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

\UseRawInputEncoding

Convex Cylinders and the Symmetric Gaussian Isoperimetric Problem

Steven Heilman Department of Mathematics, University of Southern California, Los Angeles, CA 90089-2532 stevenmheilman@gmail.com
Abstract.

Let Ω\Omega be a measurable Euclidean set in n\mathbb{R}^{n} that is symmetric, i.e. Ω=Ω\Omega=-\Omega, such that Ω×\Omega\times\mathbb{R} has the smallest Gaussian surface area among all measurable symmetric sets of fixed Gaussian volume. We conclude that either Ω\Omega or Ωc\Omega^{c} is convex. Moreover, except for the case H(x)=x,N(x)+λH(x)=\langle x,N(x)\rangle+\lambda with H0H\geq 0 and λ<0\lambda<0, we show there exist a radius r>0r>0 and an integer 0kn10\leq k\leq n-1 such that after applying a rotation, the boundary of Ω\Omega must satisfy Ω=rSk×nk1\partial\Omega=rS^{k}\times\mathbb{R}^{n-k-1}, with n1rn+1\sqrt{n-1}\leq r\leq\sqrt{n+1} when k1k\geq 1. Here SkS^{k} denotes the unit sphere of k+1\mathbb{R}^{k+1} centered at the origin, and n1n\geq 1 is an integer. One might say this result nearly resolves the symmetric Gaussian conjecture of Barthe from 2001.

Key words and phrases:
symmetric, Gaussian, minimal surface, calculus of variations
2010 Mathematics Subject Classification:
60E15, 49J40, 53A10, 60G15
Supported by NSF Grant CCF 1911216.

1. Introduction

In [Bar01], Barthe asked: which symmetric measurable Euclidean set of fixed Gaussian measure minimizes its Gaussian surface area? A subset Ωn+1\Omega\subseteq\mathbb{R}^{n+1} is symmetric if Ω=Ω\Omega=-\Omega. It is well-known that a half space minimizes Gaussian surface area, among Euclidean sets of fixed Gaussian measure [Bor75, SC74]. Despite many proofs of this so-called Gaussian isoperimetric inequality [Bor85, Led94, Led96, Bob97, BS01, Bor03, MN15a, MN15b, Eld15, MR15, BBJ17], additionally restricting to symmetric sets causes additional difficulties. Indeed, not much progress seemed to be made on Barthe’s question from 2001 until [Hei21], where various candidate optimal sets were ruled out, following [Man17], where the local minimality and non-minimality of balls centered at the origin was demonstrated. Also, as suggested by Morgan [Hei21, Conjecture 1.14], the answer to Barthe’s question should depend on the Gaussian volume restriction. For example, among symmetric Euclidean sets with Gaussian measure very close to 11 or 0, a one-dimensional slab (or its complement) should minimize Gaussian surface area. This fact was demonstrated in [BJ20], proving the first known case of Barthe’s question. The proof of [BJ20] considers minimizing the Gaussian surface area plus a “penalty term,” penalizing sets that have a nonzero Gaussian center of mass. Considerable effort is then required to show that minimizers of this functional are symmetric, and they are slabs (or complements of slabs), for an appropriately chosen constant in front of the penalty term.

In this paper, we demonstrate that convex cylinders centered at the origin are the only sets that are stable for the Gaussian surface area, after taking a product with \mathbb{R}. Our approach is rather different than that of [BJ20]. We begin by adapting the Colding-Minicozzi theory of entropy from [CM12]. In [CM12], it is shown that for self-shrinkers, i.e. surfaces Σ\Sigma satisfying H(x)x,N(x)=0H(x)-\langle x,N(x)\rangle=0 for all xΣx\in\Sigma, HH is an eigenfunction of the second variation operator LL. (Here H(x)H(x) denotes the mean curvature of Σ\Sigma at xx, i.e. the divergence of the exterior pointing unit normal vector N(x)N(x) at xx. Also LL is defined in (11).) Namely, LH=2HLH=2H (see (14)). This explicit eigenfunction with eigenvalue 22 crucially allows a stability analysis of self-shrinkers with respect to the Colding-Minicozzi entropy functional. The Colding-Minicozzi entropy of a surface Σ\Sigma is the supremum over translations and dilations of the Gaussian surface area. Critical points of this functional are self-shrinkers [CM12], hence the focus of [CM12] on these surfaces.

1.1. Adapting the Colding-Minicozzi Theory

In this paper, we are concerned with a stability analysis of the Gaussian surface area. Critical points of the Gaussian surface area functional satisfy the more general condition that there exists some λ\lambda\in\mathbb{R} such that H(x)x,N(x)=λH(x)-\langle x,N(x)\rangle=\lambda for all xΣx\in\Sigma. In this setting, the mean curvature HH is no longer an eigenfunction of the second variation operator LL (it is an almost eigenfunction though; see (14).) So, one cannot directly use the stability analysis of [CM12] for the Gaussian surface area itself. However, appropriate modifications of the arguments of [CM12] can be made successful in certain cases.

Our general strategy is to show that the second variation operator has an eigenvalue larger than 22. Finding such an eigenvalue is insufficient to classify symmetric sets with minimal Gaussian surface area. However, if we take a product of an eigenfunction multiplied by a function xn+121x_{n+1}^{2}-1 in an orthogonal direction, as an input to the second variation formula, then we obtain a Gaussian volume preserving and Gaussian surface area decreasing perturbation of Ω×\Omega\times\mathbb{R} (see Lemma 5.4). So, the main task is to look for (approximate) eigenfunctions of LL with eigenvalue larger than 22. To accomplish this task, it seems necessary to split into a few different cases.

In [CM12], the stability of the entropy is split into two cases, according to whether or not the surface Σ\Sigma is mean convex (i.e. if HH changes sign on Σ\Sigma). In the case that HH changes sign on Σ\Sigma, since HH itself is an eigenfunction of LL with eigenvalue 22, another eigenfunction of LL with a larger eigenvalue must exist. In our more general setting, since HH is no longer an eigenfunction of LL, the observation that there exists another eigenfunction with a larger eigenvalue of [CM12] no longer applies. So, we instead use a product of max(H,0)\max(H,0) multiplied by a function xn+121x_{n+1}^{2}-1 in an orthogonal direction, as an input to the second variation formula.

In the case that HH does not change sign and H(x)x,N(x)=0H(x)-\langle x,N(x)\rangle=0 on Σ\Sigma, a curvature bound of the second fundamental form is proven in [CM12] that allows other eigenfunctions of LL to be used in the second variation formula, and no curvature bound needs to be proven. In our more general setting where H(x)x,N(x)=λH(x)-\langle x,N(x)\rangle=\lambda, this curvature bound seems difficult to prove in general, so that some functions might not be usable in the second variation formula. Fortunately, we can split into two sub-cases. In the case that HH and λ\lambda have the same sign, we can use HH itself in the second variation formula. In the case that HH and λ\lambda have opposite signs, the curvature bound of [CM12] can be proven, allowing us to use a function of the second fundamental form AA in the second variation formula. However, in this case, the second variation of HH itself is not helpful, so we needed to come up with another function to input into the second variation formula, namely the largest eigenvalue of AA.

It turns out that the largest eigenvalue α\alpha of the second fundamental form AA is an almost eigenfunction of LL (see (13)). (To avoid issues with differentiability and integration by parts, we actually use a smoothed version of the largest eigenvalue of AA; see Lemma 5.3.) So, max(α,0)\max(\alpha,0) can be used in our second variation formula as long as α>0\alpha>0 on a set of positive measure on Σ\Sigma. If no such set exists where α\alpha is positive, then all eigenvalues of AA are negative everywhere, so that Σ\Sigma is the boundary of a convex set.

In the case that the mean curvature HH does not change sign, there are two sub-cases to consider. If H0H\geq 0 and λ>0\lambda>0, then we can either use HH itself in the second variation formula, or use a Huisken-type classification from [Hei21] (which does not use any second variation computations). However, if H0H\geq 0 and λ<0\lambda<0, then we are only able to deduce convexity of Ω\Omega or Ωc\Omega^{c}. In fact, no Huisken-type classification can occur in this case. We will discuss this case further in Section 1.2.

There is still one final case we have not mentioned, namely H(x)=x,N(x)=0H(x)=\langle x,N(x)\rangle=0 on Σ\Sigma, i.e. that Σ\Sigma is a Gaussian minimal cone. In this last case, these sets cannot minimize Barthe’s problem. This follows by adapting an argument of [Zhu20], which itself adapted an argument of Simons [Sim68]. We use a perturbation of the surface that is a product of a radial and angular component. The radial component is chosen to preserve Gaussian volume, and the angular component is chosen to have a large eigenvalue of the second variation operator.

The Colding-Minicozzi theory [CM12, CIMW13] was originally designed to use the Gaussian surface area to investigate singularities of mean-curvature flows. A connection of Gaussian surface area to mean-curvature flow was established by Huisken [Hui90, Hui93], and [CM12] greatly extended this connection. As a continuation of [Hei21], this paper instead applies the Colding-Minicozzi theory to a Gaussian isoperimetric conjecture.

1.2. The case H0H\geq 0 and λ<0\lambda<0

As mentioned above, in the case H0H\geq 0 and λ<0\lambda<0, we can only conclude that Ω\Omega or Ωc\Omega^{c} is convex, unlike in other cases where we can show that Ω\partial\Omega is a round cylinder centered at the origin. The compact version of this convexity statement was proven in [Lee22], though compactness was crucially used there, so it is unclear if the argument there generalizes to the noncompact case. Part of the difficulty of this case is that Huisken’s classification no longer holds [Hui90, Hui93]. Indeed, it is known that, for every integer m2m\geq 2, there exists λ=λm<0\lambda=\lambda_{m}<0 and there exists a convex embedded curve Γm2\Gamma_{m}\subseteq\mathbb{R}^{2} satisfying H(x)=x,N(x)+λH(x)=\langle x,N(x)\rangle+\lambda as in Lemma 4.1, and such that Γm\Gamma_{m} is symmetric with respect to a rotation by an angle 2π/m2\pi/m (and Γm1Γm2\Gamma_{m_{1}}\neq\Gamma_{m_{2}} if m1m2m_{1}\neq m_{2}, and also Γm\Gamma_{m} is not symmetric with respect to a rotation by an angle smaller than 2π/m2\pi/m) [Cha17, Theorem 1.2, Theorem 1.3, Proposition 3.2]. Consequently, Γm×n2n+1\Gamma_{m}\times\mathbb{R}^{n-2}\subseteq\mathbb{R}^{n+1} also satisfies H(x)=x,N(x)+λH(x)=\langle x,N(x)\rangle+\lambda. So, Huisken’s classification cannot possibly hold, at least when λ<0\lambda<0.

In the case of even m6m\geq 6, the curve Γm\Gamma_{m} can be shown to be unstable by [Hei21, Corollary 11.9], since there are more than four nodal domains corresponding to an infinitesimal rotation of Γm\Gamma_{m}. However, it is unclear how to prove instability for the cases m=2m=2 and m=4m=4. As shown at the end of [Cha17], there appears to be an infinite family of curves with a single symmetry by a rotation of an angle π\pi, and it is unclear if any of our arguments can show these curves are unstable for the Gaussian perimeter.

1.3. Future Directions and Remaining Cases of Barthe’s Problem

The ultimate goal of Barthe’s Problem 2.1 is to find the minimum Gaussian perimeter of a set in n+1\mathbb{R}^{n+1}, where we take the infimum over all dimensions n0n\geq 0. That is, we would like to determine the isoperimetric profile II_{\infty} depicted in Figure 2.

Theorem 2.4 classifies those sets of the form Ω×\Omega\times\mathbb{R} that are stable for the Gaussian surface area. If a set Ω\Omega minimizes Problem 2.1 and its surface area is achieved in the definition of II_{\infty}, then the set Ω×\Omega\times\mathbb{R} must also be stable. However, a priori, it could occur that for each n0n\geq 0, there exists a set Ωnn+1\Omega_{n}\subseteq\mathbb{R}^{n+1} that minimizes Problem 2.1 for a measure constraint γn+1(Ωn)=c(0,1)\gamma_{n+1}(\Omega_{n})=c\in(0,1), but such that Ωn×\Omega_{n}\times\mathbb{R} is unstable. Put another way, it could occur that a value I(c)I_{\infty}(c) is not achieved for any set of finite dimension. In such a case, Theorem 2.4 does not say anything about Problem 2.1. And in fact, we do expect this to happen, but only when c=1/2c=1/2. It is conjectured that, for any c1/2c\neq 1/2, I(c)I_{\infty}(c) is achieved by a set of finite dimension. And in the case c=1/2c=1/2, I(c)I_{\infty}(c) is achieved by a limit of spheres SnS^{n} of radius approximately n\sqrt{n}. With this picture in mind, Theorem 2.4 should cover many cases of Problem 2.1 (except when c=1/2c=1/2). It was conjectured by Morgan that, if Ωn+1\Omega\subseteq\mathbb{R}^{n+1} minimizes Problem 2.1, there exists r>0r>0 and there exists 0kn0\leq k\leq n such that

Ω=rSk×nk.\partial\Omega=rS^{k}\times\mathbb{R}^{n-k}.

We additionally conjecture that rr satisfies nrn+2\sqrt{n}\leq r\leq\sqrt{n+2} when k1k\geq 1.

Refer to caption
Figure 1. Gaussian surface area of B(0,s)B(0,s)\subseteq\mathbb{R} and of B(0,r)n+1B(0,r)\subseteq\mathbb{R}^{n+1}, where nrn+2\sqrt{n}\leq r\leq\sqrt{n+2} for each 1n41\leq n\leq 4, together with their complements.

2. Statement of Results

We begin by defining the Gaussian density:

γn(x)\colonequals(2π)n/2ex2/2,x2\colonequalsi=1mxi2,x=(x1,,xm)m.\gamma_{n}(x)\colonequals(2\pi)^{-n/2}e^{-\left\|x\right\|^{2}/2},\qquad\left\|x\right\|^{2}\colonequals\sum_{i=1}^{m}x_{i}^{2},\qquad\forall\,x=(x_{1},\ldots,x_{m})\in\mathbb{R}^{m}.

We also denote γn+1()\gamma_{n+1}(\cdot) as a measure of Lebesgue measurable sets in n+1\mathbb{R}^{n+1}. For a set Σn+1\Sigma\subseteq\mathbb{R}^{n+1} with Hausdorff dimension nn, we denote its Gaussian surface area as

Σγn(x)dx\colonequalslim infε0+12ε{xn+1:yΣ,xy<ε}γn(x)dx.\int_{\Sigma}\gamma_{n}(x)\,\mathrm{d}x\colonequals\liminf_{\varepsilon\to 0^{+}}\frac{1}{2\varepsilon}\int_{\{x\in\mathbb{R}^{n+1}\colon\exists\,y\in\Sigma,\,\left\|x-y\right\|<\varepsilon\}}\gamma_{n}(x)\,\mathrm{d}x.

Our main problem of interest is the following.

Problem 2.1 (Symmetric Gaussian Problem, [Bar01]).

Fix 0<c<10<c<1. Minimize

Ωγn(x)dx\int_{\partial\Omega}\gamma_{n}(x)\,\mathrm{d}x

over all (measurable) subsets Ωn+1\Omega\subseteq\mathbb{R}^{n+1} satisfying Ω=Ω\Omega=-\Omega and γn+1(Ω)=c\gamma_{n+1}(\Omega)=c.

Unless otherwise stated, all sets discussed in this paper will be Lebesgue measurable.

Remark 2.2.

If Ω\Omega minimizes Problem 2.1, then Ωc\Omega^{c} also minimizes Problem 2.1, with cc replaced by 1c1-c.

Definition 2.3.

For any integer k0k\geq 0, define the kk-dimensional sphere of radius one centered at the origin to be

Sk\colonequals{xk+1:x=1}.S^{k}\colonequals\{x\in\mathbb{R}^{k+1}\colon\left\|x\right\|=1\}.
Refer to caption
11
1/21/2
Gaussian volume cc
Gaussian
surface
area
I(c)I_{\infty}(c)
S0S^{0}
S1S^{1}
S2S^{2}
S3S^{3}
Figure 2. The conjectured form of the symmetric Gaussian isoperimetric profile.
I(c)\colonequalsinfn0infΩn+1Ω=Ωγn+1(Ω)=cΩγn(x)dxI_{\infty}(c)\colonequals\inf_{n\geq 0}\inf_{\begin{subarray}{c}\Omega\subseteq\mathbb{R}^{n+1}\\ \Omega=-\Omega\\ \gamma_{n+1}(\Omega)=c\end{subarray}}\int_{\partial\Omega}\gamma_{n}(x)\,\mathrm{d}x
Theorem 2.4 (Main Theorem).

