(y)=+.\liminf_{y\to x}u(y)=+\infty.

Such behavior of uu has been established by [Sim08] for minimizing hypersurfaces, or more generally, for submanifolds belonging to a regular multiplicity 1 class. A phenomenon of similar spirit was also exploited in [SY17] for minimal slicings and by [Wan20, Lemma 2.14] to study stable minimal hypersurfaces lying on one side of a regular cone near infinity. In this paper, we prove it for general stable minimal hypersurfaces, following the strategy of [Sim08] but more directly by first proving a Harnack inequality for stable minimal hypersurfaces.

Sketch of the Proof of Theorem 1.1 and 1.2.

Let C=En+1C=\partial E\subset\mathbb{R}^{n+1} be a minimizing hypercone. First one can approximate E𝕊nE\cap\mathbb{S}^{n} from interior by mean convex domains j\mathcal{E}_{j} with optimal regularity. Then consider in n+1\mathbb{R}^{n+1} the Plateau problem of minimizing area among integral currents with boundary j\partial\mathcal{E}_{j}. Using a similar argument as [HS85], one can show that such area-minimizer SjS_{j} is a smooth radial graph over j\mathcal{E}_{j} and (after subtracting the boundary) lies in the interior of the cone over j\mathcal{E}_{j}. Moreover, when jj\to\infty, SjS_{j} converges to the truncated cone C𝔹1C\cap\mathbb{B}_{1}. Hence if one rescale SjS_{j} by their distant to the origin, then the rescaled minimizing hypersurfaces will subconverge to some hypersurface SS minimizing in n+1\mathbb{R}^{n+1} and lying in EE, with distant 11 to the origin. To see SS is smooth, consider the Jacobi field ϕ=XνS\phi=X\cdot\nu_{S} on regular part of SS induced by rescaling, here XX is the position vector and νS\nu_{S} is the unit normal field of SS pointing away from the cone. SS being limit of rescaling of radial graphs SjS_{j} guarantees that ϕ0\phi\geq 0; And ϕ\phi can’t be identically zero since otherwise SS must be a cone and can’t have distant 11 to the origin. Thus by strong maximum principle, ϕ\phi is everywhere positive; Also by definition, ϕ\phi is bounded on each ball in n+1\mathbb{R}^{n+1}. Therefore, Theorem 1.4 implies that the singular set of SS is empty.

When EE is viscosity mean convex and not perimeter minimizing in EE, one seeks to minimizing 𝐄\mathbf{E}-functional introduced by [Ilm98, Lecture 2 C] (where it’s called 𝐊\mathbf{K}-functional) to find a self-expander SS^{\prime} asymptotic to E\partial E near infinity. However, since the rescaling of a self-expander is usually not a self-expander, one can not use maximum principle to conclude that SS^{\prime} is disjoint from its rescalings. Instead, we consider the level set flow t0Dt×{t}n+1×\bigsqcup_{t\geq 0}D_{t}\times\{t\}\subset\mathbb{R}^{n+1}\times\mathbb{R} starting from (Int(E))c(\text{Int}(E))^{c}. We still first perturb E𝕊nE\cap\mathbb{S}^{n} into a sequence of optimally regular mean convex domain j\mathcal{E}_{j} in the interior of E𝕊nE\cap\mathbb{S}^{n}, and then solve Plateau problem to find self-expanders SjS^{\prime}_{j} lying on one side of the cone CjC_{j} over j\partial\mathcal{E}_{j}, and asymptotic to CjC_{j} near infinity. Then using avoidance principle for weak mean curvature flow, we argue that the rescalings λSj\lambda\cdot S^{\prime}_{j} are disjoint from Int(D1)\text{Int}(D_{1}) for every λ1\lambda\geq 1. Then by taking jj\to\infty, we obtain an 𝐄\mathbf{E}-minimizing self-expander SES^{\prime}\subset E asymptotic to E\partial E near infinity, and disjoint from Int(D1)\text{Int}(D_{1}). Using definition of level set flow, we conclude that SS^{\prime} must coincide with D1\partial D_{1}, and the rescalings λS\lambda\cdot S^{\prime} are still all disjoint from Int(D1)\text{Int}(D_{1}), λ1\forall\lambda\geq 1. Moreover, SES^{\prime}\neq\partial E since E\partial E is not area-minimizing in EE. These altogether imply that the eigenfunction ϕ:=XνS=2HS\phi:=X\cdot\nu_{S^{\prime}}=2H_{S^{\prime}} of Jacobi operator of 𝐄\mathbf{E}-functional on SS^{\prime} induced by rescaling is non-negative and not vanishing identically. Repeat the process above, we get ϕ>0\phi>0 everywhere and SS^{\prime} has no singularity.

Organization of the Paper.

Section 2 contains the basic notations we use in this paper as well as a brief review of geometric measure theory, geometry of self-expanders and weak notions of mean curvature flow. In Section 3, we prove a multiplicity 11 result for stable minimal hypersurfaces, which enable us to derive a Neumann-Sobolev inequality and a Harnack inequality for stable minimal hypersurfaces following the same argument as [BG72] for minimizing boundaries. Using this, in Section 4 we prove a more precise asymptotic lower bound for super-solution of Jacobi field equations on a stable minimal hypersurface, and derive Theorem 1.4 as a corollary. Finally, we state and prove a more concrete version of Theorem 1.1, 1.2 and finish the proof of Theorem 1.3 in Section 5.

Acknowledgement

I am grateful to my advisor Fernando Codá Marques for his constant support and guidance.

2. Preliminaries

Throughout this paper, let n+1\mathbb{R}^{n+1} be the Euclidean space of dimension n+1n+1, n1n\geq 1; Let

  • 𝔹rn+1(x)\mathbb{B}_{r}^{n+1}(x) be the open ball of radius rr in n+1\mathbb{R}^{n+1} centered at xx; We may omit the superscript n+1n+1 if there’s no confusion about dimension; We may write 𝔹r:=𝔹r(𝟎)\mathbb{B}_{r}:=\mathbb{B}_{r}(\mathbf{0}) to be the ball centered at the origin 𝟎\mathbf{0};

  • 𝕊n:=𝔹1n+1\mathbb{S}^{n}:=\partial\mathbb{B}_{1}^{n+1} be the unit sphere;

  • 𝔸(x;r,s)\mathbb{A}(x;r,s) be the open annuli 𝔹r(x)Clos(𝔹s(x))\mathbb{B}_{r}(x)\setminus\text{Clos}(\mathbb{B}_{s}(x)) centered at xx;

  • C():={rω:ω𝕊n,r0}n+1C(\mathcal{E}):=\{r\omega:\omega\in\mathcal{E}\subset\mathbb{S}^{n},r\geq 0\}\subset\mathbb{R}^{n+1} be the cone generated by subset 𝕊n\mathcal{E}\subset\mathbb{S}^{n};

  • ηx,r\eta_{x,r} be the map between n+1\mathbb{R}^{n+1}, maps yy to r(yx)r(y-x); We may omit subscript xx if x=𝟎x=\mathbf{0};

  • k\mathscr{H}^{k} be the kk-dimensional Hausdorff measure;

  • gEucg_{Euc} be the Euclidean metric on n+1\mathbb{R}^{n+1}.

For any smooth oriented Riemannian manifold (M,g)(M,g) of dimension n+1n+1, write

  • Int(A)\text{Int}(A)\ be the interior of a subset AMA\subset M;

  • Clos(A)\text{Clos}(A)\ be the closure of a subset AMA\subset M;

  • A:=Clos(A)Int(A)\partial A:=\text{Clos}(A)\setminus\text{Int}(A)\ be the topological boundary of AA;

  • distgdist_{g}\ be the distant function over (M,g)(M,g);

  • Br(A):={xM:distg(x,A)<r}B_{r}(A):=\{x\in M:dist_{g}(x,A)<r\} be the open rr-neighborhood of subset AMA\subset M;

  • g\nabla_{g}\ (or g\nabla^{g})  be the Levi-Civita connection with respect to gg;

  • expxg\exp^{g}_{x}\ be the exponential map of (M,g)(M,g) on the tangent space TxMT_{x}M;

  • injrad(x;M,g)injrad(x;M,g)\ be the injectivity radius of (M,g)(M,g) at xx;

  • RicgRic_{g}\ be the Ricci curvature tensor of (M,g)(M,g);

  • 𝒳c(U)\mathscr{X}_{c}(U)\ be the space of compactly supported smooth vector field on an open subset UMU\subset M;

  • etXe^{tX}\ be the one-parameter family of diffeomorphism generated by X𝒳c(U)X\in\mathscr{X}_{c}(U).

We may omit the super or subscript gg if there’s no confusion. Also for two subsets U,VMU,V\subset M, write UVU\subset\subset V if Clos(U)\text{Clos}(U) is a compact subset of VV.

For a sequence of closed subset {Ej}1j\{E_{j}\}_{1\leq j\leq\infty} of MM, call EjE_{j} converges to EE_{\infty} locally in the Hausdorff distant sense, if for every compact subset KMK\subset M and every ϵ>0\epsilon>0, there exists j(ϵ,K)>>1j(\epsilon,K)>>1 such that jj(ϵ,K)\forall j\geq j(\epsilon,K),

EjKBϵ(E),EKBϵ(Ej).E_{j}\cap K\subset B_{\epsilon}(E_{\infty}),\ \ \ \ \ E_{\infty}\cap K\subset B_{\epsilon}(E_{j}).

Let ΣM\Sigma\subset M be a two-sided hypersurface, two-sided means it admits a global normal field ν\nu. When Σ\Sigma is portion of the boundary of some specified domain, unless otherwise mentioned, we use the convention that ν\nu is chosen to be the outward pointed normal field. In this article, unless otherwise stated, every hypersurface is assumed to be two-sided and optimally regular, i.e. n2(Clos(Σ)Σ)=0\mathscr{H}^{n-2}(\text{Clos}(\Sigma)\setminus\Sigma)=0 and nΣ\mathscr{H}^{n}\llcorner\Sigma is locally finite. Let

  • Reg(Σ):={xClos(Σ):Clos(Σ) is a C1,1 embedded hypersurafce near x}\mathrm{Reg}(\Sigma):=\{x\in\text{Clos}(\Sigma):\text{Clos}(\Sigma)\text{ is a }C^{1,1}\text{ embedded hypersurafce near }x\} be the regular part of Σ\Sigma;

  • Sing(Σ):=Clos(Σ)Reg(Σ)\mathrm{Sing}(\Sigma):=\text{Clos}(\Sigma)\setminus\mathrm{Reg}(\Sigma) be the singular part.

By adding points to Σ\Sigma if necessary, we may identify Σ=Reg(Σ)\Sigma=\mathrm{Reg}(\Sigma) in this paper. Call Σ\Sigma regular in an open subset UU if Sing(Σ)U=\mathrm{Sing}(\Sigma)\cap U=\emptyset. We shall work with the following function spaces on Σ\Sigma:

  • Lp(Σ)L^{p}(\Sigma) measurable functions ff with Σ|f|p<+\int_{\Sigma}|f|^{p}<+\infty;

  • Wloc1,2(Σ)W^{1,2}_{loc}(\Sigma) measurable functions which restricts to W1,2W^{1,2}-function on each compact smooth sub-domain in Σ\Sigma;

  • Clock(Σ)C^{k}_{loc}(\Sigma) functions on Σ\Sigma which admit up to kk-th order continuous derivatives, possibly unbounded on Σ\Sigma.

Also write

  • HΣ:=divΣ(ν)H_{\Sigma}:=-div_{\Sigma}(\nu) be the scalar mean curvature of Σ\Sigma, and HΣ:=HΣν\vec{H}_{\Sigma}:=H_{\Sigma}\cdot\nu be the mean curvature vector;

  • AΣ:=νA_{\Sigma}:=-\nabla\nu be the second fundamental form of Σ\Sigma.

Note that under this convention, the mean curvature vector does not depend on the choice of normal field, and the scalar mean curvature for unit sphere with respect to the outward pointed normal field is negative. Recall Σ\Sigma is minimal if and only if HΣ0H_{\Sigma}\equiv 0.

Call hypersurface Σn+1\Sigma\subset\mathbb{R}^{n+1} a hypercone if it’s invariant under dilation ηλ\eta_{\lambda}, λ>0\forall\lambda>0.

A minimal hypersurface Σ(M,g)\Sigma\subset(M,g) is called stable in an open subset UU, if

d2ds2|s=0n(esX(Σ))0,X𝒳c(U).\frac{d^{2}}{ds^{2}}\Big{|}_{s=0}\mathscr{H}^{n}(e^{sX}(\Sigma))\geq 0,\ \ \ \forall X\in\mathscr{X}_{c}(U).

By [SS81], this is equivalent to that n7+ϵ(Sing(Σ))=\mathscr{H}^{n-7+\epsilon}(\mathrm{Sing}(\Sigma))=\emptyset, ϵ>0\forall\epsilon>0 and that

(2.1) QΣ(ϕ,ϕ):=Σ|ϕ|2(|AΣ|2+Ricg(ν,ν))ϕ20,ϕCc1(ΣU).\displaystyle Q_{\Sigma}(\phi,\phi):=\int_{\Sigma}|\nabla\phi|^{2}-(|A_{\Sigma}|^{2}+Ric_{g}(\nu,\nu))\phi^{2}\geq 0,\ \ \ \forall\phi\in C_{c}^{1}(\Sigma\cap U).

Let

(2.2) LΣ:=ΔΣ+|AΣ|2+Ricg(ν,ν),\displaystyle L_{\Sigma}:=\Delta_{\Sigma}+|A_{\Sigma}|^{2}+Ric_{g}(\nu,\nu),

be the Euler-Lagrangian operator associated to QΣQ_{\Sigma}, known as the Jacobi operator. Every uCloc2(Σ)u\in C^{2}_{loc}(\Sigma) solving LΣu=0L_{\Sigma}u=0 on Σ\Sigma is called a Jacobi field.

2.1. Basics in geometric measure theory

We recall some basic notions from geometric measure theory and refer the readers to [Sim83, Fed69] for details. For 1kn+11\leq k\leq n+1, in an n+1n+1 dimensional manifold (M,g)(M,g) (not necessarily complete, isometrically embedded in some L\mathbb{R}^{L} if necessary), write

  • 𝐈k(M)\mathbf{I}_{k}(M) be the space of integral kk currents on MM;

  • 𝒵k(M):={T𝐈k(M):T=0}\mathcal{Z}_{k}(M):=\{T\in\mathbf{I}_{k}(M):\partial T=0\} be the space of integral kk-cycles, where :𝐈k𝐈k1\partial:\mathbf{I}_{k}\to\mathbf{I}_{k-1} be the boundary operator;

  • 𝒱k(M)\mathcal{I}\mathcal{V}_{k}(M) be the space of integral kk varifolds on (M,g)(M,g);

  • spt(V)\text{spt}(V) be the support of a varifold V𝒱k(M)V\in\mathcal{I}\mathcal{V}_{k}(M);

  • ff_{\sharp} be the push forward of currents or varifolds associated to a proper Lipschitz map ff.

For a hypersurface Σ(M,g)\Sigma\subset(M,g), let |Σ|g𝒱(M)|\Sigma|_{g}\in\mathcal{I}\mathcal{V}(M) be the integral varifold associated to Σ\Sigma; If further Σ\Sigma is oriented, denote [Σ]𝐈n(M)[\Sigma]\in\mathbf{I}_{n}(M) to be the associated integral current. More generally, for T𝐈n(M)T\in\mathbf{I}_{n}(M), denote |T|g|T|_{g}, Tg\|T\|_{g} to be the associated integral nn-varifold and Radon measure on MM correspondingly. The metric subscript gg above will be omit if there’s no confusion.

For an open subset UMU\subset M, let U\mathcal{F}_{U} and 𝐅U\mathbf{F}_{U} be the flat metric on 𝐈n(M)\mathbf{I}_{n}(M) and varifold metric on 𝒱n(M)\mathcal{I}\mathcal{V}_{n}(M) in UU, and let 𝐌U\mathbf{M}_{U} be the mass for an integral current; Omit UU if U=MU=M.

For a smooth vector field X𝒳c(U)X\in\mathscr{X}_{c}(U) and V𝒱n(U)V\in\mathcal{I}\mathcal{V}_{n}(U), the first variation of the area of VV with respect to XX is

δV(X)=ddt|t=0V(U)=𝑑ivπX𝑑V(x,π),\delta V(X)=\frac{d}{dt}\Big{|}_{t=0}\|V\|(U)=\int div^{\pi}X\ dV(x,\pi),

Recall VV is called stationary in UU if and only if δV(X)=0\delta V(X)=0, X𝒳c(U)\forall X\in\mathscr{X}_{c}(U). By constructing appropriate test vector fields, whenever VV is stationary in MM, for every xMx\in M, the following

rθg(x,r;V):=enλrωnrnV(Br(x))r\mapsto\theta_{g}(x,r;\|V\|):=\frac{e^{n\lambda r}}{\omega_{n}r^{n}}\|V\|(B_{r}(x))

is monotone non-decreasing in r(0,injrad(x;M,g))r\in(0,injrad(x;M,g)), where

λ=λ(x,r):=inf{C0:g2(ρx2)2(1Cρx2) as a quadratic form, on Br(x)};\lambda=\lambda(x,r):=\inf\{C\geq 0:\nabla_{g}^{2}(\rho_{x}^{2})\geq 2(1-C\rho_{x}^{2})\text{ as a quadratic form, on }B_{r}(x)\};

where ρx:=distg(,x)\rho_{x}:=dist_{g}(\cdot,x) is the distant function to xx. And when g=gEucg=g_{Euc}, θg(x,r;V)\theta_{g}(x,r;\|V\|) is constant in rr if and only if VV is a cone. Also, denote θ(x;V):=limr0θg(x,r;V)\theta(x;\|V\|):=\lim_{r\searrow 0}\theta_{g}(x,r;\|V\|) to be the density of VV at xx.

Following [DG61, Sim83], call a Lebesgue measurable subset PMP\subset M a Caccioppoli set, if for every open subset WMW\subset\subset M and every smooth vector field X𝒳c(W)X\in\mathscr{X}_{c}(W), there exists some constant C(P,W)C(P,W) such that

|P𝑑iv(X)𝑑n+1|C(P,W)XC0.|\int_{P}div(X)\ d\mathscr{H}^{n+1}|\leq C(P,W)\|X\|_{C^{0}}.

By [DG61, Sim83], a Caccioppoli set PP is naturally an integral n+1n+1-current [P]𝐈n+1(M)[P]\in\mathbf{I}_{n+1}(M). [P](W)\|\partial[P]\|(W) is called the perimeter of PP in WW, and is usually denoted as 𝒫(P;W)\mathcal{P}(P;W). Also define the 𝐅\mathbf{F}-metric among Caccioppoli sets by

𝐅(P1,P2):=([P1],[P2])+𝐅([P1],[P2]).\mathbf{F}(P_{1},P_{2}):=\mathcal{F}(\partial[P_{1}],\partial[P_{2}])+\mathbf{F}(\|\partial[P_{1}]\|,\|\partial[P_{2}]\|).

When EME\subset M is a measurable subset, call a Caccioppoli set PMP\subset M (homologically) perimeter minimizing in EE if PEP\subset E and for every open subset WWMW^{\prime}\subset\subset W\subset M and every Caccioppoli set QMQ\subset M with PΔQEWP\Delta Q\subset E\cap W^{\prime}, we have

𝒫(P;W)𝒫(Q;W).\mathcal{P}(P;W)\leq\mathcal{P}(Q;W).

If EME\subset M is a closed subset, PEP\subset E is minimizing in EE and |[P]||\partial[P]| is stationary in MM, then the regular part of [P]\partial[P] is a multiplicity one and stable; Being minimizing in EE guarantees that there’s no singularity of [P]\partial[P] near which spt([P])\text{spt}(\partial[P]) is a union of embedded C1,αC^{1,\alpha} hypersurface-with-boundary meeting along their common boundary. Hence, by [Wic14, Theorem 18.1], Reg([P])\mathrm{Reg}(\partial[P]) is a stable minimal hypersurface in MM with optimal regularity.

The compactness of perimeter minimizing Caccioppoli sets in closed subsets is a little bit subtle, due to the lack of direct cut-pasting argument as in [Sim83, Theorem 34.5].

Lemma 2.1 (Compactness).

Let {Ej}1j\{E_{j}\}_{1\leq j\leq\infty} be a sequence of closed subset in MM such that EjEE_{j}\to E_{\infty} and EjE\partial E_{j}\to\partial E_{\infty} both in locally Hausdorff distant sense in MM. Let {gj}1j\{g_{j}\}_{1\leq j\leq\infty} be a sequence of smooth metric on MM such that gjg_{j} smoothly converges to gg_{\infty}. For 1j<1\leq j<\infty, let PjEjP_{j}\subset E_{j} be a Caccioppoli set in MM minimizing perimeter in EjE_{j}, with |[Pj]||\partial[P_{j}]| stationary in MM.

Then there exists some Caccioppoli set PEP_{\infty}\subset E_{\infty} with Reg([P])\mathrm{Reg}(\partial[P_{\infty}]) an optimally regular stable minimal hypersurface in MM, such that after passing to a subsequence of jj\to\infty, 𝐅(Pj,P)0\mathbf{F}(P_{j},P_{\infty})\to 0.

Moreover, if EjEE_{j}\equiv E_{\infty}, then PP_{\infty} is also perimeter minimizing in EE_{\infty}.

Proof.

For each smooth domain WMW\subset\subset M, by comparing the perimeter of PjP_{j} with PjWP_{j}\setminus W, we know that the lim sup[Pj](W)<+\limsup\|\partial[P_{j}]\|(W)<+\infty. Hence by Fleming-Federer compactness Theorem, [Pj]\partial[P_{j}] subconverges in flat topology to some boundary of Caccioppoli set [P]\partial[P_{\infty}]; And by [SS81], the stable minimal hypersurfaces |[Pj]||\partial[P_{j}]| subconverges to some stationary integral varifold VV_{\infty} supported on a stable minimal hypersurace in EE_{\infty}.

