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

A two-piece property for free boundary minimal hypersurfaces in the (n+1)(n+1)-dimensional ball

Vanderson Lima and Ana Menezes Instituto de Matemática e Estatística
Universidade Federal do Rio Grande do Sul
Brazil
[email protected] Department of Mathematics
Princeton University
USA
[email protected]
Abstract.

We prove that every hyperplane passing through the origin in n+1\operatorname{\mathbb{R}}^{n+1} divides an embedded compact free boundary minimal hypersurface of the euclidean (n+1)(n+1)-ball in exactly two connected hypersurfaces. We also show that if a region in the (n+1)(n+1)-ball has mean convex boundary and contains a nullhomologous (n1)(n-1)-dimensional equatorial disk, then this region is a closed halfball. Our first result gives evidence to a conjecture by Fraser and Li in any dimension.

1. Introduction

Inspired by the work of Ros [26] for closed minimal surfaces in S3\operatorname{S}^{3}, the authors proved in [25] the two-piece property for free boundary minimal surfaces in the unit ball of 3\operatorname{\mathbb{R}}^{3}. This result gives evidence to a conjecture by Fraser and Li [10] concerning the first Steklov eigenvalue of free boundary minimal surfaces in B3.\operatorname{B}^{3}. Also, Kusner and McGrath [21] used our result of two-piece property in the free boundary context to prove the uniqueness of the critical catenoid among embedded minimal annuli invariant under the antipodal map. This settles a case of another well-known conjecture of [10] on the uniqueness of the critical catenoid.

In the present paper we prove that the two-piece property holds in any dimension. More precisely, we prove the following.

Theorem A (The two-piece property).

Every hyperplane in n+1\mathbb{R}^{n+1} passing through the origin divides an embedded compact free boundary minimal hypersurface of the unit (n+1)(n+1)-ball Bn+1\operatorname{B}^{n+1} in exactly two connected components.

We also prove the following result which can be seen as a strong version of the analog of the result by Solomon [29] in the free boundary context.

Theorem B.

Let WBn+1W\subset\operatorname{B}^{n+1} be a connected closed region with mean convex boundary such that W\partial W meets Sn\operatorname{S}^{n} orthogonally along its boundary and W\partial W is smooth. Suppose WW contains a set of the form PBn+1P\cap\operatorname{B}^{n+1} which is nullhomologous in WW (see Definition 3), where PP is a (n1)(n-1)-dimensional plane in n+1\mathbb{R}^{n+1} passing through the origin. Then WW is a closed halfball.

Let us remark that exactly as in the case n=2n=2, Theorem A can be proved by assuming the conjecture by Fraser and Li [10] on the the first Steklov eigenvalue of free boundary minimal hypersurfaces in Bn+1\operatorname{B}^{n+1}; hence, Theorem A gives evidence to this conjecture (see [25, Remark 2]).

The strategy to prove Theorem A and Theorem B is similar to the case n=2n=2 and uses Geometric Measure Theory to analyze the minimizers of a partially free boundary problem for the area functional. However, in higher dimensions the situation is more delicate since the hypersurfaces obtained as minimizers can have a singular set (see Theorem 1).

Motivated mainly by the celebrated work of Fraser and Schoen [11, 12], the study of free boundary minimal surfaces in B3\operatorname{B}^{3} saw a rapid development in the last few years, see for instance [22] and the references therein. However, the case of free boundary minimal hypersurfaces in Bn+1\operatorname{B}^{n+1} is not so well-studied. Concerning examples, some free boundary minimal hypersurfaces with symmetry were constructed in [14], and a variational theory has been developed in [23, 31, 32].

Regarding some properties of free boundary minimal hypersurfaces, we can mention that the asymptotic properties of the index of higher-dimensional free boundary minimal catenoids were studied in [28], and in [1] it was proved that the index of a properly embedded free boundary minimal hypersurface in Bn+1,3n+17,\operatorname{B}^{n+1},3\leq n+1\leq 7, grows linearly with the dimension of its first relative homology group. In [24] the first author proved the index can be controlled from above by a function of the LnL^{n} norm of the second fundamental form. Also, compactness results for the space of free boundary minimal hypersurfaces were obtained in [10, 2, 17].

2. Preliminary

2.1. Free boundary minimal hypersurfaces

Let Bn+1n+1\operatorname{B}^{n+1}\subset\operatorname{\mathbb{R}}^{n+1} be the unit ball of dimension n+1n+1 with boundary Bn+1=Sn\partial\operatorname{B}^{n+1}=\operatorname{S}^{n}. Throughout this paper we will denote by DnD^{n} the nn-dimensional equatorial disk which is the intersection of Bn+1\operatorname{B}^{n+1} with a hyperplane passing through the origin. In the following, s\mathcal{H}^{s} denotes the ss-dimensional Hausdorff measure, where s>0s>0.

Let Σn+1\Sigma\subset\mathbb{R}^{n+1}. Along this section we will use the following notation/assumptions:

  • Σ¯\overline{\Sigma} is compact and it is contained in Bn+1\operatorname{B}^{n+1}.

  • Σ\Sigma is an embedded orientable smooth hypersurface with boundary.

  • The singular set 𝒮Σ\mathcal{S}_{\Sigma} is the complement of Σ\Sigma in Σ¯\overline{\Sigma}. We suppose n(𝒮Σ)=0\mathcal{H}^{n}(\mathcal{S}_{\Sigma})=0.

  • The boundary of Σ\Sigma satisfies Σ=ΓIΓS\partial\Sigma=\Gamma_{I}\cup\Gamma_{S}, where int(ΓI)int(Bn+1)\mathrm{int}(\Gamma_{I})\subset\mathrm{int}(\operatorname{B}^{n+1}) and ΓSSn.\Gamma_{S}\subset\operatorname{S}^{n}. We have that, away from the singular set, Σ\partial\Sigma is an embedded smooth submanifold of dimension n1n-1.

Definition 1.

Let Σ\Sigma be as above. We say that Σ\Sigma is a minimal hypersurface with free boundary if the mean curvature vector of Σ\Sigma vanishes and Σ\Sigma meets Sn\operatorname{S}^{n} orthogonally along Σ\partial\Sigma (in particular, ΓI=).\Gamma_{I}=\emptyset). We say that Σ\Sigma is a minimal hypersurface with partially free boundary if the mean curvature vector of Σ\Sigma vanishes and its boundary ΓIΓS\Gamma_{I}\cup\Gamma_{S} satisfies that ΓI\Gamma_{I}\neq\emptyset and Σ\Sigma meets Sn\operatorname{S}^{n} orthogonally along ΓS\Gamma_{S}.

From now on, given a (partially) free boundary minimal hypersurface ΣBn+1\Sigma\subset\operatorname{B}^{n+1} with boundary Σ=ΓIΓS\partial\Sigma=\Gamma_{I}\cup\Gamma_{S}, we will call ΓI\Gamma_{I} its fixed boundary and ΓS\Gamma_{S} its free boundary.

Definition 2.

Let Σ\Sigma be a partially free boundary minimal hypersurface in Bn+1\operatorname{B}^{n+1}. We say that Σ\Sigma is stable if for any function fC(Σ)f\in C^{\infty}(\Sigma) such that f|ΓI0f|_{\Gamma_{I}}\equiv 0 and supp(f)\textrm{supp}(f) is away from the singular set Σ¯Σ\overline{\Sigma}\setminus\Sigma, we have

Σ(fΔΣf+|AΣ|2f2)𝑑n+ΓS(ffνf2)𝑑n10,-\int_{\Sigma}(f\Delta_{\Sigma}f+|A_{\Sigma}|^{2}f^{2})\,d\mathcal{H}^{n}+\int_{\Gamma_{S}}\left(f\frac{\partial f}{\partial\nu}-f^{2}\right)d\mathcal{H}^{n-1}\geq 0, (2.1)

or equivalently

Σ(|Σf|2|AΣ|2f2)𝑑nΓSf2𝑑n10,\int_{\Sigma}(|\nabla_{\Sigma}f|^{2}-|A_{\Sigma}|^{2}f^{2})\,d\mathcal{H}^{n}-\int_{\Gamma_{S}}f^{2}d\mathcal{H}^{n-1}\geq 0, (2.2)

where ν\nu is the outward normal vector field to ΓS\Gamma_{S}.

