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

Wulff inequality for minimal submanifolds in Euclidean space

Wenkui Du Department of Mathematics, Massachusetts Institute of Technology duwenkui@mit.edu Yuchao Yi Department of Mathematics, University of California San Diego yuyi@ucsd.edu  and  Ziyi Zhao Institute for Theoretical Sciences, Westlake University zhaoziyi@westlake.edu.cn
Abstract.

In this paper, we prove a Wulff inequality for nn-dimensional minimal submanifolds with boundary in n+m\mathbb{R}^{n+m}, where we associate a nonnegative anisotropic weight Φ:Sn+m1+\Phi:S^{n+m-1}\to\mathbb{R}^{+} to the boundary of minimal submanifolds. The Wulff inequality constant depends only on mm and nn, and is independent of the weights. The inequality is sharp if m=1,2m=1,2 and Φ\Phi is the support function of ellipsoids or certain type of centrally symmetric long convex bodies.

1. Introduction

The isoperimetric inequality is a fundamental geometric inequality that has been extensively studied throughout history since the ancient era of Queen Dido. It states that the round ball minimizes the boundary hypersurface area among all shapes with a given volume. There are several proofs of the isoperimetric inequality and readers can refer to the books and papers [BZ13, Oss78, Cha01]. Later, the isoperimetric inequality for general minimal submanifolds with nonzero codimension in Euclidean space was considered. It was conjectured in [AHO74] that

Conjecutre 1.1.

For any nn dimensional smooth minimal submanifold Σn+m\Sigma\subset\mathbb{R}^{n+m} with smooth boundary Σ\partial\Sigma, the following sharp isoperimetric inequality holds

(1.1) |Σ|n|Bn|1n|Σ|n1n,\displaystyle|\partial\Sigma|\geq n|B^{n}|^{\frac{1}{n}}|\Sigma|^{\frac{n-1}{n}},

and equality holds if and only if Σ\Sigma is an nn-dimensional ball in n\mathbb{R}^{n}. Here, |Σ||\Sigma| denotes the nn-dimensional volume (area) of Σ\Sigma, |Σ||\partial\Sigma| denotes the (n1)(n-1)-dimensional volume (area) of the boundary Σ\partial\Sigma, BnB^{n} is the open unit ball in n\mathbb{R}^{n}, and |Bn||B^{n}| denotes its volume.

Carleman [Car21] proved that the conjecture holds for 22-dimensional minimal surfaces diffeomorphic to a disk in 3\mathbb{R}^{3} via Wirtinger inequality and Almgren [Alm86] showed that the conjecture holds for arbitrary codimensional area minimizing submanifolds in Euclidean space. In a recent breakthrough, Brendle [Bre21] verified the sharp isoperimetric inequality conjecture for minimal submanifolds in the cases of codimension one and two, adapting ideas from optimal mass transportation [MG13] and ABP method of Cabre [Cab08].

A natural generalization of the isoperimetric inequality considered by Wulff [Wul01] is the isoperimetric problem with weighted boundary density.

Definition 1.2.

Let Φ\Phi be a positively one-homogeneous convex function in n\mathbb{R}^{n}, and we define the Φ\Phi-anisotropic perimeter

(1.2) PΦ(Ω)=ΩΦ(ν(x))𝑑S.P_{\Phi}(\Omega)=\int_{\partial\Omega}\Phi(\nu(x))dS.

The set

(1.3) 𝒲={xn:νSn1,xνΦ(ν)}\mathcal{W}=\{x\in\mathbb{R}^{n}:\forall\nu\in S^{n-1},x\cdot\nu\leq\Phi(\nu)\}

is called the corresponding Wulff shape. The support function of 𝒲\mathcal{W} is as follows:111It is well-known that support function and the Wulff shape are mutually determined and the Wulff shape induced by the support function of convex body is precisely the original convex body [Tay78].

(1.4) Ψ(y)=sup{xy:x𝒲}.\Psi(y)=\sup\{x\cdot y:x\in\mathcal{W}\}.

Wulff considered the following question: Given a positive function Φ\Phi, what shapes minimize the Φ\Phi-anisotropic perimeter among the sets of finite perimeter Ωn\Omega\subset\mathbb{R}^{n} with fixed volume. Wulff conjectured that the corresponding minimizers are homothetic to the Wulff shape 𝒲Φ\mathcal{W}_{\Phi} generated by Φ\Phi and the isoperimetric inequality can be extended to the following celebrated Wulff inequality.