Moreover, VV_{\infty} must be of multiplicity 11, since otherwise, by [SS81, Theorem 1], near some regular point xx of higher multiplicity, [Pj]\|\partial[P_{j}]\| are multiple graphs over some hyperplane through xspt(V)Ex\in\text{spt}(V_{\infty})\subset E_{\infty}. When xInt(E)x\in\text{Int}(E_{\infty}), since Ej\partial E_{j} converges to E\partial E_{\infty} locally in the Hausdorff distant sense, this is impossible by the local perimeter-minimizing property of PjP_{j} near xx; When xEx\in\partial E_{\infty}, we know that for j>>1j>>1, there must be two adjacent graphs bounding a slab in PjP_{j} near xx. This violates that PjP_{j} is perimeter-minimizing in EjE_{j} by subtracting a small ball from this slab.

Using Schoen-Simon’s epsilon regularity theorem [SS81, Theorem 1], VV_{\infty} being of multiplicity 11 implies that V=|[P]|V_{\infty}=|\partial[P_{\infty}]|, hence PjP_{j} subconverges to PP_{\infty} in 𝐅\mathbf{F}-metric.

When EjEE_{j}\equiv E_{\infty}, the cut-pasting argument in [Sim83, Theorem 34.5] works here to show that PP_{\infty} is perimeter minimizing in EE_{\infty}. ∎

Definition 2.2.

Call a closed subset EME\subset M viscosity mean convex, if E=Clos(Int(E))E=\text{Clos}(\text{Int}(E)) and for every smooth open domain ΩE\Omega\subset E with xΩEx\in\partial\Omega\cap\partial E, the mean curvature with respect to the outward pointed normal field of Ω\partial\Omega at xx is HΩ(x)0H_{\partial\Omega}(x)\leq 0.

Call a closed viscosity mean convex subset EMn+1E\subset M^{n+1} C1,1C^{1,1}-optimally regular, if there exists a closed subset 𝒮E\mathcal{S}\subset\partial E with Hausdorff dimension n7\leq n-7, such that ES\partial E\setminus S is a C1,1C^{1,1} embedded hypersurface and x𝒮\forall x\in\mathcal{S}, E\partial E is a stable minimal hypersurface with optimal regularity near xx. Call the smallest such 𝒮E\mathcal{S}\subset\partial E the singular set of E\partial E.

Without assuming E=Clos(Int(E))E=\text{Clos}(\text{Int}(E)), the notion of viscosity mean convex was introduced and studied in [ISZ98, Definition 3.1], where it was called barrier for minimal surface equation.

Remark 2.3.
  1. (1)

    By [SW89], the support of every Caccioppoli set in MM with stationary boundary is viscosity mean convex. In particular, every stable minimal hypercone in n+1\mathbb{R}^{n+1} bounds a viscosity mean convex subset.

  2. (2)

    It follows directly from the definition that if EME\subset M is a viscosity mean convex closed subset, then for every connected component UU of Int(E)\text{Int}(E), Clos(U)\text{Clos}(U) is also viscosity mean convex. Hence, in the following discussion, we may additionally assume that a viscosity mean convex subset has connected interior.

  3. (3)

    When a closed viscosity mean convex set EME\subset M is C1,1C^{1,1}-optimally regular, on its regular part, we can define the mean curvature vector HEL(E)\vec{H}_{\partial E}\in L^{\infty}(\partial E) almost everywhere. And EE being viscosity mean convex guarantees that HE,ν0\langle\vec{H}_{\partial E},\nu\rangle\leq 0, where ν\nu is the outward pointed normal field. In Appendix B, it is shown that any compact viscosity mean convex subset can be approximated by C1,1C^{1,1}-optimally regular mean convex subset.

2.2. Self-expanders

For a hypersurface Σn+1\Sigma\subset\mathbb{R}^{n+1} and a bounded open subset Wn+1W\subset\mathbb{R}^{n+1}, define the 𝐄\mathbf{E}-functional of Σ\Sigma in WW to be,

𝐄[Σ;W]:=We|x|2/4dΣ(x).\mathbf{E}[\Sigma;W]:=\int_{W}e^{|x|^{2}/4}\ d\|\Sigma\|(x).

Such 𝐄\mathbf{E}-functional is introduced in [Ilm98, Lecture 2 C] in which it’s named 𝐊\mathbf{K}-functional, and frequently studied by [BW18, BW19a, BW19b, BW20, BW21a, BW21b] and by [Din20]. Here we keep the notations with Bernstein-Wang. Critical points Σ\Sigma of 𝐄\mathbf{E} are known to be self-expanders, i.e. it satisfies the following equation

HΣX2=0,\vec{H}_{\Sigma}-\frac{X^{\perp}}{2}=0,

where XX^{\perp} denotes the projection of position vector onto normal direction of Σ\Sigma. Equivalently, by [EH89], {ηt(Σ)}t>0\{\eta_{\sqrt{t}}(\Sigma)\}_{t>0} is a family of hypersurfaces flowing by mean curvature.

Clearly, the 𝐄\mathbf{E}-functional can be extended to any Radon measure μ\mu on n+1\mathbb{R}^{n+1},

𝐄[μ;W]:=We|x|2/4𝑑μ(x).\mathbf{E}[\mu;W]:=\int_{W}e^{|x|^{2}/4}\ d\mu(x).

We also denote for simplicity that for a Caccioppoli sets Pn+1P\subset\mathbb{R}^{n+1}, 𝐄[P;W]:=𝐄[[P];W]\mathbf{E}[P;W]:=\mathbf{E}[\|\partial[P]\|;W] and for an nn-varifold VV, 𝐄[V;W]:=𝐄[V;W]\mathbf{E}[V;W]:=\mathbf{E}[\|V\|;W].

For a bounded open subset Wn+1W\subset\mathbb{R}^{n+1}, call a Caccioppoli set Pn+1P\subset\mathbb{R}^{n+1} 𝐄\mathbf{E}-minimizing in WW, if for every Caccioppoli set QΔPWWQ\Delta P\subset\subset W^{\prime}\subset\subset W, we have

𝐄[Q;W]𝐄[P;W].\mathbf{E}[Q;W^{\prime}]\geq\mathbf{E}[P;W^{\prime}].

And call an integral nn-varifold V𝒱n(W)V\in\mathcal{I}\mathcal{V}_{n}(W) 𝐄\mathbf{E}-stationary in WW, if for every X𝒳c(W)X\in\mathscr{X}_{c}(W),

δ𝐄[V;W](X):=ddt|t=0𝐄[etXV;W]=0;\delta\mathbf{E}[V;W](X):=\frac{d}{dt}\Big{|}_{t=0}\mathbf{E}[e^{tX}{}_{\sharp}V;W]=0;

Call VV 𝐄\mathbf{E}-stable if it’s 𝐄\mathbf{E}-stationary and for every X𝒳c(W)X\in\mathscr{X}_{c}(W),

δ2𝐄[V;W](X,X):=d2dt2|t=0𝐄[etXV;W]0.\delta^{2}\mathbf{E}[V;W](X,X):=\frac{d^{2}}{dt^{2}}\Big{|}_{t=0}\mathbf{E}[e^{tX}{}_{\sharp}V;W]\geq 0.

Note that for an nn-varifold, 𝐄\mathbf{E}-functional is the nn-th area-functional under the metric g^:=e|x2|/2ngEuc\hat{g}:=e^{|x^{2}|/2n}\cdot g_{Euc} on n+1\mathbb{R}^{n+1}. Hence, being 𝐄\mathbf{E}-stationary (resp. 𝐄\mathbf{E}-stable, 𝐄\mathbf{E}-minimizing) is equivalent to being stationary (resp. stable, minimizing) with respect to the area functional under g^\hat{g}.

For a 2-sided hypersurface ΣW\Sigma\subset W, by [Din20, (3.16)], being 𝐄\mathbf{E}-stable is equivalent to that for every ϕCc1(ΣW)\phi\in C_{c}^{1}(\Sigma\cap W),

(2.3) W[|ϕ|2(|AΣ|212)ϕ2]e|x|2/4dΣ(x)0.\displaystyle\int_{W}[|\nabla\phi|^{2}-(|A_{\Sigma}|^{2}-\frac{1}{2})\phi^{2}]\cdot e^{|x|^{2}/4}d\|\Sigma\|(x)\geq 0.

The Euler-Lagrangian operator of the quadratic form on the left hand side of (2.3) is,

(2.4) Σ:=ΔΣ+X2Σ+|AΣ|212.\displaystyle\mathcal{L}_{\Sigma}:=\Delta_{\Sigma}+\frac{X}{2}\cdot\nabla_{\Sigma}+|A_{\Sigma}|^{2}-\frac{1}{2}.

Moreover, if denote Δ^Σ,A^Σ\hat{\Delta}_{\Sigma},\hat{A}_{\Sigma} to be the laplacian and second fundamental form of Σ\Sigma under metric g^\hat{g}, then a simple computation shows that for every function uu on Σ\Sigma,

(2.5) Σu=e|x|24n(Δ^Σ+|A^Σ|2e|x|22n(1+n14n2|XΣ|2))(e|x|24nu),\displaystyle\mathcal{L}_{\Sigma}u=e^{\frac{|x|^{2}}{4n}}\cdot\big{(}\hat{\Delta}_{\Sigma}+|\hat{A}_{\Sigma}|^{2}-e^{-\frac{|x|^{2}}{2n}}(1+\frac{n-1}{4n^{2}}|X^{\Sigma}|^{2})\big{)}(e^{\frac{|x|^{2}}{4n}}u),

where XΣX^{\Sigma} is the projection of the position vector XX onto TΣT\Sigma.

Let 𝕊n\mathcal{E}\subset\mathbb{S}^{n} be a closed subset; Let E=C()E=C(\mathcal{E}) be the closed cone over \mathcal{E} in n+1\mathbb{R}^{n+1}. Call a closed set Pn+1P\subset\mathbb{R}^{n+1} distant asymptotic to EE near infinity if locally in the Hausdorff distant sense, when R+R\to+\infty, η1/R(P)E\eta_{1/R}(P)\to E and η1/R(P)E\eta_{1/R}(\partial P)\to\partial E.

For a unit vector v𝕊nv\in\mathbb{S}^{n} and α(0,π)\alpha\in(0,\pi), define

𝒞(v;α):={xn+1{𝟎}:x/|x|,v>cosα},\mathcal{C}(v;\alpha):=\{x\in\mathbb{R}^{n+1}\setminus\{\mathbf{0}\}:\langle x/|x|,v\rangle>\cos\alpha\},

be the open cone in vv direction with open angle α\alpha. O(n)O(n)-invariant self-expanders distant asymptotic to Clos(𝒞(v;α))\text{Clos}(\mathcal{C}(v;\alpha)) are constructed in [AIC95, Lemma 2], and also studied using variational approach in [Din20, Section 4]. We summarize the results here. By a rotation, WLOG v=en+1v=e_{n+1}.

Theorem 2.4.

[AIC95, Din20] For each α(0,π/2)\alpha\in(0,\pi/2), there exists a unique O(n)O(n)-symmetric convex closed smooth domain F(en+1;α)𝒞(en+1,α)F(e_{n+1};\alpha)\subset\mathcal{C}(e_{n+1},\alpha) minimizing 𝐄\mathbf{E}-functional in n+1\mathbb{R}^{n+1} and distant asymptotic to Clos(𝒞(en+1;α))\text{Clos}(\mathcal{C}(e_{n+1};\alpha)) near infinity.

Moreover, F(en+1,α)F(e_{n+1},\alpha) is an O(n)O(n)-symmetric super graph, i.e. there exists uαC([0,+))u_{\alpha}\in C^{\infty}([0,+\infty)) such that

F(en+1,α)={(x,z):xn,zuα(|x|)},F(e_{n+1},\alpha)=\{(x,z):x\in\mathbb{R}^{n},z\geq u_{\alpha}(|x|)\},

and uαu_{\alpha} satisfies

  1. (i)

    uα(0)>0u_{\alpha}(0)>0, uα(0)=0u_{\alpha}^{\prime}(0)=0, uα(r),uα′′(r)>0u_{\alpha}^{\prime}(r),u_{\alpha}^{\prime\prime}(r)>0 for r>0r>0 and u(r)/rcotαu(r)/r\to\cot\alpha when r+r\to+\infty;

  2. (ii)

    uαu_{\alpha} is continuous and strictly monotone in α\alpha and for fixed r0r\geq 0,

    uα(r)+ when α0,uα(r)0 when απ/2.u_{\alpha}(r)\to+\infty\ \text{ when }\alpha\to 0,\ \ \ \ \ u_{\alpha}(r)\to 0\ \text{ when }\alpha\to\pi/2.

    In particular, {F(en+1,α)}α(0,π/2)\{\partial F(e_{n+1},\alpha)\}_{\alpha\in(0,\pi/2)} foliates n×(0,+)\mathbb{R}^{n}\times(0,+\infty).

Proof.

For each α(0,π/2)\alpha\in(0,\pi/2), in [AIC95] and [Din20], a unique O(n)O(n)-invariant entire convex graph Sα𝒞(en+1,α)S_{\alpha}\subset\mathcal{C}(e_{n+1},\alpha) solving self-expanding equation and asymptotic to 𝒞(en+1;α)\partial\mathcal{C}(e_{n+1};\alpha) is constructed, where the graphical function uαu_{\alpha} satisfies (i). Moreover, the uniqueness of SαS_{\alpha} guarantees they varies continuously in α\alpha. Let F(en+1,α)F(e_{n+1},\alpha) be the closed domain above SαS_{\alpha}.

As an entire graph, by [Din20, Lemma 3.1], SαS_{\alpha} are 𝐄\mathbf{E}-minimizing in n+1\mathbb{R}^{n+1}. Since when α0\alpha\searrow 0, Clos(𝒞(en+1,α)){ren+1:r0}\text{Clos}(\mathcal{C}(e_{n+1},\alpha))\to\{r\cdot e_{n+1}:r\geq 0\} locally in the Hausdorff distant sense, by compactness of minimizing boundary we must have, dist(𝟎,Sα)+dist(\mathbf{0},S_{\alpha})\to+\infty when α0\alpha\searrow 0, hence for each r0r\geq 0, uα(r)+u_{\alpha}(r)\to+\infty when α0\alpha\searrow 0. This also implies that for each α(0,π/2)\alpha\in(0,\pi/2), when α\alpha^{\prime} sufficiently close to 0, SαF(en+1,α)S_{\alpha^{\prime}}\subset F(e_{n+1},\alpha). Let α¯:=sup{β(0,α]:StF(en+1,α) for every t(0,β)}\bar{\alpha}:=\sup\{\beta\in(0,\alpha]:S_{t}\subset F(e_{n+1},\alpha)\text{ for every }t\in(0,\beta)\}. By strong maximum principle for 𝐄\mathbf{E}-functional, α¯=α\bar{\alpha}=\alpha and StIntF(en+1,α)S_{t}\subset\text{Int}F(e_{n+1},\alpha) whenever t(0,α)t\in(0,\alpha). In other words, uαu_{\alpha} is strictly monotone in α\alpha.

Note that (i) in the Theorem also implies that 0uα<cotα0\leq u_{\alpha}^{\prime}<\cot\alpha on [0,+)[0,+\infty) and is monotone increasing. Hence when απ/2\alpha\nearrow\pi/2, uαu_{\alpha} converges to a constant function whose graph is also a self-expander, hence must be 0. This finishes the proof of (ii). ∎

For every unit vector v𝕊nv\in\mathbb{S}^{n} and α(0,π/2)\alpha\in(0,\pi/2), let TO(n+1)T\in O(n+1) be an orthogonal transform mapping en+1e_{n+1} to vv, and denote F(v;α):=T(F(en+1,α))𝒞(v;α)F(v;\alpha):=T(F(e_{n+1},\alpha))\subset\mathcal{C}(v;\alpha) to be the domain with 𝐄\mathbf{E}-minimizing boundary distant asymptotic to 𝒞(v;α)\mathcal{C}(v;\alpha) near infinity, where F(en+1,α)F(e_{n+1},\alpha) is given by Theorem 2.4. This family of self-expanders are good barriers for general asymptotic conic self-expanders.

Corollary 2.5.

Let 𝕊n\mathcal{E}\subset\mathbb{S}^{n} be a closed subset such that Clos(Int())=\text{Clos}(\text{Int}(\mathcal{E}))=\mathcal{E}, let 0<R+0<R\leq+\infty. Suppose F=spt[F]n+1F=\text{spt}[F]\subset\mathbb{R}^{n+1} is a closed subset distant asymptotic to C()C(\mathcal{E}) near infinity with |[F]||\partial[F]| being 𝐄\mathbf{E}-stationary in (n+1C())𝔹R(\mathbb{R}^{n+1}\setminus\partial C(\mathcal{E}))\cup\mathbb{B}_{R} and coincide with C()C(\mathcal{E}) outside 𝔹R\mathbb{B}_{R}, i.e. F𝔹R=C()𝔹RF\setminus\mathbb{B}_{R}=C(\mathcal{E})\setminus\mathbb{B}_{R}. (We use the convention that 𝔹R=n+1\mathbb{B}_{R}=\mathbb{R}^{n+1} if R=+R=+\infty.) Then,

{F(v,α)F, if 𝒞(v;α)𝕊n;F(v,α)F=, if Clos(𝒞(v;α)𝕊n)=.\displaystyle\begin{cases}F(v,\alpha)\subset F,&\ \text{ if }\ \mathcal{C}(v;\alpha)\cap\mathbb{S}^{n}\subset\mathcal{E};\\ F(v,\alpha)\cap F=\emptyset,&\ \text{ if }\ \text{Clos}(\mathcal{C}(v;\alpha)\cap\mathbb{S}^{n})\cap\mathcal{E}=\emptyset.\end{cases}
Proof.

We prove for R<+R<+\infty; The case when R=+R=+\infty will be similar. Since F𝔹R=C()𝔹RF\setminus\mathbb{B}_{R}=C(\mathcal{E})\setminus\mathbb{B}_{R}, by Theorem 2.4 (ii), for sufficiently small β(0,α)\beta\in(0,\alpha), F(v,β)FF(v,\beta)\subset F. Consider

α¯:=sup{β(0,α):F(v,β)F},\bar{\alpha}:=\sup\{\beta\in(0,\alpha):F(v,\beta)\subset F\},

If α¯<α\bar{\alpha}<\alpha, then by continuity of F(v,β)F(v,\beta) in β\beta from Theorem 2.4, F(v,α¯)FF(v,\bar{\alpha})\subset F and F(v,α¯)spt([F])\partial F(v,\bar{\alpha})\cap\text{spt}(\|\partial[F]\|)\neq\emptyset. Since |[F]||\partial[F]| and |F(v,α¯)||\partial F(v,\bar{\alpha})| are both 𝐄\mathbf{E}-stationary in (n+1C())𝔹R(\mathbb{R}^{n+1}\setminus\partial C(\mathcal{E}))\cup\mathbb{B}_{R}, by strong maximum principle [SW89], F(v,α¯)spt([F])\partial F(v,\bar{\alpha})\subset\text{spt}(\partial[F]), contradicts to that Clos(𝒞(v;α¯))𝔹RInt(F)𝔹R\text{Clos}(\mathcal{C}(v;\bar{\alpha}))\setminus\mathbb{B}_{R}\subset\text{Int}(F)\setminus\mathbb{B}_{R}. Hence α¯=α\bar{\alpha}=\alpha and F(v,α)FF(v,\alpha)\subset F.

Similarly, when Clos(𝒞(v;α)𝕊n)=\text{Clos}(\mathcal{C}(v;\alpha)\cap\mathbb{S}^{n})\cap\mathcal{E}=\emptyset, the argument above provides,

sup{β(0,π/2):F(v,β)F=}>α.\sup\{\beta\in(0,\pi/2):F(v,\beta)\cap F=\emptyset\}>\alpha.

Hence by Theorem 2.4, F(v,α)F=F(v,\alpha)\cap F=\emptyset. ∎

Note that using Corollary 2.5, one can also prove the existence of 𝐄\mathbf{E}-minimizing closed subset distant asymptotic to C()C(\mathcal{E}) near infinity, as is sketched in [Ilm98, Lecture 2 F] and proved for C2C^{2}-domains \mathcal{E} in [Din20, Theorem 6.3].

In Section 5, we need the following existence result with constraint.

Lemma 2.6.

Let En+1E\subset\mathbb{R}^{n+1} be a closed C1,1C^{1,1}-optimally regular mean convex cone, which is not perimeter minimizing itself in EE. Then there exists a closed Caccioppoli set FInt(E)F\subset\text{Int}(E) such that F=spt[F]F=\text{spt}[F] is distant asymptotic to EE near \infty, and that [F]\partial[F] minimizes 𝐄\mathbf{E}-functional among

{[Q]:QΔFE}.\{\partial[Q]:Q\Delta F\subset E\}.

In particular, [F]\partial[F] is a self-expanding integral cycle with optimal regularity.

Proof.

For R1R\geq 1, minimize 𝐄\mathbf{E}-functional in 𝔹2R\mathbb{B}_{2R} among Caccioppoli sets

{Qn+1:E𝔹RQE}\{Q\subset\mathbb{R}^{n+1}:E\setminus\mathbb{B}_{R}\subset Q\subset E\}

to find a Caccioppoli set FREF_{R}\subset E. Since 𝔹R\partial\mathbb{B}_{R} and Reg(E)\mathrm{Reg}(\partial E) are both expander mean convex (i.e. HX/2\vec{H}-X^{\perp}/2 either vanishes or points inward), by [LIN85, Chapter 2], |[FR]||\partial[F_{R}]| is 𝐄\mathbf{E}-stationary in (n+1E)𝔹R(\mathbb{R}^{n+1}\setminus\partial E)\cup\mathbb{B}_{R}. WLOG FR=spt(FR)F_{R}=\text{spt}(F_{R}). When R+R\to+\infty, by Lemma 2.1 and Corollary 2.5, FRF_{R} subconverges in 𝐅\mathbf{F}-metric to some closed Caccioppoli set FEF\subset E which is distant asymptotic to EE near infinity, 𝐄\mathbf{E}-minimizing in EE and 𝐄\mathbf{E}-stationary in n+1\mathbb{R}^{n+1}. In particular, F\partial F is optimally regular.

By strong maximum principle [SW89, Ilm96], since FEF\neq E (Otherwise EE is 𝐄\mathbf{E}-minimizing in EE, and by blowing up EE at 𝟎\mathbf{0}, EE is perimeter minimizing in EE, which is a contradiction), we must have FInt(E)F\subset\text{Int}(E). ∎

2.3. Weak set flow and level set flow

Level set flow in the Euclidean space is introduced by [ES91] and independently by [CGG91], and is later generalized to defined on manifolds by [Ilm92, Ilm93]. Here, we shall use the notion of weak (set) flow and level set flow possibly with boundary, developed by [Whi95].