Observe that if Σ\Sigma is stable then, by an approximation argument, the inequality (2.2) holds for any function fH1(Σ)f\in H^{1}(\Sigma) such that f(p)=0f(p)=0 for a.e. pΓIp\in\Gamma_{I} and supp(f)\textrm{supp}(f) is away from the singular set. In particular (2.2) holds for any Lipschitz function satisfying the boundary condition.

Lemma 1.

Let Σ\Sigma be a partially free boundary minimal hypersurface in Bn+1\operatorname{B}^{n+1} of finite area and such that the singular set 𝒮Σ=Σ¯Σ\mathcal{S}_{\Sigma}=\overline{\Sigma}\setminus\Sigma satisfies 𝒮Σ=𝒮0𝒮1\mathcal{S}_{\Sigma}=\mathcal{S}_{0}\cup\mathcal{S}_{1}, where 𝒮0ΓI¯\mathcal{S}_{0}\subset\overline{\Gamma_{I}} and n2(𝒮1)=0\mathcal{H}^{n-2}\big{(}\mathcal{S}_{1}\big{)}=0. If ΓI\Gamma_{I} is contained in an n-dimensional equatorial disk, then Σ\Sigma is totally geodesic.

Proof.

Let Σ\Sigma be as in the hypotheses and denote by DnD^{n} the equatorial disk that contains ΓI¯\overline{\Gamma_{I}}. Let vSnv\in\operatorname{S}^{n} be a vector orthogonal to the disk DnD^{n} and consider the function f(x)=x,vf(x)=\langle x,v\rangle, xΣ¯.x\in\overline{\Sigma}. By hypothesis, we know that f|ΓI0f|_{\Gamma_{I}}\equiv 0. A standard calculation using that Σ\Sigma is minimal and free boundary yields

ΔΣf=0,fν=f.\Delta_{\Sigma}f=0,\quad\frac{\partial f}{\partial\nu}=f.

Fix ϵ>0\epsilon>0 and consider a smooth function ηϵ:[1,1][0,1]\eta_{\epsilon}:[-1,1]\to[0,1] so that

  • ηϵ(s)=0\eta_{\epsilon}(s)=0 for |s|<ϵ|s|<\epsilon,

  • ηϵ(s)=1\eta_{\epsilon}(s)=1 for |s|>2ϵ|s|>2\epsilon,

  • |ηϵ|<Cϵ|\eta^{\prime}_{\epsilon}|<\displaystyle\frac{C}{\epsilon}, for some constant C>0C>0.

Define ϕ0,ϵ:Σ¯[0,1]\phi_{0,\epsilon}:\overline{\Sigma}\to[0,1] as ϕ0,ϵ(x)=ηϵ(f(x))\phi_{0,\epsilon}(x)=\eta_{\epsilon}(f(x)). In particular, we have |Σϕ0,ϵ|<C/ϵ|\nabla_{\Sigma}\phi_{0,\epsilon}|<C/\epsilon in Σ.\Sigma. Observe that the set S𝒮1S\subset\mathcal{S}_{1} where ϕ0,ϵ\phi_{0,\epsilon} is not smooth satisfies n2(S)=0.\mathcal{H}^{n-2}(S)=0.

Since 𝒮1{|f(x)|ϵ2}\mathcal{S}_{1}\cap\{|f(x)|\geq\frac{\epsilon}{2}\} is compact and n2(𝒮1)=0\mathcal{H}^{n-2}\big{(}\mathcal{S}_{1}\big{)}=0, for any ϵ>0\epsilon^{\prime}>0 there exist balls Bri(pi)n+1,i=1,,m,B_{r_{i}}(p_{i})\subset\mathbb{R}^{n+1},\,i=1,\cdots,m, such that

𝒮1{|f(x)|ϵ2}i=1mBri(pi),i=1mrin2<ϵ,i=1,,m.\mathcal{S}_{1}\cap\left\{|f(x)|\geq\frac{\epsilon}{2}\right\}\subset\bigcup_{i=1}^{m}B_{r_{i}}(p_{i}),\quad\sum_{i=1}^{m}r_{i}^{n-2}<\epsilon^{\prime},\ i=1,\cdots,m.

For each i=1,,mi=1,\cdots,m, consider a smooth function ϕi:Σ¯[0,1]\phi_{i}:\overline{\Sigma}\to[0,1] such that

  • ϕi(s)=0\phi_{i}(s)=0 in Bri(pi)B_{r_{i}}(p_{i}),

  • ϕi(s)=1\phi_{i}(s)=1 in n+1B2ri(pi)\mathbb{R}^{n+1}\setminus B_{2r_{i}}(p_{i}),

  • |Σϕi|<2ri,xΣ|\nabla_{\Sigma}\phi_{i}|<\displaystyle\frac{2}{r_{i}},\,\forall\,x\in\Sigma.

Define ϕϵ,fϵ:Σ¯[0,1]\phi_{\epsilon},f_{\epsilon}:\overline{\Sigma}\to[0,1] by ϕϵ(x)=min0imϕi\phi_{\epsilon}(x)=\displaystyle\min_{0\leq i\leq m}\phi_{i}, where ϕ0=ϕ0,ϵ,\phi_{0}=\phi_{0,\epsilon}, and fϵ=ϕϵff_{\epsilon}=\phi_{\epsilon}f. We have that fϵf_{\epsilon} is Lipschitz and fϵ|ΓI0f_{\epsilon}|_{\Gamma_{I}}\equiv 0, hence (2.2) holds. Moreover

|Σfϵ|2=ϕϵ2|Σf|2+2fϕϵΣf,Σϕϵ+f2|Σϕϵ|2|\nabla_{\Sigma}f_{\epsilon}|^{2}=\phi_{\epsilon}^{2}|\nabla_{\Sigma}f|^{2}+2f\phi_{\epsilon}\langle\nabla_{\Sigma}f,\nabla_{\Sigma}\phi_{\epsilon}\rangle+f^{2}|\nabla_{\Sigma}\phi_{\epsilon}|^{2}

and

Σϕϵ2|Σf|2𝑑n=Σfϕϵ2ΔΣf𝑑nΣ2fϕϵΣϕϵ,Σf𝑑n+Σϕϵ2ffν𝑑n1=Σ2fϕϵΣϕϵ,Σf𝑑n+ΓSϕϵ2f2𝑑n1,\begin{array}[]{rcl}\displaystyle\int_{\Sigma}\phi^{2}_{\epsilon}|\nabla_{\Sigma}f|^{2}d\mathcal{H}^{n}&=&\displaystyle-\int_{\Sigma}f\phi_{\epsilon}^{2}\Delta_{\Sigma}fd\mathcal{H}^{n}-\int_{\Sigma}2f\phi_{\epsilon}\langle\nabla_{\Sigma}\phi_{\epsilon},\nabla_{\Sigma}f\rangle d\mathcal{H}^{n}+\int_{\partial\Sigma}\phi_{\epsilon}^{2}f\frac{\partial f}{\partial\nu}d\mathcal{H}^{n-1}\\ \\ \displaystyle\hfil&=&\displaystyle-\int_{\Sigma}2f\phi_{\epsilon}\langle\nabla_{\Sigma}\phi_{\epsilon},\nabla_{\Sigma}f\rangle d\mathcal{H}^{n}+\int_{\Gamma_{S}}\phi_{\epsilon}^{2}f^{2}d\mathcal{H}^{n-1},\end{array}

since ΔΣf0\Delta_{\Sigma}f\equiv 0, fν=f\frac{\partial f}{\partial\nu}=f and f|ΓI0.f|_{\Gamma_{I}}\equiv 0. Hence, applying it to (2.2), we get

Σ(f2|Σϕϵ|2|AΣ|2ϕϵ2f2)𝑑n0.\int_{\Sigma}(f^{2}|\nabla_{\Sigma}\phi_{\epsilon}|^{2}-|A_{\Sigma}|^{2}\phi_{\epsilon}^{2}f^{2})\,d\mathcal{H}^{n}\geq 0. (2.3)

On the other hand, along Σ¯\overline{\Sigma} we have f21f^{2}\leq 1. Since ff has support away from the singular set, by the classical monotonicity formula at the interior and at the free boundary, there is CΣ,ϵ>0C_{\Sigma,\epsilon}>0 such that

n(B2ri(pi)Σ)CΣ,ϵrin.\mathcal{H}^{n}\bigl{(}B_{2r_{i}}(p_{i})\cap\Sigma\bigr{)}\leq C_{\Sigma,\epsilon}\,r_{i}^{n}.