Theorem 1.3 (Wulff Theorem [Wul01]).

Let 𝒲\mathcal{W} be an nn-dimensional centrally symmetric convex body in n\mathbb{R}^{n}, and Φ\Phi be the corresponding support function. Then for any set Ωn\Omega\subset\mathbb{R}^{n} of finite perimeter with |Ω|<|\Omega|<\infty, we have

(1.5) PΦ(Ω)|Ω|n1nPΦ(𝒲)|𝒲|n1n.\frac{P_{\Phi}({\Omega})}{|\Omega|^{\frac{n-1}{n}}}\geq\frac{P_{\Phi}({\mathcal{W}})}{|\mathcal{W}|^{\frac{n-1}{n}}}.

Moreover, equality holds if and only if Ω\Omega is homothetic to 𝒲\mathcal{W}, namely Ω=a𝒲+b\Omega=a\mathcal{W}+b for some a>0a>0 and bnb\in\mathbb{R}^{n} up to a set of measure zero.

This result was first stated without proof by Wulff in 1901 [Wul01]. A complete proof of Theorem 1.3 can be found in Taylor’s articles [Tay74, Tay75, Tay78]. Cabré, Ros-Oton and Serra in [CROS16] gave a new proof of Theorem 1.3 via the ABP method. Figalli, Maggi and Pratelli [FMP10] studied the quantitative version of the codimension zero anisotropic isoperimetric inequality. De Rosa, Kolasiński and Santilli [DRKS20] considered the uniqueness of critical points of the codimensional zero anisotropic isoperimetric problem and established Heintze-Karcher type inequality.

A natural question is whether the Wulff inequality holds for minimal submanifolds with nonzero codimension in Euclidean space 222Recently, De Philippis and Pigati [DPP] considered the Michael-Simon inequality for nonzero codimensional anisotropic minimal submanifolds in Euclidean space via ideas from multilinear Kakeya estimates, where the weight is put on the normal vectors of the submanifolds instead of relative normal vectors on the boundary of submanifolds.. To state the inequality, for each nn-dimensional affine subspace PP, we denote projP𝒲\text{proj}_{P}\mathcal{W} the projection of 𝒲\mathcal{W} to PP and we let P¯\bar{P} be any nn-dimensional affine subspace such that |projP¯𝒲|=min{|projP𝒲|:PGrn(n+m)}|\text{proj}_{\bar{P}}\mathcal{W}|=\min\{|\text{proj}_{P}\mathcal{W}|:P\in Gr_{n}(\mathbb{R}^{n+m})\} and denote any of them by W=projP¯𝒲W^{*}=\text{proj}_{\bar{P}}\mathcal{W}.

Conjecutre 1.4 (Wulff inequality for minimal submanifolds).

For any nn-dimensional minimal submanifold Σ\Sigma in n+m\mathbb{R}^{n+m} the following Wulff inequality

(1.6) PΦ(Σ)|Σ|n1nPΦ(W)|W|n1n\frac{P_{\Phi}({\Sigma})}{|\Sigma|^{\frac{n-1}{n}}}\geq\frac{P_{\Phi}({W^{*}})}{|W^{*}|^{\frac{n-1}{n}}}

holds, and equality holds if and only if Σ\Sigma is homothetic to some WW^{*} 333Note that by Remark 2.2, the right hand side of (1.6) is the same for any projection with minimal projection area, hence it is well-defined..

In this paper, by adapting arguments of Brendle and Cabre, we first prove a boundary weighted isoperimetric inequality for minimal submanifolds in Euclidean space:

Theorem 1.5.

Let 𝒲\mathcal{W} be centrally symmetric n+mn+m dimensional convex body in n+m\mathbb{R}^{n+m}, and Φ\Phi be the corresponding support function. Let Σ\Sigma be a minimal hypersurface with boundary Σ\partial\Sigma embedded in n+m\mathbb{R}^{n+m}. We have

(1.7) PΦ(Σ)n(supn,m𝒲f)1n|Σ|n1nP_{\Phi}({\Sigma})\geq n(\sup_{\mathcal{F}_{n,m}}\int_{\mathcal{W}}f)^{\frac{1}{n}}|\Sigma|^{\frac{n-1}{n}}

where

(1.8) n,m={fL1(n+m):supp(f)𝒲,f0,Pf1forPGraffm(n+m)}\mathcal{F}_{n,m}=\{f\in L^{1}(\mathbb{R}^{n+m}):\textrm{supp}(f)\subset\mathcal{W},f\geq 0,\int_{P}f\leq 1\,\,\text{for}\,\,P\in\textrm{Graff}_{m}(\mathbb{R}^{n+m})\}

and Graffm(n+m)\textrm{Graff}_{m}(\mathbb{R}^{n+m}) is the affine Grassmannian consisting of all mm-dimensional affine subspaces in n+m\mathbb{R}^{n+m}

Remark 1.6.