Definition 2.7.

[Whi95] Let MM be a compact nn-manifold, possibly with boundary, f:M×[a,b]n+1f:M\times[a,b]\to\mathbb{R}^{n+1} be a continuous map. Call :={(f(t,x),t):t[a,b],xM}n+1×\mathcal{M}:=\{(f(t,x),t):t\in[a,b],x\in M\}\subset\mathbb{R}^{n+1}\times\mathbb{R} a classical flow if,

  • ff is smooth on Int(M)×(a,b]\text{Int}(M)\times(a,b];

  • f(,t)f(\cdot,t) is one-to-one on MM for each t[a,b]t\in[a,b] and is a smooth embedding on Int(M)\text{Int}(M) for each t(a,b]t\in(a,b];

  • For each (x,t)Int(M)×(a,b](x,t)\in\text{Int}(M)\times(a,b], we have

    (tf(x,t))=H(x,t),(\frac{\partial}{\partial t}f(x,t))^{\perp}=\vec{H}(x,t),

    where means projection onto normal direction of f(M,t)f(M,t), and H(x,t)\vec{H}(x,t) denotes the mean curvature vector of f(M,t)f(M,t) at f(x,t)f(x,t).

Call

h:={(f(x,t),t):either t=a and xM, or t[a,b] and xM},\partial_{h}\mathcal{M}:=\{(f(x,t),t):\text{either }t=a\text{ and }x\in M,\text{ or }t\in[a,b]\text{ and }x\in\partial M\},

the heat boundary of \mathcal{M}.

Given a closed subset Γn+1×+\Gamma\subset\mathbb{R}^{n+1}\times\mathbb{R}_{+}, call a closed subset n+1×+\mathcal{M}\subset\mathbb{R}^{n+1}\times\mathbb{R}_{+} a weak (set) flow generated by Γ\Gamma at t=0t=0, if

  • (0)=Γ(0)\mathcal{M}(0)=\Gamma(0);

  • For every classical flow \mathcal{M}^{\prime} with h=\partial_{h}\mathcal{M}^{\prime}\cap\mathcal{M}=\emptyset and Γ=\mathcal{M}^{\prime}\cap\Gamma=\emptyset, we always have \mathcal{M}^{\prime}\cap\mathcal{M}.

In this paper, we need the following

Theorem 2.8.

[Whi95, Theorem 4.1] Let i=1,2i=1,2, i\mathcal{M}_{i} be weak flow generated by Γi\Gamma_{i} at t=0t=0. Suppose for each τ0\tau\geq 0, {(t,x)a:tτ}\{(t,x)\in\mathcal{M}_{a}:t\leq\tau\} is compact. Suppose also that at some T>0T>0,

distT(1,2)<minijdistT(i,Γj).dist_{T}(\mathcal{M}_{1},\mathcal{M}_{2})<min_{i\neq j}dist_{T}(\mathcal{M}_{i},\Gamma_{j}).

Then distt(1,2)distT(1,2)dist_{t}(\mathcal{M}_{1},\mathcal{M}_{2})\geq dist_{T}(\mathcal{M}_{1},\mathcal{M}_{2}) for tt on some interval [T,T+ϵ][T,T+\epsilon].

Here distt(1,2):=distn+1(1(t),2(t))dist_{t}(\mathcal{M}_{1},\mathcal{M}_{2}):=dist_{\mathbb{R}^{n+1}}(\mathcal{M}_{1}(t),\mathcal{M}_{2}(t)), where i(t):={xn+1:(x,t)i}\mathcal{M}_{i}(t):=\{x\in\mathbb{R}^{n+1}:(x,t)\in\mathcal{M}_{i}\}. In particular, this Theorem implies that whenever 1Γ2=2Γ1=\mathcal{M}_{1}\cap\Gamma_{2}=\mathcal{M}_{2}\cap\Gamma_{1}=\emptyset in Theorem 2.8, we always have 12=\mathcal{M}_{1}\cap\mathcal{M}_{2}=\emptyset.

[Whi95] also provide the construction of level set flow under this definition.

Proposition 2.9.

[Whi95, Proposition 5.1] Given a closed subset Γn+1×+\Gamma\subset\mathbb{R}^{n+1}\times\mathbb{R}_{+}, there’s a unique weak flow \mathcal{M} generated by Γ\Gamma at t=0t=0 (known as the level set flow) such that it contains every weak flow generated by Γ\Gamma at t=0t=0.

We end up this section with an example.

Lemma 2.10.

Let Wn+1W\subset\mathbb{R}^{n+1} be an open subset; ΣW\Sigma\subset W be an 𝐄\mathbf{E}-stable hypersurface with optimal regularity. Then for every WWW^{\prime}\subset\subset W and every λ0\lambda\geq 0,

:=tληt(WClos(Σ))×{tλ},\mathcal{M}:=\bigsqcup_{t\geq\lambda}\eta_{\sqrt{t}}(W^{\prime}\cap\text{Clos}(\Sigma))\times\{t-\lambda\},

is a weak (set) flow generated at t=0t=0 by

Γ=ηλ(WClos(Σ))×{0}tληt(WClos(Σ))×{tλ}.\Gamma=\eta_{\sqrt{\lambda}}(W^{\prime}\cap\text{Clos}(\Sigma))\times\{0\}\cup\bigsqcup_{t\geq\lambda}\eta_{\sqrt{t}}(\partial W^{\prime}\cap\text{Clos}(\Sigma))\times\{t-\lambda\}.
Proof.

Since the restriction of weak flow onto a sub-interval is also a weak flow, we can assume WLOG that λ=0\lambda=0. Suppose for contradiction that there exists a classical flow \mathcal{M}^{\prime} over [a,b](0,+)[a,b]\subset(0,+\infty) such that Γ=h=\mathcal{M}^{\prime}\cap\Gamma=\partial_{h}\mathcal{M}^{\prime}\cap\mathcal{M}=\emptyset but \mathcal{M}^{\prime}\cap\mathcal{M}\neq\emptyset. By definition,

𝐝(t):=distn+1((t),(t)),\mathbf{d}(t):=dist_{\mathbb{R}^{n+1}}(\mathcal{M}(t),\mathcal{M}^{\prime}(t)),

is a continuous function in t[a,b]t\in[a,b] and vanishes at some point in (a,b](a,b]; And

𝐝(t):=min{distn+1((t),h(t)),distn+1(Γ(t),(t))},\mathbf{d}_{\partial}(t):=min\{dist_{\mathbb{R}^{n+1}}(\mathcal{M}(t),\partial_{h}\mathcal{M}^{\prime}\ (t)),dist_{\mathbb{R}^{n+1}}(\Gamma(t),\mathcal{M}^{\prime}(t))\},

has infimum >0>0, and 𝐝(a)=𝐝(a)\mathbf{d}(a)=\mathbf{d}_{\partial}(a). Let

T:=inf{s[a,b]:𝐝(s)=0}(a,b],T:=\inf\{s\in[a,b]:\mathbf{d}(s)=0\}\in(a,b],

Then since 𝐝(T)>0\mathbf{d}_{\partial}(T)>0 and Int((t))\text{Int}(\mathcal{M}^{\prime}(t)) is smooth, we know that :=(T)(T)\mathcal{R}:=\mathcal{M}(T)\cap\mathcal{M}^{\prime}(T)\neq\emptyset is a closed subset contained in Reg((T))Γ(T)=ηT(Int(W)Σ)\mathrm{Reg}(\mathcal{M}(T))\setminus\Gamma(T)=\eta_{\sqrt{T}}(\text{Int}(W^{\prime})\cap\Sigma). Take a smooth closed neighborhood 𝒩η1/T()\mathcal{N}\supset\eta_{1/\sqrt{T}}(\mathcal{R}) in ΣInt(W)\Sigma\cap\text{Int}(W^{\prime}), we know that t[a,T]ηt(𝒩)×{t}\sqcup_{t\in[a,T]}\eta_{\sqrt{t}}(\mathcal{N})\times\{t\} and restriction of \mathcal{M}^{\prime} on [a,T][a,T] are two classical flows over [a,T][a,T] which intersects along t=Tt=T but not intersects along the heat boundary. This violates the maximum principle for classical flows [Whi95, Lemma 3.1]. ∎

Note that a more general result can be proved for integral Brakke motions with codimension 11 restricted to a spacetime closed subset, using the same argument as [Ilm94, 10.5]. But we don’t need it in this paper.

3. Harnack Inequality for Minimal Hypersurfaces

By [SS81], the convergence of stable minimal hypersurfaces might result in multiplicity in the limit. The goal of this section is to show that if a stable minimal hypersurface is away from higher multiplicity at scale 11, then so is it in every smaller scales. n3n\geq 3.

Denote 2\mathscr{M}^{\geq 2} to be the space of hypersurfaces with at least one piece of higher multiplicity , more precisely,

2:={V𝒱n(𝔹4n+1):\displaystyle\mathscr{M}^{\geq 2}:=\Big{\{}V\in\mathcal{I}\mathcal{V}_{n}(\mathbb{B}_{4}^{n+1}): V=i1mi|Σi| for some disjoint minimal hypersurfaces\displaystyle\ V=\sum_{i\geq 1}m_{i}|\Sigma_{i}|\text{ for some disjoint minimal hypersurfaces }
Σi(𝔹4,g) and mi such that ggEucC3,1/21/10,\displaystyle\ \Sigma_{i}\subset(\mathbb{B}_{4},g)\text{ and }m_{i}\in\mathbb{N}\text{ such that }\|g-g_{Euc}\|_{C^{3,1/2}}\leq 1/10,
m12 and Σ1Clos(𝔹2)}.\displaystyle\ m_{1}\geq 2\text{ and }\Sigma_{1}\cap\text{Clos}(\mathbb{B}_{2})\neq\emptyset\ \Big{\}}.
Proposition 3.1.

Let Λ>0\Lambda>0, ϵ>0\epsilon>0. There exists δ1(n,Λ)>0\delta_{1}(n,\Lambda)>0 and Ψ(ϵ|n,Λ)(0,1)\Psi(\epsilon|n,\Lambda)\in(0,1) with the following property. If Σ(𝔹4n+1,g)\Sigma\subset(\mathbb{B}_{4}^{n+1},g) is a stable minimal hypersurface with

(3.1) ggEucC4δ1,\displaystyle\|g-g_{Euc}\|_{C^{4}}\leq\delta_{1}, 𝐌(Σ)Λ,\displaystyle\mathbf{M}(\Sigma)\leq\Lambda, 𝐅(|Σ|,2)ϵ.\displaystyle\mathbf{F}(|\Sigma|,\mathscr{M}^{\geq 2})\geq\epsilon.

Then for every r(0,1/4)r\in(0,1/4) and xspt(Σ)𝔹1x\in\text{spt}(\Sigma)\cap\mathbb{B}_{1}, we have

𝐅𝔹4(|ηx,r1(Σ)|,2)Ψ(ϵ|n,Λ).\mathbf{F}_{\mathbb{B}_{4}}(|\eta_{x,r^{-1}}(\Sigma)|,\mathscr{M}^{\geq 2})\geq\Psi(\epsilon|n,\Lambda).

Here, Ψ(ϵ|n,Λ)\Psi(\epsilon|n,\Lambda) denotes a function in ϵ,n,Λ\epsilon,n,\Lambda, which tends to 0 when n,Λn,\Lambda are fixed and ϵ0\epsilon\to 0.

To prove Proposition 3.1, we need the following preparations.

Lemma 3.2.

For every Λ>0\Lambda>0, there exists δ2(n,Λ)>0\delta_{2}(n,\Lambda)>0 such that for every stable minimal hypercone Cn+1C\subset\mathbb{R}^{n+1} with C(𝔹1)Λ\|C\|(\mathbb{B}_{1})\leq\Lambda, we have

𝐅𝔹4(|C|,2)10δ2.\mathbf{F}_{\mathbb{B}_{4}}(|C|,\mathscr{M}^{\geq 2})\geq 10\delta_{2}.
Proof.

Suppose otherwise, there exist stable minimal hypercones CjC_{j} with uniformly bounded density at origin such that 𝐅𝔹4(|Cj|,2)0\mathbf{F}_{\mathbb{B}_{4}}(|C_{j}|,\mathscr{M}^{\geq 2})\to 0. Then by [SS81] and definition of 2\mathscr{M}^{\geq 2} we have, |Cj|m|C||C_{j}|\to m|C_{\infty}| for some stable minimal hypercone CC_{\infty} and integer m2m\geq 2. Also by the same argument as [Sha17, Section 4, Claim 5 & 6] we see that CjC_{j} induces a positive homogeneous degree 11 Jacobi field uCloc(C)u\in C_{loc}^{\infty}(C_{\infty}).

The goal next is to show that such Jacobi field do not exist. Let Γ:=C𝕊n\Gamma:=C_{\infty}\cap\mathbb{S}^{n}, and since uu is homogeneous degree 11, we can write u(rω)=rv(ω)u(r\omega)=r\cdot v(\omega), r>0r>0, ωΓ\omega\in\Gamma. By rewriting the Jacobi field equation in polar coordinates we derive

0=(ΔC+|AC|2)u\displaystyle 0=(\Delta_{C_{\infty}}+|A_{C_{\infty}}|^{2})u =[r2+n1rr+1r2(ΔΓ+|AΓ|2)](rv)\displaystyle=[\partial_{r}^{2}+\frac{n-1}{r}\partial_{r}+\frac{1}{r^{2}}(\Delta_{\Gamma}+|A_{\Gamma}|^{2})](r\cdot v)
=r1[ΔΓv+(|AΓ|2+n1)v],\displaystyle=r^{-1}\cdot[\Delta_{\Gamma}v+(|A_{\Gamma}|^{2}+n-1)v],

where ΔΓv+|AΓ|2+n1\Delta_{\Gamma}v+|A_{\Gamma}|^{2}+n-1 is the Jacobi operator of Γ𝕊n\Gamma\subset\mathbb{S}^{n}. Since v>0v>0, this implies Γ𝕊n\Gamma\subset\mathbb{S}^{n} is a stable minimal hypersurface, which is impossible. ∎

The following Almost Cone Rigidity Lemma follows directly from Schoen-Simon’s Compactness Theorem [SS81, Theorem 2] and monotonicity of area.

Lemma 3.3 (Almost Cone Rigidity).

For every ϵ>0\epsilon>0 and Λ>0\Lambda>0, there exists δ3(ϵ,Λ,n)(0,1)\delta_{3}(\epsilon,\Lambda,n)\in(0,1) such that if Σ(𝔹4n+1,g)\Sigma\subset(\mathbb{B}_{4}^{n+1},g) is a stable minimal hypersurface with Σ𝔹1\Sigma\cap\mathbb{B}_{1}\neq\emptyset and

ggEucC4δ3,\displaystyle\|g-g_{Euc}\|_{C^{4}}\leq\delta_{3}, Σ(𝔹4)Λ,\displaystyle\|\Sigma\|(\mathbb{B}_{4})\leq\Lambda, θg(𝟎,4;Σ)θg(𝟎,1;Σ)δ3.\displaystyle\theta_{g}(\mathbf{0},4;\|\Sigma\|)-\theta_{g}(\mathbf{0},1;\|\Sigma\|)\leq\delta_{3}.

Then there exists a stable minimal hypercone Cn+1C\subset\mathbb{R}^{n+1} and mm\in\mathbb{N} such that 𝐅𝔹2(|Σ|,m|C|)ϵ\mathbf{F}_{\mathbb{B}_{2}}(|\Sigma|,m|C|)\leq\epsilon.

Proof of Proposition 3.1.

Take δ1=δ3(δ2,n,Λ)/10\delta_{1}=\delta_{3}(\delta_{2},n,\Lambda)/10, given by Lemma 3.2 and 3.3.

Suppose for contradiction, there exists some ϵ>0\epsilon>0 and stable minimal hypersurfaces Σj(𝔹4,gj)\Sigma_{j}\subset(\mathbb{B}_{4},g_{j}), xjspt(Σj)𝔹1x_{j}\in\text{spt}(\Sigma_{j})\cap\mathbb{B}_{1} and rj(0,1/4)r_{j}\in(0,1/4) such that gjgC4δ1\|g_{j}-g\|_{C^{4}}\leq\delta_{1}, 𝐅𝔹4(|Σj|,2)ϵ\mathbf{F}_{\mathbb{B}_{4}}(|\Sigma_{j}|,\mathscr{M}^{\geq 2})\geq\epsilon but 𝐅𝔹4(|ηxj,rj1(Σj)|,2)0\mathbf{F}_{\mathbb{B}_{4}}(|\eta_{x_{j},r_{j}^{-1}}(\Sigma_{j})|,\mathscr{M}^{\geq 2})\to 0 when jj\to\infty. By unique continuation for stable minimal hypersurfaces, rj0r_{j}\to 0 as jj\to\infty. By possibly replace ϵ\epsilon by a smaller ϵ\epsilon^{\prime} and translate-rescale Σ\Sigma, WLOG xj𝟎x_{j}\equiv\mathbf{0}.

For each j1j\geq 1, let

Ij:={α:θgj(𝟎,4α;Σj)θgj(𝟎,4α1;Σj)δ1}{+,0}I_{j}:=\{\alpha\in\mathbb{N}:\theta_{g_{j}}(\mathbf{0},4^{-\alpha};\|\Sigma_{j}\|)-\theta_{g_{j}}(\mathbf{0},4^{-\alpha-1};\|\Sigma_{j}\|)\geq\delta_{1}\}\cup\{+\infty,0\}

Since θgj(𝟎,r;Σ)\theta_{g_{j}}(\mathbf{0},r;\|\Sigma\|) is non-decreasing in rr, by the mass upper bound on Σj\Sigma_{j}, we have IjC(Λ,n)\sharp I_{j}\leq C(\Lambda,n). Write

Ij=:{0=αj(0)<αj(1)<<αj(Ij1)=+};\displaystyle I_{j}=:\{0=\alpha_{j}^{(0)}<\alpha_{j}^{(1)}<...<\alpha_{j}^{(\sharp I_{j}-1)}=+\infty\}; αjk:=+, for kIj.\displaystyle\alpha_{j}^{k}:=+\infty,\ \text{ for }k\geq\sharp I_{j}.

Consider

i0:=sup{i0: for every sequence\displaystyle i_{0}:=\sup\Big{\{}i\in\mathbb{Z}_{\geq 0}:\text{ for every sequence } {0<sj[4αj(i)1,1/2)}j1,\displaystyle\{0<s_{j}\in[4^{-\alpha_{j}^{(i)}-1},1/2)\}_{j\geq 1},
lim infj𝐅𝔹4(|η𝟎,sj1(Σj)|,2)>0}.\displaystyle\liminf_{j\to\infty}\mathbf{F}_{\mathbb{B}_{4}}(|\eta_{\mathbf{0},s_{j}^{-1}}(\Sigma_{j})|,\mathscr{M}^{\geq 2})>0\Big{\}}.

By our contradiction assumption, 0i0<C(Λ,n)0\leq i_{0}<C(\Lambda,n) and for infinitely many jj, αji0<+\alpha_{j}^{i_{0}}<+\infty. Also by passing to a subsequence in jj, suppose 0<sj[4αj(i0+1)1,4αj(i0)1)0<s_{j}\in[4^{-\alpha_{j}^{(i_{0}+1)}-1},4^{-\alpha_{j}^{(i_{0})}-1}) such that 𝐅𝔹4(|ηsj1(Σj)|,2)0\mathbf{F}_{\mathbb{B}_{4}}(|\eta_{s_{j}^{-1}}(\Sigma_{j})|,\mathscr{M}^{\geq 2})\to 0 when jj\to\infty.

Denote for simplicity Σ^j:=η4αj(i0)+1(Σj)\hat{\Sigma}_{j}:=\eta_{4^{\alpha_{j}^{(i_{0})}+1}}(\Sigma_{j}), by definition of i0i_{0} and Schoen-Simon’s Compactness Theorem [SS81], we have

(3.2) |Σ^j||Σ^| in 𝔹2,\displaystyle|\hat{\Sigma}_{j}|\to|\hat{\Sigma}_{\infty}|\ \ \ \text{ in }\mathbb{B}_{2},

for some multiplicity 11 stable minimal hypersurface Σ^\hat{\Sigma}_{\infty} in 𝔹2\mathbb{B}_{2}; Also, s^j:=sj4αj(i0)+1(0,1)\hat{s}_{j}:=s_{j}\cdot 4^{\alpha_{j}^{(i_{0})}+1}\in(0,1) satisfies

(3.3) 𝐅(|ηs^j1(Σ^j)|,2)0 as j.\displaystyle\mathbf{F}(|\eta_{\hat{s}_{j}^{-1}}(\hat{\Sigma}_{j})|,\mathscr{M}^{\geq 2})\to 0\ \ \ \text{ as }j\to\infty.

And by definition of IjI_{j}, for every integer l[0,log4(s^j)2)l\in[0,-\log_{4}(\hat{s}_{j})-2), we have

(3.4) θgj(𝟎,4l;Σ^j)θgj(𝟎,4l1;Σ^j)δ1.\displaystyle\theta_{g_{j}}(\mathbf{0},4^{-l};\|\hat{\Sigma}_{j}\|)-\theta_{g_{j}}(\mathbf{0},4^{-l-1};\|\hat{\Sigma}_{j}\|)\leq\delta_{1}.

By (3.2) and (3.3), s^j0\hat{s}_{j}\searrow 0; By (3.4) and Lemma 3.3, for every j1j\geq 1 and l[1,log4(s^j)1)l\in[1,-\log_{4}(\hat{s}_{j})-1), there exists some stable minimal hypercone Cj(l)C_{j}^{(l)} and integer mj(l)m_{j}^{(l)}\in\mathbb{N} such that

(3.5) 𝐅𝔹4(|η4l(Σ^j)|,mj(l)|Cj(l)|)δ2.\displaystyle\mathbf{F}_{\mathbb{B}_{4}}(|\eta_{4^{l}}(\hat{\Sigma}_{j})|,m^{(l)}_{j}|C_{j}^{(l)}|)\leq\delta_{2}.