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1} be a measurable set. Assume that Ω×\Omega\times\mathbb{R} minimizes minimizing Problem 2.1. Then Ω\Omega or Ωc\Omega^{c} is convex.

Moreover, unless H0H\geq 0 and λ<0\lambda<0 (or H0H\leq 0 and λ>0\lambda>0), \exists r>0r>0, \exists 0kn0\leq k\leq n such that

Ω=rSk×nk,\partial\Omega=rS^{k}\times\mathbb{R}^{n-k},

after rotating Ω\Omega if necessary. And if k1k\geq 1, then nrn+2\sqrt{n}\leq r\leq\sqrt{n+2}.

Theorem 2.4 will be split into several cases, resulting from the combination of Theorems 6.1, 8.5, 10.1 and 7.1.

It is shown in [Man17] that the ball centered at the origin of n+1\mathbb{R}^{n+1} is a local minimum of Problem 2.1, when the ball’s radius rr satisfies r<n+2r<\sqrt{n+2}. (Consequently, the complement of a ball centered at the origin is a local minimum of Problem 2.1, when the ball’s radius rr satisfies r<n+2r<\sqrt{n+2}, by Remark 2.2.) It is also shown in [Man17] that a ball of radius r>n+2r>\sqrt{n+2} is not a local minimum of Problem 2.1. Our results (e.g. Case 2 of Theorem 7.1) show, for a ball B(0,r)n+1B(0,r)\subseteq\mathbb{R}^{n+1} of radius rr centered at the origin, B(0,r)×B(0,r)\times\mathbb{R} does not minimize problem 2.1 when rnr\leq\sqrt{n}. So, if n1n\geq 1, the Gaussian surface area of B(0,r)n+1B(0,r)\subseteq\mathbb{R}^{n+1} should only appear as a value in Figure 2 when nrn+2\sqrt{n}\leq r\leq\sqrt{n+2}.

Remark 2.5.

As shown e.g. in the introduction of [Hei21], n+1Sn\sqrt{n+1}S^{n} has asymptotic Gaussian surface area 2\sqrt{2}.

limn{xn+1:x=n+1}γn(x)dx=2.\lim_{n\to\infty}\int_{\{x\in\mathbb{R}^{n+1}\colon\left\|x\right\|=\sqrt{n+1}\}}\gamma_{n}(x)\,\mathrm{d}x=\sqrt{2}.

In contrast, a half space with Gaussian measure 1/21/2 has Gaussian surface area γ0(0)=1\gamma_{0}(0)=1.

Remark 2.6.

For any k1k\geq 1, denote ck\colonequals{xk+1:x=k}γk(x)dxc_{k}\colonequals\int_{\{x\in\mathbb{R}^{k+1}\colon\left\|x\right\|=\sqrt{k}\}}\gamma_{k}(x)\,\mathrm{d}x. As shown in [Sto94, Lemma A.4], this sequence is decreasing: c1>c2>c3>c_{1}>c_{2}>c_{3}>\cdots.

3. Preliminaries

We say that Σn+1\Sigma\subseteq\mathbb{R}^{n+1} is an nn-dimensional CC^{\infty} manifold if Σ\Sigma can be locally written as the graph of a CC^{\infty} function. For any (n+1)(n+1)-dimensional CC^{\infty} manifold Ωn+1\Omega\subseteq\mathbb{R}^{n+1} with boundary, we denote

C0(Ω;n+1)\displaystyle C_{0}^{\infty}(\Omega;\mathbb{R}^{n+1}) \colonequals{f:Ωn+1:fC(Ω;n+1),f(Ω)=0,\displaystyle\colonequals\{f\colon\Omega\to\mathbb{R}^{n+1}\colon f\in C^{\infty}(\Omega;\mathbb{R}^{n+1}),\,f(\partial\Omega)=0, (1)
r>0,f(Ω(B(0,r))c)=0}.\displaystyle\qquad\qquad\qquad\exists\,r>0,\,f(\Omega\cap(B(0,r))^{c})=0\}.

We also denote C0(Ω)\colonequalsC0(Ω;)C_{0}^{\infty}(\Omega)\colonequals C_{0}^{\infty}(\Omega;\mathbb{R}). We let div\mathrm{div} denote the divergence of a vector field in n+1\mathbb{R}^{n+1}. For any r>0r>0 and for any xn+1x\in\mathbb{R}^{n+1}, we let B(x,r)\colonequals{yn+1:xyr}B(x,r)\colonequals\{y\in\mathbb{R}^{n+1}\colon\left\|x-y\right\|\leq r\} be the closed Euclidean ball of radius rr centered at xn+1x\in\mathbb{R}^{n+1}.

Definition 3.1 (Reduced Boundary).

A measurable set Ωn+1\Omega\subseteq\mathbb{R}^{n+1} has locally finite surface area if, for any r>0r>0,

sup{Ωdiv(X(x))dx:XC0(B(0,r),n+1),supxn+1X(x)1}<.\sup\left\{\int_{\Omega}\mathrm{div}(X(x))\,\mathrm{d}x\colon X\in C_{0}^{\infty}(B(0,r),\mathbb{R}^{n+1}),\,\sup_{x\in\mathbb{R}^{n+1}}\left\|X(x)\right\|\leq 1\right\}<\infty.

Equivalently, Ω\Omega has locally finite surface area if 1Ω\nabla 1_{\Omega} is a vector-valued Radon measure such that, for any xn+1x\in\mathbb{R}^{n+1}, the total variation

1Ω(B(x,1))\colonequalssuppartitionsC1,,CmofB(x,1)m1i=1m1Ω(Ci)\left\|\nabla 1_{\Omega}\right\|(B(x,1))\colonequals\sup_{\begin{subarray}{c}\mathrm{partitions}\\ C_{1},\ldots,C_{m}\,\mathrm{of}\,B(x,1)\\ m\geq 1\end{subarray}}\sum_{i=1}^{m}\left\|\nabla 1_{\Omega}(C_{i})\right\|

is finite [CL12].

If Ωn+1\Omega\subseteq\mathbb{R}^{n+1} has locally finite surface area, we define the reduced boundary Ω\partial^{*}\Omega of Ω\Omega to be the set of points xn+1x\in\mathbb{R}^{n+1} such that

N(x)\colonequalslimr0+1Ω(B(x,r))1Ω(B(x,r))N(x)\colonequals-\lim_{r\to 0^{+}}\frac{\nabla 1_{\Omega}(B(x,r))}{\left\|\nabla 1_{\Omega}\right\|(B(x,r))}

exists, and it is exactly one element of SnS^{n}.

For more background on the reduced boundary and its regularity, we refer to the discussion in Section 2 of [BBJ17], [AFP00] and [Mag12]. The following argument is essentially identical to [BBJ17, Proposition 1], so we omit the proof.

Lemma 3.2 (Existence).

There exists a set Ωn+1\Omega\subseteq\mathbb{R}^{n+1} minimizing Problem 2.1.

3.1. Submanifold Curvature

Here we cover some basic definitions from differential geometry of submanifolds of Euclidean space.

Let \nabla denote the standard Euclidean connection, so that if X,YC0(n+1,n+1)X,Y\in C_{0}^{\infty}(\mathbb{R}^{n+1},\mathbb{R}^{n+1}), if Y=(Y1,,Yn+1)Y=(Y_{1},\ldots,Y_{n+1}), and if u1,,un+1u_{1},\ldots,u_{n+1} is the standard basis of n+1\mathbb{R}^{n+1}, then XY\colonequalsi=1n+1(X(Yi))ui\nabla_{X}Y\colonequals\sum_{i=1}^{n+1}(X(Y_{i}))u_{i}. Let NN be the outward pointing unit normal vector of an nn-dimensional hypersurface Σn+1\Sigma\subseteq\mathbb{R}^{n+1}. For any vector xΣx\in\Sigma, we write x=xT+xNx=x^{T}+x^{N}, so that xN\colonequalsx,NNx^{N}\colonequals\langle x,N\rangle N is the normal component of xx, and xTx^{T} is the tangential component of xΣx\in\Sigma. We let Σ\colonequals()T\nabla^{\Sigma}\colonequals(\nabla)^{T} denote the tangential component of the Euclidean connection.

Let e1,,ene_{1},\ldots,e_{n} be an orthonormal frame of Σn+1\Sigma\subseteq\mathbb{R}^{n+1}. That is, for a fixed xΣx\in\Sigma, there exists a neighborhood UU of xx such that e1,,ene_{1},\ldots,e_{n} is an orthonormal basis for the tangent space of Σ\Sigma, for every point in UU [Lee03, Proposition 11.17].

Define the mean curvature

H\colonequalsdiv(N)=i=1neiN,ei.H\colonequals\mathrm{div}(N)=\sum_{i=1}^{n}\langle\nabla_{e_{i}}N,e_{i}\rangle. (2)

Define the second fundamental form A=(aij)1i,jnA=(a_{ij})_{1\leq i,j\leq n} so that

aij=eiej,N, 1i,jn.a_{ij}=\langle\nabla_{e_{i}}e_{j},N\rangle,\qquad\forall\,1\leq i,j\leq n. (3)

Compatibility of the Riemannian metric says aij=eiej,N=ej,eiN+eiN,ej=ej,eiNa_{ij}=\langle\nabla_{e_{i}}e_{j},N\rangle=-\langle e_{j},\nabla_{e_{i}}N\rangle+e_{i}\langle N,e_{j}\rangle=-\langle e_{j},\nabla_{e_{i}}N\rangle, \forall 1i,jn1\leq i,j\leq n. So, multiplying by eje_{j} and summing this equality over jj gives

eiN=j=1naijej, 1in.\nabla_{e_{i}}N=-\sum_{j=1}^{n}a_{ij}e_{j},\qquad\forall\,1\leq i\leq n. (4)

Using NN,N=0\langle\nabla_{N}N,N\rangle=0,

H=(2)i=1neiN,ei=(4)i=1naii.H\stackrel{{\scriptstyle\eqref{three0.5}}}{{=}}\sum_{i=1}^{n}\langle\nabla_{e_{i}}N,e_{i}\rangle\stackrel{{\scriptstyle\eqref{three2}}}{{=}}-\sum_{i=1}^{n}a_{ii}. (5)

3.2. First and Second Variation

We will apply the calculus of variations to solve Problem 2.1. Here we present the rudiments of the calculus of variations.

The results of this section are well known to experts in the calculus of variations, and many of these results were re-proven in [BBJ17].

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1} be an (n+1)(n+1)-dimensional C2C^{2} submanifold with reduced boundary Σ\colonequalsΩ\Sigma\colonequals\partial^{*}\Omega. Let N:ΩSnN\colon\partial^{*}\Omega\to S^{n} denote the unit exterior normal to Ω\partial^{*}\Omega. Let X:n+1n+1X\colon\mathbb{R}^{n+1}\to\mathbb{R}^{n+1} be a vector field. Unless otherwise stated, we assume that X(x)X(x) is parallel to N(x)N(x) for all xΩx\in\partial^{*}\Omega, i.e.

X(x)=X(x),N(x)N(x),xΩ.X(x)=\langle X(x),N(x)\rangle N(x),\qquad\forall\,x\in\partial^{*}\Omega. (6)

Let div\mathrm{div} denote the divergence of a vector field. We write XX in its components as X=(X1,,Xn+1)X=(X_{1},\ldots,X_{n+1}), so that divX=i=1n+1xiXi\mathrm{div}X=\sum_{i=1}^{n+1}\frac{\partial}{\partial x_{i}}X_{i}. Let Ψ:n+1×(1,1)n+1\Psi\colon\mathbb{R}^{n+1}\times(-1,1)\to\mathbb{R}^{n+1} such that

Ψ(x,0)=x,ddsΨ(x,s)=X(Ψ(x,s)),xn+1,s(1,1).\Psi(x,0)=x,\qquad\qquad\frac{\mathrm{d}}{\mathrm{d}s}\Psi(x,s)=X(\Psi(x,s)),\quad\forall\,x\in\mathbb{R}^{n+1},\,s\in(-1,1). (7)

For any s(1,1)s\in(-1,1), let Ω(s)\colonequalsΨ(Ω,s)\Omega^{(s)}\colonequals\Psi(\Omega,s). Note that Ω0=Ω\Omega_{0}=\Omega. Let Σ(s)\colonequalsΩ(s)\Sigma^{(s)}\colonequals\partial^{*}\Omega^{(s)}.

Definition 3.3.

We call {Ω(s)}s(1,1)\{\Omega^{(s)}\}_{s\in(-1,1)} as defined above a normal variation of Ωn+1\Omega\subseteq\mathbb{R}^{n+1}. We also call {Σ(s)}s(1,1)\{\Sigma^{(s)}\}_{s\in(-1,1)} a normal variation of Σ=Ω\Sigma=\partial\Omega.

Lemma 3.4 (First Variation, [Hei21, Lemma 2.4]).

Let XC0(n+1,n+1)X\in C_{0}^{\infty}(\mathbb{R}^{n+1},\mathbb{R}^{n+1}). Let f(x)=X(x),N(x)f(x)=\langle X(x),N(x)\rangle for any xΩx\in\partial^{*}\Omega. Then

dds|s=0γn+1(Ω(s))=Ωf(x)γn+1(x)dx.\frac{\mathrm{d}}{\mathrm{d}s}\Big{|}_{s=0}\gamma_{n+1}(\Omega^{(s)})=\int_{\partial^{*}\Omega}f(x)\gamma_{n+1}(x)\,\mathrm{d}x. (8)
dds|s=0Ω(s)γn(x)dx=Ω(H(x)N(x),x)f(x)γn(x)dx.\frac{\mathrm{d}}{\mathrm{d}s}\Big{|}_{s=0}\int_{\partial^{*}\Omega^{(s)}}\gamma_{n}(x)\,\mathrm{d}x=\int_{\partial^{*}\Omega}(H(x)-\langle N(x),x\rangle)f(x)\gamma_{n}(x)\,\mathrm{d}x. (9)

4. Variations and Regularity

In this section, we show that a minimizer of Problem 2.1 exists, and the boundary of the minimizer is CC^{\infty} except on a set of Hausdorff dimension at most n7n-7.

The results below are repeated from [Hei21].

Unless otherwise stated, all sets Ωn+1\Omega\subseteq\mathbb{R}^{n+1} below are assumed to be measurable sets of locally finite surface area, such that the Gaussian surface area of Ω\Omega, Ωγn(x)dx\int_{\partial^{*}\Omega}\gamma_{n}(x)\,\mathrm{d}x is finite.

Lemma 4.1 (First Variation for Minimizers, [Hei21, Lemma 3.1]).

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1} minimize Problem 2.1. Let Σ\colonequalsΩ\Sigma\colonequals\partial^{*}\Omega. Then there exists λ\lambda\in\mathbb{R} such that, for any xΩx\in\partial^{*}\Omega, H(x)x,N(x)=λH(x)-\langle x,N(x)\rangle=\lambda.

Lemma 4.2 (Second Variation for Minimizers, [Hei21, Lemma 2.5, Lemma 3.2]).

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1} minimize Problem 2.1. Let Σ\colonequalsΩ\Sigma\colonequals\partial^{*}\Omega. Then, for any fC0(Σ)f\in C_{0}^{\infty}(\Sigma) such that Σf(x)γn(x)dx=0\int_{\Sigma}f(x)\gamma_{n}(x)\,\mathrm{d}x=0, and such that f(x)=f(x)f(x)=f(-x) for all xΣx\in\Sigma, there exists a vector field XC0(n+1,n+1)X\in C_{0}^{\infty}(\mathbb{R}^{n+1},\mathbb{R}^{n+1}) with f(x)=X(x),N(x)f(x)=\langle X(x),N(x)\rangle for all xΣx\in\Sigma such that X(x)=X(x)X(x)=X(-x) for all xn+1x\in\mathbb{R}^{n+1}, and such that the variation of {Ω(s)}s(1,1)\{\Omega^{(s)}\}_{s\in(-1,1)} corresponding to XX satisfies

d2ds2|s=0Ω(s)γn(x)dx=Σf(x)Lf(x)γn(x)dx0.\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}\Big{|}_{s=0}\int_{\partial^{*}\Omega^{(s)}}\gamma_{n}(x)\,\mathrm{d}x=\int_{\Sigma}f(x)Lf(x)\gamma_{n}(x)\,\mathrm{d}x\leq 0.
Lemma 4.3 (Existence and Regularity, [Hei21, Lemma 3.3]).

The minimum value of Problem 2.1 exists. That is, there exists a measurable set Ωn+1\Omega\subseteq\mathbb{R}^{n+1} such that Ω\Omega achieves the minimum value of Problem 2.1. Also, Ω\partial^{*}\Omega is a CC^{\infty} manifold. Moreover, if n<7n<7, then ΩΩ=\partial\Omega\setminus\partial^{*}\Omega=\emptyset, and if n7n\geq 7, then the Hausdorff dimension of ΩΩ\partial\Omega\setminus\partial^{*}\Omega is at most n7n-7.