By Fubini’s theorem we have the trivial upper bound

(1.9) supn,m𝒲f|W|.\sup_{\mathcal{F}_{n,m}}\int_{\mathcal{W}}f\leq|W^{*}|.

It is also clear that the supremum is always positive, as the function f=χBn+m(δ)f=\chi_{B^{n+m}(\delta)} for δ>0\delta>0 sufficiently small is in n,m\mathcal{F}_{n,m}. Later in (5.6), we will provide a better estimate of the quantity on left hand side of (1.9) for general Wulff shapes 𝒲\mathcal{W}, and show that in some special cases the equality in (1.9) holds.

To state our next result, we define a class of convex bodies generated from ellipsoids using gluing and cutting operations. We refer to this class of convex bodies as long convex bodies, since a typical example of such shape is given by 𝒲=K×[T,T]mn+m\mathcal{W}=K\times[-T,T]^{m}\subset\mathbb{R}^{n+m} where KK is any nn-dimensional centrally symmetric convex body and 2Tdiam(K)2T\geq\text{diam}(K).

Definition 1.7 (long convex body).

Let n+m\mathcal{E}_{n+m} be the set of all (n+m)(n+m)-dimensional ellipsoids centering at origin. For i=1,2i=1,2, let 𝒲i\mathcal{W}_{i} be centrally symmetric convex bodies in n+m\mathbb{R}^{n+m}, and WiW_{i}^{*} some area-minimizing projection for 𝒲i\mathcal{W}_{i}. Let PP be an nn-dimensional affine subspace such that |projP𝒲1|=|W1||\text{proj}_{P}\mathcal{W}_{1}|=|W_{1}^{*}|. The set of long convex bodies n,m\mathcal{L}_{n,m} is the smallest set of centrally symmetric convex bodies satisfying the following conditions:

  • n+mn,m\mathcal{E}_{n+m}\subset\mathcal{L}_{n,m}.

  • Gluing: if 𝒲1n,m\mathcal{W}_{1}\in\mathcal{L}_{n,m}, 𝒲2𝒲1\mathcal{W}_{2}\supset\mathcal{W}_{1} and projP𝒲1=projP𝒲2\text{proj}_{P}\mathcal{W}_{1}=\text{proj}_{P}\mathcal{W}_{2}, then 𝒲2n,m\mathcal{W}_{2}\in\mathcal{L}_{n,m}.

  • Cutting: if 𝒲1n,m\mathcal{W}_{1}\in\mathcal{L}_{n,m}, and 𝒲2=(projP)1(projP𝒲2)𝒲1\mathcal{W}_{2}=(\text{proj}_{P})^{-1}(\text{proj}_{P}\mathcal{W}_{2})\cap\mathcal{W}_{1}, then 𝒲2n,m\mathcal{W}_{2}\in\mathcal{L}_{n,m}.

Remark 1.8.

It is clear that n,m\mathcal{L}_{n,m} consists of all centrally symmetric convex bodies that can be obtained by performing a finite number of gluing and cutting operations on an ellipsoid. Performing these operations repeatedly elongates the shape in the normal directions of the area-minimizing projection. These two operations provide necessary conditions for equality in (1.9), in the sense that performing gluing or cutting along some area-minimizing projection direction should preserve the fact that it is an area-minimizing projection direction. Specifically, if the equality in (1.9) holds for some 𝒲1\mathcal{W}_{1}, and 𝒲2\mathcal{W}_{2} (not necessarily centrally symmetric here) is generated from 𝒲1\mathcal{W}_{1} by gluing or cutting with respect to some PP such that projP𝒲1\text{proj}_{P}\mathcal{W}_{1} is an area-minimizing projection for 𝒲1\mathcal{W}_{1}, then projP𝒲2\text{proj}_{P}\mathcal{W}_{2} must also be an area-minimizing projection for 𝒲2\mathcal{W}_{2}, and the equality in (1.9) still holds for 𝒲2\mathcal{W}_{2} (see Lemma 4.3). For example, when 𝒲\mathcal{W} is an (n+1)(n+1)-dimensional cube or a short cylinder of the form Bn×[ε2,ε2]B^{n}\times[-\frac{\varepsilon}{2},\frac{\varepsilon}{2}] for small ε>0\varepsilon>0 in Figure 1, it is possible to cut along some area-minimizing projection direction, such that after the cut, it is no longer the area-minimizing projection direction for the resulting shape. Hence, the equality in (1.9) must be strict for the cube and short cylinder.