On the other hand, recall by [Ilm96], each tangent cone of a stable minimal hypersurface is a multiplicity 11 stable hypercone. Hence, by Lemma 3.2, there exists some integer l0=l0(Σ^)>>1l_{0}=l_{0}(\hat{\Sigma}_{\infty})>>1 such that

𝐅𝔹4(|η4l0(Σ^)|,2)9δ2.\mathbf{F}_{\mathbb{B}_{4}}(|\eta_{4^{l_{0}}}(\hat{\Sigma}_{\infty})|,\mathscr{M}^{\geq 2})\geq 9\delta_{2}.

And since |Σ^j||Σ^||\hat{\Sigma}_{j}|\to|\hat{\Sigma}_{\infty}|, we can choose j0>>1j_{0}>>1 such that for every jj0j\geq j_{0}, sj<4l03s_{j}<4^{-l_{0}-3} and

𝐅𝔹4(|η4l0(Σ^j)|,2)8δ2\mathbf{F}_{\mathbb{B}_{4}}(|\eta_{4^{l_{0}}}(\hat{\Sigma}_{j})|,\mathscr{M}^{\geq 2})\geq 8\delta_{2}

Combine this with (3.5), we derive 𝐅𝔹2(mj(l0)|Cj(l0)|,2)7δ2\mathbf{F}_{\mathbb{B}_{2}}(m^{(l_{0})}_{j}|C^{(l_{0})}_{j}|,\mathscr{M}^{\geq 2})\geq 7\delta_{2}, and hence mj(l0)=1m^{(l_{0})}_{j}=1. Moreover, recall for every two integral varifold V1,V2𝒱n(𝔹2n+1)V_{1},V_{2}\in\mathcal{I}\mathcal{V}_{n}(\mathbb{B}_{2}^{n+1}) and every λ>1\lambda>1, we have

𝐅𝔹2((ηλ)V1,(ηλ)V2)λ𝐅𝔹2(V1,V2).\mathbf{F}_{\mathbb{B}_{2}}((\eta_{\lambda})_{\sharp}V_{1},(\eta_{\lambda})_{\sharp}V_{2})\leq\lambda\cdot\mathbf{F}_{\mathbb{B}_{2}}(V_{1},V_{2}).

Thus combine with (3.5) we see,

𝐅𝔹2(mj(l)|Cj(l)|,mj(l+1)|Cj(l+1)|)\displaystyle\ \mathbf{F}_{\mathbb{B}_{2}}(m^{(l)}_{j}|C^{(l)}_{j}|,m^{(l+1)}_{j}|C^{(l+1)}_{j}|)
\displaystyle\leq 𝐅𝔹2(mj(l)|Cj(l)|,|η4l+1(Σ^j)|)+𝐅𝔹2(|η4l+1(Σ^j)|,mj(l+1)|Cj(l+1)|)4δ2+δ2=5δ2.\displaystyle\ \mathbf{F}_{\mathbb{B}_{2}}(m^{(l)}_{j}|C^{(l)}_{j}|,|\eta_{4^{l+1}}(\hat{\Sigma}_{j})|)+\mathbf{F}_{\mathbb{B}_{2}}(|\eta_{4^{l+1}}(\hat{\Sigma}_{j})|,m^{(l+1)}_{j}|C^{(l+1)}_{j}|)\leq 4\delta_{2}+\delta_{2}=5\delta_{2}.

Therefore, using Lemma 3.2 and starting at l0l_{0}, we can prove inductively that mj(l)=1m^{(l)}_{j}=1 for integer l[l0,log4(s^j)1)l\in[l_{0},-\log_{4}(\hat{s}_{j})-1) and again by Lemma 3.2 and (3.5), we have lim supj𝐅𝔹4(|ηs^j1(Σ^j)|,2)δ2>0\limsup_{j\to\infty}\mathbf{F}_{\mathbb{B}_{4}}(|\eta_{\hat{s}_{j}^{-1}}(\hat{\Sigma}_{j})|,\mathscr{M}^{\geq 2})\geq\delta_{2}>0. This contradicts to (3.3). ∎

Following the argument in [BG72], an immediate corollary of Proposition 3.1 is the Neumann-Sobolev type inequality below for stable minimal hypersurfaces,

Corollary 3.4 (Neumann-Sobolev Inequality).

Let ϵ,Λ>0\epsilon,\Lambda>0; δ1\delta_{1} be in Proposition 3.1. Then there exists β(n,ϵ,Λ)(0,1/2)\beta(n,\epsilon,\Lambda)\in(0,1/2), C(n,ϵ,Λ)>0C(n,\epsilon,\Lambda)>0 with the following property. If Σ\Sigma is a stable minimal hypersurface in (𝔹4n+1,g)(\mathbb{B}_{4}^{n+1},g) satisfying (3.1). Then for every fW1,1(Σ)f\in W^{1,1}(\Sigma), we have

infkfkLnn1(𝔹β;Σ)C𝔹4|Σf|dΣ.\inf_{k\in\mathbb{R}}\|f-k\|_{L^{\frac{n}{n-1}}(\mathbb{B}_{\beta};\|\Sigma\|)}\leq C\cdot\int_{\mathbb{B}_{4}}|\nabla_{\Sigma}f|\ d\|\Sigma\|.
Proof.

We shall first show the following,

Claim. There exists β(n,ϵ,Λ)(0,1/2)\beta(n,\epsilon,\Lambda)\in(0,1/2), C~(n,ϵ,Λ)>0\tilde{C}(n,\epsilon,\Lambda)>0 with the following property. If Σ\Sigma is a stable minimal hypersurface in (𝔹4n+1,g)(\mathbb{B}_{4}^{n+1},g) satisfying (3.1); T1,T2𝐈n(𝔹4)T_{1},T_{2}\in\mathbf{I}^{n}(\mathbb{B}_{4}) are integral currents such that

(3.6) [Σ]=T1+T2;\displaystyle[\Sigma]=T_{1}+T_{2}; Σ=T1+T2.\displaystyle\|\Sigma\|=\|T_{1}\|+\|T_{2}\|.

Then we have for every xspt(Σ)𝔹1x\in\text{spt}(\Sigma)\cap\mathbb{B}_{1},

(3.7) 𝐌𝔹4(T1)=𝐌𝔹4(T2)C~1min{𝐌𝔹β(x)(T1),𝐌𝔹β(x)(T2)}11/n.\displaystyle\mathbf{M}_{\mathbb{B}_{4}}(\partial T_{1})=\mathbf{M}_{\mathbb{B}_{4}}(\partial T_{2})\geq\tilde{C}^{-1}\cdot min\{\mathbf{M}_{\mathbb{B}_{\beta}(x)}(T_{1}),\mathbf{M}_{\mathbb{B}_{\beta}(x)}(T_{2})\}^{1-1/n}.

Proof of Claim. By Lemma 3.2, by a translation and rescaling of Σ\Sigma and considering a smaller ϵ\epsilon, WLOG x=𝟎x=\mathbf{0}. Suppose the Claim fails, then there are stable minimal hypersurfaces Σj\Sigma^{j} satisfying (3.1) and integral currents T1j,T2j𝐈n(𝔹4)T^{j}_{1},T^{j}_{2}\in\mathbf{I}_{n}(\mathbb{B}_{4}) such that [Σj]=T1j+T2j[\Sigma^{j}]=T^{j}_{1}+T^{j}_{2}, Σj=T1j+T2j\|\Sigma^{j}\|=\|T^{j}_{1}\|+\|T^{j}_{2}\| and that

(3.8) 𝐌𝔹4(T1j)=𝐌𝔹4(T2j)<1jmin{𝐌𝔹j1(T1j),𝐌𝔹j1(T2j)}11/n.\displaystyle\mathbf{M}_{\mathbb{B}_{4}}(\partial T^{j}_{1})=\mathbf{M}_{\mathbb{B}_{4}}(\partial T^{j}_{2})<\frac{1}{j}\cdot min\{\mathbf{M}_{\mathbb{B}_{j^{-1}}}(T^{j}_{1}),\mathbf{M}_{\mathbb{B}_{j^{-1}}}(T^{j}_{2})\}^{1-1/n}.

Also by slicing Theorem [Sim83], for each T=TijT=T^{j}_{i} (j1j\geq 1, i=1,2i=1,2) and a.e. s(j1,4)s\in(j^{-1},4),

𝐌𝔹4((T𝔹s))𝐌𝔹s(T)+dds𝐌𝔹s(T).\mathbf{M}_{\mathbb{B}_{4}}(\partial(T\llcorner\mathbb{B}_{s}))\leq\mathbf{M}_{\mathbb{B}_{s}}(\partial T)+\frac{d}{ds}\mathbf{M}_{\mathbb{B}_{s}}(T).

Combined with (3.8) and Micheal-Simon inequality [Sim83, Theorem 18.6], this implies

(3.9) 𝐌𝔹s(T)n1nS(n)j𝐌𝔹s(T)n1n+dds𝐌𝔹s(T), for a.e. t(j1,1).\displaystyle\mathbf{M}_{\mathbb{B}_{s}}(T)^{\frac{n-1}{n}}\leq\frac{S(n)}{j}\mathbf{M}_{\mathbb{B}_{s}}(T)^{\frac{n-1}{n}}+\frac{d}{ds}\mathbf{M}_{\mathbb{B}_{s}}(T),\ \ \ \text{ for a.e. }t\in(j^{-1},1).

Take j10S(n)j\geq 10S(n) in (3.9) and integrate over (j1,t)(j^{-1},t), we derive

(3.10) 𝐌𝔹t(Tij)δ(n)(tj1)n,t(j1,1).\displaystyle\mathbf{M}_{\mathbb{B}_{t}}(T^{j}_{i})\geq\delta(n)\cdot(t-j^{-1})^{n},\ \ \ \forall t\in(j^{-1},1).

By Proposition 3.1 and Schoen-Simon compactness, when jj\to\infty, |η𝟎,j(Σj)||Σ||\eta_{\mathbf{0},j}(\Sigma^{j})|\to|\Sigma^{\infty}| in 𝒱n(n)\mathcal{I}\mathcal{V}_{n}(\mathbb{R}^{n}) for some multiplicity 11 stable minimal hypersurface Σ\Sigma^{\infty}. Moreover, since the tangent cone CC of Σ\Sigma^{\infty} at infinity is also some rescaled limit of Σj\Sigma^{j}, we must have by Proposition 3.1 that CC is of multiplicity 11 and hence Σ\Sigma^{\infty} is connected.

By (3.8), Fleming-Federer Compactness Theorem and area monotonicity formula for minimal hypersurfaces,

𝐅((η𝟎,j)Tij,Ti)0,\mathbf{F}((\eta_{\mathbf{0},j})_{\sharp}T^{j}_{i},T^{\infty}_{i})\to 0,

for some Ti𝐈n(n+1)T^{\infty}_{i}\in\mathbf{I}_{n}(\mathbb{R}^{n+1}) (i=1,2i=1,2) such that

T1=T2=0;\displaystyle\partial T^{\infty}_{1}=\partial T^{\infty}_{2}=0; T1+T2=[Σ],\displaystyle T^{\infty}_{1}+T^{\infty}_{2}=[\Sigma^{\infty}], T1+T2=Σ.\displaystyle\|T^{\infty}_{1}\|+\|T^{\infty}_{2}\|=\|\Sigma^{\infty}\|.

And by (3.10), spt(T1)𝔹2\text{spt}(T^{\infty}_{1})\cap\mathbb{B}_{2}\neq\emptyset, spt(T2)𝔹2\text{spt}(T^{\infty}_{2})\cap\mathbb{B}_{2}\neq\emptyset. But by constancy Theorem [Sim83, Theorem 26.27], this is impossible for a connected stable minimal hypersurface Σ\Sigma^{\infty}. Hence the Claim is proved.

Now to prove Corollary 3.4, suppose WLOG fC1(Σ)f\in C^{1}(\Sigma), and denote for simplicity p:=n/(n1)p:=n/(n-1). Choose kk\in\mathbb{R} such that

(3.11) 𝐌𝔹β({f>k}),𝐌𝔹β({f<k})12𝐌𝔹β(Σ).\displaystyle\mathbf{M}_{\mathbb{B}_{\beta}}(\{f>k\}),\ \mathbf{M}_{\mathbb{B}_{\beta}}(\{f<k\})\leq\frac{1}{2}\mathbf{M}_{\mathbb{B}_{\beta}}(\Sigma).

Let f±:=(fk)±=max{±(fk),0}f^{\pm}:=(f-k)^{\pm}=max\{\pm(f-k),0\}; A±(t):=[Σ]{f±>t}A^{\pm}(t):=[\Sigma]\llcorner\{f^{\pm}>t\}. Then by co-area formula, we have

𝔹4Σ|Σf+|dΣ\displaystyle\int_{\mathbb{B}_{4}\cap\Sigma}|\nabla_{\Sigma}f^{+}|\ d\|\Sigma\| =0𝐌𝔹4(A+(t))𝑑t\displaystyle=\int_{0}^{\infty}\mathbf{M}_{\mathbb{B}_{4}}(\partial A^{+}(t))\ dt
C~10𝐌𝔹β(A+(t))1/p𝑑t\displaystyle\geq\tilde{C}^{-1}\int_{0}^{\infty}\mathbf{M}_{\mathbb{B}_{\beta}}(A^{+}(t))^{1/p}\ dt
C~1(p0𝐌𝔹β(A+(t))tp1𝑑t)1/p\displaystyle\geq\tilde{C}^{-1}\Big{(}p\cdot\int_{0}^{\infty}\mathbf{M}_{\mathbb{B}_{\beta}}(A^{+}(t))t^{p-1}\ dt\Big{)}^{1/p}
=C1f+Lp(𝔹β;Σ),\displaystyle=C^{-1}\|f^{+}\|_{L^{p}(\mathbb{B}_{\beta};\|\Sigma\|)},

where the first inequality follows from the Claim and choice of kk; the second inequality follows from the following Hardy-Littlewood-Polya Inequality: If p>1p>1 and η0\eta\geq 0 is a non-increasing function on (0,+)(0,+\infty), then

(0η1/p𝑑t)pp0η(t)tp1𝑑t.\Big{(}\int_{0}^{\infty}\eta^{1/p}\ dt\Big{)}^{p}\geq p\int_{0}^{\infty}\eta(t)t^{p-1}\ dt.

Similarly one derive 𝔹4|Σf|dΣC1fLp(𝔹β;Σ)\int_{\mathbb{B}_{4}}|\nabla_{\Sigma}f^{-}|\ d\|\Sigma\|\geq C^{-1}\|f^{-}\|_{L^{p}(\mathbb{B}_{\beta};\|\Sigma\|)}. Hence Corollary 3.4 is proved. ∎

By De Giorgi-Moser iteration process [GT01] and the abstract John-Nirenberg Ineq [BG72, Section 4], Corollary 3.4 implies the following Harnack inequality for stable minimal hypersurfaces.

Corollary 3.5 (Harnack Inequality).

Let Σ\Sigma be in Corollary 3.4. There exists η(ϵ,n,Λ)(0,1)\eta(\epsilon,n,\Lambda)\in(0,1) such that if 0<uWloc1,2(Σ)0<u\in W^{1,2}_{loc}(\Sigma) satisfies ΔΣuΛu0\Delta_{\Sigma}u-\Lambda u\leq 0 on Σ\Sigma in the distribution sense, then for every p(0,nn2)p\in(0,\frac{n}{n-2}), every xspt(Σ)𝔹1x\in\text{spt}(\Sigma)\cap\mathbb{B}_{1} and every r(0,η)r\in(0,\eta), we have

(1Σ(𝔹r(x))Σ𝔹r(x)|u|pdΣ)1/pC(p,ϵ,Λ,n)infΣ𝔹r(x)u.\big{(}\frac{1}{\|\Sigma\|(\mathbb{B}_{r}(x))}\int_{\Sigma\cap\mathbb{B}_{r}(x)}|u|^{p}\ d\|\Sigma\|\big{)}^{1/p}\leq C(p,\epsilon,\Lambda,n)\cdot\inf_{\Sigma\cap\mathbb{B}_{r}(x)}u.
Remark 3.6.

By a slight modification, a similar Neumann-Sobolev Inequality and Harnack inequality still holds for minimal hypersurfaces with bounded index, where the constant will also depend on the index upper bound.

On the other hand, the assumption in (3.1) that Σ\Sigma is away from varifolds with multiplicity 2\geq 2 at scale 11 CANNOT be dropped. A typical counterexample is a sequence of pairs of parallel hyperplanes Vj:=|Lj(1)|+|Lj(2)|V_{j}:=|L^{(1)}_{j}|+|L^{(2)}_{j}| converging to a multiplicity 22 hyperplane, and a sequence of functions on spt(Vj)\text{spt}(V_{j}) which is 11 on Lj(1)L^{(1)}_{j} and 1/j1/j on Lj(2)L^{(2)}_{j}. Counterexample in unstable situation could be a blow down sequence of catenoids converging to a multiplicity 22 hyperplane and a positive harmonic function which approaches 11 on one end and 0 on the other end.

4. Growth Rate Lower Bound for Positive Jacobi Fields

The goal of this section is to prove the following.

Proposition 4.1.

Let ϵ(0,1)\epsilon\in(0,1), Λ>0\Lambda>0, n7n\geq 7;

γn:=n22+(n22)2n+1\gamma_{n}:=-\frac{n-2}{2}+\sqrt{(\frac{n-2}{2})^{2}-n+1}

For every γ>γn\gamma>\gamma_{n}, there exists C=C(ϵ,Λ,n,γ)>0C=C(\epsilon,\Lambda,n,\gamma)>0 and τ0(ϵ,Λ,n,γ)(0,1/2)\tau_{0}(\epsilon,\Lambda,n,\gamma)\in(0,1/2) such that the following hold.

Let Σ(𝔹4n+1,g)\Sigma\subset(\mathbb{B}_{4}^{n+1},g) be a two-sided stable minimal hypersurface satisfying (3.1); uWloc1,2(Σ)u\in W^{1,2}_{loc}(\Sigma) be a positive super-solution of (ΔΣ+|AΣ|2Λ)u=0(\Delta_{\Sigma}+|A_{\Sigma}|^{2}-\Lambda)u=0, i.e.

ΣΣuΣξ+(Λ|AΣ|2)uξ0,ξCc1(Σ,0).\int_{\Sigma}\nabla_{\Sigma}u\cdot\nabla_{\Sigma}\xi+(\Lambda-|A_{\Sigma}|^{2})u\xi\ \geq 0,\ \ \ \forall\xi\in C^{1}_{c}(\Sigma,\mathbb{R}_{\geq 0}).

Then for every pSing(Σ)𝔹1p\in\mathrm{Sing}(\Sigma)\cap\mathbb{B}_{1} and every r(0,τ0)r\in(0,\tau_{0}), we have

(4.1) infΣ𝔹r(p)uCuL1(Σ𝔹τ0(p))rγ.\displaystyle\inf_{\Sigma\cap\mathbb{B}_{r}(p)}u\geq C\cdot\|u\|_{L^{1}(\Sigma\cap\mathbb{B}_{\tau_{0}}(p))}\cdot r^{\gamma}.

Note that the exponent γn\gamma_{n} is optimal and can be realized by a homogeneous positive Jacobi field on Simons cone.

The key for the proof is the following Growth Lemma. Let

𝒞n=1(Λ):={|C|:Cn+1 stable minimal hypercone with C(𝔹4)Λ}{hyperplanes},\mathscr{C}_{n}^{=1}(\Lambda):=\{|C|:C\subset\mathbb{R}^{n+1}\text{ stable minimal hypercone with }\|C\|(\mathbb{B}_{4})\leq\Lambda\}\setminus\{\text{hyperplanes}\},

be the space of multiplicity 11 varifolds associated to nontrivial stable minimal hypercones. For a function ϕ\phi over Σ\Sigma, p>0p>0 and a domain ΩΣ\Omega\subset\Sigma, let

ϕLp(Ω;Σ):=(Σ(Ω)1Ω|ϕ|pdΣ)1/p.\|\phi\|_{L^{p}_{*}(\Omega;\|\Sigma\|)}:=\big{(}\|\Sigma\|(\Omega)^{-1}\cdot\int_{\Omega}|\phi|^{p}\ d\|\Sigma\|\big{)}^{1/p}.
Lemma 4.2 (Growth Lemma).

Let Λ>0\Lambda>0, γ>γn\gamma>\gamma_{n} be fixed as in Prop 4.1. There exists δ(n,Λ,γ)(0,1)\delta(n,\Lambda,\gamma)\in(0,1) and r0(n,Λ,γ)(0,1/2)r_{0}(n,\Lambda,\gamma)\in(0,1/2) with the following property.

Let Σ(𝔹4r01,g)\Sigma\subset(\mathbb{B}_{4r_{0}^{-1}},g) be a two-sided stable minimal hypersurface with normal field ν\nu, satisfying

(4.2) Sing(Σ)𝟎,\displaystyle\mathrm{Sing}(\Sigma)\ni\mathbf{0}, ggEucC4δ,\displaystyle\|g-g_{Euc}\|_{C^{4}}\leq\delta, 𝐅𝔹4(|Σ|,𝒞n=1(Λ))δ.\displaystyle\mathbf{F}_{\mathbb{B}_{4}}(|\Sigma|,\mathscr{C}_{n}^{=1}(\Lambda))\leq\delta.

Let 0<uWloc1,2(Σ)0<u\in W^{1,2}_{loc}(\Sigma) be a positive supersolution of ΔΣu+(|AΣ|2δ)u=0\Delta_{\Sigma}u+(|A_{\Sigma}|^{2}-\delta)u=0 in the distribution sense on Σ\Sigma, in other words,

(4.3) ΣuΣξ+(δ|AΣ|2)uξdΣ0,ξCc1(Σ,0).\displaystyle\int\nabla_{\Sigma}u\cdot\nabla_{\Sigma}\xi+(\delta-|A_{\Sigma}|^{2})u\xi\ d\|\Sigma\|\geq 0,\ \ \ \forall\xi\in C^{1}_{c}(\Sigma,\mathbb{R}_{\geq 0}).

Then uL1(𝔹1;Σ)u\in L^{1}(\mathbb{B}_{1};\|\Sigma\|) and

uL1(𝔹r;Σ)uL1(𝔹1;Σ)rγ,r[r0/2,r0].\|u\|_{L^{1}_{*}(\mathbb{B}_{r};\|\Sigma\|)}\geq\|u\|_{L^{1}_{*}(\mathbb{B}_{1};\|\Sigma\|)}\cdot r^{\gamma},\ \ \ \forall r\in[r_{0}/2,r_{0}].