5. Eigenfunctions of L

Let e1,,ene_{1},\ldots,e_{n} be an orthonormal frame for an orientable nn-dimensional hypersurface Σn+1\Sigma\subseteq\mathbb{R}^{n+1} with Σ=\partial\Sigma=\emptyset. Let Δ\colonequalsi=1neiei\Delta\colonequals\sum_{i=1}^{n}\nabla_{e_{i}}\nabla_{e_{i}} be the Laplacian associated to Σ\Sigma. Let \colonequalsi=1neiei\nabla\colonequals\sum_{i=1}^{n}e_{i}\nabla_{e_{i}} be the gradient associated to Σ\Sigma. (The symbol ()\nabla_{\cdot}(\cdot) still denotes the Euclidean connection, and the meaning of the symbol \nabla should be clear from context.) For any n×nn\times n matrix B=(bij)1i,jnB=(b_{ij})_{1\leq i,j\leq n}, define B2\colonequalsi,j=1nbij2\left\|B\right\|^{2}\colonequals\sum_{i,j=1}^{n}b_{ij}^{2}. For any fC(Σ)f\in C^{\infty}(\Sigma), define

f\colonequalsΔfx,f.\mathcal{L}f\colonequals\Delta f-\langle x,\nabla f\rangle. (10)
Lf\colonequalsΔfx,f+f+A2f.Lf\colonequals\Delta f-\langle x,\nabla f\rangle+f+\left\|A\right\|^{2}f. (11)

Note that there is a factor of 22 difference between our definition of LL and the definition of LL in [CM12]. Below we often remove the xx arguments of the functions for brevity. We extend LL to matrices so that (LB)ij\colonequalsL(Bij)(LB)_{ij}\colonequals L(B_{ij}) for all 1i,jn1\leq i,j\leq n, and we can similarly extend \nabla to matrices.

Remark 5.1.

Let f,gC(Σ)f,g\in C^{\infty}(\Sigma). Using (11), we get the following product rule for LL.

L(fg)\displaystyle L(fg) =fΔg+gΔf+2f,gfx,ggx,f+A2fg+fg\displaystyle=f\Delta g+g\Delta f+2\langle\nabla f,\nabla g\rangle-f\langle x,\nabla g\rangle-g\langle x,\nabla f\rangle+\left\|A\right\|^{2}fg+fg
=fLg+gf+2f,g.\displaystyle=fLg+g\mathcal{L}f+2\langle\nabla f,\nabla g\rangle.
Lemma 5.2 (HH is almost an eigenfunction of LL [CM15, Proposition 1.2] [Gua18, Lemma 2.1], [Hei21, Lemma 4.1]).

Let Σn+1\Sigma\subseteq\mathbb{R}^{n+1} be an orientable hypersurface. Let λ\lambda\in\mathbb{R}. If

H(x)=x,N(x)+λ,xΣ.H(x)=\langle x,N(x)\rangle+\lambda,\qquad\forall\,x\in\Sigma. (12)

Then

LA=2AλA2.LA=2A-\lambda A^{2}. (13)
LH=2H+λA2.LH=2H+\lambda\left\|A\right\|^{2}. (14)

Equation (13) suggests that an eigenvalue α\alpha of AA should satisfy an equation similar to (13). That is, if we choose an orthonormal frame such that AA is diagonal at a point, then it should be the case that LαL\alpha is nearly equal to 2αλα22\alpha-\lambda\alpha^{2}. (Near an umbilic point, the largest eigenvalue of AA might change quickly, so we should not expect LαL\alpha to be equal to 2αλα22\alpha-\lambda\alpha^{2}.) Such a computation was done in [Lee22]. Also, there is a technical issue, that an eigenvalue of AA might not be differentiable, or more importantly, we might not be able to integrate by parts using α\alpha, i.e. we cannot use α\alpha in the second variation formula, Lemma 4.2.

For these technical reasons, we instead use a softmax function in (15) to approximate the largest eigenvalue of AA. The function in (15) will be smooth, so we can integrate by parts with it and use it in Lemma 4.2. Moreover, since the function (15) does not rely on the choice of orthonormal frame, we can substitute (13) into the differentiation formula for (15) in (16) below. There is an extra term that appears in (16) that results from the smoothing in (15), but the sign of this extra term is determined, so it does not affect our calculations later very much.

Let β>0\beta>0. Define the (β\beta-smoothed) maximum eigenvalue of AA by

maxβ(A)\colonequals1βlog(TreβA).\mathrm{max}_{\beta}(A)\colonequals\frac{1}{\beta}\log\Big{(}\mathrm{Tr}\,e^{\beta A}\Big{)}. (15)

Here Tr\mathrm{Tr} denotes the trace of a square matrix.

Lemma 5.3.
maxβ(A)=Tr(eβATr(eβA)A)+βi=1nTr(eβATr(eβA)[eiAInTr(eβATr(eβA)eiA)]2).\mathcal{L}\mathrm{max}_{\beta}(A)=\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\mathcal{L}A\Big{)}+\beta\sum_{i=1}^{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Big{[}\nabla_{e_{i}}A-I_{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\nabla_{e_{i}}A\Big{)}\Big{]}^{2}\Big{)}. (16)

Consequently, if β>0\beta>0,

maxβ(A)j=1npj(A)ajj,\mathcal{L}\mathrm{max}_{\beta}(A)\geq\sum_{j=1}^{n}p_{j}(A)\mathcal{L}a_{jj},
Proof.

Fix i,k{1,,n}i,k\in\{1,\ldots,n\}. Then

eimaxβ(A)=(15)eiTr(eβA)βTr(eβA)=Tr(eieβA)βTr(eβA)=Tr(eβAeiA)Tr(eβA).\nabla_{e_{i}}\mathrm{max}_{\beta}(A)\stackrel{{\scriptstyle\eqref{smaxdef}}}{{=}}\frac{\nabla_{e_{i}}\mathrm{Tr}(e^{\beta A})}{\beta\mathrm{Tr}(e^{\beta A})}=\frac{\mathrm{Tr}(\nabla_{e_{i}}e^{\beta A})}{\beta\mathrm{Tr}(e^{\beta A})}=\frac{\mathrm{Tr}(e^{\beta A}\nabla_{e_{i}}A)}{\mathrm{Tr}(e^{\beta A})}. (17)
eieimaxβ(A)=(17)Tr(eβA)ei[Tr(eβAeiA)]Tr(eβAeiA)ei[Tr(eβA)][Tr(eβA)]2\displaystyle\nabla_{e_{i}}\nabla_{e_{i}}\mathrm{max}_{\beta}(A)\stackrel{{\scriptstyle\eqref{chat3}}}{{=}}\frac{\mathrm{Tr}(e^{\beta A})\nabla_{e_{i}}[\mathrm{Tr}(e^{\beta A}\nabla_{e_{i}}A)]-\mathrm{Tr}(e^{\beta A}\nabla_{e_{i}}A)\nabla_{e_{i}}[\mathrm{Tr}(e^{\beta A})]}{[\mathrm{Tr}(e^{\beta A})]^{2}} (18)
=(17)Tr(eβA)[Tr(eβAeieiA)+βTr(eβA(eiA)2]Tr(eβAeiA)βTr(eβAeiA)[Tr(eβA)]2.\displaystyle\stackrel{{\scriptstyle\eqref{chat3}}}{{=}}\frac{\mathrm{Tr}(e^{\beta A})[\mathrm{Tr}(e^{\beta A}\nabla_{e_{i}}\nabla_{e_{i}}A)+\beta\mathrm{Tr}(e^{\beta A}(\nabla_{e_{i}}A)^{2}]-\mathrm{Tr}(e^{\beta A}\nabla_{e_{i}}A)\beta\mathrm{Tr}(e^{\beta A}\nabla_{e_{i}}A)}{[\mathrm{Tr}(e^{\beta A})]^{2}}.

Summing (18) over 1in1\leq i\leq n, we obtain

Δmaxβ(A)=Tr(eβAΔA)Tr(eβA)+βi=1nTr(eβA(eiA)2)Tr(eβA)βi=1n(Tr(eβAeiA)Tr(eβA))2.\Delta\mathrm{max}_{\beta}(A)=\frac{\mathrm{Tr}(e^{\beta A}\Delta A)}{\mathrm{Tr}(e^{\beta A})}+\beta\sum_{i=1}^{n}\frac{\mathrm{Tr}(e^{\beta A}(\nabla_{e_{i}}A)^{2})}{\mathrm{Tr}(e^{\beta A})}-\beta\sum_{i=1}^{n}\Big{(}\frac{\mathrm{Tr}(e^{\beta A}\nabla_{e_{i}}A)}{\mathrm{Tr}(e^{\beta A})}\Big{)}^{2}. (19)

Note that C\colonequalseβA/Tr(eβA)C\colonequals e^{\beta A}/\mathrm{Tr}(e^{\beta A}) is a symmetric positive definite matrix with trace 11, so the last two terms can be written as a sum of terms of the form Tr(C(DInTr(CD))2)\mathrm{Tr}(C(D-I_{n}\mathrm{Tr}(CD))^{2}), i.e.

Δmaxβ(A)=Tr(eβATr(eβA)ΔA)+βi=1nTr(eβATr(eβA)[eiAInTr(eβATr(eβA)eiA)]2).\Delta\mathrm{max}_{\beta}(A)=\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Delta A\Big{)}+\beta\sum_{i=1}^{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Big{[}\nabla_{e_{i}}A-I_{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\nabla_{e_{i}}A\Big{)}\Big{]}^{2}\Big{)}. (20)

Recall: if P,QP,Q are symmetric positive semidefinite real matrices, Tr(PQ)=Tr(P1/2QP1/2)\mathrm{Tr}(PQ)=\mathrm{Tr}(P^{1/2}QP^{1/2}), and P1/2QP1/2P^{1/2}QP^{1/2} is itself a symmetric positive semidefinite matrix, so that Tr(PQ)0\mathrm{Tr}(PQ)\geq 0. Since AA is symmetric, so is eiA\nabla_{e_{i}}A, so that (20) implies

Δmaxβ(A)Tr(eβATr(eβA)ΔA).\Delta\mathrm{max}_{\beta}(A)\geq\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Delta A\Big{)}. (21)

Finally, using (17), if β>0\beta>0, then

maxβ(A)\displaystyle\mathcal{L}\mathrm{max}_{\beta}(A) =(10)Δmaxβ(A)x,maxβ(A)\displaystyle\stackrel{{\scriptstyle\eqref{three4.3}}}{{=}}\Delta\mathrm{max}_{\beta}(A)-\langle x,\nabla\mathrm{max}_{\beta}(A)\rangle
(21)Tr(eβATr(eβA)[ΔAx,A])=(10)Tr(eβATr(eβA)A).\displaystyle\stackrel{{\scriptstyle\eqref{chat7}}}{{\geq}}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}[\Delta A-\langle x,\nabla A\rangle]\Big{)}\stackrel{{\scriptstyle\eqref{three4.3}}}{{=}}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\mathcal{L}A\Big{)}.

The main strategy used to prove Theorem 2.4 is to demonstrate that Ω\partial\Omega has an eigenvalue larger than 22, and then to use Lemma 5.4 below to conclude that Ω×\Omega\times\mathbb{R} is unstable for Problem 2.1. Proving that the hypothesis of Lemma 5.4 holds seems to require breaking into several different cases, as we do below in the various cases of the proof of Theorem 2.4.

Lemma 5.4.

Let Ωn\Omega\subseteq\mathbb{R}^{n}. Let f:Σf\colon\Sigma\to\mathbb{R} satisfy f(x)=f(x)f(x)=f(-x) for all xΣx\in\Sigma, Σ(f2+f2+|ff|)γn(x)dx<\int_{\Sigma}(f^{2}+\left\|\nabla f\right\|^{2}+\left|f\mathcal{L}f\right|)\gamma_{n}(x)\,\mathrm{d}x<\infty, and Lf=δfLf=\delta f, or ΣfLfγn(x)dxδΣf2γn(x)dx\int_{\Sigma}fLf\gamma_{n}(x)\,\mathrm{d}x\geq\delta\int_{\Sigma}f^{2}\gamma_{n}(x)\,\mathrm{d}x. Let g:n+1g\colon\mathbb{R}^{n+1}\to\mathbb{R} defined by

g(x1,,xn+1)\colonequals(xn+121)f(x1,,xn),(x1,,xn,xn+1)Σ×.g(x_{1},\ldots,x_{n+1})\colonequals(x_{n+1}^{2}-1)f(x_{1},\ldots,x_{n}),\qquad\forall\,(x_{1},\ldots,x_{n},x_{n+1})\in\Sigma\times\mathbb{R}.

Then g(x)=g(x)g(-x)=g(x) for all xn+1x\in\mathbb{R}^{n+1}, Σ×g(x)γn(x)dx=0\int_{\Sigma\times\mathbb{R}}g(x)\gamma_{n}(x)\,\mathrm{d}x=0, and the corresponding variation of Σ×\Sigma\times\mathbb{R} satisfies

d2ds2|s=0(Σ×)(s)γn(x)𝑑x(δ2)Σ×|g(x)|2γn(x)𝑑x.\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}\Big{|}_{s=0}\int_{(\Sigma\times\mathbb{R})^{(s)}}\gamma_{n}(x)dx\leq-(\delta-2)\int_{\Sigma\times\mathbb{R}}\left|g(x)\right|^{2}\gamma_{n}(x)dx.
Proof.

The property g(x)=g(x)g(x)=g(-x) for all xn+1x\in\mathbb{R}^{n+1} is evident. By Fubini’s Theorem, gg automatically preserves the Gaussian volume of Ω\Omega, i.e.

Σ(xn+121)fγn(x)𝑑x=(xn+121)γ1(xn+1)Σfγn1(x1,,xn)𝑑x1𝑑xn=0.\int_{\Sigma}(x_{n+1}^{2}-1)f\gamma_{n}(x)dx=\int_{\mathbb{R}}(x_{n+1}^{2}-1)\gamma_{1}(x_{n+1})\int_{\Sigma^{\prime}}f\gamma_{n-1}(x_{1},\ldots,x_{n})dx_{1}\cdots dx_{n}=0.

Since ff does not depend on xn+1x_{n+1}, f,(xn+121)=0\langle\nabla f,\nabla(x_{n+1}^{2}-1)\rangle=0, and by the product rule for LL (Remark 5.1) we have

Lg\displaystyle Lg =(11)(xn+121)Lf+f(xn+121)+f,(xn+121)\displaystyle\stackrel{{\scriptstyle\eqref{three4.5}}}{{=}}(x_{n+1}^{2}-1)Lf+f\mathcal{L}(x_{n+1}^{2}-1)+\langle\nabla f,\nabla(x_{n+1}^{2}-1)\rangle (22)
=(xn+121)Lf+f(22xn+12)=(11)(xn+121)Lf2g.\displaystyle=(x_{n+1}^{2}-1)Lf+f(2-2x_{n+1}^{2})\stackrel{{\scriptstyle\eqref{three4.5}}}{{=}}(x_{n+1}^{2}-1)Lf-2g.

Finally, from Lemma 4.2,

d2ds2|s=0(Σ×)(s)γn(x)𝑑x\displaystyle\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}\Big{|}_{s=0}\int_{(\Sigma\times\mathbb{R})^{(s)}}\gamma_{n}(x)dx =Σ×gLgγn(x)dx=(22)Σ×(xn+121)2[fLf2f2]γn(x)dx\displaystyle=\int_{\Sigma\times\mathbb{R}}-gLg\gamma_{n}(x)dx\stackrel{{\scriptstyle\eqref{six2}}}{{=}}\int_{\Sigma\times\mathbb{R}}-(x_{n+1}^{2}-1)^{2}[fLf-2f^{2}]\gamma_{n}(x)dx
(δ2)Σ×g2γn(x)𝑑x.\displaystyle\leq-(\delta-2)\int_{\Sigma\times\mathbb{R}}g^{2}\gamma_{n}(x)dx.

(The integration by parts here is justified by Lemma 8.2.) ∎

The following Lemma says that a minimizer of Problem 2.1 satisfies an a priori bound on the Euclidean volume growth of its boundary.

Lemma 5.5.

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1} minimize Problem 2.1 with n1n\geq 1. Then Σ\colonequalsΩ\Sigma\colonequals\partial^{*}\Omega has polynomial volume growth, i.e. there exists a constant b>0b>0 such that the Euclidean surface area of Σ\Sigma satisfies

Vol({xΣ:x>r})brn+1,r>1.\mathrm{Vol}(\{x\in\Sigma\colon\left\|x\right\|>r\})\leq br^{n+1},\qquad\forall\,r>1.
Proof.