Refer to caption
Figure 1. Projections of short cylinder and cube. For short cylinder, projection along any horizontal direction is area-minimizing. However, cutting along a horizontal direction sufficiently close to boundary gives a shape UU such that horizontal projection is no longer area-minimizing, since the vertical projection has area roughly k3/2k^{3/2} and the horizontal projection has area εk\varepsilon k. Similarly, the vertical projection was area-minimizing for cube, but such a vertical cut gives UU, whose vertical projection has area roughly 2k\sqrt{2}k and horizontal projection has area kk.

In the cases of codimension one and two, by designing good candidates on ellipsoids and long convex bodies, we verify the above two conjectures when 𝒲n,m\mathcal{W}\in\mathcal{L}_{n,m} (m=1,2m=1,2) and the weights are support functions of 𝒲\mathcal{W}.

Theorem 1.9.

For m=1,2m=1,2, let 𝒲n,m\mathcal{W}\in\mathcal{L}_{n,m} and Φ\Phi be the corresponding support function, then for Σ\Sigma being a minimal submanifold with boundary Σ\partial\Sigma embedded in n+m\mathbb{R}^{n+m}, we have

(1.10) PΦ(Σ)|Σ|n1nPΦ(W)|W|n1n,\frac{P_{\Phi}({\Sigma})}{|\Sigma|^{\frac{n-1}{n}}}\geq\frac{P_{\Phi}({W^{*}})}{|W^{*}|^{\frac{n-1}{n}}},

In particular, equality holds if and only if Σ\Sigma is homothetic to WW^{*}.

Remark 1.10.

In higher codimensional case when 𝒲\mathcal{W} is an ellipsoid centering at origin and m>2m>2, we will show in Proposition 4.2 that

(1.11) PΦ(Σ)|Σ|n1n((n+m)|Bn+m|m|Bm||Bn|)1nPΦ(W)|W|n1n.\frac{P_{\Phi}({\Sigma})}{|\Sigma|^{\frac{n-1}{n}}}\geq\left(\frac{(n+m)|B^{n+m}|}{m|B^{m}||B^{n}|}\right)^{\frac{1}{n}}\frac{P_{\Phi}({W^{*}})}{|W^{*}|^{\frac{n-1}{n}}}.

Combine with the sharp cases in Theorem 1.9, the isotropic (i.e. 𝒲=Bn+m\mathcal{W}=B^{n+m} and Φ(x)=|x|\Phi(x)=|x|) results in [Bre21] can be generalized to any ellipsoid.

More generally we prove a Wulff inequality for minimal submanifolds in the Euclidean space, where the constant depends only on dimension and codimension.

Theorem 1.11.

Let 𝒲\mathcal{W} be centrally symmetric (n+m)(n+m)-dimensional convex body in n+m\mathbb{R}^{n+m} and Φ\Phi be the corresponding support function. Then for a minimal submanifold Σ\Sigma with boundary Σ\partial\Sigma embedded in n+m\mathbb{R}^{n+m}, we have

(1.12) PΦ(Σ)|Σ|n1ncn,mPΦ(W)|W|n1n,\frac{P_{\Phi}({\Sigma})}{|\Sigma|^{\frac{n-1}{n}}}\geq c_{n,m}\frac{P_{\Phi}({W^{*}})}{|W^{*}|^{\frac{n-1}{n}}},

where

(1.13) cn,m={1n+mm=1,2,1n+m((n+m)|Bn+m|m|Bm||Bn|)1nm3c_{n,m}=\begin{cases}\frac{1}{\sqrt{n+m}}&m=1,2,\\ \frac{1}{\sqrt{n+m}}\left(\frac{(n+m)|B^{n+m}|}{m|B^{m}||B^{n}|}\right)^{\frac{1}{n}}&m\geq 3\end{cases}

is a constant only depending on mm and nn but independent of the weight Φ\Phi.555The Wulff inequality is scaling-invariant for weights or (n+m)(n+m)-dimensional Wulff shape, but the space of (n+m)(n+m)-dimensional Wulff shapes, after taking quotient by the scaling symmetry, is non-compact.