In [Sim08, Theorem 1], a general Growth Theorem is established for multiplicity 11 class of minimal submanifolds. It follows directly from Proposition 3.1 that minimal hypersurfaces in Euclidean balls together with their blow-up limits forms a multiplicity 11 class of hypersurfaces. Hence Simon’s result applies here to obtain Lemma 4.2. For sake of completeness, we include in the Appendix A a direct proof using Harnack Inequality in Corollary 3.5.

Proof of Proposition 4.1.

WLOG p=𝟎p=\mathbf{0}. Let τ:=min{r0N+1:r0Nη(ϵ,Λ,n)/4}\tau:=min\{r_{0}^{N+1}:r_{0}^{N}\geq\eta(\epsilon,\Lambda,n)/4\} be given by Lemma 4.2 and Corollary 3.5; τ0:=τ2\tau_{0}:=\tau^{2}. By definition, τ,τ0\tau,\tau_{0} only depend on ϵ,Λ,n\epsilon,\Lambda,n. Let

J:={j2:ητj1(Σ)𝔹4r01(𝔹4r01,τ2jητjg) satisfies (4.2),uητj satisfies (4.3)}.J:=\{j\in\mathbb{Z}_{\geq 2}:\eta_{\tau^{j}}^{-1}(\Sigma)\cap\mathbb{B}_{4r_{0}^{-1}}\subset(\mathbb{B}_{4r_{0}^{-1}},\tau^{-2j}\eta_{\tau^{j}}^{*}g)\text{ satisfies }(\ref{Equ_Almost Cone Assumption}),u\circ\eta_{\tau^{j}}\text{ satisfies }(\ref{Equ_u satisf Perturb Jac supersol})\}.

By area monotonicity, Lemma 3.3 and Proposition 3.1, we see that

(4.4) (J)N(n,Λ,ϵ,γ)<+.\displaystyle\sharp(\mathbb{N}\setminus J)\leq N(n,\Lambda,\epsilon,\gamma)<+\infty.

Now for every r(0,τ0)r\in(0,\tau_{0}), let l2l\in\mathbb{Z}_{\geq 2} be such that τl+1r<τl\tau^{l+1}\leq r<\tau^{l}. Denote for simplicity Iu(r):=uL1(𝔹r;Σ)I_{u}(r):=\|u\|_{L^{1}_{*}(\mathbb{B}_{r};\|\Sigma\|)}. Note that by Corollary 3.5, for every j1j\geq 1,

(4.5) Iu(τj+1)1Σ(𝔹τj+1)inf𝔹τjuc(n,Λ,ϵ)Iu(τj).\displaystyle I_{u}(\tau^{j+1})\geq\frac{1}{\|\Sigma\|(\mathbb{B}_{\tau^{j+1}})}\inf_{\mathbb{B}_{\tau^{j}}}u\geq c(n,\Lambda,\epsilon)I_{u}(\tau^{j}).

And for every jJj\in J, by Lemma 4.2, we have better estimate,

(4.6) Iu(τj+1)Iu(τj)τγ.\displaystyle I_{u}(\tau^{j+1})\geq I_{u}(\tau^{j})\cdot\tau^{\gamma}.

Combine Corollary 3.5, (4.4), (4.5) and (4.6) we have,

infΣ𝔹ru\displaystyle\inf_{\Sigma\cap\mathbb{B}_{r}}u c(n,Λ,ϵ)uL1(𝔹τl;Σ)\displaystyle\geq c(n,\Lambda,\epsilon)\|u\|_{L^{1}_{*}(\mathbb{B}_{\tau^{l};\|\Sigma\|})}
=c(n,Λ,ϵ)(j=2l1Iu(τj+1)Iu(τj))uL1(𝔹τ0;Σ)\displaystyle=c(n,\Lambda,\epsilon)\Big{(}\prod_{j=2}^{l-1}\frac{I_{u}(\tau^{j+1})}{I_{u}(\tau^{j})}\Big{)}\cdot\|u\|_{L^{1}_{*}(\mathbb{B}_{\tau_{0}};\|\Sigma\|)}
=c(n,Λ,ϵ)(2jl1;jJIu(τj+1)Iu(τj))(2jl1;jJIu(τj+1)Iu(τj))uL1(𝔹τ0;Σ)\displaystyle=c(n,\Lambda,\epsilon)\Big{(}\prod_{2\leq j\leq l-1;j\in J}\frac{I_{u}(\tau^{j+1})}{I_{u}(\tau^{j})}\Big{)}\cdot\Big{(}\prod_{2\leq j\leq l-1;j\notin J}\frac{I_{u}(\tau^{j+1})}{I_{u}(\tau^{j})}\Big{)}\cdot\|u\|_{L^{1}_{*}(\mathbb{B}_{\tau_{0}};\|\Sigma\|)}
c(n,Λ,ϵ)(τγ)l(J)c(n,Λ,ϵ)(J)uL1(𝔹τ0;Σ)\displaystyle\geq c(n,\Lambda,\epsilon)\cdot(\tau^{\gamma})^{l-\sharp(\mathbb{N}\setminus J)}\cdot c(n,\Lambda,\epsilon)^{\sharp(\mathbb{N}\setminus J)}\cdot\|u\|_{L^{1}_{*}(\mathbb{B}_{\tau_{0}};\|\Sigma\|)}
c(n,Λ,ϵ,γ)uL1(𝔹τ0;Σ)rγ.\displaystyle\geq c(n,\Lambda,\epsilon,\gamma)\|u\|_{L^{1}_{*}(\mathbb{B}_{\tau_{0}};\|\Sigma\|)}\cdot r^{\gamma}.

This finishes the proof of the Proposition. ∎

Since γn<1<0\gamma_{n}<-1<0, a direct corollary of Proposition 4.1 is,

Corollary 4.3.

Let (M,g)(M,g) be a smooth Riemannian manifold (not necessarily complete), Λ>0\Lambda>0. Let ΣM\Sigma\subset M be a two-sided stable minimal hypersurface with optimal regularity. Let uCloc2(Σ)u\in C^{2}_{loc}(\Sigma) be a positive function such that

ΔΣu+|AΣ|2uΛu0,\Delta_{\Sigma}u+|A_{\Sigma}|^{2}u-\Lambda u\leq 0,

on Σ\Sigma. Then for every xSing(Σ)x\in\mathrm{Sing}(\Sigma),

lim infyxu(y)=+.\liminf_{y\to x}u(y)=+\infty.

5. Proof of Main Theorem

Throughout this section, let En+1E\subset\mathbb{R}^{n+1} be a closed viscosity mean convex cone with connected nonempty interior. Let :=E𝕊n\mathcal{E}:=E\cap\mathbb{S}^{n} be a closed viscosity mean convex domain in 𝕊n\mathbb{S}^{n}. Let’s start with the following regularity Lemma.

Lemma 5.1.

Suppose [E]\partial[E] is not minimizing in EE (This is automatically true if EE is not a Caccioppoli set). Then there exists a closed subset PEP\subset E such that,

  1. (i)

    PE𝔹1P\supset E\setminus\mathbb{B}_{1} and P𝔹1Int(E)P\cap\mathbb{B}_{1}\subset\text{Int}(E);

  2. (ii)

    P𝔹1\partial P\cap\mathbb{B}_{1} is a smooth radial graph over Int(E)𝕊n\text{Int}(E)\cap\mathbb{S}^{n}, i.e., there exists vC(Int();(0,1))v\in C^{\infty}(\text{Int}(\mathcal{E});(0,1)) such that

    P𝔹1={rω:ωInt(),v(ω)r<1};P\cap\mathbb{B}_{1}=\{r\omega:\omega\in\text{Int}(\mathcal{E}),v(\omega)\leq r<1\};
  3. (iii)

    [P]𝔹1\partial[P]\llcorner\mathbb{B}_{1} is area-minimizing in 𝔹1Int(E)\mathbb{B}_{1}\cap\text{Int}(E).

Proof.

Let’s first further assume that E𝕊n\partial E\cap\mathbb{S}^{n} is C1,1C^{1,1} optimally regular mean convex in 𝕊n\mathbb{S}^{n}. Consider minimize perimeter in 𝔹2\mathbb{B}_{2} among family of Caccioppoli sets

{Q:E𝔹1QE},\{Q:E\setminus\mathbb{B}_{1}\subset Q\subset E\},

to get a Caccioppoli set PEP\subset E. Since [E]\partial[E] is not minimizing in EE, we must have PEP\neq E. By [LIN85, Chapter 2], since EE and Clos(𝔹1)\text{Clos}(\mathbb{B}_{1}) are C1,1C^{1,1} mean convex away from a low dimensional subset, we know that [P]\|\partial[P]\| is a stationary integral varifold in (n+1E)𝔹1(\mathbb{R}^{n+1}\setminus\partial E)\cup\mathbb{B}_{1}. Then by Solomon-White’s strong maximum principle [SW89] and Ilmanen’s strong-maximum principle [Ilm96], either E=PE=P (which is impossible) or spt(P)E𝔹1{𝟎}\text{spt}(P)\cap\partial E\cap\mathbb{B}_{1}\subset\{\mathbf{0}\}. If 𝟎spt(P)\mathbf{0}\in\text{spt}(P), then by Lemma 2.1, the tangent cone of PP at 𝟎\mathbf{0} is also perimeter-minimizing in EE, hence cannot coincide with EE, this also violates the strong maximum principle for cross sections of EE and PP in SSnSS^{n}.

Now we have shown that spt(P)𝔹1Int(E)\text{spt}(P)\cap\mathbb{B}_{1}\subset\text{Int}(E), since PP minimize perimeter in E𝔹1E\cap\mathbb{B}_{1}, we have dimSing([P])n7\dim\mathrm{Sing}(\partial[P])\leq n-7. In particular, WLOG P=spt[P]P=\text{spt}[P]. Consider the rescalings of PP and

(5.1) r:=sup{s>0:ηt(P)P,t(0,s)}(0,1].\displaystyle r:=\sup\{s>0:\eta_{t}(P)\supset P,\forall t\in(0,s)\}\in(0,1].

If r<1r<1, then there exists y𝔹1Pηr(P)y\in\mathbb{B}_{1}\cap\partial P\cap\eta_{r}(\partial P) being a touching point of Pηr(P)P\subset\eta_{r}(P) on the boundary, which violates strong maximum principle [Ilm96]. Hence r=1r=1.

Let νP\nu_{P} be the inward pointed normal field defined on Reg(P)\mathrm{Reg}(\partial P). Consider the Jacobi field ϕ:=XνP\phi:=X\cdot\nu_{P} on Reg(P)\mathrm{Reg}(\partial P) induced by rescalings, where XX is the position vector. Since r=1r=1 in (5.1), we know that ϕ0\phi\geq 0 on Reg(P𝔹1)\mathrm{Reg}(\partial P\cap\mathbb{B}_{1}). And since ϕ\phi satisfies the Jacobi field equation

Δϕ+|AP|2ϕ=0,\Delta\phi+|A_{\partial P}|^{2}\phi=0,

by strong maximum principle, either ϕ0\phi\equiv 0 (which then implies PP being a cone, contradicts to that 𝟎P\mathbf{0}\notin P), or ϕ>0\phi>0 on regular part of P\partial P. Then by Corollary 4.3 and that ϕ1\phi\leq 1, we know that Sing(P𝔹1)=\mathrm{Sing}(\partial P\cap\mathbb{B}_{1})=\emptyset. And thus, by ϕ>0\phi>0 and r=1r=1 in (5.1), we know that the projection xx/|x|x\mapsto x/|x| is injective restricting to P𝔹1\partial P\cap\mathbb{B}_{1} and onto 𝕊n\mathcal{E}\subset\mathbb{S}^{n}, which indicates that P𝔹1\partial P\cap\mathbb{B}_{1} is a radial graph.

Now for a general EE, by Corollary B.3, we can approximate \mathcal{E} from interior by a family of connected C1,1C^{1,1}-optimally regular mean convex closed subset j𝕊n\mathcal{E}_{j}\subset\mathbb{S}^{n}. Let PjC(j)P_{j}\subset C(\mathcal{E}_{j}) be the corresponding closed subset given above, which is perimeter minimizing in C(j)𝔹1C(\mathcal{E}_{j})\cap\mathbb{B}_{1} and has stationary boundary in (n+1C(j))𝔹1(\mathbb{R}^{n+1}\setminus\partial C(\mathcal{E}_{j}))\cup\mathbb{B}_{1}, with inward pointed normal vector νj\nu_{j} on Pj𝔹1\partial P_{j}\cap\mathbb{B}_{1}, and ϕj:=rrνj\phi_{j}:=r\partial_{r}\cdot\nu_{j} be the Jacobi field on Pj𝔹1\partial P_{j}\cap\mathbb{B}_{1} induced by rescaling. Let jj\to\infty, by Lemma 2.1, after passing to a subsequence, PjP_{j} converges to some closed subset PEP\subset E locally in the Hausdorff distant sense, where PE𝔹1P\supset E\setminus\mathbb{B}_{1} and |P||\partial P| is a stable minimal hypersurface in (n+1E)𝔹1(\mathbb{R}^{n+1}\setminus\partial E)\cup\mathbb{B}_{1}. Thus by mean convexity of Clos(𝔹1)\text{Clos}(\mathbb{B}_{1}), we have Clos(P𝔹1)𝕊n=\text{Clos}(P\cap\mathbb{B}_{1})\cap\mathbb{S}^{n}=\mathcal{E}. Also, on Reg(P𝔹1)\mathrm{Reg}(\partial P\cap\mathbb{B}_{1}), ϕj\phi_{j} subconverges to ϕ=XνP0\phi=X\cdot\nu_{\partial P}\geq 0, which is a Jacobi field on P𝔹1\partial P\cap\mathbb{B}_{1}. By strong maximum principle, either ϕ0\phi\equiv 0 (which implies that PP is a cone, and then E=PE=P with boundary being a stable minimal hypercone, reduces to the previous situation), or ϕ>0\phi>0 everywhere. Repeat the argument above using Corollary 4.3, we know that PP is also a smooth radial graph over 𝕊n\mathcal{E}\subset\mathbb{S}^{n}. ∎

Now we are ready to prove the following stronger version of Theorem 1.1.

Theorem 5.2.

Let En+1E\subset\mathbb{R}^{n+1} be closed and invariant under scaling. Suppose C=[E]C=\partial[E] is a stationary integral cone and is area-minimizing in EE. Then there exists SInt(E)S\subset\text{Int}(E) minimizing in EE and foliates Int(E)\text{Int}(E) by rescaling.

Proof.

By Corollary B.3, there exists a increasing sequence of connected closed C1,1C^{1,1} optimally regular mean convex domain jInt()\mathcal{E}_{j}\subset\text{Int}(\mathcal{E}) approximating \mathcal{E} such that lim supjj(𝕊n)=(𝕊n)\limsup_{j}\|\partial\mathcal{E}_{j}\|(\mathbb{S}^{n})=\|\partial\mathcal{E}\|(\mathbb{S}^{n}). And since jInt(j+1)\mathcal{E}_{j}\subset\text{Int}(\mathcal{E}_{j+1}), by Frankel property of minimal hypersurfaces in 𝕊n\mathbb{S}^{n}, for j>>1j>>1, Ej:=C(j)E_{j}:=C(\mathcal{E}_{j}) do not have stationary boundary, hence are not perimeter minimizing in EjE_{j}. By Lemma 5.1, there exist closed subsets PjC(j)P_{j}\subset C(\mathcal{E}_{j}) such that Pj𝔹1P_{j}\cap\mathbb{B}_{1} are smooth minimizing radial graphs in Int(C(j))𝔹1\text{Int}(C(\mathcal{E}_{j}))\cap\mathbb{B}_{1}. When jj\to\infty, by Lemma 2.1, PjP_{j} 𝐅\mathbf{F}-subconverges to some Caccioppoli set PEP_{\infty}\subset E with P\partial P_{\infty} a stable minimal hypersurface in (n+1E)𝔹1(\mathbb{R}^{n+1}\setminus\partial E)\cup\mathbb{B}_{1} and PP_{\infty} coincide with EE on 𝔹1c\mathbb{B}_{1}^{c}. Moreover, as a minimizing boundary in C(j)C(\mathcal{E}_{j}), [Pj]\partial[P_{j}] satisfies

Pj(𝔹1)[Ej](𝔹1).\|\partial P_{j}\|(\mathbb{B}_{1})\leq\|\partial[E_{j}]\|(\mathbb{B}_{1}).

Hence by taking jj\to\infty we have,

(5.2) [P](𝔹1)[E](𝔹1).\displaystyle\|\partial[P_{\infty}]\|(\mathbb{B}_{1})\leq\|\partial[E]\|(\mathbb{B}_{1}).

Since EE is perimeter minimizing in EE, (5.2) implies that PP_{\infty} also minimizes perimeter in 𝔹2E\mathbb{B}_{2}\cap E. Hence by unique continuation of stable minimal hypersurfaces, P=EP_{\infty}=E. Therefore, when jj\to\infty,

dj:=dist(𝟎,Pj)0.d_{j}:=dist(\mathbf{0},P_{j})\searrow 0.

By Lemma 2.1, when jj\to\infty, P^j:=η1/dj(Pj)\hat{P}_{j}:=\eta_{1/d_{j}}(P_{j}) are perimeter-minimizing in 𝔹1/djEj\mathbb{B}_{1/d_{j}}\cap E_{j} and hence 𝐅\mathbf{F}-subconverges to some P^E\hat{P}_{\infty}\subset E, minimizing perimeter in Int(E)\text{Int}(E) and has P^\partial\hat{P}_{\infty} a stable minimal hypersurface in n+1\mathbb{R}^{n+1}. Since

dsit(𝟎,spt(P^))=limjdist(𝟎,P^j)=1,dsit(\mathbf{0},\text{spt}(\hat{P}_{\infty}))=\lim_{j\to\infty}dist(\mathbf{0},\hat{P}_{j})=1,

we know that P^\partial\hat{P}_{\infty} is not supported on E\partial E. Hence by strong maximum principle [SW89, Ilm96], spt(P^)Int(E)\text{spt}(\hat{P}_{\infty})\subset\text{Int}(E) and P^\hat{P}_{\infty} also minimize perimeter in EE.

Let ν^j\hat{\nu}_{j} be the inward pointed normal field of P^j\hat{P}_{j} on P^j𝔹1/dj\partial\hat{P}_{j}\cap\mathbb{B}_{1/d_{j}}, 1j1\leq j\leq\infty; ψj:=Xν^j\psi_{j}:=X\cdot\hat{\nu}_{j} be the Jacobi field induced by rescaling defined on Reg(P^j)𝔹1/dj\mathrm{Reg}(\partial\hat{P}_{j})\cap\mathbb{B}_{1/d_{j}}. Since P^j\hat{P}_{j} are also radial graphs, when j<j<\infty, ψj>0\psi_{j}>0. When jj\to\infty, ψj\psi_{j} subconverges to ψ\psi_{\infty} on Reg(P^)\mathrm{Reg}(\partial\hat{P}_{\infty}). Hence ψ0\psi_{\infty}\geq 0. Again by strong maximum principle for Jacobi field equation, either ψ0\psi_{\infty}\equiv 0 (which implies P^\hat{P}_{\infty} is a cone and violates that dist(𝟎,P^)=1dist(\mathbf{0},\hat{P}_{\infty})=1), or ψ>0\psi_{\infty}>0 everywhere. Then by Corollary 4.3, Sing(P^)=\mathrm{Sing}(\partial\hat{P}_{\infty})=\emptyset and S=P^Int(E)S=\partial\hat{P}_{\infty}\subset\text{Int}(E) is area-minimizing in EE. Furthermore, ψ>0\psi_{\infty}>0 everywhere implies, by the same argument as Lemma 5.1, that SS is a smooth radial graph over \mathcal{E}, hence foliates Int(E)\text{Int}(E) by rescalings. ∎

Now we turn to the non-minimizing case,

Theorem 5.3.

Let En+1E\subset\mathbb{R}^{n+1} be a closed, viscosity mean convex cone with connected interior. Suppose C=[E]C=\partial[E] is NOT minimizing in EE (This is the case when EE is not a Caccioppoli set).

Let 𝒟:=t0Dt×{t}n+1×\mathcal{D}:=\bigcup_{t\geq 0}D_{t}\times\{t\}\subset\mathbb{R}^{n+1}\times\mathbb{R} be the level set flow of D0:=n+1Int(E)D_{0}:=\mathbb{R}^{n+1}\setminus\text{Int}(E). Then Dt=ηt(D1)D_{t}=\eta_{\sqrt{t}}(D_{1}), t>0\forall t>0; D1D_{1} is a smooth domain and D1\partial D_{1} is a self-expander supported in Int(E)\text{Int}(E) and minimizing 𝐄\mathbf{E}-functional in Int(E)\text{Int}(E).

Moreover, D1\partial D_{1} is a radial graph over \mathcal{E}, i.e. there exists vC(;(0,+))v\in C^{\infty}(\mathcal{E};(0,+\infty)) which diverges near \partial\mathcal{E} such that

D1c={rω:ω,r>v(ω)}.D_{1}^{c}=\{r\omega:\omega\in\mathcal{E},r>v(\omega)\}.

In particular, the mean curvature of D1\partial D_{1} with respect to outward pointed normal field is positive along D1\partial D_{1}.

Proof.

That Dt=ηt(D1)D_{t}=\eta_{\sqrt{t}}(D_{1}) follows from scaling invariance of D0D_{0} and the uniqueness of level set flow, see Proposition 2.9;

Now let jInt()\mathcal{E}_{j}\subset\text{Int}(\mathcal{E}) be the C1,1C^{1,1}-optimally regular closed connected mean convex approximation given by Corollary B.3; Thus Ej:=C(j)E_{j}:=C(\mathcal{E}_{j}) satisfies the assumption in Lemma 2.6. Let D0j:=n+1Int(Ej)D_{0}^{j}:=\mathbb{R}^{n+1}\setminus\text{Int}(E_{j}). We know that Int(D0j)D0{𝟎}\text{Int}(D_{0}^{j})\supset D_{0}\setminus\{\mathbf{0}\} and D0jD0\partial D_{0}^{j}\to\partial D_{0} locally in Hausdorff distant sense as jj\to\infty.