Thus

Σf2|Σϕϵ|2𝑑n\displaystyle\int_{\Sigma}f^{2}|\nabla_{\Sigma}\phi_{\epsilon}|^{2}d\mathcal{H}^{n} i=0mΣf2|Σϕi,ϵ|2𝑑n\displaystyle\leq\sum_{i=0}^{m}\int_{\Sigma}f^{2}|\nabla_{\Sigma}\phi_{i,\epsilon}|^{2}d\mathcal{H}^{n}
=Σf2|Σϕ0,ϵ|2𝑑n+i=1m(B2ri(pi)Bri(pi))Σf2|Σϕi,ϵ|2𝑑n\displaystyle=\int_{\Sigma}f^{2}|\nabla_{\Sigma}\phi_{0,\epsilon}|^{2}d\mathcal{H}^{n}+\sum_{i=1}^{m}\int_{\big{(}B_{2r_{i}}(p_{i})\setminus B_{r_{i}}(p_{i})\big{)}\cap\Sigma}f^{2}|\nabla_{\Sigma}\phi_{i,\epsilon}|^{2}d\mathcal{H}^{n}
4Cn(Σ{|f|1(ϵ,2ϵ)})+i=1m4ri2n(B2ri(pi)Σ)\displaystyle\leq 4C\mathcal{H}^{n}\bigl{(}\Sigma\cap\{|f|^{-1}(\epsilon,2\epsilon)\}\bigr{)}+\sum_{i=1}^{m}\frac{4}{r_{i}^{2}}\,\mathcal{H}^{n}\bigl{(}B_{2r_{i}}(p_{i})\cap\Sigma\bigr{)}
4Cn(Σ{|f|1(ϵ,2ϵ)})+CΣ,ϵi=1mrin2\displaystyle\leq 4C\mathcal{H}^{n}\bigl{(}\Sigma\cap\{|f|^{-1}(\epsilon,2\epsilon)\}\bigr{)}+C^{\prime}_{\Sigma,\epsilon}\sum_{i=1}^{m}r_{i}^{n-2}
4Cn(Σ{|f|1(ϵ,2ϵ)})+CΣ,ϵϵ.\displaystyle\leq 4C\mathcal{H}^{n}\bigl{(}\Sigma\cap\{|f|^{-1}(\epsilon,2\epsilon)\}\bigr{)}+C^{\prime}_{\Sigma,\epsilon}\,\epsilon^{\prime}.

If we let ϵ0\epsilon^{\prime}\to 0 first and then ϵ0\epsilon\to 0 we obtain

Σ|AΣ|2f2𝑑n=0.\int_{\Sigma}|A_{\Sigma}|^{2}f^{2}d\mathcal{H}^{n}=0.

If |AΣ|0|A_{\Sigma}|\equiv 0 then Σ\Sigma is totally geodesic and we are done. If |AΣ|(x)>0|A_{\Sigma}|(x)>0 for some xΣ,x\in\Sigma, then we can find a neighborhood UU of xx in Σ\Sigma such that |AΣ||A_{\Sigma}| is strictly positive. This implies y,v=0\langle y,v\rangle=0 for any yUy\in U. Therefore, Σ\Sigma is entirely contained in the disk Dn;D^{n}; in particular, it is totally geodesic. ∎

An equatorial disk DnD^{n} divides the ball Bn+1\operatorname{B}^{n+1} into two (open) halfballs. We will denote these two halfballs by B+\operatorname{B}^{+} and B,\operatorname{B}^{-}, and we have Bn+1Dn=B+B.\operatorname{B}^{n+1}\setminus D^{n}=\operatorname{B}^{+}\cup\operatorname{B}^{-}.

In the next proposition we will summarize some facts about partially free boundary minimal surfaces in Bn+1\operatorname{B}^{n+1} which we will use in the proof of Theorem 3.

Proposition 1.
  1. (i)

    Let DnD^{n} be an equatorial disk and let Σ\Sigma be a smooth partially free boundary minimal hypersurface in Bn+1\operatorname{B}^{n+1} contained in one of the closed halfballs determined by DnD^{n}, say B+¯,\overline{\operatorname{B}^{+}}, and such that ΣB+¯\partial\Sigma\subset\partial\overline{\operatorname{B}^{+}}. If Σ\Sigma is not contained in an equatorial disk, then Σ\Sigma has necessarily nonempty fixed boundary and nonempty free boundary.

  2. (ii)

    The only smooth (partially) free boundary minimal hypersurface that contains a (n1)(n-1)-dimensional piece of the free boundary of a nn-dimensional equatorial disk is (contained in) this equatorial disk itself.

Proof.

(i){\it(i)} If the free boundary were empty, we could apply the (interior) maximum principle with the family of hyperplanes parallel to the disk DnD^{n} and conclude that Σ\Sigma should be contained in the disk DnD^{n}. On the other hand, if the fixed boundary were empty, then we would have a minimal hypersurface entirely contained in a halfball without fixed boundary; hence, we could apply the (interior or free boundary version of) maximum principle with the family of equatorial disks that are rotations of DnD^{n} around a (n1)(n-1)-dimensional equatorial disk and conclude that Σ\Sigma should be an equatorial disk.

(𝑖𝑖){\it(ii)} Let DnD^{n} be an equatorial disk and suppose that Σ\Sigma is a (partially) free boundary minimal hypersurface such that ΣDn\Sigma\cap D^{n} contains a (n1)(n-1)-dimensional piece Υ\Upsilon of the free boundary of DnD^{n} in Sn.\operatorname{S}^{n}. Assume, without loss of generality, Dn{xn+1=0}D^{n}\subset\{x_{n+1}=0\}.

Observe that since Σ\Sigma is free boundary we know that xn+1η|Υ=xn+1|Υ=0,\frac{\partial x_{n+1}}{\partial\eta}{\big{|}}_{\Upsilon}=x_{n+1}{\big{|}}_{\Upsilon}=0, where η\eta is the conormal vector to Υ;\Upsilon; and since Σ\Sigma is a minimal hypersurface in n+1\operatorname{\mathbb{R}}^{n+1} we have that xn+1|Σx_{n+1}{\big{|}}_{\Sigma} is harmonic.

We will show that xn+1|Σ0.x_{n+1}{\big{|}}_{\Sigma}\equiv 0.

Consider an extension Σ^\hat{\Sigma} of Σ\Sigma along Υ\Upsilon such that Υint(Σ^)\Upsilon\subset\mbox{int}(\hat{\Sigma}) and define x^n+1\hat{x}_{n+1} on Σ^\hat{\Sigma} as

{x^n+1=xn+1onΣx^n+1=0onΣ^Σ\Big{\{}\begin{array}[]{ll}\hat{x}_{n+1}=x_{n+1}&\ \mbox{on}\ \ \Sigma\\ \hat{x}_{n+1}=0&\ \mbox{on}\ \ \hat{\Sigma}\setminus\Sigma\end{array}

Observe that x^n+1|Υ=xn+1|Υ0\hat{x}_{n+1}{\big{|}}_{\Upsilon}=x_{n+1}{\big{|}}_{\Upsilon}\equiv 0, x^n+1η^|Υ=0\frac{\partial\hat{x}_{n+1}}{\partial\hat{\eta}}{\big{|}}_{\Upsilon}=0 and x^n+1η|Υ=xn+1η|Υ=0,\frac{\partial\hat{x}_{n+1}}{\partial\eta}{\big{|}}_{\Upsilon}=\frac{\partial{x}_{n+1}}{\partial\eta}{\big{|}}_{\Upsilon}=0, where η^\hat{\eta} is the conormal to Υ\Upsilon pointing towards Σ\Sigma and η\eta is the conormal to Υ\Upsilon pointing towards Σ^Σ;\hat{\Sigma}\setminus\Sigma; hence, x^n+1\hat{x}_{n+1} is C1C^{1} in a neighborhood of Υ\Upsilon in Σ^.\hat{\Sigma}.