Remark 1.12.

It would be very interesting to improve the constant cn,mc_{n,m} in (1.12) to some constant independent of mm and nn, which is expected to be 1. Another interesting direction is to prove sharp minimal submanifolds Wulff inequality for more general class of weights or the associated convex bodies.

The paper is organized as follows: in Section 2, we prove Theorem 1.5; in Section 3, we prove Theorem 1.9; and in Section 4, we prove Theorem 1.11.

Acknowledgments. The authors appreciate the funding and research environment support from MIT, UCSD, and ITS Westlake University, respectively.

2. Basic Convex Geometry

In this section, we first recall some basic properties of convex geometry that will be used in later sections.

Let PP be an nn-dimensional subspace passing through the origin. We now consider the following two sets in PP:

  1. (1)

    WPW_{P} is the Wulff shape induced by Φ|P\Phi|_{P},

  2. (2)

    projP𝒲\text{proj}_{P}\mathcal{W} is the projection of 𝒲\mathcal{W} onto PP;

The following lemma shows that the two objects are the same.

Lemma 2.1.

If Φ\Phi is the support function of 𝒲\mathcal{W}, then WP=projP𝒲W_{P}=\text{proj}_{P}\mathcal{W}, and Φ|P\Phi|_{P} is the support function of WPW_{P} in PP.

Proof.

It suffices to show that Φ|P\Phi|_{P} is the support function of projP𝒲\text{proj}_{P}\mathcal{W}, since support functions and Wulff shapes uniquely determine each other (see [CROS16, page 2976]). For any νPSn+m1\nu\in P\cap S^{n+m-1} and x𝒲x\in\mathcal{W}, we have xν=projPxνx\cdot\nu=\text{proj}_{P}x\cdot\nu, hence

(2.1) Φ|P(ν)=sup{xν:x𝒲}=sup{projPxν:x𝒲}=sup{yν:yprojP𝒲}.\Phi|_{P}(\nu)=\sup\{x\cdot\nu:x\in\mathcal{W}\}=\sup\{\text{proj}_{P}x\cdot\nu:x\in\mathcal{W}\}=\sup\{y\cdot\nu:y\in\text{proj}_{P}\mathcal{W}\}.

Remark 2.2.

Let PP be any nn-dimensional affine subspace such that |projP𝒲||\text{proj}_{P}\mathcal{W}| is minimized, and let PP^{\prime} be the nn-dimensional subspace parallel to PP and passing through the origin. As a consequence of the lemma, Φ|P\Phi|_{P^{\prime}} is also the support function of projP𝒲\text{proj}_{P}\mathcal{W} in PP, so

PΦ(projP𝒲)\displaystyle P_{\Phi}(\text{proj}_{P}\mathcal{W}) =projP𝒲Φ|P(ν(x))dS\displaystyle=\int_{\partial\text{proj}_{P}\mathcal{W}}\Phi|_{P^{\prime}}(\nu(x))dS
=projP𝒲xν(x)𝑑S\displaystyle=\int_{\partial\text{proj}_{P}\mathcal{W}}x\cdot\nu(x)dS
=projP𝒲div(x)𝑑x\displaystyle=\int_{\text{proj}_{P}\mathcal{W}}\text{div}(x)dx
(2.2) =n|projP𝒲|,\displaystyle=n|\text{proj}_{P}\mathcal{W}|,

where we used the fact that for almost every xprojP𝒲x\in\partial\text{proj}_{P}\mathcal{W}, the unit outward normal ν(x)\nu(x) exists, and xν(x)=Φ|P(x)x\cdot\nu(x)=\Phi|_{P^{\prime}}(x) because it Φ|P\Phi|_{P^{\prime}} is the support function. According to (2.2), for any P1P_{1} and P2P_{2} such that both |projP1𝒲||\text{proj}_{P_{1}}\mathcal{W}| and |projP2𝒲||\text{proj}_{P_{2}}\mathcal{W}| are minimized, it follows that

(2.3) PΦ(projP1𝒲)|projP1𝒲|n1n=PΦ(projP2𝒲)|projP2𝒲|n1n.\displaystyle\frac{P_{\Phi}(\text{proj}_{P_{1}}\mathcal{W})}{|\text{proj}_{P_{1}}\mathcal{W}|^{\frac{n-1}{n}}}=\frac{P_{\Phi}(\text{proj}_{P_{2}}\mathcal{W})}{|\text{proj}_{P_{2}}\mathcal{W}|^{\frac{n-1}{n}}}.