Denote B(0,r)\colonequals{xn+1:x<r}B(0,r)\colonequals\{x\in\mathbb{R}^{n+1}\colon\left\|x\right\|<r\}. Let Ωr\colonequalsΩB(0,r)c\Omega_{\geq r}\colonequals\Omega\cap B(0,r)^{c}. Let Cn+1C\subseteq\mathbb{R}^{n+1} be the complement of a ball centered at the origin such that γn+1(Ωr)=γn+1(C)\gamma_{n+1}(\Omega_{\geq r})=\gamma_{n+1}(C). Define Ω\colonequalsC(ΩΩr)\Omega^{\prime}\colonequals C\cup(\Omega\setminus\Omega_{\geq r}). Then γn+1(Ω)=γn+1(Ω)\gamma_{n+1}(\Omega)=\gamma_{n+1}(\Omega^{\prime}), and since Ω\Omega minimizes Problem 2.1, we have

B(0,r)cΣγn(x)+B(0,r)Σγn(x)dx=Σγn(x)dx\displaystyle\int_{B(0,r)^{c}\cap\Sigma}\gamma_{n}(x)+\int_{B(0,r)\cap\Sigma}\gamma_{n}(x)\,\mathrm{d}x=\int_{\Sigma}\gamma_{n}(x)\,\mathrm{d}x
Σγn(x)dx=B(0,r)cΣγn(x)dx+B(0,r)Σγn(x)dx\displaystyle\qquad\leq\int_{\Sigma^{\prime}}\gamma_{n}(x)\,\mathrm{d}x=\int_{B(0,r)^{c}\cap\Sigma^{\prime}}\gamma_{n}(x)\,\mathrm{d}x+\int_{B(0,r)\cap\Sigma^{\prime}}\gamma_{n}(x)\,\mathrm{d}x
=B(0,r)cΣγn(x)dx+B(0,r)Σγn(x)dx.\displaystyle\qquad=\int_{B(0,r)^{c}\cap\Sigma}\gamma_{n}(x)\,\mathrm{d}x+\int_{B(0,r)\cap\Sigma^{\prime}}\gamma_{n}(x)\,\mathrm{d}x.

The last line used the definition of Ω\Omega^{\prime}. Canceling like terms, we then get

B(0,r)cΣγn(x)B(0,r)cΣγn(x)dx2B(0,r)cγn(x)dx.\int_{B(0,r)^{c}\cap\Sigma}\gamma_{n}(x)\leq\int_{B(0,r)^{c}\cap\Sigma^{\prime}}\gamma_{n}(x)\,\mathrm{d}x\leq 2\int_{\partial B(0,r)^{c}}\gamma_{n}(x)\,\mathrm{d}x.

The last inequality used the definition of Ω\Omega^{\prime}. So, there exists b=bn+1b=b_{n+1} such that, \forall r>1r>1

B(0,r)cΣγn(x)brn+1er2/2.\int_{B(0,r)^{c}\cap\Sigma}\gamma_{n}(x)\leq br^{n+1}e^{-r^{2}/2}. (23)

It then follows that Ω\partial\Omega has polynomial volume growth. Denote ΣB(0,r)c(B(0,N))\equalscolonΣrN\Sigma\cap B(0,r)^{c}\cap(B(0,N))\equalscolon\Sigma_{r\leq N}, so that

γn(ΣrN)\displaystyle\gamma_{n}(\Sigma_{r\leq N}) =rNet2/2ddtVol(ΣtN+1)dt(2π)n2rNet2+ε/2ddtVol(ΣtN+1)dt(2π)n2\displaystyle=\int_{r}^{N}e^{-t^{2}/2}\frac{\mathrm{d}}{\mathrm{d}t}\mathrm{Vol}(\Sigma_{t\leq N+1})\frac{\mathrm{d}t}{(2\pi)^{\frac{n}{2}}}\geq\int_{r}^{N}e^{-t^{2+\varepsilon}/2}\frac{\mathrm{d}}{\mathrm{d}t}\mathrm{Vol}(\Sigma_{t\leq N+1})\frac{\mathrm{d}t}{(2\pi)^{\frac{n}{2}}} (24)
=er2+ε/2Vol(ΣrN+1)1(2π)n2eN2+ε/2Vol(ΣNN+1)1(2π)n2\displaystyle=e^{-r^{2+\varepsilon}/2}\mathrm{Vol}(\Sigma_{r\leq N+1})\frac{1}{(2\pi)^{\frac{n}{2}}}-e^{-N^{2+\varepsilon}/2}\mathrm{Vol}(\Sigma_{N\leq N+1})\frac{1}{(2\pi)^{\frac{n}{2}}}
+rNtet2+ε/2Vol(ΣtN+1)dt(2π)n2.\displaystyle\qquad\qquad\qquad\qquad+\int_{r}^{N}te^{-t^{2+\varepsilon}/2}\mathrm{Vol}(\Sigma_{t\leq N+1})\frac{\mathrm{d}t}{(2\pi)^{\frac{n}{2}}}.

The last term is nonnegative and finite. Letting NN\to\infty, the second term goes to zero by (23). Then letting ε0+\varepsilon\to 0^{+}, (24) becomes

Vol(ΣB(0,r)c)1(2π)n2(23)er2/2brn+1er2/2=brn+1.\mathrm{Vol}(\Sigma\cap B(0,r)^{c})\frac{1}{(2\pi)^{\frac{n}{2}}}\stackrel{{\scriptstyle\eqref{cdb}}}{{\leq}}e^{r^{2}/2}br^{n+1}e^{-r^{2}/2}=br^{n+1}.

6. Classification of Stable Self-Shrinkers

Theorem 6.1.

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1} minimize Problem 2.1. Assume that Σ\colonequalsΩ\Sigma\colonequals\partial^{*}\Omega satisfies

H(x)=x,N(x),xΣ.H(x)=\langle x,N(x)\rangle,\qquad\forall\,x\in\Sigma. (25)

Then there exists an integer 0kn0\leq k\leq n such that, after rotating Ω\Omega,

Σ=kSk×nk.\Sigma=\sqrt{k}S^{k}\times\mathbb{R}^{n-k}.
Proof.

Case 1. Assume Ω\partial\Omega is compact.

We repeat the proof of [Hei21, Proposition 1.5]. Let HH be the mean curvature of Σ\Sigma. If H0H\geq 0 on Σ\Sigma, then Huisken’s classification [Hui90, Hui93] [CM12, Theorem 0.17] of compact surfaces satisfying (25) implies that Σ\Sigma is a round sphere (Σ=nSn\Sigma=\sqrt{n}S^{n}). So, we may assume that HH changes sign on Σ\Sigma. From (14) with λ=0\lambda=0, LH=2HLH=2H. Since HH changes sign, 22 is not the largest eigenvalue of LL, by spectral theory [Zhu20, Lemma 6.5] (e.g. using that (LA22)1(L-\left\|A\right\|^{2}-2)^{-1} is a compact operator). That is, there exists a C2C^{2} function g:Σg\colon\Sigma\to\mathbb{R} and there exists δ>2\delta>2 such that Lg=δgLg=\delta g. Moreover, g>0g>0 on Σ\Sigma. Since g>0g>0 and Σ=Σ\Sigma=-\Sigma, it follows by (11) that g(x)+g(x)g(x)+g(-x) is an eigenfunction of LL with eigenvalue δ\delta. That is, we may assume that g(x)=g(x)g(x)=g(-x) for all xΣx\in\Sigma.

From the second variation formula, Lemma 4.2,

d2ds2|s=0Σ(s)γn(x)dx=Σf(x)Lf(x)γn(x)dx.\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}|_{s=0}\int_{\Sigma^{(s)}}\gamma_{n}(x)\,\mathrm{d}x=-\int_{\Sigma}f(x)Lf(x)\gamma_{n}(x)\,\mathrm{d}x.

So, to complete the proof, it suffices by Lemma 4.2 to find a C2C^{2} function ff such that

  • f(x)=f(x)f(x)=f(-x) for all xΣx\in\Sigma. (ff preserves symmetry.)

  • Σf(x)γn(x)dx=0\int_{\Sigma}f(x)\gamma_{n}(x)\,\mathrm{d}x=0. (ff preserves Gaussian volume.)

  • Σf(x)Lf(x)γn(x)dx>0\int_{\Sigma}f(x)Lf(x)\gamma_{n}(x)\,\mathrm{d}x>0. (ff decreases Gaussian surface area.)

We choose gg as above so that Lg=δgLg=\delta g, δ>2\delta>2 and so that Σ(H(x)+g(x))γn(x)dx=0\int_{\Sigma}(H(x)+g(x))\gamma_{n}(x)\,\mathrm{d}x=0. (Since HH changes sign and g0g\geq 0, gg can satisfy the last equality by multiplying it by an appropriate constant.) We then define f\colonequalsH+gf\colonequals H+g. Then ff satisfies the first two properties. So, it remains to show that ff satisfies the last property. Note that, since HH and gg have different eigenvalues, they are orthogonal, i.e. ΣH(x)g(x)γn(x)dx=0\int_{\Sigma}H(x)g(x)\gamma_{n}(x)\,\mathrm{d}x=0. Therefore,

Σf(x)Lf(x)γn(x)dx\displaystyle\int_{\Sigma}f(x)Lf(x)\gamma_{n}(x)\,\mathrm{d}x =Σ(H(x)+g(x))(2H(x)+δg(x))γn(x)dx\displaystyle=\int_{\Sigma}(H(x)+g(x))(2H(x)+\delta g(x))\gamma_{n}(x)\,\mathrm{d}x
=2Σ(H(x))2γn(x)dx+δΣ(g(x))2γn(x)dx>0.\displaystyle=2\int_{\Sigma}(H(x))^{2}\gamma_{n}(x)\,\mathrm{d}x+\delta\int_{\Sigma}(g(x))^{2}\gamma_{n}(x)\,\mathrm{d}x>0.

(Since H(x)=x,N(x)H(x)=\langle x,N(x)\rangle for all xΣx\in\Sigma, and Σ\Sigma is compact, both Σ(H(x))2γn(x)dx\int_{\Sigma}(H(x))^{2}\gamma_{n}(x)\,\mathrm{d}x and ΣH(x)γn(x)dx\int_{\Sigma}H(x)\gamma_{n}(x)\,\mathrm{d}x are finite a priori.) We conclude that HH cannot change sign, i.e. we must have Σ=nSn\Sigma=\sqrt{n}S^{n} in this case.

Case 2. Assume Ω\partial\Omega is not compact, and HH is not identically zero.

This case follows the argument of Case 1 [CM12, Lemmas 9.44 and 9.45], [Zhu20, Proposition 6.11]. If HH does not change sign, Huisken’s classification [Zhu20, Theorem 0.4] now implies that there exists 0kn0\leq k\leq n such that, after rotating Ω\Omega, Ω=kSk×nk\partial\Omega=\sqrt{k}S^{k}\times\mathbb{R}^{n-k}. If HH changes sign, then instead of asserting the existence of gg, one approximates gg by a sequence of Dirichlet eigenfunctions on the intersection of Σ\Sigma with large compact balls. We then repeat the proof of Case 1. The eigenfunctions HH and gg are not necessarily orthogonal in the non-compact case, but their inner product ΣH(x)g(x)γn(x)dx\int_{\Sigma}H(x)g(x)\gamma_{n}(x)\,\mathrm{d}x can be made to satisfy |ΣH(x)g(x)γn(x)dx|ε(ΣH2γn(x)dx)1/2(Σg2γn(x)dx)1/2\left|\int_{\Sigma}H(x)g(x)\gamma_{n}(x)\,\mathrm{d}x\right|\leq\varepsilon(\int_{\Sigma}H^{2}\gamma_{n}(x)\,\mathrm{d}x)^{1/2}(\int_{\Sigma}g^{2}\gamma_{n}(x)\,\mathrm{d}x)^{1/2} for arbitrary ε>0\varepsilon>0 [Zhu20, Equation 6.41], so the above argument works, with this change. Also, the integration by parts is justified in Lemmas 8.2 and 8.4.

Case 3. Assume HH is identically zero.

In this final remaining case, we have from (25) that H(x)=x,N(x)=0H(x)=\langle x,N(x)\rangle=0 for all xΣx\in\Sigma. That is, Σ\Sigma is a (minimal) cone. We eliminate this case in Section 11.

7. Classification of Stable Convex Sets

Theorem 7.1.

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1}. Assume that Ω×\Omega\times\mathbb{R} minimizes Problem 2.1. (From Lemma 4.1, there exists λ\lambda\in\mathbb{R} such that H(x)=x,N(x)+λH(x)=\langle x,N(x)\rangle+\lambda \forall xΣx\in\Sigma.) Assume that H0H\geq 0 on Σ\colonequalsΩ\Sigma\colonequals\partial^{*}\Omega and λ0\lambda\geq 0. Then, after rotating Ω\Omega,

Ω=λS0×n.\partial\Omega=\lambda S^{0}\times\mathbb{R}^{n}.
Proof.

By (14),

ΩHLHγn(x)dx=Ω[2H2+λHA2]γn(x)dx.\int_{\partial\Omega}HLH\gamma_{n}(x)\,\mathrm{d}x=\int_{\partial\Omega}[2H^{2}+\lambda H\left\|A\right\|^{2}]\gamma_{n}(x)\,\mathrm{d}x. (26)

(The quantity ΩHLHγn(x)dx\int_{\partial\Omega}HLH\gamma_{n}(x)\,\mathrm{d}x is finite a priori by Lemma 8.4; note also Ω\partial\Omega has polynomial volume growth by Lemma 5.5.) So,

ΩHLHγn(x)dxΩ2H2γn(x)dx,\int_{\partial\Omega}HLH\gamma_{n}(x)\,\mathrm{d}x\geq\int_{\partial\Omega}2H^{2}\gamma_{n}(x)\,\mathrm{d}x, (27)

with equality only when HA=0H\left\|A\right\|=0 identically, i.e. when H=0H=0 identically (since A=0\left\|A\right\|=0 implies H=0H=0). If H=0H=0 on all of Σ\Sigma, then by (14),

0=LH=2H+λA2=λA2.0=LH=2H+\lambda\left\|A\right\|^{2}=\lambda\left\|A\right\|^{2}.

Since λ>0\lambda>0 we conclude that A=0\left\|A\right\|=0 on all of Σ\Sigma. In summary, the inequality in (27) is strict, unless A=0\left\|A\right\|=0 on Σ\Sigma.

In the former case of strict inequality in (27), let f\colonequalsHf\colonequals H in Lemma 5.4. Note that H(x)=H(x)H(x)=H(-x) for all xΣx\in\Sigma. Then Lemma 5.4 and (27) imply that the variation of Ω×\Omega\times\mathbb{R} correspond to gg satisfies

d2ds2|s=0(Σ×)(s)γn+1(x)𝑑x<0.\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}\Big{|}_{s=0}\int_{(\Sigma\times\mathbb{R})^{(s)}}\gamma_{n+1}(x)dx<0.

This inequality violates the minimality of Ω\Omega, achieving a contradiction, demonstrating that we must have A=0\left\|A\right\|=0 on Σ\Sigma. In the latter case, we must then have Σ=rS0×n\Sigma=rS^{0}\times\mathbb{R}^{n} for some r>0r>0. Since 0=H=x,N+λ0=H=\langle x,N\rangle+\lambda, we conclude that r=λr=\lambda. ∎

Note that Theorem 7.1 proved a stronger result than we need to prove Theorem 2.4. Theorem 7.1 shows that the only Ω\Omega with Ω×\Omega\times\mathbb{R} stable for Gaussian surface area with H0H\geq 0 and λ>0\lambda>0 (or H0H\leq 0 and λ<0\lambda<0) is a symmetric slab or its complement.

8. Non-Mean Convex Sets are Unstable

In order to prove instability of noncompact sets, we will need to integrate by parts in the second variation formula in Lemma 4.2. The noncompactness, together with the low-dimensional singularities on the surface (see Lemma 4.3) mean that integrating by parts is nontrivial, hence the results below.

Lemma 8.1 (Integration by Parts, [CM12, Corollary 3.10], [Zhu20, Lemma 5.4], [Hei21, Lemma 4.4]).

Let Σn+1\Sigma\subseteq\mathbb{R}^{n+1} be an nn-dimensional hypersurface. Let f,g:Σf,g\colon\Sigma\to\mathbb{R}. Assume that ff is a C2C^{2} function and gg is a C2C^{2} function with compact support. Then

Σfgγn(x)𝑑x=Σgfγn(x)𝑑x=Σf,gγn(x)𝑑x.\int_{\Sigma}f\mathcal{L}g\gamma_{n}(x)dx=\int_{\Sigma}g\mathcal{L}f\gamma_{n}(x)dx=-\int_{\Sigma}\langle\nabla f,\nabla g\rangle\gamma_{n}(x)dx.
Corollary 8.2 (Integration by Parts, Version 2 [Zhu20, Lemma 5.4], [Hei21, Corollary 5.10]).