Claim 1.

x^n+1\hat{x}_{n+1} is a weak solution to the Laplacian equation Δu=0.\Delta u=0.

Observe that x^n+1\hat{x}_{n+1} is a harmonic function on Σ^Υ\hat{\Sigma}\setminus\Upsilon, so we just need to show the claim in a neighborhood of Υ.\Upsilon.

Consider a domain Ω=Ω1Ω2\Omega=\Omega_{1}\cup\Omega_{2} where Ωi¯=Γi(Ω¯Υ)\partial\overline{\Omega_{i}}=\Gamma_{i}\cup(\overline{\Omega}\cap\Upsilon) with Ω1Σ^Σ\Omega_{1}\subset\hat{\Sigma}\setminus\Sigma and Ω2Σ\Omega_{2}\subset\Sigma (see Figure 1), and let ϕ:Σ^\phi:\hat{\Sigma}\to\operatorname{\mathbb{R}} be a smooth function with compact support contained in Ω.\Omega.

Refer to caption
Figure 1. Ω=Ω1Ω2\Omega=\Omega_{1}\cup\Omega_{2}.

Integration by parts gives us

Ωϕ,x^n+1dσ=Ωx^n+1Δϕdσ+Ωx^n+1ϕ,νdL=Ωx^n+1Δϕdσ,\begin{array}[]{rcl}\displaystyle\int_{\Omega}\langle\nabla\phi,\nabla\hat{x}_{n+1}\rangle{\rm d}\sigma&=&\displaystyle-\int_{\Omega}\hat{x}_{n+1}\Delta\phi\,{\rm d}\sigma+\int_{\partial\Omega}\hat{x}_{n+1}\langle\nabla\phi,\nu\rangle{\rm d}L\\ &=&\displaystyle-\int_{\Omega}\hat{x}_{n+1}\Delta\phi\,{\rm d}\sigma,\end{array}

since supp(ϕ)Ω,(\phi)\subset\subset\Omega, where ν\nu is the outward conormal to Ω.\partial\Omega.

Then,

Ωx^n+1Δϕdσ=Ωϕ,x^n+1dσ=Ω1ϕ,x^n+1dσ+Ω2ϕ,x^n+1dσ.\begin{array}[]{rcl}-\displaystyle\int_{\Omega}\hat{x}_{n+1}\Delta\phi\,{\rm d}\sigma&=&\displaystyle\int_{\Omega}\langle\nabla\phi,\nabla\hat{x}_{n+1}\rangle{\rm d}\sigma\\ &=&\displaystyle\int_{\Omega_{1}}\langle\nabla\phi,\nabla\hat{x}_{n+1}\rangle{\rm d}\sigma+\displaystyle\int_{\Omega_{2}}\langle\nabla\phi,\nabla\hat{x}_{n+1}\rangle{\rm d}\sigma.\\ \end{array}

We have

Ω1ϕ,x^n+1dσ=Ω1ϕΔx^n+1dσ+Ω1ϕx^n+1,ν1dL=Υϕx^n+1,ν1dL=0,\begin{array}[]{rcl}\displaystyle\int_{\Omega_{1}}\langle\nabla\phi,\nabla\hat{x}_{n+1}\rangle{\rm d}\sigma&=&-\displaystyle\int_{\Omega_{1}}\phi\Delta\hat{x}_{n+1}{\rm d}\sigma+\displaystyle\int_{\partial\Omega_{1}}\phi\langle\nabla\hat{x}_{n+1},\nu_{1}\rangle{\rm d}L\\ &=&\displaystyle\int_{\Upsilon}\phi\langle\nabla\hat{x}_{n+1},\nu_{1}\rangle{\rm d}L\\ &=&0,\end{array}

where in the first equality we used that x^n+1|Ω10\hat{x}_{n+1}{\big{|}}_{\Omega_{1}}\equiv 0 and in the second equality we used the fact that x^n+1ν1|Υ=0,\frac{\partial\hat{x}_{n+1}}{\partial\nu_{1}}{\big{|}}_{\Upsilon}=0, where ν1\nu_{1} is the outward conomal to Υ\Upsilon with respect to Ω1.\Omega_{1}.

Analogously, we have

Ω2ϕ,x^n+1dσ=Ω2ϕΔx^n+1dσ+Ω2ϕx^n+1,ν1dL=Υϕx^n+1,ν2dL=0,\begin{array}[]{rcl}\displaystyle\int_{\Omega_{2}}\langle\nabla\phi,\nabla\hat{x}_{n+1}\rangle{\rm d}\sigma&=&-\displaystyle\int_{\Omega_{2}}\phi\Delta\hat{x}_{n+1}{\rm d}\sigma+\displaystyle\int_{\partial\Omega_{2}}\phi\langle\nabla\hat{x}_{n+1},\nu_{1}\rangle{\rm d}L\\ &=&\displaystyle\int_{\Upsilon}\phi\langle\nabla\hat{x}_{n+1},\nu_{2}\rangle{\rm d}L\\ &=&0,\end{array}

where in the first equality we used that x^n+1|Ω2=xn+1|Ω2\hat{x}_{n+1}{\big{|}}_{\Omega_{2}}=x_{n+1}{\big{|}}_{\Omega_{2}} is harmonic and in the second equality we used the fact that x^n+1ν2|Υ=0,\frac{\partial\hat{x}_{n+1}}{\partial\nu_{2}}{\big{|}}_{\Upsilon}=0, where ν2\nu_{2} is the outward conomal to Υ\Upsilon with respect to Ω2.\Omega_{2}.

Therefore, the claim follows and, by the Elliptic theory, x^n+1\hat{x}_{n+1} has to be a (strong) solution to the Laplacian equation. Moreover, since x^n+1\hat{x}_{n+1} vanishes on an open set, the unique continuation result implies that x^n+10\hat{x}_{n+1}\equiv 0 on Σ^,\hat{\Sigma}, that is, Σ\Sigma is (contained in) the equatorial disk Dn.D^{n}.

2.2. Integer rectifiable varifolds

A set Mn+1M\subset\mathbb{R}^{n+1} is called countably kk-rectifiable if MM is k\mathcal{H}^{k}-measurable and if

Mj=0Mj,M\subset\bigcup_{j=0}^{\infty}M_{j},

where k(M0)=0\mathcal{H}^{k}(M_{0})=0 and for j1j\geq 1, MjM_{j} is an kk-dimensional C1C^{1}-submanifold of n+1\mathbb{R}^{n+1}. Such MM possesses k\mathcal{H}^{k}-a.e. an approximate tangent space TxMT_{x}M.

Let G(n+1,k)G(n+1,k) be the Grassmannian of kk-hyperplanes in n+1\mathbb{R}^{n+1}. An integer multiplicity rectifiable kk-varifold 𝒱=v(M,θ)\mathcal{V}=v(M,\theta) is a Radon measure on U×G(n+1,k)U\times G(n+1,k), defined by

𝒱(f)=Mf(x,TxM)θ(x)𝑑k,fC0c(U×G(n+1,k)),\mathcal{V}(f)=\int_{M}f(x,T_{x}M)\,\theta(x)\,d\mathcal{H}^{k},\ f\in C_{0}^{c}\big{(}U\times G(n+1,k)\big{)},

where MUM\subset U is countably kk-rectifiable and θ>0\theta>0 is a locally k\mathcal{H}^{k}-integrable integer valued function. Also, we say 𝒱=v(M,θ)\mathcal{V}=v(M,\theta) is stationary if

M(divMζ)θ𝑑k=0,\int_{M}\big{(}\operatorname{div}_{M}\zeta\big{)}\theta\,d\mathcal{H}^{k}=0, (2.4)

for any C1C^{1}-vector field ζ\zeta of compact support.