Thus, the right-hand sides of (1.6), (1.10) and (1.12) are well defined for given 𝒲\mathcal{W}.

3. Proof of Theorem 1.5

In this section, we prove Theorem 1.5. Let Σn+m\Sigma\subset\mathbb{R}^{n+m} be a compact nn-dimensional minimal submanifold with boundary Σ\partial\Sigma. We first consider the special case that Σ\Sigma is connected with boundary Σ\partial\Sigma. By solving u:Σu:\Sigma\to\mathbb{R} from the following Poisson equation with Neumann boundary condition

(3.1) {ΔΣu=PΦ(Σ)|Σ| in Σ,Σu,ν=Φ(ν) on Σ,\displaystyle\begin{cases}\Delta_{\Sigma}u=\frac{P_{\Phi}(\Sigma)}{|\Sigma|}&\text{ in }\Sigma,\\ \langle\nabla^{\Sigma}u,\nu\rangle=\Phi(\nu)&\text{ on }\partial\Sigma,\end{cases}

we can construct the map:

(3.2) T:NΣn+1,(x,y)Σu(x)+yT:N\Sigma\xrightarrow{}\mathbb{R}^{n+1},(x,y)\mapsto\nabla^{\Sigma}u(x)+y

where NΣN\Sigma is the normal bundle of Σ\Sigma. As in [Bre21], we define the following set:

(3.3) A={(x,y)N(Σ\Σ):det(DΣ2uI​IΣ(x),y)0},A=\{(x,y)\in N(\Sigma\backslash\partial\Sigma):\det(D^{2}_{\Sigma}u-\langle\textrm{I\!I}_{\Sigma}(x),y\rangle)\geq 0\},

where I​IΣ\textrm{I\!I}_{\Sigma} is the second fundamental form of Σn+m\Sigma\subset\mathbb{R}^{n+m}

Lemma 3.1.

We have T(A)𝒲intT(A)\supset\mathcal{W}^{int}, thus TT is a surjective map from AA onto the interior of the Wulff shape 𝒲\mathcal{W}.

Proof.

Let ξ𝒲int\xi\in\mathcal{W}^{int} and we consider the map w:Σw:\Sigma\xrightarrow{}\mathbb{R} defined by

(3.4) w(x)=u(x)x,ξ.\displaystyle w(x)=u(x)-\langle x,\xi\rangle.

Denote ν(x)\nu(x) the unit outward normal vector on the boundary Σ\partial\Sigma relative to Σ\Sigma. Then, for any xΣx\in\partial\Sigma,

(3.5) Σw(x),ν(x)=Σu(x),ν(x)ξ,ν(x)=Φ(ν)ξ,ν(x)>0.\langle\nabla^{\Sigma}w(x),\nu(x)\rangle=\langle\nabla^{\Sigma}u(x),\nu(x)\rangle-\langle\xi,\nu(x)\rangle=\Phi(\nu)-\langle\xi,\nu(x)\rangle>0.

Thus ww attains its minimum at some interior point x¯\bar{x}. It follows that Σw(x¯)=0\nabla^{\Sigma}w(\bar{x})=0, which implies

(3.6) ξ=Σu(x¯)+y\displaystyle\xi=\nabla^{\Sigma}u(\bar{x})+y

for some yNx¯Σy\in N_{\bar{x}}\Sigma. It remains to compute the determinant. Since ww achieves its minimum, we have

(3.7) 0DΣ2w(x¯)=DΣ2u(x)I​IΣ(x¯),ξ=DΣ2u(x)I​IΣ(x¯),y,0\leq D^{2}_{\Sigma}w(\bar{x})=D^{2}_{\Sigma}u(x)-\langle\textrm{I\!I}_{\Sigma}(\bar{x}),\xi\rangle=D^{2}_{\Sigma}u(x)-\langle\textrm{I\!I}_{\Sigma}(\bar{x}),y\rangle,

where the last equality holds from (3.6). Therefore, we obtain T(x¯,y)=ξT(\bar{x},y)=\xi and (x¯,y)A(\bar{x},y)\in A. ∎

Lemma 3.2.

For all (x,y)A(x,y)\in A, the determinant of transport map TT satisfies

(3.8) 0detDT(x,y)(PΦ(Σ)n|Σ|)n.0\leq\det DT(x,y)\leq\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}.
Proof.