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1}. Let f,g:Ωf,g\colon\partial^{*}\Omega\to\mathbb{R} be C2C^{2} functions. Suppose the Hausdorff dimension of ΩΩ\partial\Omega\setminus\partial^{*}\Omega is at most n7n-7. Assume that

Σfgγn(x)𝑑x<,Σ|fg|γn(x)dx<,Σfg6/5γn(x)𝑑x<.\int_{\Sigma}\left\|\nabla f\right\|\left\|\nabla g\right\|\gamma_{n}(x)dx<\infty,\qquad\int_{\Sigma}\left|f\mathcal{L}g\right|\gamma_{n}(x)dx<\infty,\qquad\int_{\Sigma}\left\|f\nabla g\right\|^{6/5}\gamma_{n}(x)dx<\infty.

Then

Σfgγn(x)𝑑x=Σf,gγn(x)𝑑x.\int_{\Sigma}f\mathcal{L}g\gamma_{n}(x)dx=-\int_{\Sigma}\langle\nabla f,\nabla g\rangle\gamma_{n}(x)dx.
Corollary 8.3 ([Zhu20, Lemma 6.2], [Hei21, Corollary 5.5]).

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1} and let Σ\colonequalsΩ\Sigma\colonequals\partial\Omega. Assume δ(Σ)<\delta(\Sigma)<\infty. Suppose g:Σg\colon\Sigma\to\mathbb{R} is a C2C^{2} function with g>0g>0 and Lg=δgLg=\delta g. Assume that the Hausdorff dimension of ΩΩ\partial\Omega\setminus\partial^{*}\Omega is at most n7n-7. Then for any k0k\geq 0, AxkL2(Σ,γn)\left\|A\right\|\left\|x\right\|^{k}\in L_{2}(\Sigma,\gamma_{n})

Lemma 8.4 ([CM12, Theorem 9.36], [Hei21, Lemma 5.6]).

Let Σn+1\Sigma\subseteq\mathbb{R}^{n+1} be a connected, orientable CC^{\infty} hypersurface with polynomial volume growth and with possibly nonempty boundary. Assume \exists λ\lambda\in\mathbb{R} such that H(x)=x,N(x)+λH(x)=\langle x,N(x)\rangle+\lambda for all xΣx\in\Sigma. Let δ\colonequalsδ(Σ)\delta\colonequals\delta(\Sigma). Assume δ(Σ)<\delta(\Sigma)<\infty. Then

HL2(Σ,γn),A|H|L2(Σ,γn).\left\|\nabla H\right\|\in L_{2}(\Sigma,\gamma_{n}),\qquad\left\|A\right\|\left|H\right|\in L_{2}(\Sigma,\gamma_{n}).
Σ|HH|γn(x)𝑑x<.\int_{\Sigma}\left|H\mathcal{L}H\right|\gamma_{n}(x)dx<\infty.
Theorem 8.5.

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1}. Suppose HH changes sign on Σ\colonequalsΩ\Sigma\colonequals\partial^{*}\Omega (that is, Σ\Sigma is not mean convex). Then Ω×\Omega\times\mathbb{R} does not minimize Problem 2.1.

Proof.

Assume for the sake of contradiction that Ω×\Omega\times\mathbb{R} minimizes Problem 2.1. From Lemma 4.1, there exists λ\lambda\in\mathbb{R} such that

H(x)=x,N(x)+λ,xΣ.H(x)=\langle x,N(x)\rangle+\lambda,\qquad\forall\,x\in\Sigma. (28)

The case λ=0\lambda=0 was already treated in Theorem 6.1. We may therefore assume that λ0\lambda\neq 0. Replacing Ω\Omega with Ωc\Omega^{c} if necessary (i.e. changing the direction of NN, which changes the sign of HH and NN and therefore of λ\lambda), we may assume that λ>0\lambda>0.

Since HH is not mean convex, HH changes sign, i.e. there exists a set of positive measure on Σ\Sigma where H>0H>0. Let h\colonequalsmax(H,0)h\colonequals\max(H,0). By (14),

ΣhLhγn(x)dx=Σ[2h2+λhA2]γn(x)dx.\int_{\Sigma}hLh\gamma_{n}(x)\,\mathrm{d}x=\int_{\Sigma}[2h^{2}+\lambda h\left\|A\right\|^{2}]\gamma_{n}(x)\,\mathrm{d}x. (29)

(Note that Σ\Sigma has polynomial volume growth by Lemma 5.5. Then, by Lemma 8.4 and Corollary 8.2, HH and hh are well-defined, the integrals in (29) are finite, and we can integrate by parts with HH or hh.) Since λ>0\lambda>0,

ΩhLhγn(x)dx>Ω2h2γn(x)dx.\int_{\partial\Omega}hLh\gamma_{n}(x)\,\mathrm{d}x>\int_{\partial\Omega}2h^{2}\gamma_{n}(x)\,\mathrm{d}x. (30)

Note that equality cannot occur in (30) since equality would only occur when A=0\left\|A\right\|=0 on the set where H>)H>), but this cannot happen since A=0\left\|A\right\|=0 implies H=0H=0).

Note that hh satisfies h(x)=h(x)h(x)=h(-x) for all xΣx\in\Sigma and h0h\neq 0. Then Lemma 5.4 and (30) imply that the variation of Ω×\Omega\times\mathbb{R} correspond to hh satisfies

d2ds2|s=0(Σ×)(s)γn(x)𝑑x<0.\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}\Big{|}_{s=0}\int_{(\Sigma\times\mathbb{R})^{(s)}}\gamma_{n}(x)dx<0.

This inequality violates the minimality of Ω×\Omega\times\mathbb{R}, achieving a contradiction, and completing the proof. ∎

9. Curvature Bounds

We will eventually reduce the mean-convex case of Theorem 2.4 to the convex case. Doing so requires using a (smoothed version of) the largest eigenvalue of the second fundamental form AA (see (15)) in the second variation formula from Lemma 4.2. In order to use (15) in Lemma 4.2, we need to integrate by parts, which then requires an a priori curvature bound to hold. This curvature bound (see Lemma 9.4) was known to hold for mean-convex self-shrinkers, i.e. when H=x,N>0H=\langle x,N\rangle>0 [CM12], and a few small adjustments to their argument allow the bound to apply in the case H=x,N+λH=\langle x,N\rangle+\lambda with λ<0\lambda<0 and H0H\geq 0.

For any hypersurface Σn+1\Sigma\subseteq\mathbb{R}^{n+1}, we define

δ=δ(Σ)\colonequalssupfC0(Σ)ΣfLfγn(x)𝑑xΣf2γn(x)𝑑x.\delta=\delta(\Sigma)\colonequals\sup_{f\in C_{0}^{\infty}(\Sigma)}\frac{\int_{\Sigma}fLf\gamma_{n}(x)dx}{\int_{\Sigma}f^{2}\gamma_{n}(x)dx}. (31)
Lemma 9.1 ([Hei21, Lemma 5.1]).

If Ωn+1\Omega\subseteq\mathbb{R}^{n+1} minimizes Problem 2.1, then δ(Ω)<\delta(\partial^{*}\Omega)<\infty.

Lemma 9.2 (Simons-type inequality, [Sim68, CM12, CW18, Zhu20]).

Let Σ\Sigma be a CC^{\infty} orientable hypersurface. Let λ\lambda\in\mathbb{R}. Suppose H(x)=x,N(x)+λH(x)=\langle x,N(x)\rangle+\lambda, xΣ\forall\,x\in\Sigma. Then

ALA\displaystyle\left\|A\right\|L\left\|A\right\| =2A2λA2,A+A2A2\displaystyle=2\left\|A\right\|^{2}-\lambda\langle A^{2},A\rangle+\left\|\nabla A\right\|^{2}-\left\|\nabla\left\|A\right\|\right\|^{2} (32)
2A2λA2,A.\displaystyle\geq 2\left\|A\right\|^{2}-\lambda\langle A^{2},A\rangle.
Lemma 9.3 ([CM12, Lemma 10.2]).

Let Σn+1\Sigma\subseteq\mathbb{R}^{n+1} be any nn-dimensional hypersurface. Then

(1+2n+1)A2A2+2nn+1H2.\Big{(}1+\frac{2}{n+1}\Big{)}\left\|\nabla\left\|A\right\|\right\|^{2}\leq\left\|\nabla A\right\|^{2}+\frac{2n}{n+1}\left\|\nabla H\right\|^{2}. (33)
Lemma 9.4 ([CM12, Proposition 10.14], [Hei21, Lemma 5.9]).

Let Ω\Omega minimize problem 2.1. Assume δ(Ω)<\delta(\partial^{*}\Omega)<\infty. Assume that H0H\geq 0 on Σ\Sigma (From Lemma 4.1, H(x)=x,N(x)+λH(x)=\langle x,N(x)\rangle+\lambda for all xΣx\in\Sigma). Assume λ<0\lambda<0. Then

Σ(A2+A4+A2+A2)γn(x)𝑑x<.\int_{\Sigma}(\left\|A\right\|^{2}+\left\|A\right\|^{4}+\left\|\nabla\left\|A\right\|\right\|^{2}+\left\|\nabla A\right\|^{2})\gamma_{n}(x)dx<\infty.
Proof.

Since H0H\geq 0, and

(Δx,1)H=(11)(L2A2)H=(14)(λH)A20,(\Delta-\langle x,\nabla\rangle-1)H\stackrel{{\scriptstyle\eqref{three4.5}}}{{=}}(L-2-\left\|A\right\|^{2})H\stackrel{{\scriptstyle\eqref{three9}}}{{=}}(\lambda-H)\left\|A\right\|^{2}\leq 0,

the maximum principle [Eva98, Theorem 6.4.2] (or in this case, the minimum principle applied to L2A2L-2-\left\|A\right\|^{2}) implies that H>0H>0 on Σ\Sigma. Then logH\log H is well-defined, so

logH=(10)i=1nei(eiHH)x,HH=i=1n(eiH)2H2+HH=logH2+HH\displaystyle\mathcal{L}\log H\stackrel{{\scriptstyle\eqref{three4.3}}}{{=}}\sum_{i=1}^{n}\nabla_{e_{i}}\left(\frac{\nabla_{e_{i}}H}{H}\right)-\frac{\langle x,\nabla H\rangle}{H}=\sum_{i=1}^{n}\frac{-(\nabla_{e_{i}}H)^{2}}{H^{2}}+\frac{\mathcal{L}H}{H}=-\left\|\nabla\log H\right\|^{2}+\frac{\mathcal{L}H}{H} (34)
=(11)logH2+LHA2HHH=(14)logH2+2H+λA2A2HHH\displaystyle\stackrel{{\scriptstyle\eqref{three4.5}}}{{=}}-\left\|\nabla\log H\right\|^{2}+\frac{LH-\left\|A\right\|^{2}H-H}{H}\stackrel{{\scriptstyle\eqref{three9}}}{{=}}-\left\|\nabla\log H\right\|^{2}+\frac{2H+\lambda\left\|A\right\|^{2}-\left\|A\right\|^{2}H-H}{H}
=logH2+(1A2)+λA2H.\displaystyle\,\,=-\left\|\nabla\log H\right\|^{2}+(1-\left\|A\right\|^{2})+\lambda\frac{\left\|A\right\|^{2}}{H}.

Note that ΣA2γn(x)𝑑x<\int_{\Sigma}\left\|A\right\|^{2}\gamma_{n}(x)dx<\infty by Corollary 8.3. Let ϕC0(Σ)\phi\in C_{0}^{\infty}(\Sigma). Integrating by parts with Lemma 8.1,

Σϕ2,logHγn(x)𝑑x=Σϕ2logHγn(x)𝑑x\displaystyle\int_{\Sigma}\langle\nabla\phi^{2},\nabla\log H\rangle\gamma_{n}(x)dx=-\int_{\Sigma}\phi^{2}\mathcal{L}\log H\gamma_{n}(x)dx
=(34)Σϕ2(logH2+(1+A2)λA2H)γn(x)𝑑x.\displaystyle\qquad\qquad\qquad\stackrel{{\scriptstyle\eqref{logeq}}}{{=}}\int_{\Sigma}\phi^{2}\Big{(}\left\|\nabla\log H\right\|^{2}+(-1+\left\|A\right\|^{2})-\lambda\frac{\left\|A\right\|^{2}}{H}\Big{)}\gamma_{n}(x)dx.

From the AMGM inequality, |ϕ2,logH|ϕ2+ϕ2logH2\left|\langle\nabla\phi^{2},\nabla\log H\rangle\right|\leq\left\|\nabla\phi\right\|^{2}+\phi^{2}\left\|\nabla\log H\right\|^{2}, so that

Σϕ2(A2λA2H)γn(x)𝑑xΣ[ϕ2+ϕ2]γn(x)𝑑x.\int_{\Sigma}\phi^{2}\Big{(}\left\|A\right\|^{2}-\lambda\frac{\left\|A\right\|^{2}}{H}\Big{)}\gamma_{n}(x)dx\leq\int_{\Sigma}\Big{[}\left\|\nabla\phi\right\|^{2}+\phi^{2}\Big{]}\gamma_{n}(x)dx. (35)

Recall that we assumed λ<0\lambda<0 so the integral on the left is nonnegative. Therefore,

Σϕ2A2γn(x)𝑑xΣ[ϕ2+ϕ2]γn(x)𝑑x.\int_{\Sigma}\phi^{2}\left\|A\right\|^{2}\gamma_{n}(x)dx\leq\int_{\Sigma}\Big{[}\left\|\nabla\phi\right\|^{2}+\phi^{2}\Big{]}\gamma_{n}(x)dx. (36)

Let 0<ε<1/20<\varepsilon<1/2 to be chosen later. Using now ϕ\colonequalsηA\phi\colonequals\eta\left\|A\right\| in (36), where ηC0(Σ)\eta\in C_{0}^{\infty}(\Sigma), η0\eta\geq 0, and using the AMGM inequality in the form 2abεa2+b2/ε2ab\leq\varepsilon a^{2}+b^{2}/\varepsilon, a,b>0a,b>0,

Ση2A4γn(x)𝑑x\displaystyle\int_{\Sigma}\eta^{2}\left\|A\right\|^{4}\gamma_{n}(x)dx (37)
Σ[η2A2+2ηAAη+A2η2+η2A2]γn(x)𝑑x\displaystyle\leq\int_{\Sigma}\Big{[}\eta^{2}\left\|\nabla\left\|A\right\|\right\|^{2}+2\eta\left\|A\right\|\left\|\nabla\left\|A\right\|\right\|\left\|\nabla\eta\right\|+\left\|A\right\|^{2}\left\|\nabla\eta\right\|^{2}+\eta^{2}\left\|A\right\|^{2}\Big{]}\gamma_{n}(x)dx
Σ[(1+ε)η2A2+A2η2(1+1/ε)+η2A2]γn(x)𝑑x.\displaystyle\leq\int_{\Sigma}\Big{[}(1+\varepsilon)\eta^{2}\left\|\nabla\left\|A\right\|\right\|^{2}+\left\|A\right\|^{2}\left\|\nabla\eta\right\|^{2}(1+1/\varepsilon)+\eta^{2}\left\|A\right\|^{2}\Big{]}\gamma_{n}(x)dx.

Using the product rule for \mathcal{L}, and that =LA21\mathcal{L}=L-\left\|A\right\|^{2}-1

12A2\displaystyle\frac{1}{2}\mathcal{L}\left\|A\right\|^{2} =A2+AA=A2+A(LAA3A)\displaystyle=\left\|\nabla\left\|A\right\|\right\|^{2}+\left\|A\right\|\mathcal{L}\left\|A\right\|=\left\|\nabla\left\|A\right\|\right\|^{2}+\left\|A\right\|\Big{(}L\left\|A\right\|-\left\|A\right\|^{3}-\left\|A\right\|\Big{)}
=(32)A2+2A2λA2,A+A2A2A4A2\displaystyle\stackrel{{\scriptstyle\eqref{three30}}}{{=}}\left\|\nabla\left\|A\right\|\right\|^{2}+2\left\|A\right\|^{2}-\lambda\langle A^{2},A\rangle+\left\|\nabla A\right\|^{2}-\left\|\nabla\left\|A\right\|\right\|^{2}-\left\|A\right\|^{4}-\left\|A\right\|^{2}
=A2+A2A4λA2,A\displaystyle=\left\|\nabla A\right\|^{2}+\left\|A\right\|^{2}-\left\|A\right\|^{4}-\lambda\langle A^{2},A\rangle
(33)(1+2n+1)A22nn+1H2+A2A4λA2,A.\displaystyle\stackrel{{\scriptstyle\eqref{two12}}}{{\geq}}\Big{(}1+\frac{2}{n+1}\Big{)}\left\|\nabla\left\|A\right\|\right\|^{2}-\frac{2n}{n+1}\left\|\nabla H\right\|^{2}+\left\|A\right\|^{2}-\left\|A\right\|^{4}-\lambda\langle A^{2},A\rangle.