Then we have the following result, see [20].

Lemma 2.

Let Σn+1\Sigma\subset\mathbb{R}^{n+1} be an embedded C1C^{1}-hypersurface such that n1((Σ¯Σ)U)=0\mathcal{H}^{n-1}\big{(}(\overline{\Sigma}\setminus\Sigma)\cap U\big{)}=0, for every open set Un+1U\subset\mathbb{R}^{n+1} with compact closure. Let θ>0\theta>0 be a integer valued function which is locally constant. Then, the following conditions are equivalent:

  1. (1)

    𝒱=v(Σ,θ)\mathcal{V}=v(\Sigma,\theta) is stationary.

  2. (2)

    HΣ=0\vec{H}_{\Sigma}=0, and there is CΣ>0C_{\Sigma}>0 such that for any ball Br(p)n+1B_{r}(p)\subset\mathbb{R}^{n+1} we have

    n(Br(p)Σ)CΣrn.\mathcal{H}^{n}\bigl{(}B_{r}(p)\cap\Sigma\bigr{)}\leq C_{\Sigma}\,r^{n}.

2.3. Minimizing Currents with Partially Free Boundary

In this section we will use the following notation.

  • Un+1U\subset\mathbb{R}^{n+1} is an open set;

  • 𝒟k(U)={C-k-formsω;sptωU}\mathcal{D}^{k}(U)=\{C^{\infty}\textrm{-}\ k\textrm{-}\textrm{forms}\ \omega;\ \operatorname{spt}\ \omega\subset U\};

  • 𝒟k(U)\mathcal{D}_{k}(U) denotes the dual of 𝒟k(U)\mathcal{D}^{k}(U), and its elements are called kk-currents with support in UU;

  • The mass of T𝒟k(U)T\in\mathcal{D}_{k}(U) in WW is defined by

    𝐌W(T):=sup{T(ω);ω𝒟k(U),sptωW,|ω|1}+;{\bf M}_{W}(T):=\sup\{T(\omega);\ \omega\in\mathcal{D}^{k}(U),\ \operatorname{spt}\omega\subset W,\ |\omega|\leq 1\}\leq+\infty;
  • The boundary of T𝒟k(U)T\in\mathcal{D}_{k}(U) is the (k1)(k-1)-current T𝒟k1(U)\partial T\in\mathcal{D}_{k-1}(U) given by

    T(ω):=T(dω),\partial T(\omega):=T(d\omega),

    where dd denotes the exterior derivative operator.

Consider a compact domain Wn+1W\subset\operatorname{\mathbb{R}}^{n+1} such that W=SM\partial W=S\cup M, where SS is a compact C2C^{2}-hypersurface (not necessarily connected) with boundary, MM is a smooth compact mean convex hypersurface with boundary, which intersects SS orthogonally along S\partial S, and such that int(S)int(M)=\mathrm{int}(S)\cap\mathrm{int}(M)=\emptyset.

Let ΩW\Omega\subset W be a compact hypersurface with boundary Γ=Ω\Gamma=\partial\Omega. We assume that Γint(W)\Gamma\cap\mathrm{int}(W) is an embedded C2C^{2}-submanifold of dimension n1n-1 away from a singular set 𝒮0\mathcal{S}_{0} such that n1(𝒮0)=0\mathcal{H}^{n-1}(\mathcal{S}_{0})=0.

Define the class \mathfrak{C} of admissible currents by

={T𝒟n(n+1);Tis integer multiplicity rectifiable,\displaystyle\mathfrak{C}=\{T\in\mathcal{D}_{n}(\operatorname{\mathbb{R}}^{n+1});\ T\ \mbox{is integer multiplicity rectifiable},
sptTWand is compact,andspt(ΓT)S},\displaystyle\operatorname{spt}T\subset W\ \mbox{and is compact},\ \mbox{and}\ \operatorname{spt}\bigl{(}\llbracket\Gamma\rrbracket-\partial T\bigr{)}\subset S\},

where Γ\llbracket\Gamma\rrbracket is the current associated to Γ\Gamma with multiplicity one. We want to minimize area in \mathfrak{C}, that is, we are looking for TT\in\mathfrak{C} such that

𝐌(T)=inf{𝐌(T~);T~}.{\bf M}(T)=\inf\{{\bf M}(\tilde{T});\ \tilde{T}\in\mathfrak{C}\}. (2.5)

Observe that \mathfrak{C}\neq\emptyset since Ω\llbracket\Omega\rrbracket\in\mathfrak{C}. Hence, it follows from  [8, 5.1.6(1)5.1.6(1)], that the variational problem (2.5) has a solution (see also [15]). If TT\in\mathfrak{C} is a solution we have

𝐌(T)\displaystyle{\bf M}(T) \displaystyle\leq 𝐌(T+X),\displaystyle{\bf M}(T+X), (2.6)
sptT\displaystyle\operatorname{spt}T \displaystyle\subset W,\displaystyle W, (2.7)
μT(S)\displaystyle\mu_{T}(S) =\displaystyle= 0,\displaystyle 0, (2.8)

for any integer multiplicity current X𝒟n(n+1)X\in\mathcal{D}_{n}(\mathbb{R}^{n+1}) with compact support such that sptXW\operatorname{spt}X\subset W and sptXS\operatorname{spt}\partial X\subset S.

In order to apply the known regularity theory for TT we need the following result, whose proof is the same as that of the case n=2n=2 (see Section 3 in [25]).

Proposition 2.

If TT is a solution of (2.5), then either sptTΓWM\operatorname{spt}T\setminus\Gamma\subset W\setminus M or sptTM\operatorname{spt}T\subset M.

For any given nn-dimensional compact set KWK\subset W we call corners the set of points of K\partial K which also belong to SS. We then have the following regularity result.

Theorem 1.

Let TT be a solution of (2.5). Then there is a set 𝒮1sptT\mathcal{S}_{1}\subset\operatorname{spt}T such that, away from 𝒮0𝒮1Γ\mathcal{S}_{0}\cup\mathcal{S}_{1}\cup\Gamma, TT is supported in a oriented embedded minimal C2C^{2}-hypersurface, which meets SS orthogonally along spt(ΓT)\operatorname{spt}\bigl{(}\llbracket\Gamma\rrbracket-\partial T\bigr{)}. Moreover

{𝒮1=,ifn6,𝒮1is discrete,ifn=7,n7+δ(𝒮1)=0,δ>0,ifn>7.\left\{\begin{array}[]{rl}&\mathcal{S}_{1}=\emptyset,\quad\text{if}\ n\leq 6,\\ \\ &\mathcal{S}_{1}\ \textrm{is discrete},\quad\text{if}\ n=7,\\ \\ &\mathcal{H}^{n-7+\delta}\big{(}\mathcal{S}_{1}\big{)}=0,\ \forall\,\delta>0,\quad\text{if}\ n>7.\end{array}\right. (2.9)
Proof.

Let us write

sptT(𝒮0)=𝒮1,\operatorname{spt}T\setminus(\mathcal{S}_{0})=\mathcal{R}\cup\mathcal{S}_{1},

where the union is disjoint and \mathcal{R} consists of the points xsptTx\in\operatorname{spt}T such that there is a neighborhood Un+1U\subset\mathbb{R}^{n+1} of xx where T  UT\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}U is given by mm-times (mm\in\mathbb{N}) integration over an embedded C2C^{2}-hypersurface with boundary. To complete the proof we will prove the following:

  • Regularity at the interior: (sptTsptT)(\operatorname{spt}T\setminus\operatorname{spt}\partial T)\cap\mathcal{R}\neq\emptyset and (sptTsptT)𝒮1(\operatorname{spt}T\setminus\operatorname{spt}\partial T)\cap\mathcal{S}_{1} satisfies (2.9);

  • Regularity at the free-boundary: spt(ΓT)\operatorname{spt}\bigl{(}\llbracket\Gamma\rrbracket-\partial T\bigr{)}\cap\mathcal{R}\neq\emptyset and spt(ΓT)𝒮1\operatorname{spt}\bigl{(}\llbracket\Gamma\rrbracket-\partial T\bigr{)}\cap\mathcal{S}_{1} satisfies (2.9);

The interior regularity is a classical result, see [8, Section 5.3]. Since by Proposition 2 the free part of the boundary is contained in SSS\setminus\partial S, we can use the result by Grüter [16] to conclude the regularity at the free boundary (away from the corners). ∎

3. The two-piece property and other results

Definition 3.