As shown in [Bre21, Lemma 5], one can show that

(3.9) detDT(x,y)=det(DΣ2uI​IΣ(x),y).\displaystyle\det DT(x,y)=\det(D^{2}_{\Sigma}u-\langle\textrm{I\!I}_{\Sigma}(x),y\rangle).

By the definition of AA, we have det(DΣ2uI​IΣ(x),y)0\det(D^{2}_{\Sigma}u-\langle\textrm{I\!I}_{\Sigma}(x),y\rangle)\geq 0. Thus, by applying the arithmetic-geometric mean inequality, the vanishing mean curvature property of minimal submanifolds, and (3.1), we obtain

(3.10) det(DΣ2uI​IΣ(x),y)(tr(DΣ2uI​IΣ(x),y)n)n=(PΦ(Σ)n|Σ|)n.\det(D^{2}_{\Sigma}u-\langle\textrm{I\!I}_{\Sigma}(x),y\rangle)\leq\left(\frac{\text{tr}(D^{2}_{\Sigma}u-\langle\textrm{I\!I}_{\Sigma}(x),y\rangle)}{n}\right)^{n}=\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}.

Proof of Theorem 1.5.

We recall

(3.11) fn,m={fL1(n+m):supp(f)𝒲,f0,Pf1forPGraffm(n+m)},f\in\mathcal{F}_{n,m}=\{f\in L^{1}(\mathbb{R}^{n+m}):\textrm{supp}(f)\subset\mathcal{W},f\geq 0,\int_{P}f\leq 1\,\,\text{for}\,\,P\in\textrm{Graff}_{m}(\mathbb{R}^{n+m})\},

and by Lemma 3.1 and Lemma 3.2, we have

(3.12) 𝒲f(ξ)𝑑ξ\displaystyle\int_{\mathcal{W}}f(\xi)d\xi ΣNxΣf(T(x,y))|detDT(x,y)|1A(x,y)𝑑y𝑑x\displaystyle\leq\int_{\Sigma}\int_{N_{x}\Sigma}f(T(x,y))|\det DT(x,y)|1_{A}(x,y)dydx
(PΦ(Σ)n|Σ|)nΣNxΣf(Σu(x)+y)𝑑y𝑑x\displaystyle\leq\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}\int_{\Sigma}\int_{N_{x}\Sigma}f(\nabla^{\Sigma}u(x)+y)dydx
(PΦ(Σ)n|Σ|)n|Σ|\displaystyle\leq\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}|\Sigma|
=(PΦ(Σ)n|Σ|n1n)n.\displaystyle=\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|^{\frac{n-1}{n}}}\right)^{n}.

In particular, the above inequality gives

(3.13) supn,m𝒲f(PΦ(Σ)n|Σ|n1n)n,\sup_{\mathcal{F}_{n,m}}\int_{\mathcal{W}}f\leq\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|^{\frac{n-1}{n}}}\right)^{n},

which implies (1.7). It remains to consider the case where Σ\Sigma is disconnected. In that case, we apply the inequality to each individual connected component of Σ\Sigma, sum over all connected components, and use the strict inequality

(3.14) an1n+bn1n>a(a+b)1n+b(a+b)1n=(a+b)n1n\displaystyle a^{\frac{n-1}{n}}+b^{\frac{n-1}{n}}>a(a+b)^{-\frac{1}{n}}+b(a+b)^{-\frac{1}{n}}=(a+b)^{\frac{n-1}{n}}

for all a,b>0a,b>0.

4. Proof of Theorem 1.9

In this section, we prove Theorem 1.9. By Theorem 1.5 and (2.2), it suffices to show that there exists some f~n,m\tilde{f}\in\mathcal{F}_{n,m} such that

(4.1) 𝒲f~=|W|,\displaystyle\int_{\mathcal{W}}\tilde{f}=|W^{*}|,

or there exists a sequence fσn,mf_{\sigma}\in\mathcal{F}_{n,m} such that

(4.2) limσ1𝒲fσ=|W|.\displaystyle\lim_{\sigma\to 1}\int_{\mathcal{W}}f_{\sigma}=|W^{*}|.

In Proposition 4.1, Proposition 4.2 and Proposition 4.4, we demonstrate that when m=1,2m=1,2, equation (4.1) or (4.2) holds when 𝒲\mathcal{W} is a centrally symmetric (n+m)(n+m)-dimensional ellipsoid and long convex body. In Proposition 4.2 we also obtain a weaker estimate for m>2m>2. Then in Proposition 4.5, we discuss the sharpness and rigidity when equality holds for ellipsoids and long bodies and m=1,2m=1,2.