Fix j1j\geq 1, by Lemma 2.6, there exists a closed Caccioppoli set Fj(D0j)c=Int(Ej)F^{j}\subset(D^{j}_{0})^{c}=\text{Int}(E_{j}) minimizing 𝐄\mathbf{E}-functional in EjE_{j} and distant asymptotic to EjE_{j} near infinity. In particular, Fj\partial F^{j} is a self-expander in n+1\mathbb{R}^{n+1} with optimal regularity. We now Claim that

(5.3) ηλ(Fj)Int(D1)=;ηλ(Int(Fj))D1=,λ1.\displaystyle\eta_{\lambda}(F^{j})\cap\text{Int}(D_{1})=\emptyset;\ \ \eta_{\lambda}(\text{Int}(F^{j}))\cap D_{1}=\emptyset,\ \ \ \forall\lambda\geq 1.

To see this, recall that by avoidance principle of level set flow, we have, locally in the Hausdorff distant sense,

limλ0Dλ=limλ0ηλ(D1)D0Int(D0j){𝟎}.\lim_{\lambda\searrow 0}D_{\lambda}=\lim_{\lambda\searrow 0}\eta_{\lambda}(D_{1})\subset D_{0}\subset\text{Int}(D_{0}^{j})\cup\{\mathbf{0}\}.

Hence there exists ϵj>0\epsilon_{j}>0 and λj(0,1)\lambda_{j}\in(0,1) such that for every 0<λλj0<\lambda\leq\lambda_{j} and every μ(0,1]\mu\in(0,1],

(5.4) dist(Dλμ𝒜,ηλ(Fj)𝒜)12dist(D0𝒜,(D0j)c𝒜)ϵj,\displaystyle dist(D_{\lambda\mu}\cap\mathcal{A},\eta_{\sqrt{\lambda}}(F^{j})\cap\mathcal{A})\geq\frac{1}{2}dist(D_{0}\cap\mathcal{A},(D_{0}^{j})^{c}\cap\mathcal{A})\geq\epsilon_{j},

where 𝒜:=𝔸(𝟎;1/2,1)\mathcal{A}:=\mathbb{A}(\mathbf{0};1/2,1). In particular, (5.4) implies that for every μ(0,1]\mu\in(0,1],

FjDμ𝔹1/λj=.F^{j}\cap D_{\mu}\setminus\mathbb{B}_{1/\sqrt{\lambda_{j}}}=\emptyset.

Hence (5.3) follows from Theorem 2.8 and Lemma 2.10 by comparing 𝒟\mathcal{D} with the weak set flow

tηt+ϵ(spt([Fj])Clos(𝔹2/λj)),t\mapsto\eta_{\sqrt{t+\epsilon}}(\text{spt}(\partial[F^{j}])\cap\text{Clos}(\mathbb{B}_{2/\sqrt{\lambda_{j}}})),

for every ϵ>0\epsilon>0 and then send ϵ0\epsilon\searrow 0.

Now let jj\to\infty in (5.3), by 𝐄\mathbf{E}-minimizing property of FjF^{j} and Corollary 2.5 and Lemma 2.1, we have FjFF^{j}\to F^{\infty} in F-metric, where F=spt[F]EF^{\infty}=\text{spt}[F^{\infty}]\subset E is a closed Caccioppoli set distant asymptotic to EE near infinity, and has S:=FS^{\infty}:=\partial F^{\infty} a 𝐄\mathbf{E}-stable hypersurface with optimal regularity, minimizing 𝐄\mathbf{E}-functional in Int(E)\text{Int}(E). By (5.3), we have

(5.5) ηλ(F)Int(D1)=,ηλ(Int(F))D1=,λ1.\displaystyle\eta_{\lambda}(F^{\infty})\cap\text{Int}(D_{1})=\emptyset,\ \eta_{\lambda}(\text{Int}(F^{\infty}))\cap D_{1}=\emptyset,\ \ \ \forall\lambda\geq 1.

Also, we assert that FF^{\infty} is not a cone. In fact otherwise, E=FE=F^{\infty} whose boundary is a stable minimal hypercone. In particular, 𝟎F\mathbf{0}\in F^{\infty}. Hence when jj\to\infty, dj:=distn+1(𝟎,Fj)0d_{j}:=dist_{\mathbb{R}^{n+1}}(\mathbf{0},F^{j})\to 0, and by Lemma 2.1, η1/j(Fj)\eta_{1/j}(F^{j}) 𝐅\mathbf{F}-subconverges to some Caccioppoli set F^=spt[F^]E\hat{F}^{\infty}=\text{spt}[\hat{F}^{\infty}]\subset E with distant 11 to the origin and an optimally regular stable minimal boundary. By maximum principle [SW89, Ilm96], F^Int(E)\hat{F}^{\infty}\subset\text{Int}(E). But since EE is not minimizing, by comparing rescalings of F^\hat{F}^{\infty} with the Caccioppoli set constructed in Lemma 5.1 and using maximum principle [SW89, Ilm96], we get a contradiction.

On the other hand, by Lemma 2.10, tn+1ηt(Int(F))t\mapsto\mathbb{R}^{n+1}\setminus\eta_{\sqrt{t}}(\text{Int}(F^{\infty})) is a weak-set flow from D0D_{0}. Hence by Proposition 2.9, we have D1n+1Int(F)D_{1}\supset\mathbb{R}^{n+1}\setminus\text{Int}(F^{\infty}). Combined with (5.5) when λ=1\lambda=1, we have D1=F\partial D_{1}=\partial F^{\infty} and Int(D1)=n+1F\text{Int}(D_{1})=\mathbb{R}^{n+1}\setminus F^{\infty}.

Finally by (5.5), we have ηλ(F)F\eta_{\lambda}(F^{\infty})\subset F^{\infty} for λ1\lambda\geq 1. Hence ϕ:=2HS=XνS0\phi:=2H^{S^{\infty}}=X\cdot\nu_{S^{\infty}}\geq 0 on regular part of SS^{\infty} and is not identically 0 since SS^{\infty} is not a cone. Recall that by self-expanding equation [Din20, (3.9)], we have (S+1)ϕ=0(\mathcal{L}_{S^{\infty}}+1)\phi=0, where S\mathcal{L}_{S^{\infty}} is the Jacobi operator for 𝐄\mathbf{E}-functional defined in (2.4). By (2.5), this is also a Jacobi-type equation. Hence by strong maximum principle, ϕ>0\phi>0 and is bounded on regular part of SS^{\infty} in each ball. And therefore by Corollary 4.3, Sing(S)=\mathrm{Sing}(S^{\infty})=\emptyset and spt(S)\text{spt}(S^{\infty}) is a radial graph over Int()\text{Int}(\mathcal{E}). ∎

Proof of Theorem 1.3.

Let En+1E\subset\mathbb{R}^{n+1} be a closed cone bounded by CC. Recall by Remark 2.3 (1), EE is viscosity mean convex. If EE is not perimeter minimizing in EE, then Theorem 5.3 provides entirely smooth, strictly mean convex self-expander SInt(E)S\subset\text{Int}(E) and distant asymptotic to EE near infinity. Thus the blow-down of SS restricted to 𝔹1\mathbb{B}_{1} will be strictly mean convex hypersurface approximating C𝔹1C\cap\mathbb{B}_{1} in the Hausdorff distant sense.

If EE minimizes perimeter in EE, Theorem 5.2 provides a smooth minimizing hypersurface SInt(E)S\subset\text{Int}(E) with CC to be the tangent cone at infinity. Blow-down ηλ(S)\eta_{\lambda}(S) (λ0\lambda\searrow 0) also converges to CC locally in the Hausdorff distant sense, and restricting to 𝔹1\mathbb{B}_{1}, ηλ(S)\eta_{\lambda}(S) are strictly stable. Let ϕ1>0\phi_{1}>0 be the first eigenfunction of Jacobi operator Lηλ(S)L_{\eta_{\lambda}(S)} on the domain ηλ(S)𝔹1\eta_{\lambda}(S)\cap\mathbb{B}_{1}, then for sufficiently small ε\varepsilon,

{xεϕ1(x)νηλ(S):xηλ(S)𝔹1},\{x-\varepsilon\phi_{1}(x)\cdot\nu_{\eta_{\lambda}(S)}:x\in\eta_{\lambda}(S)\cap\mathbb{B}_{1}\},

is a properly embedded smooth hypersurface in 𝔹1\mathbb{B}_{1} with positive mean curvature, and still approximate C𝔹1C\cap\mathbb{B}_{1} in Hausdorff distant sense. ∎

Appendix A Growth Lemma for Positive Jacobi Fields

The goal of this section is to prove Lemma 4.2. We begin with the following uniform control of first eigenfunction for Jacobi operator of cross section. For every C𝒞n=1(Λ)C\in\mathscr{C}_{n}^{=1}(\Lambda), let Γ=Γ(C):=C𝕊n\Gamma=\Gamma(C):=C\cap\mathbb{S}^{n} be the cross section of CC, and for each domain ΩΓ\Omega\subset\Gamma, let

μ1(Ω):=inf{Γ|Γϕ|2|AΓ|2ϕ2:ϕCc1(Ω),ϕL2(Γ)=1}.\mu_{1}(\Omega):=\inf\{\int_{\Gamma}|\nabla_{\Gamma}\phi|^{2}-|A_{\Gamma}|^{2}\phi^{2}:\phi\in C_{c}^{1}(\Omega),\|\phi\|_{L^{2}(\Gamma)}=1\}.

Denote for simplicity μ1(C):=μ1(Γ(C))\mu_{1}(C):=\mu_{1}(\Gamma(C)). Recall by [Law69, Per02, Zhu18], μ1(C)(n1)\mu_{1}(C)\leq-(n-1), and equality holds if and only if CC is a quadratic cone.

For each xΓ(C)x\in\Gamma(C), following [CN13], let

rΓ(x):=sup{r(0,1):|AΓ|r1 on 𝔹r(x)Γ}r_{\Gamma}(x):=\sup\{r\in(0,1):|A_{\Gamma}|\leq r^{-1}\text{ on }\mathbb{B}_{r}(x)\cap\Gamma\}

be the regularity scale. Clearly, rΓr_{\Gamma} is continuous in xx and in smooth convergence of Γ\Gamma. For each δ>0\delta>0, denote Ωδ(C):={xΓ:rΓ(x)>δ}\Omega_{\delta}(C):=\{x\in\Gamma:r_{\Gamma}(x)>\delta\}.

Lemma A.1.

There exists ρ0(n,Λ)>0\rho_{0}(n,\Lambda)>0 such that μ1(Ωρ0(C))(n1)\mu_{1}(\Omega_{\rho_{0}}(C))\leq-(n-1) for every C𝒞n=1(Λ)C\in\mathscr{C}_{n}^{=1}(\Lambda).

Proof.

Suppose otherwise, there are Cj𝒞n=1(Λ)C_{j}\in\mathscr{C}_{n}^{=1}(\Lambda) such that μ1(Ω1/j(Cj))>(n1)\mu_{1}(\Omega_{1/j}(C_{j}))>-(n-1).

Suppose CjC𝒞n=1(Λ)C_{j}\to C_{\infty}\in\mathscr{C}_{n}^{=1}(\Lambda) and write Γj:=Cj𝕊n\Gamma_{j}:=C_{j}\cap\mathbb{S}^{n}, 1j1\leq j\leq\infty. Then we must have μ1(C)=(n1)\mu_{1}(C_{\infty})=-(n-1). Since otherwise by [Zhu18], μ1(C)<(n1)\mu_{1}(C_{\infty})<-(n-1) and thus there exists ψCc1(Γ(C))\psi\in C_{c}^{1}(\Gamma(C_{\infty})) such that

Γ|Γψ|2|AΓ|2ψ2<(n1)Γψ2.\int_{\Gamma_{\infty}}|\nabla_{\Gamma_{\infty}}\psi|^{2}-|A_{\Gamma_{\infty}}|^{2}\psi^{2}<-(n-1)\int_{\Gamma_{\infty}}\psi^{2}.

Since the convergence of Γj\Gamma_{j} is smooth near spt(ψ)\text{spt}(\psi) and rΓjr_{\Gamma_{j}} is continuous in jj and xx near spt(ψ)\text{spt}(\psi), this is a contradiction to μ1(Ω1/j(Cj))>(n1)\mu_{1}(\Omega_{1/j}(C_{j}))>-(n-1).

Now that μ1(C)=(n1)\mu_{1}(C_{\infty})=-(n-1). By [Zhu18], CC_{\infty} is a quadratic cone, which has smooth cross section; Then by Allard Regularity, for j>>1j>>1, μ1(Ω1/j(Cj))=μ1(Cj)(n1)\mu_{1}(\Omega_{1/j}(C_{j}))=\mu_{1}(C_{j})\leq-(n-1), that’s a contradiction. ∎

Lemma A.2.

For every Λ>1\Lambda>1, there exists C(n,Λ)>1C(n,\Lambda)>1 with the following property. Let δ2(n,Λ)>0\delta_{2}(n,\Lambda)>0 be in Lemma 3.2 and η(n,Λ)=η(δ2/2,n,Λ)(0,1)\eta(n,\Lambda)=\eta(\delta_{2}/2,n,\Lambda)\in(0,1) be in Corollary 3.5; Also let γn\gamma_{n} be in Proposition 4.1.

If C𝒞n=1(Λ)C\in\mathscr{C}_{n}^{=1}(\Lambda) is a stable minimal hypercone; 0<uCloc(C𝔹4η1)0<u\in C_{loc}^{\infty}(C\cap\mathbb{B}_{4\eta^{-1}}) is a Jacobi field, i.e. ΔCu+|AC|2u=0\Delta_{C}u+|A_{C}|^{2}u=0. Then,

uL1(𝔹r;C)C(n,Λ)1rγnuL1(𝔹1;C), for every r(0,1).\|u\|_{L^{1}_{*}(\mathbb{B}_{r};\|C\|)}\geq C(n,\Lambda)^{-1}r^{\gamma_{n}}\cdot\|u\|_{L^{1}_{*}(\mathbb{B}_{1};\|C\|)},\ \ \ \text{ for every }r\in(0,1).
Proof.

Let Γ:=C𝕊n\Gamma:=C\cap\mathbb{S}^{n} be the cross section of CC in Lemma A.2; ρ0\rho_{0} be in Lemma A.1. Let Ω\Omega be a smooth domain in Γ\Gamma such that Ωρ0(C)ΩΩρ0/2(C)\Omega_{\rho_{0}}(C)\subset\Omega\subset\Omega_{\rho_{0}/2}(C), and 0<φΩ0<\varphi_{\Omega} be a first Dirichlet eigenfunction of ΔΓ|AΓ|2-\Delta_{\Gamma}-|A_{\Gamma}|^{2}, i.e.

(A.1) {(ΔΓ+|AΓ|2)φΩ=μ1(Ω)φΩ, on Ω;φΩ=0, on Ω if Ω.\displaystyle\begin{cases}-(\Delta_{\Gamma}+|A_{\Gamma}|^{2})\varphi_{\Omega}=\mu_{1}(\Omega)\varphi_{\Omega},&\ \text{ on }\Omega;\\ \varphi_{\Omega}=0,&\ \text{ on }\partial\Omega\text{ if }\partial\Omega\neq\emptyset.\end{cases}

with eigenvalue μ1(Ω)μ1(Ωρ0(C))(n1)\mu_{1}(\Omega)\leq\mu_{1}(\Omega_{\rho_{0}}(C))\leq-(n-1) by Lemma A.1. By definition of Ω\Omega, |AΓ|2ρ01|A_{\Gamma}|\leq 2\rho_{0}^{-1} on 𝔹ρ0/2(Ω)\mathbb{B}_{\rho_{0}/2}(\Omega). Hence by covering Ω\Omega with ball of radius ρ0\rho_{0} and applying classical elliptic estimates [GT01] we see

(A.2) supΩ|φΩ|C(n,Λ)φΩL1\displaystyle\sup_{\Omega}|\varphi_{\Omega}|\leq C(n,\Lambda)\|\varphi_{\Omega}\|_{L^{1}}

Now let U(r):=ΩφΩ(ω)u(rω)𝑑ωU(r):=\int_{\Omega}\varphi_{\Omega}(\omega)u(r\omega)\ d\omega. Since

0=(ΔC+|AC|2)u=r2u+n1rru+1r2(ΔΓ+|AΓ|2)u.0=(\Delta_{C}+|A_{C}|^{2})u=\partial_{r}^{2}u+\frac{n-1}{r}\partial_{r}u+\frac{1}{r^{2}}(\Delta_{\Gamma}+|A_{\Gamma}|^{2})u.

We then have for r(0,1]r\in(0,1],

(A.3) U′′(r)+n1rU(r)=1r2ΩφΩ(ω)(ΔΓ+|AΓ|2)u(rω)=1r2Ω(ΔΓ+|AΓ|2)φΩ(ω)u(rω)+1r2ΩζφΩuμ1(Ω)r2ΩφΩ(ω)u(rω)n1r2U(r),\displaystyle\begin{split}U^{\prime\prime}(r)+\frac{n-1}{r}U^{\prime}(r)&=-\frac{1}{r^{2}}\int_{\Omega}\varphi_{\Omega}(\omega)(\Delta_{\Gamma}+|A_{\Gamma}|^{2})u(r\omega)\\ &=-\frac{1}{r^{2}}\int_{\Omega}(\Delta_{\Gamma}+|A_{\Gamma}|^{2})\varphi_{\Omega}(\omega)\cdot u(r\omega)+\frac{1}{r^{2}}\int_{\partial\Omega}\partial_{\zeta}\varphi_{\Omega}\cdot u\\ &\leq\frac{\mu_{1}(\Omega)}{r^{2}}\int_{\Omega}\varphi_{\Omega}(\omega)u(r\omega)\\ &\leq-\frac{n-1}{r^{2}}U(r),\end{split}

where the first inequality follows from (A.1) as well as positivity of uu and the inward normal derivative ζφ-\partial_{\zeta}\varphi once Ω\partial\Omega\neq\emptyset; the second inequality follows from μ1(Ω)(n1)\mu_{1}(\Omega)\leq-(n-1) by Lemma A.1. Let

γn=n22+(n22)2(n1),\displaystyle\gamma_{n}=-\frac{n-2}{2}+\sqrt{(\frac{n-2}{2})^{2}-(n-1)}, βn:=2(n22)2(n1).\displaystyle\beta_{n}:=2\sqrt{(\frac{n-2}{2})^{2}-(n-1)}.

And set W(s):=U(s1/βn)sγn/βnW(s):=U(s^{-1/\beta_{n}})\cdot s^{\gamma_{n}/\beta_{n}}, s1s\geq 1, then (A.3) is equivalent to W′′(s)0W^{\prime\prime}(s)\leq 0 on s1s\geq 1. Since u>0u>0 on CC and then W>0W>0 on [1,+)[1,+\infty), we must have W0W^{\prime}\geq 0 on [1,+)[1,+\infty). Thus,

(A.4) U(r)U(1)rγn,r(0,1).\displaystyle U(r)\geq U(1)\cdot r^{\gamma_{n}},\ \ \ \forall r\in(0,1).

We then have for every r(0,1)r\in(0,1),

uL1(𝔹r;C)supΩφΩ\displaystyle\|u\|_{L^{1}(\mathbb{B}_{r};\|C\|)}\sup_{\Omega}\varphi_{\Omega} 𝔹rφΩ(x|x|)u(x)dC(x)\displaystyle\geq\int_{\mathbb{B}_{r}}\varphi_{\Omega}(\frac{x}{|x|})u(x)\ d\|C\|(x)
=0rsn1U(s)𝑑s0rsn1+γnU(1)𝑑s\displaystyle=\int_{0}^{r}s^{n-1}U(s)\ ds\geq\int_{0}^{r}s^{n-1+\gamma_{n}}U(1)\ ds
=cnrn+γnΩφΩ(ω)u(ω)𝑑ω\displaystyle=c_{n}r^{n+\gamma_{n}}\cdot\int_{\Omega}\varphi_{\Omega}(\omega)u(\omega)\ d\omega
cnrn+γnφΩL1inf𝔹1u\displaystyle\geq c_{n}r^{n+\gamma_{n}}\cdot\|\varphi_{\Omega}\|_{L^{1}}\inf_{\mathbb{B}_{1}}u
C(n,Λ)1rn+γnsupΩφΩuL1(𝔹1;C),\displaystyle\geq C(n,\Lambda)^{-1}r^{n+\gamma_{n}}\cdot\sup_{\Omega}\varphi_{\Omega}\cdot\|u\|_{L^{1}(\mathbb{B}_{1};\|C\|)},

where the second inequality follows from (A.4), the last inequality follows from (A.2) and Corollary 3.5. This finish the proof. ∎

Lemma A.3.

Let Σ(𝔹4,g)\Sigma\subset(\mathbb{B}_{4},g) be a two-sided minimal hypersurface under metric gg such that, the Hessian of distant function square satisfies g2(distg(𝟎,)2)1\nabla_{g}^{2}(dist_{g}(\mathbf{0},\cdot)^{2})\geq 1 on 𝔹2\mathbb{B}_{2}. (Note that this is automatically true when gg is C2C^{2} close to gEucg_{Euc}.)

Let uWloc1,2(Σ)u\in W^{1,2}_{loc}(\Sigma) be a positive super-solution of ΔΣu+(|AΣ|2δ)u=0\Delta_{\Sigma}u+(|A_{\Sigma}|^{2}-\delta)u=0 in distribution sense, i.e.

(A.5) ΣuΣξ+(δ|AΣ|2)uξdΣ0,ξCc1(Σ,0),\displaystyle\int\nabla_{\Sigma}u\cdot\nabla_{\Sigma}\xi+(\delta-|A_{\Sigma}|^{2})u\xi\ d\|\Sigma\|\geq 0,\ \ \ \forall\xi\in C^{1}_{c}(\Sigma,\mathbb{R}_{\geq 0}),

where δn/2\delta\leq n/2. Then there exists vCloc2(Σ𝔹1)v\in C^{2}_{loc}(\Sigma\cap\mathbb{B}_{1}) solving ΔΣv+(|AΣ|2δ)v=0\Delta_{\Sigma}v+(|A_{\Sigma}|^{2}-\delta)v=0 on Σ𝔹1\Sigma\cap\mathbb{B}_{1} such that

12inf𝔹1uvu\frac{1}{2}\inf_{\mathbb{B}_{1}}u\leq v\leq u
Proof.