Let WW be a region in Bn+1\operatorname{B}^{n+1} and let ΥW\Upsilon\subset W be a (n1)(n-1)-dimensional equatorial disk (that is, the intersection of Bn+1\operatorname{B}^{n+1} with an (n1)(n-1)-dimensional plane passing through the origin). We say that Υ\Upsilon is nullhomologous in WW if there exists a compact hypersurface MWM\subset W such that M=ΥΓ\partial M=\Upsilon\cup\Gamma, where Γ\Gamma is a (n1)(n-1)-dimensional compact set contained in Sn\operatorname{S}^{n} (see Figure 2).

Refer to caption
Figure 2. In this region WW, any (n1)(n-1)-dimensional equatorial disk ΥW\Upsilon\subset W is nullhomologous.

The boundary of the region WW can be written as UVU\cup V, where int(U)int(Bn+1)\mathrm{int}(U)\subset\mathrm{int}(\operatorname{B}^{n+1}) and VSn.V\subset\operatorname{S}^{n}. In the next theorem we will denote by W\partial W the closure of the component U,U, that is, W=U¯.\partial W=\overline{U}.

Theorem 2.

Let WBn+1W\subset\operatorname{B}^{n+1} be a connected closed region with (not necessarily strictly) mean convex boundary such that W\partial W meets Sn\operatorname{S}^{n} orthogonally along its boundary and W\partial W is smooth. If WW contains a (n1)(n-1)-dimensional equatorial disk Υ\Upsilon, and Υ\Upsilon is nullhomologous in WW, then WW is a closed (n+1)(n+1)-dimensional halfball.

Proof.

Up to a rotation of Υ\Upsilon around the origin, we can assume that ΥW\Upsilon\cap\partial W is nonempty. Since Υ\Upsilon is nullhomologous in WW, there exists a compact hypersurface MM contained in WW such that M=ΥΓ\partial M=\Upsilon\cup\Gamma, where Γ\Gamma is a (n1)(n-1)-dimensional compact set contained in Sn.\operatorname{S}^{n}. We consider the class of admissible currents

={T𝒟n(n+1);Tis integer multiplicity rectifiable,\displaystyle\mathfrak{C}=\{T\in\mathcal{D}_{n}(\mathbb{R}^{n+1});\ T\ \mbox{is integer multiplicity rectifiable},
sptTWand is compact,andspt(MT)SnW},\displaystyle\operatorname{spt}T\subset W\ \mbox{and is compact},\ \mbox{and}\ \operatorname{spt}\bigl{(}\llbracket\partial M\rrbracket-\partial T\bigr{)}\subset\operatorname{S}^{n}\cap W\},

where M\llbracket\partial M\rrbracket is the current associated to M\partial M with multiplicity one, and we minimize area (mass) in .\mathfrak{C}. Then, by the results presented in Section 2.3, we get a compact embedded (orientable) partially free boundary minimal hypersurface ΣW\Sigma\subset W which minimizes area among compact hypersurfaces in WW with boundary on the class Γ=ΥΓ~;\Gamma=\Upsilon\cup\tilde{\Gamma}; in particular, its fixed boundary is exactly Υ.\Upsilon. Moreover, by Proposition 2 in Section 2.3, either ΣW\Sigma\subset\partial W or ΣWΥ.\Sigma\cap\partial W\subset\Upsilon.

Now the same arguments we used in the proof of Theorem 1 in [25] can be applied. In fact:

Claim 2.

Σ\Sigma is stable.

In the case WΣΓ\partial W\cap\Sigma\subset\Gamma, Σ\Sigma is automatically stable in the sense of Definition 2, since it minimizes area for all local deformations.

Suppose ΣW\Sigma\subset\partial W. For any fCc(Σ)f\in C^{\infty}_{c}(\Sigma) with f|Υ0f|_{\Upsilon}\equiv 0, consider Q(f,f)Q(f,f) defined by

Q(f,f)=Σ(|Σf|2|AΣ|2f2)𝑑nΓf2𝑑n1Σf2𝑑n,Q(f,f)=\frac{\int_{\Sigma}\left(|\nabla_{\Sigma}f|^{2}-|A_{\Sigma}|^{2}f^{2}\right)d\mathcal{H}^{n}-\int_{\Gamma}f^{2}d\mathcal{H}^{n-1}}{\int_{\Sigma}f^{2}d\mathcal{H}^{n}},

and let f1f_{1} be a first eigenfunction, i.e., Q(f1,f1)=inffQ(f,f)Q(f_{1},f_{1})=\inf_{f}Q(f,f).

Observe that although differently from the classical stability quotient (we have an extra term that depends on the boundary of Σ\Sigma) we can still guarantee the existence of a first eigenfunction. In fact, since for any δ>0\delta>0 there exists Cδ>0C_{\delta}>0 such that fL2(Σ)δfL2(Σ)+CδfL2(Σ)||f||_{L^{2}(\partial\Sigma)}\leq\delta||\nabla f||_{L^{2}(\Sigma)}+C_{\delta}||f||_{L^{2}(\Sigma)}, for any fW1,2(Σ)f\in W^{1,2}(\Sigma), we can use this inequality to prove that the infimum is finite. Once this is established the classical arguments to show the existence of a first eigenfunction work.

Since ||f1||=|f1||\nabla|f_{1}||=|\nabla f_{1}| a.e., we have Q(f1,f1)=Q(|f1|,|f1|)Q(f_{1},f_{1})=Q(|f_{1}|,|f_{1}|), that is, |f1||f_{1}| is also a first eigenfunction. Since |f1|0|f_{1}|\geq 0, the maximum principle implies that |f1|>0|f_{1}|>0 in ΣΣ\Sigma\setminus\partial\Sigma, in particular, f1f_{1} does not change sign in ΣΣ\Sigma\setminus\partial\Sigma. Then we can assume that f1>0f_{1}>0 in ΣΣ\Sigma\setminus\partial\Sigma and, by continuity, we get f10f_{1}\geq 0 in Γ.\Gamma. Therefore, we can use f1f_{1} as a test function to our variational problem: Let ζ\zeta be a smooth vector field such that ζ(x)TxSn,\zeta(x)\in T_{x}\operatorname{S}^{n}, for all xSnx\in\operatorname{S}^{n}, ζ(x)(TxΣ),\zeta(x)\in(T_{x}\Sigma)^{\perp}, for all xΣx\in\Sigma, and ζ\zeta points towards WW along Σ\Sigma. Let Φ\Phi be the flow of ζ\zeta. For ε\varepsilon small enough the hypersurfaces Σt={Φ(x,tf1)\Sigma_{t}=\{\Phi\bigl{(}x,tf_{1}\bigr{)}; xΣx\in\Sigma, 0<t<ε}0<t<\varepsilon\} are contained in WW. Since Σ\Sigma has least area among the hypersurfaces Σt\Sigma_{t}, we know that

0d2dt2|t=0+|Σt|=Σ(|Σf1|2|AΣ|2f12)dnΓf12dn1,0\leq\frac{d^{2}}{dt^{2}}\biggl{|}_{t=0^{+}}|\Sigma_{t}|=\int_{\Sigma}(|\nabla_{\Sigma}f_{1}|^{2}-|A_{\Sigma}|^{2}f_{1}^{2})\,d\mathcal{H}^{n}-\int_{\Gamma}f_{1}^{2}d\mathcal{H}^{n-1},

which implies that Q(f1,f1)0Q(f_{1},f_{1})\geq 0. Since f1f_{1} is a first eigenfunction, we get that Q(f,f)0Q(f,f)\geq 0 for any fCc(Σ)f\in C_{c}^{\infty}(\Sigma) with f|Υ0f|_{\Upsilon}\equiv 0. Therefore, we have stability for Σ\Sigma.