Multiplying this inequality by η2\eta^{2} and integrating by parts with Lemma 8.1,

2ΣηAη,Aγn(x)𝑑x=12Ση2,A2γn(x)𝑑x=12Ση2A2γn(x)𝑑x\displaystyle-2\int_{\Sigma}\eta\left\|A\right\|\langle\nabla\eta,\nabla\left\|A\right\|\rangle\gamma_{n}(x)dx=-\frac{1}{2}\int_{\Sigma}\langle\nabla\eta^{2},\nabla\left\|A\right\|^{2}\rangle\gamma_{n}(x)dx=\frac{1}{2}\int_{\Sigma}\eta^{2}\mathcal{L}\left\|A\right\|^{2}\gamma_{n}(x)dx
Ση2((1+2n+1)A22nn+1H2A4λA2,A)γn(x)𝑑x.\displaystyle\qquad\geq\int_{\Sigma}\eta^{2}\Big{(}\Big{(}1+\frac{2}{n+1}\Big{)}\left\|\nabla\left\|A\right\|\right\|^{2}-\frac{2n}{n+1}\left\|\nabla H\right\|^{2}-\left\|A\right\|^{4}-\lambda\langle A^{2},A\rangle\Big{)}\gamma_{n}(x)dx.

(We removed the A2\left\|A\right\|^{2} term since doing so only decreases the quantity on the right.) Rearranging this inequality and then using the AMGM inequality in the form b2/ε2abεa2b^{2}/\varepsilon-2ab\geq-\varepsilon a^{2},

Σ(η2A4+λη2A2,A+2nn+1η2H2+1εη2A2η2)γn(x)𝑑x\displaystyle\int_{\Sigma}\Big{(}\eta^{2}\left\|A\right\|^{4}+\lambda\eta^{2}\langle A^{2},A\rangle+\frac{2n}{n+1}\eta^{2}\left\|\nabla H\right\|^{2}+\frac{1}{\varepsilon}\eta^{2}\left\|A\right\|^{2}\left\|\nabla\eta\right\|^{2}\Big{)}\gamma_{n}(x)dx (38)
(1+2n+1ε)Ση2A2γn(x)𝑑x.\displaystyle\qquad\qquad\geq\Big{(}1+\frac{2}{n+1}-\varepsilon\Big{)}\int_{\Sigma}\eta^{2}\left\|\nabla\left\|A\right\|\right\|^{2}\gamma_{n}(x)dx.

Substituting (38) into (37),

Ση2A4γn(x)𝑑x1+ε1+2n+1εΣη2A4γn(x)𝑑x\displaystyle\int_{\Sigma}\eta^{2}\left\|A\right\|^{4}\gamma_{n}(x)dx\leq\frac{1+\varepsilon}{1+\frac{2}{n+1}-\varepsilon}\int_{\Sigma}\eta^{2}\left\|A\right\|^{4}\gamma_{n}(x)dx
+10Σ[η2H2+η2(1+1/ε)(1+η2)A2+|λ|η2(|A,A2|)]γn(x)𝑑x.\displaystyle\quad+10\int_{\Sigma}\Big{[}\eta^{2}\left\|\nabla H\right\|^{2}+\eta^{2}(1+1/\varepsilon)(1+\left\|\nabla\eta\right\|^{2})\left\|A\right\|^{2}+\left|\lambda\right|\eta^{2}\Big{(}\left|\langle A,A^{2}\rangle\right|\Big{)}\Big{]}\gamma_{n}(x)dx.

Note that |A,A2|A3\left|\langle A,A^{2}\rangle\right|\leq\left\|A\right\|^{3}. Using the AMGM inequality in the form 2a3a4ε+a2/ε2a^{3}\leq a^{4}\varepsilon+a^{2}/\varepsilon, we then get

Ση2A4γn(x)𝑑x\displaystyle\int_{\Sigma}\eta^{2}\left\|A\right\|^{4}\gamma_{n}(x)dx (10ε|λ|+1+ε1+2n+1ε)Ση2A4γn(x)𝑑x\displaystyle\leq\Big{(}10\varepsilon\left|\lambda\right|+\frac{1+\varepsilon}{1+\frac{2}{n+1}-\varepsilon}\Big{)}\int_{\Sigma}\eta^{2}\left\|A\right\|^{4}\gamma_{n}(x)dx
+10Ση2(H2+(1+(1+|λ|)/ε)(1+η2)A2)γn(x)𝑑x.\displaystyle\qquad+10\int_{\Sigma}\eta^{2}\Big{(}\left\|\nabla H\right\|^{2}+(1+(1+\left|\lambda\right|)/\varepsilon)(1+\left\|\nabla\eta\right\|^{2})\left\|A\right\|^{2}\Big{)}\gamma_{n}(x)dx.

Now, choose ε<1/(20(n+1)(|λ|+1))\varepsilon<1/(20(n+1)(\left|\lambda\right|+1)), so that 10|λ|ε+1+ε1+2n+1ε<110\left|\lambda\right|\varepsilon+\frac{1+\varepsilon}{1+\frac{2}{n+1}-\varepsilon}<1. We then can move the η2A4\eta^{2}\left\|A\right\|^{4} term on the right side to the left side to get some cε>0c_{\varepsilon}>0 such that

Ση2A4γn(x)𝑑x\displaystyle\int_{\Sigma}\eta^{2}\left\|A\right\|^{4}\gamma_{n}(x)dx cεΣη2[H2+A2(1+η2)]γn(x)𝑑x\displaystyle\leq c_{\varepsilon}\int_{\Sigma}\eta^{2}\Big{[}\left\|\nabla H\right\|^{2}+\left\|A\right\|^{2}(1+\left\|\nabla\eta\right\|^{2})\Big{]}\gamma_{n}(x)dx
cεΣη2[A2(1+x2+η2)]γn(x)𝑑x.\displaystyle\leq c_{\varepsilon}\int_{\Sigma}\eta^{2}\Big{[}\left\|A\right\|^{2}(1+\left\|x\right\|^{2}+\left\|\nabla\eta\right\|^{2})\Big{]}\gamma_{n}(x)dx.

In the last line, we used the inequality H2A2x2\left\|\nabla H\right\|^{2}\leq\left\|A\right\|^{2}\left\|x\right\|^{2}. This follows by Lemma 4.1, since H(x)=x,N+λH(x)=\langle x,N\rangle+\lambda, so for any 1in1\leq i\leq n, eiH(x)=j=1naijx,ej\nabla_{e_{i}}H(x)=-\sum_{j=1}^{n}a_{ij}\langle x,e_{j}\rangle.

We now choose a sequence of η=ηr\eta=\eta_{r} increasing to 11 as rr\to\infty so that the η2\left\|\nabla\eta\right\|^{2} term vanishes. This is possible due to the assumptions that δ(Ω)<\delta(\partial^{*}\Omega)<\infty and the Hausdorff dimension of ΩΩ\partial\Omega\setminus\partial^{*}\Omega is at most n4n-4. Such functions are constructed and this estimate is made in [Zhu20, Lemma 6.4]:

ΣA2ηr2ηr2γn(x)𝑑xc(δ)(rne(r4)2/4+r1),r>1.\int_{\Sigma}\left\|A\right\|^{2}\left\|\nabla\eta_{r}\right\|^{2}\eta_{r}^{2}\gamma_{n}(x)dx\leq c(\delta)(r^{n}e^{-(r-4)^{2}/4}+r^{-1}),\qquad\forall r>1. (39)

It therefore follows from Corollary 8.3 applied to Σ=Ω\Sigma=\partial^{*}\Omega that

ΣA4γn(x)𝑑x<.\int_{\Sigma}\left\|A\right\|^{4}\gamma_{n}(x)dx<\infty. (40)

It then follows from (38) that ΣA2γn(x)𝑑x<.\int_{\Sigma}\left\|\nabla\left\|A\right\|\right\|^{2}\gamma_{n}(x)dx<\infty.

Finally, multiplying the above equality A2=2A2+2A22A42λA2,A\mathcal{L}\left\|A\right\|^{2}=2\left\|\nabla A\right\|^{2}+2\left\|A\right\|^{2}-2\left\|A\right\|^{4}-2\lambda\langle A^{2},A\rangle by η2\eta^{2} and integrating by parts with Lemma 8.1, we get

2Ση2(A2A4λA2,A)γn(x)𝑑xΣη2A2γn(x)𝑑x\displaystyle 2\int_{\Sigma}\eta^{2}(\left\|\nabla A\right\|^{2}-\left\|A\right\|^{4}-\lambda\langle A^{2},A\rangle)\gamma_{n}(x)dx\leq\int_{\Sigma}\eta^{2}\mathcal{L}\left\|A\right\|^{2}\gamma_{n}(x)dx
=4ΣηAη,Aγn(x)dx2Ση2A2+A2η2γn(x)dx.\displaystyle\qquad=-4\int_{\Sigma}\eta\left\|A\right\|\langle\nabla\eta,\nabla\left\|A\right\|\rangle\gamma_{n}(x)dx\leq 2\int_{\Sigma}\eta^{2}\left\|\nabla\left\|A\right\|\right\|^{2}+\left\|A\right\|^{2}\left\|\nabla\eta\right\|^{2}\rangle\gamma_{n}(x)dx.

Then the A4\left\|A\right\|^{4} integral is finite by (40), the A2\left\|\nabla\left\|A\right\|\right\|^{2} integral is finite, the last term has a finite integral by (39), so the integral of A2\left\|\nabla A\right\|^{2} is also finite. ∎

Recall the definition of maxβ(A)\mathrm{max}_{\beta}(A) from (15). In order to integrate by parts with maxβ(A)\mathrm{max}_{\beta}(A) via Lemma 8.2, we need the following Lemma

Lemma 9.5.

Let β>0\beta>0. Then

supβ>0Σmaxβ(A)2γn(x)dx<,supβ>1Σ|maxβ(A)maxβ(A)|γn(x)dx<,\sup_{\beta>0}\int_{\Sigma}\left\|\nabla\mathrm{max}_{\beta}(A)\right\|^{2}\gamma_{n}(x)\,\mathrm{d}x<\infty,\qquad\qquad\sup_{\beta>1}\int_{\Sigma}\left|\mathrm{max}_{\beta}(A)\mathcal{L}\mathrm{max}_{\beta}(A)\right|\gamma_{n}(x)\,\mathrm{d}x<\infty,
supβ>1Σmaxβ(A)maxβ(A)6/5γn(x)dx<.\sup_{\beta>1}\int_{\Sigma}\left\|\mathrm{max}_{\beta}(A)\nabla\mathrm{max}_{\beta}(A)\right\|^{6/5}\gamma_{n}(x)\,\mathrm{d}x<\infty.
Proof.

From (17) and Lemma 9.4, we have for any β>0\beta>0

Σmaxβ(A)2γn(x)dxΣA2γn(x)dx<.\int_{\Sigma}\left\|\nabla\mathrm{max}_{\beta}(A)\right\|^{2}\gamma_{n}(x)\,\mathrm{d}x\leq\int_{\Sigma}\left\|\nabla A\right\|^{2}\gamma_{n}(x)\,\mathrm{d}x<\infty.

Note that maxβ(A)lognβ+A\max_{\beta}(A)\leq\frac{\log n}{\beta}+\left\|A\right\|. From (17), Hölder’s inequality with exponents 33 and 3/23/2, and Lemma 9.4,

Σmaxβ(A)maxβ(A)6/5γn(x)dx\displaystyle\int_{\Sigma}\left\|\mathrm{max}_{\beta}(A)\nabla\mathrm{max}_{\beta}(A)\right\|^{6/5}\gamma_{n}(x)\,\mathrm{d}x
(Σ(lognβ+A)18/5γn(x)dx)1/3(ΣA18/10γn(x)dx)2/3<.\displaystyle\qquad\qquad\qquad\leq\Big{(}\int_{\Sigma}\Big{(}\frac{\log n}{\beta}+\left\|A\right\|\Big{)}^{18/5}\gamma_{n}(x)\,\mathrm{d}x\Big{)}^{1/3}\Big{(}\int_{\Sigma}\left\|\nabla A\right\|^{18/10}\gamma_{n}(x)\,\mathrm{d}x\Big{)}^{2/3}<\infty.

Now, let ϕC0(Σ)\phi\in C_{0}^{\infty}(\Sigma), Let h\colonequalsmax(maxβ(A),0)h\colonequals\max(\mathrm{max}_{\beta}(A),0), and integrate by parts with Lemma 8.1,

Σϕhhγn(x)𝑑x\displaystyle\int_{\Sigma}\phi h\mathcal{L}h\gamma_{n}(x)dx ={xΣ:maxβ(A)0}ϕmaxβ(A)2γn(x)𝑑x\displaystyle=-\int_{\{x\in\Sigma\colon\mathrm{max}_{\beta}(A)\geq 0\}}\phi\left\|\nabla\mathrm{max}_{\beta}(A)\right\|^{2}\gamma_{n}(x)dx (41)
{xΣ:maxβ(A)0}maxβ(A)maxβ(A),ϕγn(x)𝑑x.\displaystyle\qquad-\int_{\{x\in\Sigma\colon\mathrm{max}_{\beta}(A)\geq 0\}}\mathrm{max}_{\beta}(A)\langle\nabla\mathrm{max}_{\beta}(A),\nabla\phi\rangle\gamma_{n}(x)dx.

Let ϕ\phi be an approximation to the identity supported on a ball of radius r>0r>0 as in Lemma 9.4. Letting rr\to\infty and using Σmaxβ(A)maxβ(A)6/5γn(x)dx<\int_{\Sigma}\left\|\mathrm{max}_{\beta}(A)\nabla\mathrm{max}_{\beta}(A)\right\|^{6/5}\gamma_{n}(x)\,\mathrm{d}x<\infty, the last integral in (41) goes to zero by e.g. [Zhu20, Corollary 5.3]. Also, maxβ(A)=(17)Tr(eβAA)/Tr(eβA)\nabla\mathrm{max}_{\beta}(A)\stackrel{{\scriptstyle\eqref{chat3}}}{{=}}\mathrm{Tr}(e^{\beta A}\nabla A)/\mathrm{Tr}(e^{\beta A}), so that maxβ(A)2A2\left\|\nabla\mathrm{max}_{\beta}(A)\right\|^{2}\leq\left\|\nabla A\right\|^{2}, so the term Σϕmaxβ(A)2γn(x)𝑑x\int_{\Sigma}\phi\left\|\nabla\mathrm{max}_{\beta}(A)\right\|^{2}\gamma_{n}(x)dx in (41) converges as rr\to\infty, since ΣA2γn(x)dx<\int_{\Sigma}\left\|\nabla A\right\|^{2}\gamma_{n}(x)\,\mathrm{d}x<\infty by Lemma 9.4. We therefore conclude that the first integral in (41) converges as rr\to\infty. Observe that

Σϕh(h)γn(x)𝑑x={xΣ:maxβ(A)0}ϕmaxβ(A)(maxβ(A))γn(x)𝑑x\displaystyle\int_{\Sigma}\phi h\mathcal{L}(h)\gamma_{n}(x)dx=\int_{\{x\in\Sigma\colon\mathrm{max}_{\beta}(A)\geq 0\}}\phi\,\mathrm{max}_{\beta}(A)\mathcal{L}(\mathrm{max}_{\beta}(A))\gamma_{n}(x)dx (42)
=(16){xΣ:maxβ(A)0}ϕmaxβ(A)[Tr(eβATr(eβA)A)\displaystyle\stackrel{{\scriptstyle\eqref{lsofeq}}}{{=}}\int_{\{x\in\Sigma\colon\mathrm{max}_{\beta}(A)\geq 0\}}\phi\,\mathrm{max}_{\beta}(A)\Big{[}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\mathcal{L}A\Big{)}
+βi=1nTr(eβATr(eβA)[eiAInTr(eβATr(eβA)eiA)]2)]γn(x)dx\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad+\beta\sum_{i=1}^{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Big{[}\nabla_{e_{i}}A-I_{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\nabla_{e_{i}}A\Big{)}\Big{]}^{2}\Big{)}\Big{]}\gamma_{n}(x)dx
=(13){xΣ:maxβ(A)0}ϕmaxβ(A)[Tr(eβATr(eβA)[2AλA2A(A2+1)])\displaystyle\stackrel{{\scriptstyle\eqref{three9p}}}{{=}}\int_{\{x\in\Sigma\colon\mathrm{max}_{\beta}(A)\geq 0\}}\phi\,\mathrm{max}_{\beta}(A)\Big{[}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}[2A-\lambda A^{2}-A(\left\|A\right\|^{2}+1)]\Big{)}
+βi=1nTr(eβATr(eβA)[eiAInTr(eβATr(eβA)eiA)]2)]γn(x)dx.\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad+\beta\sum_{i=1}^{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Big{[}\nabla_{e_{i}}A-I_{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\nabla_{e_{i}}A\Big{)}\Big{]}^{2}\Big{)}\Big{]}\gamma_{n}(x)dx.