Claim. With a positive uWloc1,2(Σ)u\in W_{loc}^{1,2}(\Sigma) satisfying (A.5), we have

|ϕ|2+(δ|AΣ|2)ϕ2dΣ0,ϕCc1(Σ),\int|\nabla\phi|^{2}+(\delta-|A_{\Sigma}|^{2})\phi^{2}\ d\|\Sigma\|\geq 0,\ \ \ \forall\phi\in C_{c}^{1}(\Sigma),

with equality if and only if ϕ0\phi\equiv 0.
Proof of the Claim. For every ϕCc1(Σ)\phi\in C_{c}^{1}(\Sigma), let ψ:=u1ϕW01,2(Σ)\psi:=u^{-1}\phi\in W_{0}^{1,2}(\Sigma). Take ξ=uψ2=ϕψ\xi=u\psi^{2}=\phi\psi in (A.5), we have

0u(ϕψ)+(δ|AΣ|2)ϕ2dΣ=|ϕ|2+(δ|AΣ|2)ϕ2u2|ψ|2dΣ.\displaystyle 0\leq\int\nabla u\cdot\nabla(\phi\psi)+(\delta-|A_{\Sigma}|^{2})\phi^{2}\ d\|\Sigma\|=\int|\nabla\phi|^{2}+(\delta-|A_{\Sigma}|^{2})\phi^{2}-u^{2}|\nabla\psi|^{2}\ d\|\Sigma\|.

This finishes the proof of Claim.

Now we turn to the proof of Lemma A.3. By a renormalization, suppose WLOG that inf𝔹1u=1\inf_{\mathbb{B}_{1}}u=1. Consider for every smooth domain ΩΣ𝔹1\Omega\subset\subset\Sigma\cap\mathbb{B}_{1}, minimize

JΩ[ϕ]:=|ϕ|2+(|AΣ|2+δ)ϕ2dΣ,J_{\Omega}[\phi]:=\int|\nabla\phi|^{2}+(-|A_{\Sigma}|^{2}+\delta)\phi^{2}\ d\|\Sigma\|,

among {ϕ:ϕ1W01,2(Ω)}\{\phi:\phi-1\in W_{0}^{1,2}(\Omega)\} to get a minimizer vΩC2(Clos(Ω))v_{\Omega}\in C^{2}(\text{Clos}(\Omega)) solving ΔΣvΩ+(|AΣ|2δ)vΩ=0\Delta_{\Sigma}v_{\Omega}+(|A_{\Sigma}|^{2}-\delta)v_{\Omega}=0 in Ω\Omega. Since u1u\geq 1 on Σ𝔹1\Sigma\cap\mathbb{B}_{1}, we know that (vΩu)+W01,2(Σ)(v_{\Omega}-u)^{+}\in W_{0}^{1,2}(\Sigma) is supported in Clos(Ω)\text{Clos}(\Omega), thus by taking ξ=(vΩu)+\xi=(v_{\Omega}-u)^{+} in (A.5) we have,

0\displaystyle 0 u(vΩu)+(δ|AΣ|2)u(vΩu)+\displaystyle\geq\int-\nabla u\cdot\nabla(v_{\Omega}-u)^{+}-(\delta-|A_{\Sigma}|^{2})u(v_{\Omega}-u)^{+}
=(vΩu)(vΩu)++(δ|AΣ|2)(vΩu)(vΩu)+\displaystyle=\int\nabla(v_{\Omega}-u)\cdot\nabla(v_{\Omega}-u)^{+}+(\delta-|A_{\Sigma}|^{2})(v_{\Omega}-u)\cdot(v_{\Omega}-u)^{+}
=|ξ|2+(δ|AΣ|2)ξ20,\displaystyle=\int|\nabla\xi|^{2}+(\delta-|A_{\Sigma}|^{2})\xi^{2}\geq 0,

where the first equality follows by integration by parts and the equation satisfied by vΩv_{\Omega}; the second equality follows from the Claim. Hence the Claim guarantees that (vΩu)+=0(v_{\Omega}-u)^{+}=0, in other words vΩuv_{\Omega}\leq u in Ω\Omega. On the other hands, let w:=(1+distg(,𝟎))/21w:=(1+dist_{g}(\cdot,\mathbf{0}))/2\leq 1 on Σ𝔹1\Sigma\cap\mathbb{B}_{1}, a simple computation shows that

ΔΣw+(|AΣ|2δ)w0 on Σ𝔹1.\Delta_{\Sigma}w+(|A_{\Sigma}|^{2}-\delta)w\geq 0\ \text{ on }\Sigma\cap\mathbb{B}_{1}.

This provides a sub-solution to the equation. Repeat the same process above gives vΩw1/2v_{\Omega}\geq w\geq 1/2 on Ω\Omega.

Now take Ωj\Omega_{j} be an increasing family of smooth domain approximating Σ𝔹1\Sigma\cap\mathbb{B}_{1}, 1/2vΩju1/2\leq v_{\Omega_{j}}\leq u be the solution constructed above relative to Ωj\Omega_{j}. By classical Harnack inequality for Elliptic equations [GT01], vΩjv_{\Omega_{j}} subconverges to some solution vCloc2(Σ𝔹1)v\in C_{loc}^{2}(\Sigma\cap\mathbb{B}_{1}) of ΔΣv+(|AΣ|2δ)v=0\Delta_{\Sigma}v+(|A_{\Sigma}|^{2}-\delta)v=0 which satisfies 1/2vu1/2\leq v\leq u, thus finishes proof of the Lemma. ∎

Proof of Lemma 4.2.

Take r0=r0(n,Λ,γ)(0,η(δ2,Λ,n)/8)r_{0}=r_{0}(n,\Lambda,\gamma)\in(0,\eta(\delta_{2},\Lambda,n)/8) TBD, where δ2=δ2(n,Λ)\delta_{2}=\delta_{2}(n,\Lambda) be given by Lemma 3.2 and η\eta is given by the Harnack inequality Corollary 3.5.

Suppose for contradiction that there exists γ>γn\gamma>\gamma_{n} and stable minimal hypersurfaces Σj(𝔹4,gj)\Sigma_{j}\subset(\mathbb{B}_{4},g_{j}) such that 𝟎Sing(Σj)\mathbf{0}\in\mathrm{Sing}(\Sigma_{j}), gjgEucg_{j}\to g_{Euc} in C4C^{4} and 𝐅𝔹4r01(|Σj|,|C|)0\mathbf{F}_{\mathbb{B}_{4}r_{0}^{-1}}(|\Sigma_{j}|,|C|)\to 0 for some stable minimal hypercone Cn+1C\subset\mathbb{R}^{n+1}; But there are ujWloc1,2(Σj)u_{j}\in W^{1,2}_{loc}(\Sigma_{j}) weak supersolution to ΔΣjuj+(|AΣj|21/j)uj=0\Delta_{\Sigma_{j}}u_{j}+(|A_{\Sigma_{j}}|^{2}-1/j)u_{j}=0 and rj[r0/2,r0]r_{j}\in[r_{0}/2,r_{0}], rjr[r0/2,r0]r_{j}\to r_{\infty}\in[r_{0}/2,r_{0}] satisfying

(A.6) ujL1(𝔹rj;Σ)<ujL1(𝔹1;Σ)rjγ\displaystyle\|u_{j}\|_{L^{1}_{*}(\mathbb{B}_{r_{j}};\|\Sigma\|)}<\|u_{j}\|_{L^{1}_{*}(\mathbb{B}_{1};\|\Sigma\|)}\cdot r_{j}^{\gamma}

Note that by Lemma 3.2 and 𝐅𝔹4r01(|Σ|,|C|)0\mathbf{F}_{\mathbb{B}_{4r_{0}^{-1}}}(|\Sigma|,|C|)\to 0, we have for j>>1j>>1, 𝐅𝔹4r01(|Σj|,2)δ2(n,2Λ)\mathbf{F}_{\mathbb{B}_{4r_{0}^{-1}}}(|\Sigma_{j}|,\mathscr{M}^{\geq 2})\geq\delta_{2}(n,2\Lambda). Hence the Harnack inequality Corollary 3.5 applies for functions on Σj\Sigma_{j}.

Also by Lemma A.3, we can take vjCloc2(Σj𝔹1)v_{j}\in C^{2}_{loc}(\Sigma_{j}\cap\mathbb{B}_{1}) such that on Σj𝔹1\Sigma_{j}\cap\mathbb{B}_{1},

(A.7) 0<12infΣj𝔹1ujvjuj;\displaystyle 0<\frac{1}{2}\inf_{\Sigma_{j}\cap\mathbb{B}_{1}}u_{j}\leq v_{j}\leq u_{j}; ΔΣjvj+(|AΣj|21/j)vj=0.\displaystyle\Delta_{\Sigma_{j}}v_{j}+(|A_{\Sigma_{j}}|^{2}-1/j)v_{j}=0.

Thus by applying Harnack inequality Corollary 3.5 on uju_{j} and (A.6), (A.7), we have

(A.8) vjL1(𝔹rj)C(n,Λ)infΣj𝔹1ujrjγC(n,Λ)infΣj𝔹1vjrjγ.\displaystyle\|v_{j}\|_{L^{1}_{*}(\mathbb{B}_{r_{j}})}\leq C(n,\Lambda)\inf_{\Sigma_{j}\cap\mathbb{B}_{1}}u_{j}\cdot r_{j}^{\gamma}\leq C(n,\Lambda)\inf_{\Sigma_{j}\cap\mathbb{B}_{1}}v_{j}\cdot r_{j}^{\gamma}.

On the other hand, fix a regular point xC𝔹η/4x_{\infty}\in C\cap\mathbb{B}_{\eta/4} and suppose Σjxjx\Sigma_{j}\ni x_{j}\to x_{\infty}. Renormalize vjv_{j} such that vj(xj)=1v_{j}(x_{j})=1. Then by classical Harnack inequality [GT01], after passing to a subsequence, vjvv_{j}\to v_{\infty} in Cloc2C^{2}_{loc} away from Sing(C)\mathrm{Sing}(C) for some positive Jacobi field vCloc(C𝔹1)v_{\infty}\in C^{\infty}_{loc}(C\cap\mathbb{B}_{1}), i.e. (ΔC+|AC|2)v=0(\Delta_{C}+|A_{C}|^{2})v_{\infty}=0. Again by Harnack inequality Corollary 3.5, for every p(1,n/(n2))p\in(1,n/(n-2)) we have,

vjLp(𝔹η/4;Σj)C(p,n,Λ)vj(xj)=C(p,n,Λ).\displaystyle\|v_{j}\|_{L^{p}}(\mathbb{B}_{\eta/4};\|\Sigma_{j}\|)\leq C(p,n,\Lambda)v_{j}(x_{j})=C(p,n,\Lambda).

Hence by Hölder inequality, for every small neighborhood 𝒩𝔹η/4\mathcal{N}\subset\mathbb{B}_{\eta/4} of Sing(C)𝔹η/8\mathrm{Sing}(C)\cap\mathbb{B}_{\eta/8}, we have

vjL1(𝒩);Σj)vjLp(𝒩;Σj)Σj(𝒩)(p1)/pC(p,n,Λ)Σj(𝒩)(p1)/p.\|v_{j}\|_{L^{1}(\mathcal{N});\|\Sigma_{j}\|)}\leq\|v_{j}\|_{L^{p}(\mathcal{N};\|\Sigma_{j}\|)}\cdot\|\Sigma_{j}\|(\mathcal{N})^{(p-1)/p}\leq C(p,n,\Lambda)\cdot\|\Sigma_{j}\|(\mathcal{N})^{(p-1)/p}.

This guarantees that vjv_{j} are not L1L^{1}-concentrating near Sing(Σj)𝔹η/8\mathrm{Sing}(\Sigma_{j})\cap\mathbb{B}_{\eta/8}. Hence vjv_{j} subconverges to vv_{\infty} also in L1(𝔹η/8)L^{1}(\mathbb{B}_{\eta/8}). By (A.8) and Lemma A.2, we derive,

vL1(𝔹r;C)\displaystyle\|v_{\infty}\|_{L^{1}_{*}(\mathbb{B}_{r_{\infty}};\|C\|)} C(n,Λ)infC𝔹η/8vrγ\displaystyle\leq C(n,\Lambda)\inf_{C\cap\mathbb{B}_{\eta/8}}v_{\infty}\cdot r_{\infty}^{\gamma}
C(n,Λ)vL1(𝔹η/8;C)rγ\displaystyle\leq C(n,\Lambda)\|v_{\infty}\|_{L^{1}_{*}(\mathbb{B}_{\eta/8};\|C\|)}\cdot r_{\infty}^{\gamma}
C¯(n,Λ)vL1(𝔹r;C)(η8r)γn,\displaystyle\leq\bar{C}(n,\Lambda)\|v_{\infty}\|_{L^{1}_{*}(\mathbb{B}_{r_{\infty}};\|C\|)}\cdot(\frac{\eta}{8r_{\infty}})^{\gamma_{n}},

which then implies

C¯(n,Λ)(η8)γnrγnγr0γnγ.\bar{C}(n,\Lambda)\cdot(\frac{\eta}{8})^{\gamma_{n}}\geq r_{\infty}^{\gamma_{n}-\gamma}\geq r_{0}^{\gamma_{n}-\gamma}.

This will be a contradiction if we take r0<(C¯(n,Λ)(η/8)γn)1/(γnγ)r_{0}<\big{(}\bar{C}(n,\Lambda)\cdot(\eta/8)^{\gamma_{n}})^{1/(\gamma_{n}-\gamma)}. ∎

Appendix B Minimization with Obstacle

Recall the notion of viscosity mean convex is introduced in Definition 2.2. Our goal is to prove the following.

Theorem B.1.

Let EE be a compact viscosity mean convex subset in a Riemannian manifold (M,g)(M,g) of dimension n+1n+1; Let ZInt(E)Z\subset\text{Int}(E) be a closed C2C^{2} domain (possibly be empty). Then there exists a closed subset PP with the following properties,

  1. (i)

    ZPEZ\subset P\subset E;

  2. (ii)

    The topological boundary P\partial P is a stable minimal hypersurface with optimally regularity in the portion MZM\setminus Z and is C1,1C^{1,1} near each point of PZ\partial P\cap\partial Z;

  3. (iii)

    For every Caccioppoli set QQ such that ZQInt(E)Z\subset Q\subset\subset\text{Int}(E), we have 𝒫(P)𝒫(Q)\mathcal{P}(P)\leq\mathcal{P}(Q);

  4. (iv)

    If (M,g)(𝕊n+1,ground)(M,g)\cong(\mathbb{S}^{n+1},g_{\text{round}}) is isometric to the round sphere, then one can choose so that PInt(E)P\subset\text{Int}(E).

In particular, PP is a C1,1C^{1,1}-optimally regular mean convex subset.

In the Euclidean space, such existence-with-barrier result was established by [ISZ98, Theorem 3.6]. Here we shall use a similar argument. The only difficulty is that in a general Riemannian manifold, there’s usually no isometric translations. So instead of minimizing area, we shall first minimize an 𝒜ϵ\mathcal{A}^{\epsilon} functional to find prescribed mean curvature approximations. Such 𝒜ϵ\mathcal{A}^{\epsilon} functional was introduced and carefully studied by [ZZ20].

Proof.

Suppose WLOG that injrad(M,g)>1injrad(M,g)>1 and every ball of radius<1<1 in (M,g)(M,g) is strictly convex. Let d(x):=distg(x,Ec)d(x):=dist_{g}(x,E^{c}) be the distant function defined in EE; Then there exists j0>>1j_{0}>>1 such that Z{d>1/j0}Z\subset\{d>1/j_{0}\}; Let ηC(M;[0,1])\eta\in C^{\infty}(M;[0,1]) be a cut-off function such that η=0\eta=0 in a neighborhood of ZZ and η=1\eta=1 on {d1/(2j0)}\{d\leq 1/(2j_{0})\}.

Let ϵ(0,1)\epsilon\in(0,1), consider the the following 𝒜ϵ\mathcal{A}^{\epsilon}-functional [ZZ20] for a Caccioppoli set QQ,

𝒜ϵ(Q):=𝒫(Q)+Qϵη(x)𝑑n+1(x).\displaystyle\mathcal{A}^{\epsilon}(Q):=\mathcal{P}(Q)+\int_{Q}\epsilon\cdot\eta(x)\ d\mathscr{H}^{n+1}(x).

By first variation [ZZ20], a stationary Caccioppoli set for 𝒜ϵ\mathcal{A}^{\epsilon} has boundary mean curvature equals to ϵη\epsilon\cdot\eta.
Claim 1. There exists jϵ4j0+1/ϵj_{\epsilon}\geq 4j_{0}+1/\epsilon such that for every j>jϵj>j_{\epsilon}, there exists a closed Caccioppoli set ZQj{d>1/j}Z\subset Q_{j}\subset\{d>1/j\} which minimize 𝒜ϵ\mathcal{A}^{\epsilon} among

(B.1) 𝒯j:={Q:ZQ{d1/j}}.\displaystyle\mathscr{T}_{j}:=\{Q:Z\subset Q\subset\{d\geq 1/j\}\}.

In particular, Qj\partial Q_{j} is optimally regular and stable in MZM\setminus Z and C1,1C^{1,1} near each point on Z\partial Z.

With this Claim 1, choose Qϵ:=Qjϵ+1Q^{\epsilon}:=Q_{j_{\epsilon}+1} and send ϵ0\epsilon\searrow 0, by spirit of Lemma 2.1 and [LIN85, Chapter 2], QϵQ^{\epsilon} 𝐅\mathbf{F}-subconverges to some closed Caccioppoli set P=spt[P]EP=\text{spt}[P]\subset E, with P\partial P stable minimal hypersuraces in MZM\setminus Z and C1,1C^{1,1} near Z\partial Z. This proves (i) and (ii). Also since PP is the 𝐅\mathbf{F}-limit of 𝒜ϵ\mathcal{A}^{\epsilon}-minimizer QϵQ^{\epsilon}, (iii) also holds for PP. When (M,g)(𝕊n+1,ground)(M,g)\cong(\mathbb{S}^{n+1},g_{\text{round}}), since Qϵ\partial Q^{\epsilon} is 𝒜ϵ\mathcal{A}^{\epsilon}-stationary and thus has mean curvature >0>0 on {d<1/(2j0)}\{d<1/(2j_{0})\}, by a rotation of 𝕊n+1\mathbb{S}^{n+1} and viscosity mean convexity of EE, we must have d1/(2j0)d\geq 1/(2j_{0}) along QϵQ^{\epsilon} and then along PP. This proves (iv), and finish the proof of Theorem B.1.

Proof of Claim 1. Let jϵ=j(ϵ,g)4j0+1/ϵj_{\epsilon}=j(\epsilon,g)\geq 4j_{0}+1/\epsilon to be determined later.

For j>jϵj>j_{\epsilon}, let Q¯j𝒯j\bar{Q}_{j}\in\mathscr{T}_{j} be a general 𝒜ϵ\mathcal{A}^{\epsilon}-minimizer among 𝒯j\mathscr{T}_{j}, where 𝒯j\mathscr{T}_{j} is defined in (B.1). For every xspt[Q¯j]{d=1/j}x\in\text{spt}[\bar{Q}_{j}]\cap\{d=1/j\}, consider minimize 𝒜ϵ\mathcal{A}^{\epsilon} among

{RM:RΔQ¯jB1/j(x)},\{R\subset M:R\Delta\bar{Q}_{j}\subset B_{1/j}(x)\},

to find a Caccioppoli set Q~j\tilde{Q}_{j}.
Claim 2. By taking jϵj_{\epsilon} large, for every yspt[Q~j]B1/j(x)y\in\text{spt}[\tilde{Q}_{j}]\cap B_{1/j}(x), we have d(y)>1/jd(y)>1/j.

With this Claim 2, Q~j𝒯j\tilde{Q}_{j}\in\mathscr{T}_{j} is also an 𝒜ϵ\mathcal{A}^{\epsilon}-minimizer among 𝒯j\mathscr{T}_{j}, and whenever Q¯j\bar{Q}_{j} is 𝒜ϵ\mathcal{A}^{\epsilon}-stationary and optimally regular in an open subset W{d>1/j}W\subset\{d>1/j\}, by unique continuation of CMC hypersurfaces [ZZ20], Q~j\tilde{Q}_{j} must be 𝒜ϵ\mathcal{A}^{\epsilon}-stationary and optimally regular in WB1/j(x)W\cup B_{1/j}(x). Hence by covering {d=1/j}\{d=1/j\} with finitely many ball of radius 1/j1/j and repeat the replacement process, we can find QjQ_{j} satisfying the assertion of Claim 1.
Proof of Claim 2. Let jϵ=j(ϵ,g)4j0+1/ϵj_{\epsilon}=j(\epsilon,g)\geq 4j_{0}+1/\epsilon TBD. By definition of Q~j\tilde{Q}_{j} and strict convexity of B1/j(x)B_{1/j}(x), we must have

spt[Q~j]B1/j(x)spt[Q¯j]B1/j(x){d1/j}.\text{spt}[\tilde{Q}_{j}]\setminus B_{1/j}(x)\subset\text{spt}[\bar{Q}_{j}]\setminus B_{1/j}(x)\subset\{d\geq 1/j\}.

If Claim 2 fails, then there exists yspt([Q~j])B1/j(x)y\in\text{spt}(\partial[\tilde{Q}_{j}])\cap B_{1/j}(x) realizing inf{d(z):zspt(Q~j)B1/j(x)}\inf\{d(z):z\in\text{spt}(\tilde{Q}_{j})\cap B_{1/j}(x)\}. Let yEy^{\prime}\in\partial E and 0<d01/j0<d_{0}\leq 1/j be such that distg(y,y)=d(y)=d0dist_{g}(y,y^{\prime})=d(y)=d_{0}. Since Bd0(y)spt(Q~j)=B_{d_{0}}(y^{\prime})\cap\text{spt}(\tilde{Q}_{j})=\emptyset, the tangent cone of Q~j\partial_{*}\tilde{Q}_{j} at yy must be a plane, and hence by 𝒜ϵ\mathcal{A}^{\epsilon}-minimizing of Q~j\tilde{Q}_{j} in B1/j(x)B_{1/j}(x), we know that yy is a regular point of [Q~j]\partial[\tilde{Q}_{j}]. Moreover, by comparing the second fundamental form A~j\tilde{A}_{j} of [Q~j]\partial[\tilde{Q}_{j}] and second fundamental form of Bd0(y)\partial B_{d_{0}}(y^{\prime}), together with the fact that the mean curvature of [Q~j]\partial[\tilde{Q}_{j}] at yy is ϵ\epsilon by first variation of 𝒜ϵ\mathcal{A}^{\epsilon}, we know that,

(B.2) |A~j(y)|C(n,g)d01.\displaystyle|\tilde{A}_{j}(y)|\leq C(n,g)\cdot d_{0}^{-1}.