Then, since Υ\Upsilon is contained in an equatorial disk DnD^{n}, Lemma 1 implies that Σ\Sigma is necessarily a half nn-dimensional equatorial disk. If ΣW\Sigma\subset\partial W, then we already conclude that WW has to be a (n+1)(n+1)-dimensional halfball.

Suppose ΣWΥ\Sigma\cap\partial W\subset\Upsilon. Rotate Σ\Sigma around Υ\Upsilon until the last time it remains in WW (this last time exists once ΣW\Sigma\cap\partial W is nonempty), and let us still denote this rotated hypersurface by Σ\Sigma. In particular, there exists a point pp where Σ\Sigma and W\partial W are tangent. We will conclude that WW is necessarily a (n+1)(n+1)-dimensional halfball.

In fact, if pint(Υ),p\in\mbox{int}(\Upsilon), we can write W\partial W locally as a graph over Σ\Sigma around pp and apply the classical Hopf Lemma; if pΥ,p\in\partial\Upsilon, we can use the Serrin’s Maximum Principle at a corner (see Appendix A in [25] for the details); and if pΣΥp\in\Sigma\setminus\Upsilon we can apply (the interior or the free boundary version of) the maximum principle. In any case, we get that WW is a (n+1)(n+1)-dimensional halfball. ∎

Now we prove the two-piece property for free boundary minimal hypersurfaces in Bn+1\operatorname{B}^{n+1}.

Theorem 3.

Let MM be a compact embedded smooth free boundary minimal hypersurface in Bn+1.\operatorname{B}^{n+1}. Then for any equatorial disk DnD^{n}, MB+M\cap\operatorname{B}^{+} and MBM\cap\operatorname{B}^{-} are connected.

Proof.

If MM is an equatorial disk, then the result is trivial. So let us assume this is not the case.

Suppose that, for some equatorial disk DnD^{n}, MB+M\cap\operatorname{B}^{+} is a disjoint union of two nonempty open hypersurfaces M1M_{1} and M2M_{2}, M1M_{1} being connected. Notice that by Proposition 1(i) both M1¯\overline{M_{1}} and (all components of) M2¯\overline{M_{2}} have nonempty fixed boundary and nonempty free boundary. Let us denote by ΓI=M1¯Dn\Gamma_{I}=\partial\overline{M_{1}}\cap D^{n} the fixed boundary of M1¯\overline{M_{1}} which might be disconnected. If M1M_{1} and DnD^{n} are transverse, then ΓI\Gamma_{I} is an embedded smooth submanifold of dimension n1n-1. If M1M_{1} and DnD^{n} are tangent, then the local description of nodal sets of elliptic PDE’s (see for instance [19]) imply that int(ΓI)\mathrm{int}(\Gamma_{I}) is an embedded smooth submanifold of dimension n1n-1, away from a singular set 𝒮0\mathcal{S}_{0} such that n1(𝒮0)=0\mathcal{H}^{n-1}(\mathcal{S}_{0})=0.

Denote by WW and WW^{\prime} the closures of the two components of Bn+1M.\operatorname{B}^{n+1}\setminus M. They are compact domains with mean convex boundary. Hence, we can minimize area for the following partially free boundary problem (see Section 2.3):

We consider the class of admissible currents

={T𝒟n(n+1);Tis integer multiplicity rectifiable,\displaystyle\mathfrak{C}=\{T\in\mathcal{D}_{n}(\mathbb{R}^{n+1});\ T\ \mbox{is integer multiplicity rectifiable},
sptTWand is compact,andspt(M1¯T)S2W},\displaystyle\operatorname{spt}T\subset W\ \mbox{and is compact},\ \mbox{and}\ \operatorname{spt}\bigl{(}\llbracket\partial\overline{M_{1}}\rrbracket-\partial T\bigr{)}\subset\operatorname{S}^{2}\cap W\},

where M1¯\llbracket\partial\overline{M_{1}}\rrbracket is the current associated to M1¯\partial\overline{M_{1}} with multiplicity one, and we minimize area (mass) in .\mathfrak{C}. Then, by the results presented in Section 2.3, we get a compact embedded (orientable) partially free boundary minimal hypersurface ΣW\Sigma\subset W which minimizes area among compact hypersurfaces in WW with the same fixed boundary as M1¯\overline{M_{1}}, which is contained in Dn.D^{n}. Moreover, by Proposition 2 in Section 2.3, either ΣW\Sigma\subset\partial W or ΣWΣ.\Sigma\cap\partial W\subset\partial\Sigma.

Arguing as in Claim 2 of Theorem 2, we can prove the stability of Σ\Sigma. Also, observe that by Theorem 1 the singular set 𝒮1\mathcal{S}_{1} of ΣD\Sigma\setminus D is empty or satisfies n7+δ(𝒮1)=0,δ>0\mathcal{H}^{n-7+\delta}(\mathcal{S}_{1})=0,\,\forall\,\delta>0, in particular n2(𝒮1)=0\mathcal{H}^{n-2}(\mathcal{S}_{1})=0. So, we can apply Lemma 1 and conclude that each component of Σ\Sigma is a piece of an equatorial disk.

The case ΣW\Sigma\subset\partial W can not happen because this would imply that MM is a disk, and we are assuming it is not. Therefore, only the second case can happen, that is, any component of Σ\Sigma meets W\partial W only at points of Σ.\partial\Sigma. Observe that each component of Σ\Sigma that is not bounded by a (n1)(n-1)-dimensional equatorial disk is necessarily contained in DnD^{n}. If some component of Σ\Sigma were bounded by a (n1)(n-1)-dimensional equatorial disk, then we could apply Theorem 2 and would conclude that MM is a nn-dimensional equatorial disk, which is not the case. Then Σ\Sigma is entirely contained in DnD^{n} and, since ΣWΣ\Sigma\cap\partial W\subset\partial\Sigma, MWM\subset\partial W and MDnM\cap D^{n} does not contain any (n1)(n-1)-dimensional piece of Dn\partial D^{n} (Proposition 1(ii)), we have ΣM=ΓI\Sigma\cap M=\Gamma_{I}.

Doing the same procedure as in the last paragraph for WW^{\prime}, we can construct another compact hypersurface Σ\Sigma^{\prime} of DnD^{n} with fixed boundary Σ=ΓI\partial\Sigma^{\prime}=\Gamma_{I} and such that ΣW\Sigma^{\prime}\subset W^{\prime} and ΣM=ΓI\Sigma^{\prime}\cap M=\Gamma_{I}. Notice that ΣΣ\Sigma\cup\Sigma^{\prime} is a hypersurface without fixed boundary of DnD^{n}, therefore ΣΣ=Dn.\Sigma\cup\Sigma^{\prime}=D^{n}. In fact, let us denote by TT and TT^{\prime} the minimizing currents associated to Σ\Sigma and Σ\Sigma^{\prime} respectively, that is, sptT=Σ\operatorname{spt}T=\Sigma and sptT=Σ\operatorname{spt}T^{\prime}=\Sigma^{\prime}. First observe that spt(TT)Dn\operatorname{spt}\partial(T-T^{\prime})\subset\partial D^{n} and (TT)=0\partial\partial(T-T^{\prime})=0; hence, by the Constancy Theorem, we know that (TT)=kDn\partial(T-T^{\prime})=k\partial D^{n}, for some interger kk. Now, since spt(TTkDn)Dn\operatorname{spt}(T-T^{\prime}-kD^{n})\subset D^{n} and (TTkDn)=0,\partial(T-T^{\prime}-kD^{n})=0, the Constancy Theorem implies that TT=kDnT-T^{\prime}=kD^{n}; but since ΓI\Gamma_{I} has multiplicity one, this also holds for TT and TT^{\prime} and therefore k=1k=1 necessarily. Hence, ΣΣ=spt(TT)=Dn\Sigma\cup\Sigma^{\prime}=\operatorname{spt}(T-T^{\prime})=D^{n}.