For any r,β>1r,\beta>1 we can bound all terms except the last one in absolute value by a constant plus

Σ(2A2+|λ|A3+A4+A3)γn(x)𝑑x.\displaystyle\int_{\Sigma}\Big{(}2\left\|A\right\|^{2}+\left|\lambda\right|\left\|A\right\|^{3}+\left\|A\right\|^{4}+\left\|A\right\|^{3}\Big{)}\gamma_{n}(x)dx.

This quantity is finite by Lemma 9.4. As observed in (41), the first term in (42) converges to a finite value as rr\to\infty. Letting then rr\to\infty, we conclude that

{xΣ:maxβ(A)0}maxβ(A)βi=1nTr(eβATr(eβA)[eiAInTr(eβATr(eβA)eiA)]2)γn(x)dx<.\int_{\{x\in\Sigma\colon\mathrm{max}_{\beta}(A)\geq 0\}}\mathrm{max}_{\beta}(A)\beta\sum_{i=1}^{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Big{[}\nabla_{e_{i}}A-I_{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\nabla_{e_{i}}A\Big{)}\Big{]}^{2}\Big{)}\gamma_{n}(x)\,\mathrm{d}x<\infty.

Repeating the above argument with h\colonequalsmax(maxβ(A),0)h\colonequals\max(-\mathrm{max}_{\beta}(A),0), we also obtain

{xΣ:maxβ(A)0}maxβ(A)βi=1nTr(eβATr(eβA)[eiAInTr(eβATr(eβA)eiA)]2)γn(x)dx<.\int_{\{x\in\Sigma\colon\mathrm{max}_{\beta}(A)\leq 0\}}-\mathrm{max}_{\beta}(A)\beta\sum_{i=1}^{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Big{[}\nabla_{e_{i}}A-I_{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\nabla_{e_{i}}A\Big{)}\Big{]}^{2}\Big{)}\gamma_{n}(x)\,\mathrm{d}x<\infty.

Therefore,

Σi=1n|maxβ(A)|βi=1nTr(eβATr(eβA)[eiAInTr(eβATr(eβA)eiA)]2)γn(x)dx<.\int_{\Sigma}\sum_{i=1}^{n}\left|\mathrm{max}_{\beta}(A)\right|\beta\sum_{i=1}^{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Big{[}\nabla_{e_{i}}A-I_{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\nabla_{e_{i}}A\Big{)}\Big{]}^{2}\Big{)}\gamma_{n}(x)\,\mathrm{d}x<\infty.

We finally can conclude that

Σ|maxβ(A)maxβ(A)|γn(x)dx<.\int_{\Sigma}\left|\mathrm{max}_{\beta}(A)\mathcal{L}\mathrm{max}_{\beta}(A)\right|\gamma_{n}(x)\,\mathrm{d}x<\infty.

All above upper bounds on these integrals did not depend on β>1\beta>1, so we can additionally take the supremum over β>1\beta>1 to conclude the proof. (Note that all quantities in (42) are finite after taking supβ>1\sup_{\beta>1}, except possible for the last term. Therefore, the last term is also bounded after taking the supremum over β>1\beta>1.)

10. Stable Mean Convex Sets are Convex

Theorem 10.1.

Let Ωn+1\Omega\subseteq\mathbb{R}^{n+1}. Assume that Ω\Omega is mean convex (i.e. HH does not change sign on Ω\partial\Omega). Assume Ω×\Omega\times\mathbb{R} minimizes Problem 2.1. Then Ω\Omega or Ωc\Omega^{c} is convex.

Proof.

From Lemma 4.1, there exists λ\lambda\in\mathbb{R} such that

H(x)=x,N(x)+λ,xΣ.H(x)=\langle x,N(x)\rangle+\lambda,\qquad\forall\,x\in\Sigma.

The case λ=0\lambda=0 was already treated in Theorem 6.1. We may therefore assume that λ0\lambda\neq 0. Replacing Ω\Omega with Ωc\Omega^{c} (i.e. changing the direction of the unit exterior normal vector, which changes the signs of HH and NN and therefore of λ\lambda), we may assume that λ<0\lambda<0. We may additionally assume that H0H\geq 0 on Σ\Sigma, since the case H0H\leq 0 and λ<0\lambda<0 was treated already in Theorem 7.1.

Since we assumed H0H\geq 0 and λ<0\lambda<0, we may freely use Lemma 9.4.

Recall the definition of maxβ(A)\mathrm{max}_{\beta}(A) in (15). Let h\colonequalsmax(maxβ(A),0)h\colonequals\max(\mathrm{max}_{\beta}(A),0). We may assume that there exists β>0\beta>0 such that h>0h>0 on a set of positive measure on Σ\Sigma, otherwise all eigenvalues of AA are negative, i.e. Ω\Omega is convex, and the proof is complete. From Lemma 5.3,

ΣhLhγn(x)dx=(10)(11)Σ(hh+h2(A2+1))γn(x)dx\displaystyle\int_{\Sigma}hLh\gamma_{n}(x)\,\mathrm{d}x\stackrel{{\scriptstyle\eqref{three4.3}\wedge\eqref{three4.5}}}{{=}}\int_{\Sigma}\Big{(}h\mathcal{L}h+h^{2}(\left\|A\right\|^{2}+1)\Big{)}\gamma_{n}(x)\,\mathrm{d}x (43)
=Σ[h2(A2+1)+hTr(eβATr(eβA)A)\displaystyle=\int_{\Sigma}\Big{[}h^{2}(\left\|A\right\|^{2}+1)+h\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\mathcal{L}A\Big{)}
+βi=1nTr(eβATr(eβA)[eiAInTr(eβATr(eβA)eiA)]2)γn(x)dx\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad+\beta\sum_{i=1}^{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Big{[}\nabla_{e_{i}}A-I_{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\nabla_{e_{i}}A\Big{)}\Big{]}^{2}\Big{)}\gamma_{n}(x)\,\mathrm{d}x
=(11)Σ[h2(A2+1)h(A2+1)Tr(eβATr(eβA)A)+hTr(eβATr(eβA)LA)\displaystyle\stackrel{{\scriptstyle\eqref{three4.5}}}{{=}}\int_{\Sigma}\Big{[}h^{2}(\left\|A\right\|^{2}+1)-h(\left\|A\right\|^{2}+1)\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}A\Big{)}+h\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}LA\Big{)}
+βi=1nTr(eβATr(eβA)[eiAInTr(eβATr(eβA)eiA)]2)γn(x)dx.\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad+\beta\sum_{i=1}^{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\Big{[}\nabla_{e_{i}}A-I_{n}\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}(e^{\beta A})}\nabla_{e_{i}}A\Big{)}\Big{]}^{2}\Big{)}\gamma_{n}(x)\,\mathrm{d}x.

(The quantity ΩhLhγn(x)dx\int_{\partial\Omega}hLh\gamma_{n}(x)\,\mathrm{d}x is finite a priori by Lemma 9.5. And when we integrate by parts with hh in the second variation formula, that will be justified by Lemma 9.5 and Lemma 8.2.) The final term is nonnegative. Using this and (13),

ΣhLhγn(x)dxΣ(h(A2+1)[hTr(AeβATreβA)]+hTr(eβATreβA[2AλA2])γn(x)dx.\int_{\Sigma}hLh\gamma_{n}(x)\,\mathrm{d}x\geq\int_{\Sigma}\Big{(}h(\left\|A\right\|^{2}+1)\Big{[}h-\mathrm{Tr}\Big{(}\frac{Ae^{\beta A}}{\mathrm{Tr}e^{\beta A}}\Big{)}\Big{]}+h\mathrm{Tr}\Big{(}\frac{e^{\beta A}}{\mathrm{Tr}e^{\beta A}}[2A-\lambda A^{2}]\Big{)}\gamma_{n}(x)\,\mathrm{d}x. (44)

We now let β\beta\to\infty. Then hh converges to the maximum of 0 and the maximum eigenvalue αmax(A)\alpha_{\mathrm{max}}(A) of AA, as does Tr(AeβATreβA)\mathrm{Tr}(\frac{Ae^{\beta A}}{\mathrm{Tr}e^{\beta A}}). So, the first term vanishes by the Dominated Convergence Theorem and Lemma 9.4. Denote g\colonequalsmax(αmax(A),0)g\colonequals\max(\alpha_{\mathrm{max}}(A),0). Similarly, the Dominated Convergence Theorem gives

limβΣhLhγn(x)dxΣg(2gλg))γn(x)dx.\lim_{\beta\to\infty}\int_{\Sigma}hLh\gamma_{n}(x)\,\mathrm{d}x\geq\int_{\Sigma}g(2g-\lambda g)\Big{)}\gamma_{n}(x)\,\mathrm{d}x. (45)

Recall that we assumed that λ<0\lambda<0 and h0h\geq 0, therefore g0g\geq 0 and

limβΣhLhγn(x)dx>Σ2g2γn(x)dx.\lim_{\beta\to\infty}\int_{\Sigma}hLh\gamma_{n}(x)\,\mathrm{d}x>\int_{\Sigma}2g^{2}\gamma_{n}(x)\,\mathrm{d}x.

(Equality cannot occur here since g=0g=0 would imply that Ω\Omega is convex, and the proof would be complete.) Since limβΣh2γn(x)dx=Σg2γn(x)dx\lim_{\beta\to\infty}\int_{\Sigma}h^{2}\gamma_{n}(x)\,\mathrm{d}x=\int_{\Sigma}g^{2}\gamma_{n}(x)\,\mathrm{d}x, we have

limβΣhLhγn(x)dxΣh2γn(x)dx>2.\lim_{\beta\to\infty}\frac{\int_{\Sigma}hLh\gamma_{n}(x)\,\mathrm{d}x}{\int_{\Sigma}h^{2}\gamma_{n}(x)\,\mathrm{d}x}>2. (46)

we can now conclude the proof as in Theorem 7.1.

Let f\colonequalshf\colonequals h in Lemma 5.4. Note that h(x)=h(x)h(x)=h(-x) for all xΣx\in\Sigma. Then Lemma 5.4 and (46) imply that the variation of Ω×\Omega\times\mathbb{R} correspond to gg satisfies

d2ds2|s=0(Σ×)(s)γn(x)𝑑x<0.\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}\Big{|}_{s=0}\int_{(\Sigma\times\mathbb{R})^{(s)}}\gamma_{n}(x)dx<0.

This inequality violates the minimality of Ω\Omega, achieving a contradiction, and completing the proof. ∎

11. Symmetric Minimal Cones are Unstable

In this Section, we prove Case 3 of Theorem 6.1. That is, we show that symmetric minimal cones are unstable for the Gaussian surface area, within the category of symmetric sets. That is, a symmetric minimal cone can be perturbed to another nearby symmetric set with smaller Gaussian surface area, in a way that preserves the Gaussian volume of the set.

Proof of Case 3 of Theorem 6.1.

In this proof, in order to match the notation and formulas from [Zhu20], we use a factor of 44 in our Gaussian density, rather than a factor of 22.

From the regularity part of Lemma 4.3, since H(x)=x,N(x)=0H(x)=\langle x,N(x)\rangle=0 for all xΣx\in\Sigma and Σ\Sigma is a cone (i.e. Σ\Sigma has a singularity at the origin), we must have n7n\geq 7. Let Sn\colonequals{xn+1:x=1}S^{n}\colonequals\{x\in\mathbb{R}^{n+1}\colon\left\|x\right\|=1\}. Let W\colonequalsΣSnW\colonequals\Sigma\cap S^{n} denote the cone Σ\Sigma intersected with the sphere, and let MM denote the regular part of WW. From the second variation formula, Lemma 4.2, recall that

d2ds2|s=0Σ(s)ex2/4(4π)n/2dx=Σf(x)Lf(x)ex2/4(4π)n/2dx,\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}\Big{|}_{s=0}\int_{\Sigma^{(s)}}e^{-\left\|x\right\|^{2}/4}(4\pi)^{-n/2}\,\mathrm{d}x=-\int_{\Sigma}f(x)L^{\prime}f(x)e^{-\left\|x\right\|^{2}/4}(4\pi)^{-n/2}\,\mathrm{d}x, (47)

where, as opposed to (11), we now have [Zhu20, Equation (1.12)],

L\colonequalsΔ12x,+12+A2.L^{\prime}\colonequals\Delta-\frac{1}{2}\langle x,\nabla\rangle+\frac{1}{2}+\left\|A\right\|^{2}. (48)

For any xn+1x\in\mathbb{R}^{n+1}, define r=r(x)\colonequalsxr=r(x)\colonequals\left\|x\right\|. Using [Zhu20, Equation (4.18)], we can decompose the operator LL^{\prime} from (48) into its radial component and its spherical component

L=r2(L~(n1)+L1).L^{\prime}=r^{-2}(\widetilde{L}-(n-1)+L_{1}). (49)

Here L~\widetilde{L} is the spherical component [Zhu20, Equation (4.13)]

L~\colonequalsΔM+A~2+(n1),\widetilde{L}\colonequals\Delta_{M}+\|\widetilde{A}\|^{2}+(n-1),

where A~\widetilde{A} denotes the second fundamental form of MM, and L1L_{1} is the radial component

L1\colonequalsr22r2+(n1)rrr32r+r22.L_{1}\colonequals r^{2}\frac{\partial^{2}}{\partial r^{2}}+(n-1)r\frac{\partial}{\partial r}-\frac{r^{3}}{2}\frac{\partial}{\partial r}+\frac{r^{2}}{2}.

We can then rewrite the quadratic form from (47) as

Σf(x)Lf(x)ex24dx(4π)n2=ΣSnr=0r=rn3f(r,θ)(L~(n1)+L1)f(r,θ)er24drdθ(4π)n2.\int_{\Sigma}f(x)L^{\prime}f(x)\frac{e^{-\frac{\left\|x\right\|^{2}}{4}}\,\mathrm{d}x}{(4\pi)^{\frac{n}{2}}}=\int_{\Sigma\cap S^{n}}\int_{r=0}^{r=\infty}r^{n-3}f(r,\theta)(\widetilde{L}-(n-1)+L_{1})f(r,\theta)e^{-\frac{r^{2}}{4}}\frac{\mathrm{d}r\mathrm{d}\theta}{(4\pi)^{\frac{n}{2}}}. (50)

We first note that

L1r=(n1)r.L_{1}r=(n-1)r. (51)

Let t(r)\colonequalsr+42nrt(r)\colonequals r+\frac{4-2n}{r}, \forall r>0r>0. Then t(r)t(r) is an eigenfunction of L1L_{1}, since

L1t(r)\displaystyle L_{1}t(r) =(51)(n1)r+(42n)(r22r3+(n1)r(r2)12r3(r2)+r22r1)\displaystyle\stackrel{{\scriptstyle\eqref{four9}}}{{=}}(n-1)r+(4-2n)\Big{(}r^{2}2r^{-3}+(n-1)r(-r^{-2})-\frac{1}{2}r^{3}(-r^{-2})+\frac{r^{2}}{2}r^{-1}\Big{)}
=(n1)r+(42n)(2r1(n1)r1+r)\displaystyle=(n-1)r+(4-2n)(2r^{-1}-(n-1)r^{-1}+r)
=(n+3)r+(42n)(3n)r1=(n3)f.\displaystyle=(-n+3)r+(4-2n)(3-n)r^{-1}=-(n-3)f.