On the other hand, let LyL_{y} be the image of Bϵ(𝟎)ν~yTyMB_{\epsilon}(\mathbf{0})\cap\tilde{\nu}_{y}^{\perp}\subset T_{y}M under exponential map under metric gg centered at yy, where ν~y\tilde{\nu}_{y} is the unit normal vector of [Q~j]\partial[\tilde{Q}_{j}] at yy pointing towards yy^{\prime}. We shall work under Fermi coordinates

Φ:Ly×(ϵ,ϵ)M,(z,t)expzg(tνL(z)),\Phi:L_{y}\times(-\epsilon,\epsilon)\to M,\ \ \ (z,t)\mapsto\exp^{g}_{z}(t\cdot\nu^{L}(z)),

where νL\nu^{L} be the unit normal field of LyL_{y} such that νL(y)=ν~y\nu^{L}(y)=\tilde{\nu}_{y}.

Consider R:=Φ({(z,t+d0):Φ(z,t)spt(Q~j)B1/j(x)})R:=\Phi(\{(z,t+d_{0}):\Phi(z,t)\in\text{spt}(\tilde{Q}_{j})\cap B_{1/j}(x)\}). Since distg(Φ(z,t),Φ(z,t+d0))d0dist_{g}(\Phi(z,t),\Phi(z,t+d_{0}))\leq d_{0}, we have RER\subset E and y=Φ(y,d0)Ry^{\prime}=\Phi(y,d_{0})\in R is a regular point of R\partial R. Moreover, by (B.2) and the following Lemma B.2, the mean curvature HR(y)H_{\partial R}(y^{\prime}) of R\partial R at yy^{\prime} satisfies,

HR(y)H~j(y)C¯(n,g)d0ϵC¯(n,g)jϵ1,H_{\partial R}(y^{\prime})\geq\tilde{H}_{j}(y)-\bar{C}(n,g)\cdot d_{0}\geq\epsilon-\bar{C}(n,g)\cdot j_{\epsilon}^{-1},

where recall the mean curvature of [Q~j]\partial[\tilde{Q}_{j}] at yy is H~j(y)=ϵ\tilde{H}_{j}(y)=\epsilon. Choosing jϵ>C¯(n,g)ϵ1j_{\epsilon}>\bar{C}(n,g)\cdot\epsilon^{-1}, we have HR(y)>0H_{\partial R}(y^{\prime})>0, violates the viscosity mean convexity of EE and hence is a contradiction. This finish the proof of Claim 2. ∎

Lemma B.2.

Let {gt}t(1,1)\{g^{t}\}_{t\in(-1,1)} be a smooth family of Riemannian metric on L:=𝔹1n(𝟎)L:=\mathbb{B}_{1}^{n}(\mathbf{0}) (not necessarily complete); Let {At:=12ddt(gt)}t(1,1)\{A^{t}:=\frac{1}{2}\frac{d}{dt}(g^{t})\}_{t\in(-1,1)} be the family of symmetric 2-tensor of derivative of gtg^{t}. Assume that A0(𝟎)=0A^{0}(\mathbf{0})=0.

Consider on the cylinder L×(1,1)L\times(-1,1), the metric g:=gtdt2g:=g^{t}\oplus dt^{2}. Assume that the Riemannian curvature tensor of gg satisfies |Rmg|1/2n|Rm_{g}|\leq 1/2n on L×(1,1)L\times(-1,1).

Let uu be a C2C^{2}-function on some small ball 𝔹r(𝟎)L\mathbb{B}_{r}(\mathbf{0})\subset L taking value in (1,1)(-1,1) such that u(𝟎)=:au(\mathbf{0})=:a and du(𝟎)=0du(\mathbf{0})=0. Let Γu:={(z,u(z)):z𝔹r}L×(1,1)\Gamma_{u}:=\{(z,u(z)):z\in\mathbb{B}_{r}\}\subset L\times(-1,1) be a C2C^{2} hypersurface. Let {zi}1in\{z^{i}\}_{1\leq i\leq n} be a normal coordinates of g0g^{0} defined near 𝟎\mathbf{0} on LL. Then the mean curvature of Γu\Gamma_{u} at (𝟎,a)(\mathbf{0},a) with respect to the upper-pointed normal field is given by

HΓu(𝟎,a)=(g02u(𝟎)Aa(𝟎))ijga(𝟎)ij,H_{\Gamma_{u}}(\mathbf{0},a)=(\nabla^{2}_{g_{0}}u(\mathbf{0})-A^{a}(\mathbf{0}))_{ij}\cdot g_{a}(\mathbf{0})^{ij},

In particular, the difference between mean curvature of Γu\Gamma_{u} at (𝟎,a)(\mathbf{0},a) and Γua\Gamma_{u-a} at (𝟎,0)(\mathbf{0},0) has estimate

|HΓu(𝟎,a)HΓua(𝟎,0)|C(n)(a2|AΓua|(𝟎,0)+a).|H_{\Gamma_{u}}(\mathbf{0},a)-H_{\Gamma_{u-a}}(\mathbf{0},0)|\leq C(n)(a^{2}|A_{\Gamma_{u-a}}|(\mathbf{0},0)+a).
Proof.

First note that {z1,,zn,t}\{z^{1},...,z^{n},t\} is a local coordinates of L×(1,1)L\times(-1,1) with the coordinate vector fields denoted by {1,,n,t}\{\partial_{1},...,\partial_{n},\partial_{t}\}. By [Pet06], under this,

2At(i,j)\displaystyle 2A^{t}(\partial_{i},\partial_{j}) =ddtgt(i,j)=tg(g(i,j))=2g(igt,j);\displaystyle=\frac{d}{dt}g^{t}(\partial_{i},\partial_{j})=\nabla^{g}_{\partial_{t}}(g(\partial_{i},\partial_{j}))=2g(\nabla^{g}_{\partial_{i}}\partial_{t},\partial_{j});
ddtAt(i,j)\displaystyle\frac{d}{dt}A^{t}(\partial_{i},\partial_{j}) =12g(tgigt,j)+12g(tgjgt,i)+g(igt,jgt)\displaystyle=\frac{1}{2}g(\nabla^{g}_{\partial_{t}}\nabla^{g}_{\partial_{i}}\partial_{t},\partial_{j})+\frac{1}{2}g(\nabla^{g}_{\partial_{t}}\nabla^{g}_{\partial_{j}}\partial_{t},\partial_{i})+g(\nabla^{g}_{\partial_{i}}\partial_{t},\nabla^{g}_{\partial_{j}}\partial_{t})
=Rmg(t,i,t,j)+At(i,k)At(j,l)gtkl,\displaystyle=Rm_{g}(\partial_{t},\partial_{i},\partial_{t},\partial_{j})+A^{t}(\partial_{i},\partial_{k})A^{t}(\partial_{j},\partial_{l})g_{t}^{kl},

where [gtkl][g_{t}^{kl}] is the inverse of matrix [gijt][g^{t}_{ij}]. In particular, by a bootstrap argument, we have

(B.3) |At(i,j)|t2,\displaystyle|A^{t}(\partial_{i},\partial_{j})|\leq\frac{t}{2}, |gijtδij|t22.\displaystyle|g^{t}_{ij}-\delta_{ij}|\leq\frac{t^{2}}{2}.

Let Φu:z(z,u(z))\Phi_{u}:z\mapsto(z,u(z)) be a parametrization of Γu\Gamma_{u}. Then the upward pointed normal field of Γu\Gamma_{u} at (𝟎,a)(\mathbf{0},a) is

νu(𝟎,a)=iuguijj+t1+iujuguij=t;\nu^{u}(\mathbf{0},a)=\frac{-\partial_{i}ug_{u}^{ij}\partial_{j}+\partial_{t}}{\sqrt{1+\partial_{i}u\partial_{j}ug_{u}^{ij}}}=\partial_{t};

The induced metric of Γu\Gamma_{u} under this parametrization is

g(iΦu,jΦu)=g(i+iut,j+jut);g(\partial_{i}\Phi_{u},\partial_{j}\Phi_{u})=g(\partial_{i}+\partial_{i}u\cdot\partial_{t},\partial_{j}+\partial_{j}u\cdot\partial_{t});

And the second fundamental form of Γu\Gamma_{u} at (𝟎,a)(\mathbf{0},a) is

AΓu(iΦu,jΦu)|(𝟎,a)=i+iutg(j+jut),νu|(𝟎,a)=Aa(𝟎)+g02u(𝟎)(i,j).A_{\Gamma_{u}}(\partial_{i}\Phi_{u},\partial_{j}\Phi_{u})|_{(\mathbf{0},a)}=\langle\nabla^{g}_{\partial_{i}+\partial_{i}u\cdot\partial_{t}}(\partial_{j}+\partial_{j}u\cdot\partial_{t}),\nu^{u}\rangle|_{(\mathbf{0},a)}=-A^{a}(\mathbf{0})+\nabla^{2}_{g^{0}}u(\mathbf{0})(\partial_{i},\partial_{j}).

Hence the mean curvature of Γu\Gamma_{u} at (𝟎,a)(\mathbf{0},a) is

HΓu(𝟎,a)=(g02u(𝟎)Aa(𝟎))ijga(𝟎)ij;H_{\Gamma_{u}}(\mathbf{0},a)=(\nabla_{g^{0}}^{2}u(\mathbf{0})-A^{a}(\mathbf{0}))_{ij}\cdot g_{a}(\mathbf{0})^{ij};

The estimate on the difference of mean curvature follows from the estimate (B.3). ∎

Corollary B.3.

Let (𝕊n+1,ground)\mathcal{E}\subset(\mathbb{S}^{n+1},g_{\text{round}}) be a closed viscosity mean convex subset with connected interior. Then there exists an increasing sequence of closed connected C1,1C^{1,1} optimally regular mean convex subsets {jInt()}j1\{\mathcal{E}_{j}\subset\text{Int}(\mathcal{E})\}_{j\geq 1} such that when jj\to\infty, j\mathcal{E}_{j}\to\mathcal{E}, j\partial\mathcal{E}_{j}\to\partial\mathcal{E} both in Hausdorff distant sense. Furthermore, if the topological boundary \partial\mathcal{E} is nn-rectifiable (In particular \mathcal{E} is a Caccioppoli set), then we can choose so that

limj𝒫(j)=𝒫().\lim_{j\to\infty}\mathcal{P}(\mathcal{E}_{j})=\mathcal{P}(\mathcal{E}).
Proof.

Let d:=dist𝕊n+1(,c)d_{\mathcal{E}}:=dist_{\mathbb{S}^{n+1}}(\cdot,\mathcal{E}^{c}) be the distant function to the boundary defined in \mathcal{E}. d~\tilde{d}_{\mathcal{E}} be a smooth function on Int()\text{Int}(\mathcal{E}) such that dd~dd_{\mathcal{E}}\leq\tilde{d}_{\mathcal{E}}\leq d_{\mathcal{E}}. Let ϵj0\epsilon_{j}\searrow 0 be a sequence of regular value of d~\tilde{d}_{\mathcal{E}}, and Zj:={d~ϵj}Int()Z_{j}:=\{\tilde{d}_{\mathcal{E}}\geq\epsilon_{j}\}\subset\text{Int}(\mathcal{E}) are compact smooth domains.

By Theorem B.1, there exists C1,1C^{1,1}-optimally regular closed mean convex subset ZjPjInt()Z_{j}\subset P_{j}\subset\text{Int}(\mathcal{E}). Clearly by definition, PjP_{j}\to\mathcal{E} and Pj\partial P_{j}\to\partial\mathcal{E} both in Hausdorff distant sense. Also, let j0>>1j_{0}>>1 such that Zj0Z_{j_{0}}\neq\emptyset and fix ω0Zj0\omega_{0}\in Z_{j_{0}}; Take j\mathcal{E}_{j} to be the connect component of PjP_{j} containing ω0\omega_{0}, we know that since Int()\text{Int}(\mathcal{E}) is connected, j\mathcal{E}_{j}\to\mathcal{E} , j\partial\mathcal{E}_{j}\to\partial\mathcal{E} also in Hausdorff distant sense.

If further \partial\mathcal{E} is nn-rectifiable, by [Fed69, Theorem 3.2.39] and co-area formula, there exists Rj:={dδj}Int()R_{j}:=\{d_{\mathcal{E}}\geq\delta_{j}\}\subset\text{Int}(\mathcal{E}) such that limj𝒫(Rj)=𝒫()\lim_{j\to\infty}\mathcal{P}(R_{j})=\mathcal{P}(\mathcal{E}). Then by Theorem B.1 (iii) and upper semi-continuity of perimeter under convergence, we have limj𝒫(j)=𝒫()\lim_{j\to\infty}\mathcal{P}(\mathcal{E}_{j})=\mathcal{P}(\mathcal{E}). ∎

References

  • [AIC95] Sigurd Angenent, Tom Ilmanen, and David L Chopp. A computed example of nonuniqueness of mean curvature flow in r3. Communications in Partial Differential Equations, 20(11-12):1937–1958, 1995.
  • [BG72] Enrico Bombieri and E Giusti. Harnack’s inequality for elliptic differential equations on minimal surfaces. Inventiones mathematicae, 15(1):24–46, 1972.
  • [Bro86] JE Brothers. Some open problems in geometric measure theory and its applications suggested by participants. Geometric Measure Theory and the Calculus of Variations, 44:441, 1986.
  • [BW18] Jacob Bernstein and Lu Wang. An integer degree for asymptotically conical self-expanders. arXiv preprint arXiv:1807.06494, 2018.
  • [BW19a] Jacob Bernstein and Lu Wang. Relative expander entropy in the presence of a two-sided obstacle and applications. arXiv preprint arXiv:1906.07863, 2019.
  • [BW19b] Jacob Bernstein and Lu Wang. Topological uniqueness for self-expanders of small entropy. arXiv preprint arXiv:1902.02642, 2019.
  • [BW20] Jacob Bernstein and Lu Wang. A mountain-pass theorem for asymptotically conical self-expanders. arXiv preprint arXiv:2003.13857, 2020.
  • [BW21a] Jacob Bernstein and Lu Wang. Smooth compactness for spaces of asymptotically conical self-expanders of mean curvature flow. International Mathematics Research Notices, 2021(12):9016–9044, 2021.
  • [BW21b] Jacob Bernstein and Lu Wang. The space of asymptotically conical self-expanders of mean curvature flow. Mathematische Annalen, 380(1):175–230, 2021.
  • [CGG91] Yun Gang Chen, Yoshikazu Giga, and Shun’ichi Goto. Uniqueness and existence of viscosity solutions of generalized mean curvature flow equations. Journal of differential geometry, 33(3):749–786, 1991.
  • [CLS20] Otis Chodosh, Yevgeny Liokumovich, and Luca Spolaor. Singular behavior and generic regularity of min-max minimal hypersurfaces. arXiv preprint arXiv:2007.11560, 2020.
  • [CN13] Jeff Cheeger and Aaron Naber. Quantitative stratification and the regularity of harmonic maps and minimal currents. Communications on Pure and Applied Mathematics, 66(6):965–990, 2013.
  • [DG61] Ennio De Giorgi. Frontiere Orientate di Misura Minima. Seminario di Matematica della Scuola Normale Superiore di Pisa, 1960-61. Editrice Tecnico Scientifica, Pisa, 1961.
  • [Din20] Qi Ding. Minimal cones and self-expanding solutions for mean curvature flows. Mathematische Annalen, 376(1):359–405, 2020.
  • [EH89] Klaus Ecker and Gerhard Huisken. Mean curvature evolution of entire graphs. Annals of Mathematics, 130(3):453–471, 1989.
  • [ES91] Lawrence C Evans and Joel Spruck. Motion of level sets by mean curvature. i. Journal of Differential Geometry, 33(3):635–681, 1991.
  • [Fed69] Herbert Federer. Geometric Measure Theory, volume 153 of Grundlehren der Math. Wiss. Springer-Verlag, New York, 1969.
  • [Fed70] Herbert Federer. The singular sets of area minimizing rectifiable currents with codimension one and of area minimizing flat chains modulo two with arbitrary codimension. Bull. Amer. Math. Soc., 76:767–771, 1970.
  • [FF60] Herbert Federer and Wendell H. Fleming. Normal and integral currents. Ann. of Math. (2), 72:458–520, 1960.
  • [GT01] David Gilbarg and Neil S. Trudinger. Elliptic Partial Differential Equations of Second Order. Springer-Verlag, New York, 2001. reprint of the 1998 edition.
  • [HS85] Robert Hardt and Leon Simon. Area minimizing hypersurfaces with isolated singularities. J. Reine. Angew. Math, 1985.
  • [Ilm92] Tom Ilmanen. Generalized flow of sets by mean curvature on a manifold. Indiana University mathematics journal, pages 671–705, 1992.
  • [Ilm93] Tom Ilmanen. The level-set flow on a manifold. Differential geometry: partial differential equations on manifolds (Los Angeles, CA, 1990), 54:193–204, 1993.
  • [Ilm94] Tom Ilmanen. Elliptic regularization and partial regularity for motion by mean curvature, volume 520. American Mathematical Soc., 1994.
  • [Ilm96] T. Ilmanen. A strong maximum principle for singular minimal hypersurfaces. Calc. Var., 1996.
  • [Ilm98] Tom Ilmanen. Lectures on mean curvature flow and related equations. 1998.
  • [ISZ98] Tom Ilmanen, Peter Sternberg, and William P Ziemer. Equilibrium solutions to generalized motion by mean curvature. The Journal of Geometric Analysis, 8(5):845–858, 1998.
  • [Law69] H Blaine Lawson. Local rigidity theorems for minimal hypersurfaces. Annals of Mathematics, pages 187–197, 1969.
  • [LIN85] FANG-HUA LIN. REGULARITY FOR A CLASS OF PARAMETRIC OBSTACLE PROBLEMS (INTEGRAND, INTEGRAL CURRENT, PRESCRIBED MEAN CURVATURE, MINIMAL SURFACE SYSTEM). PhD thesis, 1985.
  • [Lin87] Fang Hua Lin. Approximation by smooth embedded hypersurfaces with positive mean curvature. Bulletin of the Australian Mathematical Society, 36(2):197–208, 1987.
  • [Loh18] Joachim Lohkamp. Minimal smoothings of area minimizing cones. arXiv preprint arXiv:1810.03157, 2018.
  • [LW20] Yangyang Li and Zhihan Wang. Generic regularity of minimal hypersurfaces in dimension 8. arXiv preprint arXiv:2012.05401, 2020.
  • [NV20] Aaron Naber and Daniele Valtorta. The singular structure and regularity of stationary varifolds. Journal of the European Mathematical Society, 22(10):3305–3382, 2020.
  • [Per02] Oscar Perdomo. First stability eigenvalue characterization of clifford hypersurfaces. Proceedings of the American Mathematical Society, 130(11):3379–3384, 2002.
  • [Pet06] Riemannian Geometry. NY: Springer Science + Business Media, LLC, 2006.
  • [Sha17] Ben Sharp. Compactness of minimal hypersurfaces with bounded index. J. Differential Geom., 106(2):317–339, 2017.
  • [Sim68] James Simons. Minimal varieties in riemannian manifolds. Annals of Mathematics, 1968.
  • [Sim83] Leon Simon. Lectures on Geometric Measure Theory, volume 3 of Proc. Centre for Mathematical Analysis, Australian National University. Australian National University, Centre for Mathematical Analysis, Canberra, 1983.
  • [Sim93] Leon Simon. Rectifiability of the singular sets of multiplicity 1 minimal surfaces and energy minimizing maps. Surveys in differential geometry, 2(1):246–305, 1993.
  • [Sim08] Leon Simon. A general asymptotic decay lemma for elliptic problems. arXiv preprint arXiv:0806.0462, 2008.
  • [Sim21a] Leon Simon. A liouville-type theorem for stable minimal hypersurfaces. arXiv preprint arXiv:2101.06404, 2021.
  • [Sim21b] Leon Simon. Stable minimal hypersurfaces in N+1+\mathbb{R}^{N+1+\ell} with singular set an arbitrary closed kk in {0}×\{0\}\times\mathbb{R}^{\ell}. arXiv preprint arXiv:2101.06401, 2021.
  • [Sma93] Nathan Smale. Generic regularity of homologically area minimizing hypersurfaces in eight-dimensional manifolds. Comm. in Analysis and Geom., 1993.
  • [SS81] Richard Schoen and Leon Simon. Regularity of stable minimal hypersurfaces. Comm. Pure Appl. Math., 34(6):741–797, 1981.
  • [SW89] Bruce Solomon and Brian White. A strong maximum principle for varifolds that are stationary with respect to even parametric elliptic functionals. Indiana University Mathematics Journal, 38(3):683–691, 1989.
  • [SY17] Richard Schoen and Shing-Tung Yau. Positive scalar curvature and minimal hypersurface singularities. arXiv preprint arXiv:1704.05490, 2017.
  • [Wan20] Zhihan Wang. Deformations of singular minimal hypersurfaces i, isolated singularities. arXiv preprint arXiv:2011.00548, 2020.
  • [Whi95] Brian White. The topology of hypersurfaces moving by mean curvature. Communications in analysis and geometry, 3(2):317–333, 1995.
  • [Wic14] Neshan Wickramasekera. A general regularity theory for stable codimension 1 integral varifolds. Annals of mathematics, pages 843–1007, 2014.
  • [Zhu18] Jonathan J Zhu. First stability eigenvalue of singular minimal hypersurfaces in spheres. Calculus of Variations and Partial Differential Equations, 57(5):1–13, 2018.
  • [ZZ20] Xin Zhou and Jonathan J. Zhu. Existence of hypersurfaces with prescribed mean curvature i – generic min-max. Cambridge Journal of Mathematics, 2020.