In particular, MDn=ΓI,M\cap D^{n}=\Gamma_{I}, which implies that M2=MB+M1M_{2}=M\cap\operatorname{B}^{+}\setminus M_{1} has fixed boundary contained in ΓI\Gamma_{I}. For n=2n=2, since MM is embedded and ΓI\Gamma_{I} has singularities of nn-prong type (if any), we know that the fixed boundaries of M1M_{1} and M2M_{2} are disjoint, in particular, the fixed boundary of M2¯\overline{M_{2}} is necessarily empty and this yields a contradiction by Proposition 1(i). It remains to analyse the case when n3.n\geq 3.

Let us assume, without loss of generality, that Dn=B{xn+1=0}D^{n}=\operatorname{B}\cap\{x_{n+1}=0\}; hence, we have MDn=ΓI={qM;xn+1(q)=0}M\cap D^{n}=\Gamma_{I}=\{q\in M;x_{n+1}(q)=0\} which is the nodal set of the Steklov eigenfunction xn+1:M.x_{n+1}:M\to\operatorname{\mathbb{R}}.

Observe that if qΓIq\in\Gamma_{I} and Mxn+1(q)0\nabla_{M}x_{n+1}(q)\neq 0 then, since MM is embedded, we know that in a neighborhood of qq we have MDn=M1¯DnM\cap D^{n}=\overline{M_{1}}\cap D^{n}; in particular, qq can not be contained in M2¯.\partial\overline{M_{2}}.

Now let us analyse the singular set 𝒮={xn+1=0}{Mxn+1=0}ΓI.\mathcal{S}=\{x_{n+1}=0\}\cap\{\nabla_{M}x_{n+1}=0\}\subset\Gamma_{I}. By Theorem 1.7 in [19], the Hausdorff dimension of 𝒮\mathcal{S} is less than or equal to n2;n-2; in particular, n1(𝒮)=0\mathcal{H}^{n-1}(\mathcal{S})=0 and therefore by Lemma 2 M2¯\overline{M_{2}} is stationary. By [30] we can conclude that either M2¯Dn=\overline{M_{2}}\cap D^{n}=\emptyset or DnM2¯D^{n}\subset\overline{M_{2}} (which we already know is not possible). Therefore, M2¯\overline{M_{2}} has empty fixed boundary which is a contradiction by Proposition 1(i).

Therefore, the theorem is proved. ∎

References

  • [1] L. Ambrozio, A. Carlotto and B. Sharp, Index estimates for free boundary minimal hypersurfaces, Math. Ann. 370 (2018), 1063–1078.
  • [2] L. Ambrozio, A. Carlotto and B. Sharp, Compactness analysis for free boundary minimal hypersurfaces, Calc. Var. Partial Differential Equations 57 (2018), no. 1, Art. 22, 39 pp.
  • [3] W. K. Allard, On the first variation of a varifold, Ann. of Math. 95 (1972), no. 3, 417–491.
  • [4] W. K. Allard, On the first variation of a varifold: boundary behavior, Ann. of Math. 101 (1975), no. 3, 418–446.
  • [5] T. Bourni, Allard-type boundary regularity for C1,αC^{1,\alpha} boundaries, Adv. Calc. Var. 9 (2016), no. 2, 143–161.
  • [6] F. Duzaar and K. Steffan, Optimal interior and boundary regularity for almost minimizers to elliptic variational integrals, J. Reine Angew. Math. 546 (2002), 73–138.
  • [7] N. Edelen, A note on the singular set of area-minimizing hypersurfaces, Calc. Var. Partial Differential Equations 59 (2020), no. 1, Art. 18, 9 pp.
  • [8] H. Federer, Geometric measure theory, Grundlehren der math. Wiss. 153. Springer, New York 1969.
  • [9] H. 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 (1970), 767–771.
  • [10] A. Fraser and M. Li, Compactness of the space of embedded minimal surfaces with free boundary in three-manifolds with nonnegative Ricci curvature and convex boundary, J. Differential Geom. 96 (2014), 183–200.
  • [11] A. Fraser and R. Schoen, The first Steklov eigenvalue, conformal geometry, and minimal surfaces, Adv. Math. 226 (2011), no. 5, 4011–4030.
  • [12] A. Fraser and R. Schoen, Minimal surfaces and eigenvalue problems, Contemp. Math. 599 (2013), 105–121.
  • [13] A. Fraser and R. Schoen, Sharp eigenvalue bounds and minimal surfaces in the ball, Invent. Math. 203 (2016), no. 3, 823–890.
  • [14] B. Freidin, M. Gulian and P. McGrath, Free boundary minimal surfaces in the unit ball with low cohomogeneity, Proc. Amer. Math. Soc. 145 (2017), no. 4, 1671–1683.
  • [15] M. Grüter, Regularitt von minimierenden Strömen bei einer freien Randbedingung, Universitt Düsseldorf, 1985.
  • [16] M. Grüter, Optimal regularity for codimension one minimal surfaces with a free boundary, Manuscripta Math. 58 (1987), no. 3, 295–343.
  • [17] Q. Guang, X. Zhou. Compactness and generic finiteness for free boundary minimal hypersurfaces. Pacific J. Math. 310 (2021), no. 1, 85–114.
  • [18] R. Hardt and L. Simon, Boundary regularity and embedded solutions for the oriented Plateau problem, Ann. of Math. (2) 110 (1979), no. 3, 439–486.
  • [19] R. Hardt and L. Simon, Nodal sets for solutions of elliptic equations, J. Differential Geom. 30 (1989), 505–522.
  • [20] T. Ilmanen, A strong maximum principle for singular minimal hypersurfaces, Calc. Var. 4 (1996), 443–467.
  • [21] R. Kusner and P. McGrath, On free boundary minimal annuli embedded in the unit ball, Preprint at arXiv:2011.06884 (2020).
  • [22] M. Li, Free boundary minimal surfaces in the unit ball: recent advances and open questions, Preprint at arXiv:arXiv:1907.05053v3 (2020).
  • [23] M. Li and X. Zhou, Min-max theory for free boundary minimal hypersurfaces I-regularity theory, J. Differential Geom. 118 (2021), no. 3, 487–553.
  • [24] V. Lima, Bounds for the Morse index of free boundary minimal surfaces, arXiv:1710.10971 to appear on Asian Journal of Mathematics.
  • [25] V. Lima and A. Menezes, A two-piece property for free boundary minimal surfaces in the ball, Trans. Amer. Math. Soc., 374 (2021), no. 3, 1661–1686.
  • [26] A. Ros, A two-piece property for compact minimal surfaces in a three-sphere, Indiana Univ. Math. J. 44 (1995), no. 3, 841–849.
  • [27] L. Simon, Lectures on geometric measure theory, Proceedings of the Centre for Mathematical Analysis, Australian National University, vol. 3, Australian National University, Centre for Mathematical Analysis, Canberra, 1983. vii+272 pp.
  • [28] G. Smith, A. Stern, H. Tran, D. Zhou, On the Morse index of higher-dimensional free boundary minimal catenoids. Calc. Var. Partial Differential Equations 60 (2021), no. 6, Paper No. 208.
  • [29] B. Solomon, Harmonic maps to spheres, J. Differential Geom. 21 (1985), 151–162.
  • [30] B. Solomon and B. White, A strong maximum principle for varifolds that are stationary with respect to even parametric elliptic functionals, Indiana Univ. Math. J. 38 (1989), no. 3, 683– 691.
  • [31] A. Sun, Z. Wang and X. Zhou, Multiplicity one for min-max theory in compact manifolds with boundary and its applications. arXiv:2011.04136 [math.DG]
  • [32] Z. Wang, Existence of infinitely many free boundary minimal hypersurfaces, Preprint at arXiv:2001.04674 (2020).