Let g(r)\colonequalsrg(r)\colonequals r, \forall r>0r>0. Integrating by parts we get

0rn1er2/4dr=2(n2)0rn3er2/4dr.\int_{0}^{\infty}r^{n-1}e^{-r^{2}/4}\,\mathrm{d}r=2(n-2)\int_{0}^{\infty}r^{n-3}e^{-r^{2}/4}\,\mathrm{d}r. (52)

We then obtain

0t(r)g(r)rn3er2/4dr=0[rn1+(42n)rn3]er2/4dr=(52)0.\int_{0}^{\infty}t(r)g(r)r^{n-3}e^{-r^{2}/4}\,\mathrm{d}r=\int_{0}^{\infty}[r^{n-1}+(4-2n)r^{n-3}]e^{-r^{2}/4}\,\mathrm{d}r\stackrel{{\scriptstyle\eqref{four13}}}{{=}}0. (53)

(Recall n7n\geq 7 so this integral is finite a priori.) And if h(r)\colonequalsr+22nrh(r)\colonequals r+\frac{2-2n}{r}, then

0h(r)rn1er2/4dr=0[rn+(22n)rn2]er2/4dr=0.\int_{0}^{\infty}h(r)r^{n-1}e^{-r^{2}/4}\,\mathrm{d}r=\int_{0}^{\infty}[r^{n}+(2-2n)r^{n-2}]e^{-r^{2}/4}\,\mathrm{d}r=0. (54)

So, hh corresponds to a Gaussian volume-preserving perturbation. Also, for all r>0r>0,

h(r)\displaystyle h(r) =(r+42nr)22n42n+r22n42nr\displaystyle=\Big{(}r+\frac{4-2n}{r}\Big{)}\frac{2-2n}{4-2n}+r-\frac{2-2n}{4-2n}r (55)
=(r+42nr)n1n2+r242n=n1n2t(r)1n2g(r).\displaystyle=\Big{(}r+\frac{4-2n}{r}\Big{)}\frac{n-1}{n-2}+r\frac{2}{4-2n}=\frac{n-1}{n-2}t(r)-\frac{1}{n-2}g(r).

So, using (L1(n1))g=(51)0(L_{1}-(n-1))g\stackrel{{\scriptstyle\eqref{four9}}}{{=}}0, we have

(L1(n1))h(r)\displaystyle(L_{1}-(n-1))h(r) =(55)(L1(n1))n1n2t(r)=n1n2((n3)(n1))t(r)\displaystyle\stackrel{{\scriptstyle\eqref{four5}}}{{=}}(L_{1}-(n-1))\frac{n-1}{n-2}t(r)=\frac{n-1}{n-2}(-(n-3)-(n-1))t(r) (56)
=n1n2(2n+4)t(r)=2(n1)t(r).\displaystyle=\frac{n-1}{n-2}(-2n+4)t(r)=-2(n-1)t(r).

Now, we will construct a function f:Σf\colon\Sigma\to\mathbb{R} using a product of a radial function and a spherical function. Define

f(r,θ)\colonequalsϕ(θ)h(r),r>0,θM.f(r,\theta)\colonequals\phi(\theta)h(r),\qquad\forall\,r>0,\,\,\forall\,\theta\in M. (57)

Starting with the radial integral in (50), we have

r=0r=rn3f(L~(n1)+L1)fer2/4dr\displaystyle\int_{r=0}^{r=\infty}r^{n-3}f(\widetilde{L}-(n-1)+L_{1})fe^{-r^{2}/4}\,\mathrm{d}r (58)
=(56)r=0r=rn3f[L~f2(n1)t(r)ϕ]er2/4dr\displaystyle\stackrel{{\scriptstyle\eqref{four6}}}{{=}}\int_{r=0}^{r=\infty}r^{n-3}f\Big{[}\widetilde{L}f-2(n-1)t(r)\phi\Big{]}e^{-r^{2}/4}\,\mathrm{d}r
=(55)r=0r=rn3[fL~f(n1n2t(r)1n2g(r))(2(n1)[t(r)]ϕ2)]er2/4dr\displaystyle\stackrel{{\scriptstyle\eqref{four5}}}{{=}}\int_{r=0}^{r=\infty}r^{n-3}\Big{[}f\widetilde{L}f-\Big{(}\frac{n-1}{n-2}t(r)-\frac{1}{n-2}g(r)\Big{)}(2(n-1)[t(r)]\phi^{2})\Big{]}e^{-r^{2}/4}\,\mathrm{d}r
=(53)r=0r=rn3[fL~fn1n22(n1)[t(r)]2ϕ2]er2/4dr\displaystyle\stackrel{{\scriptstyle\eqref{four4}}}{{=}}\int_{r=0}^{r=\infty}r^{n-3}\Big{[}f\widetilde{L}f-\frac{n-1}{n-2}2(n-1)[t(r)]^{2}\phi^{2}\Big{]}e^{-r^{2}/4}\,\mathrm{d}r
=(53)(55)r=0r=rn3[ϕL~ϕ[(n1n2)2[t(r)]2+1(n2)2[g(r)]2]\displaystyle\stackrel{{\scriptstyle\eqref{four4}\wedge\eqref{four5}}}{{=}}\int_{r=0}^{r=\infty}r^{n-3}\Big{[}\phi\widetilde{L}\phi\Big{[}\Big{(}\frac{n-1}{n-2}\Big{)}^{2}[t(r)]^{2}+\frac{1}{(n-2)^{2}}[g(r)]^{2}\Big{]}
n1n22(n1)[t(r)]2ϕ2]er2/4dr.\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\frac{n-1}{n-2}2(n-1)[t(r)]^{2}\phi^{2}\Big{]}e^{-r^{2}/4}\,\mathrm{d}r.

(Recall n7n\geq 7 so the above integrals are finite.) (Even though hh is unbounded near the origin, we can use hh in the second variation formula by multiplying by cutoff functions, as in [Zhu20, Lemma 6.6, Proposition 6.7].)

We now use [Zhu20, Theorem 0.3] (see also the proof of Theorem 8.3 in [Zhu20]) to find a nonnegative Dirichlet eigenfunction ϕ:M\phi\colon M\to\mathbb{R} of L~\widetilde{L} such that

L~ϕ=κϕ,\widetilde{L}\phi=\kappa\phi, (59)

with κ2(n1)\kappa\geq 2(n-1), and such that ϕ(θ)=ϕ(θ)\phi(-\theta)=\phi(\theta) for all θM\theta\in M. (Since Ω\Omega is symmetric, Σ\Sigma cannot be a hyperplane through the origin, i.e. MM is not totally geodesic, so [Zhu20, Theorem 0.3] applies. Also, since ϕ\phi is nonnegative and Σ\Sigma is symmetric, ϕ()+ϕ()\phi(\cdot)+\phi(-\cdot) is a nonnegative eigenfunction of L~\widetilde{L}, i.e. we may assume a priori that ϕ\phi itself is symmetric.)

Plugging (59) into (58) and using the inequalities [g(r)]2>0[g(r)]^{2}>0 and (n1n2)2>n1n2>1\big{(}\frac{n-1}{n-2}\big{)}^{2}>\frac{n-1}{n-2}>1,

r=0r=rn3f(L~(n1)+L1)fer2/4dr\displaystyle\int_{r=0}^{r=\infty}r^{n-3}f(\widetilde{L}-(n-1)+L_{1})fe^{-r^{2}/4}\,\mathrm{d}r (60)
=r=0r=rn3[κϕ2[(n1n2)2[t(r)]2+1(n2)2[g(r)]2]\displaystyle=\int_{r=0}^{r=\infty}r^{n-3}\Big{[}\kappa\phi^{2}\Big{[}\Big{(}\frac{n-1}{n-2}\Big{)}^{2}[t(r)]^{2}+\frac{1}{(n-2)^{2}}[g(r)]^{2}\Big{]}
n1n22(n1)[t(r)]2ϕ2]er2/4dr\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\frac{n-1}{n-2}2(n-1)[t(r)]^{2}\phi^{2}\Big{]}e^{-r^{2}/4}\,\mathrm{d}r
>r=0r=rn3[κϕ2[t(r)]22(n1)[t(r)]2ϕ2]er2/4dr\displaystyle>\int_{r=0}^{r=\infty}r^{n-3}\Big{[}\kappa\phi^{2}[t(r)]^{2}-2(n-1)[t(r)]^{2}\phi^{2}\Big{]}e^{-r^{2}/4}\,\mathrm{d}r
=[κϕ22(n1)ϕ2]r=0r=rn3(t(r))2er2/4dr.\displaystyle=\Big{[}\kappa\phi^{2}-2(n-1)\phi^{2}\Big{]}\int_{r=0}^{r=\infty}r^{n-3}(t(r))^{2}e^{-r^{2}/4}\,\mathrm{d}r.

Combining (47), (50) and (60), then using κ2(n1)\kappa\geq 2(n-1) [Zhu20, Theorem 0.3],

d2ds2|s=0Σ(s)ex2/4(4π)n/2dx\displaystyle\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}\Big{|}_{s=0}\int_{\Sigma^{(s)}}e^{-\left\|x\right\|^{2}/4}(4\pi)^{-n/2}\,\mathrm{d}x (61)
<M[κ[ϕ(θ)]2+2(n1)[ϕ(θ)]2]dθr=0r=rn3(t(r))2er2/4dr(4π)n/20.\displaystyle\quad<\int_{M}\Big{[}-\kappa[\phi(\theta)]^{2}+2(n-1)[\phi(\theta)]^{2}\Big{]}\,\mathrm{d}\theta\cdot\int_{r=0}^{r=\infty}r^{n-3}(t(r))^{2}e^{-r^{2}/4}\,\mathrm{d}r\cdot(4\pi)^{-n/2}\leq 0.

Then (61) implies that we have found a function ff such that f(x)=f(x)f(-x)=f(x) for all xΣx\in\Sigma with d2ds2|s=0Σ(s)ex2/4(4π)n/2dx<0\frac{\mathrm{d}^{2}}{\mathrm{d}s^{2}}|_{s=0}\int_{\Sigma^{(s)}}e^{-\left\|x\right\|^{2}/4}(4\pi)^{-n/2}\,\mathrm{d}x<0. Moreover, from (54), (57) and Fubini’s Theorem,

Σf(x)ex2/4dx=0.\int_{\Sigma}f(x)e^{-\left\|x\right\|^{2}/4}\,\mathrm{d}x=0.

That is, we have shown that Σ\Sigma is unstable within the category of symmetric sets, i.e. Ω\Omega cannot minimize Problem 2.1. ∎

Acknowledgement. Thanks to Galyna Livshyts and Jonathan Zhu for helpful discussions.

References

  • [AFP00] Luigi Ambrosio, Nicola Fusco, and Diego Pallara, Functions of bounded variation and free discontinuity problems, Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 2000. MR 1857292
  • [Bar01] Franck Barthe, An isoperimetric result for the Gaussian measure and unconditional sets, Bulletin of the London Mathematical Society 33 (2001), 408–416.
  • [BBJ17] Marco Barchiesi, Alessio Brancolini, and Vesa Julin, Sharp dimension free quantitative estimates for the Gaussian isoperimetric inequality, Ann. Probab. 45 (2017), no. 2, 668–697. MR 3630285
  • [BJ20] Marco Barchiesi and Vesa Julin, Symmetry of minimizers of a Gaussian isoperimetric problem, Probab. Theory Related Fields 177 (2020), no. 1-2, 217–256. MR 4095016
  • [Bob97] S. G. Bobkov, An isoperimetric inequality on the discrete cube, and an elementary proof of the isoperimetric inequality in gauss space, Ann. Probab. 25 (1997), no. 1, 206–214.
  • [Bor75] Christer Borell, The Brunn-Minkowski inequality in Gauss space, Invent. Math. 30 (1975), no. 2, 207–216. MR 0399402 (53 #3246)
  • [Bor85] by same author, Geometric bounds on the Ornstein-Uhlenbeck velocity process, Z. Wahrsch. Verw. Gebiete 70 (1985), no. 1, 1–13. MR 795785 (87k:60103)
  • [Bor03] by same author, The Ehrhard inequality, C. R. Math. Acad. Sci. Paris 337 (2003), no. 10, 663–666. MR 2030108 (2004k:60102)
  • [BS01] A. Burchard and M. Schmuckenschl ger, Comparison theorems for exit times, Geometric & Functional Analysis GAFA 11 (2001), no. 4, 651–692 (English).
  • [Cha17] Jui-En Chang, 1-dimensional solutions of the λ\lambda-self shrinkers, Geometriae Dedicata (2017), 1–16.
  • [CIMW13] Tobias Holck Colding, Tom Ilmanen, William P. Minicozzi, II, and Brian White, The round sphere minimizes entropy among closed self-shrinkers, J. Differential Geom. 95 (2013), no. 1, 53–69. MR 3128979
  • [CL12] Marco Cicalese and Gian Paolo Leonardi, A selection principle for the sharp quantitative isoperimetric inequality, Arch. Ration. Mech. Anal. 206 (2012), no. 2, 617–643. MR 2980529
  • [CM12] Tobias H. Colding and William P. Minicozzi, II, Generic mean curvature flow I: generic singularities, Ann. of Math. (2) 175 (2012), no. 2, 755–833. MR 2993752
  • [CM15] Tobias Holck Colding and William P. Minicozzi, II, Uniqueness of blowups and łojasiewicz inequalities, Ann. of Math. (2) 182 (2015), no. 1, 221–285. MR 3374960
  • [CW18] Qing-Ming Cheng and Guoxin Wei, Complete λ\lambda-hypersurfaces of weighted volume-preserving mean curvature flow, Calc. Var. Partial Differential Equations 57 (2018), no. 2, Art. 32, 21. MR 3763110
  • [Eld15] Ronen Eldan, A two-sided estimate for the gaussian noise stability deficit, Inventiones mathematicae 201 (2015), no. 2, 561–624 (English).
  • [Eva98] Lawrence C. Evans, Partial differential equations, Graduate Studies in Mathematics, vol. 19, American Mathematical Society, Providence, RI, 1998. MR 1625845
  • [Gua18] Qiang Guang, Gap and rigidity theorems of λ\lambda-hypersurfaces, Proc. Amer. Math. Soc. 146 (2018), no. 10, 4459–4471. MR 3834671
  • [Hei21] Steven Heilman, Symmetric convex sets with minimal Gaussian surface area, Amer. J. Math. 143 (2021), no. 1, 53–94. MR 4201779
  • [Hui90] Gerhard Huisken, Asymptotic behavior for singularities of the mean curvature flow, J. Differential Geom. 31 (1990), no. 1, 285–299. MR 1030675
  • [Hui93] by same author, Local and global behaviour of hypersurfaces moving by mean curvature, Differential geometry: partial differential equations on manifolds (Los Angeles, CA, 1990), Proc. Sympos. Pure Math., vol. 54, Amer. Math. Soc., Providence, RI, 1993, pp. 175–191. MR 1216584
  • [Led94] Michel Ledoux, Semigroup proofs of the isoperimetric inequality in Euclidean and Gauss space, Bull. Sci. Math. 118 (1994), no. 6, 485–510. MR 1309086 (96c:49061)
  • [Led96] by same author, Isoperimetry and Gaussian analysis, Lectures on probability theory and statistics (Saint-Flour, 1994), Lecture Notes in Math., vol. 1648, Springer, Berlin, 1996, pp. 165–294. MR 1600888 (99h:60002)
  • [Lee03] John M. Lee, Introduction to smooth manifolds, Graduate Texts in Mathematics, vol. 218, Springer-Verlag, New York, 2003. MR 1930091 (2003k:58001)
  • [Lee22] Tang-Kai Lee, Convexity of λ\lambda-hypersurfaces, Proc. Amer. Math. Soc. 150 (2022), no. 4, 1735–1744. MR 4375760
  • [Mag12] Francesco Maggi, Sets of finite perimeter and geometric variational problems, Cambridge Studies in Advanced Mathematics, vol. 135, Cambridge University Press, Cambridge, 2012, An introduction to geometric measure theory. MR 2976521
  • [Man17] Domenico Angelo La Manna, Local minimality of the ball for the gaussian perimeter, Advances in Calculus of Variations 12 (2017), 193 – 210.
  • [MN15a] Elchanan Mossel and Joe Neeman, Robust dimension free isoperimetry in gaussian space, The Annals of Probability 43 (2015), no. 3, 971–991.
  • [MN15b] by same author, Robust optimality of Gaussian noise stability, J. Eur. Math. Soc. (JEMS) 17 (2015), no. 2, 433–482. MR 3317748
  • [MR15] Matthew McGonagle and John Ross, The hyperplane is the only stable, smooth solution to the isoperimetric problem in Gaussian space, Geom. Dedicata 178 (2015), 277–296. MR 3397495
  • [SC74] V. N. Sudakov and B. S. Cirelson, Extremal properties of half-spaces for spherically invariant measures, Zap. Naučn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 41 (1974), 14–24, 165, Problems in the theory of probability distributions, II. MR 0365680 (51 #1932)
  • [Sim68] James Simons, Minimal varieties in riemannian manifolds, Ann. of Math. (2) 88 (1968), 62–105. MR 0233295
  • [Sto94] Andrew Stone, A density function and the structure of singularities of the mean curvature flow, Calculus of Variations and Partial Differential Equations 2 (1994), 443–480.
  • [Zhu20] Jonathan J. Zhu, On the entropy of closed hypersurfaces and singular self-shrinkers, J. Differential Geom. 114 (2020), no. 3, 551–593. MR 4072205