Proposition 4.1.

If codimension m=1m=1 and 𝒲n+1\mathcal{W}\in\mathcal{E}_{n+1}, there exists some f~n,1\tilde{f}\in\mathcal{F}_{n,1} such that (4.1) holds. Hence

(4.3) supfn,1𝒲f=|W|.\sup_{f\in\mathcal{F}_{n,1}}\int_{\mathcal{W}}f=|W^{*}|.
Proof.

Suppose 𝒲\mathcal{W} is an ellipsoid given by

(4.4) x12λ12++xn+12λn+121\frac{x_{1}^{2}}{\lambda_{1}^{2}}+...+\frac{x_{n+1}^{2}}{\lambda_{n+1}^{2}}\leq 1

where 0<λ1λn+10<\lambda_{1}\leq...\leq\lambda_{n+1}. Denote

(4.5) Λ=(λ11λ21λn+11).\displaystyle\Lambda=\begin{pmatrix}\lambda_{1}^{-1}&&&&\\ &\lambda_{2}^{-1}&&&\\ &&\ddots&&\\ &&&&\lambda_{n+1}^{-1}\\ \end{pmatrix}.

Let the function f(x)f(x) be defined by

(4.6) f(x)=11|Λx|21𝒲(x).f(x)=\frac{1}{\sqrt{1-|\Lambda x|^{2}}}1_{\mathcal{W}}(x).

Consider a line l(t)=t𝜶+𝝎l(t)=t\boldsymbol{\alpha}+\boldsymbol{\omega} that passes through the ellipsoid along unit vector 𝜶\boldsymbol{\alpha} direction. Then the end points Ti𝜶+𝝎T_{i}\boldsymbol{\alpha}+\boldsymbol{\omega} (i=1,2i=1,2) of the chord in the ellipsoid satisfies

(4.7) |Λ(Ti𝜶+𝝎)|2=1,i=1,2.|\Lambda(T_{i}\boldsymbol{\alpha}+\boldsymbol{\omega})|^{2}=1,\quad i=1,2.

This is a quadratic equation

(4.8) aTi2+bTi+c=0,i=1,2,-aT_{i}^{2}+bT_{i}+c=0,\quad i=1,2,

where

(4.9) a=|Λ𝜶|2,b=2Λ𝜶,𝝎,c=1|Λ𝝎|2.a=|\Lambda\boldsymbol{\alpha}|^{2},\quad b=-2\langle\Lambda\boldsymbol{\alpha},\boldsymbol{\omega}\rangle,\quad c=1-|\Lambda\boldsymbol{\omega}|^{2}.

Upon solving this, we obtain

(4.10) (2aTibb2+4ac)2=1.(\frac{2aT_{i}-b}{\sqrt{b^{2}+4ac}})^{2}=1.

Thus, the integral over the chord is given by

T1T211|Λ(t𝜶+𝝎)|2𝑑t\displaystyle\int_{T_{1}}^{T_{2}}\frac{1}{\sqrt{1-|\Lambda(t\boldsymbol{\alpha}+\boldsymbol{\omega})|^{2}}}dt
=T1T21at2+bt+c𝑑t\displaystyle=\int_{T_{1}}^{T_{2}}\frac{1}{\sqrt{-at^{2}+bt+c}}dt
=1aarcsin(2atbb2+4ac)|T1T2\displaystyle=\left.\frac{1}{\sqrt{a}}\arcsin(\frac{2at-b}{\sqrt{b^{2}+4ac}})\right|_{T_{1}}^{T_{2}}
(4.11) =πa.\displaystyle=\frac{\pi}{\sqrt{a}}.

We notice that

(4.12) πa=π|Λ𝜶|λn+1π,\frac{\pi}{\sqrt{a}}=\frac{\pi}{|\Lambda\boldsymbol{\alpha}|}\leq\lambda_{n+1}\pi,

and the equality holds if and only if 𝜶=(0,,0,1)\boldsymbol{\alpha}=(0,...,0,1), that is the line is parallel to the longest axis. We now define the function f~\tilde{f} as

(4.13) f~=fπλn+10.\tilde{f}=\frac{f}{\pi\lambda_{n+1}}\geq 0.

It satisfies its integral over any line that passes through the ellipsoid is at most 1, and achieves 1 when the line is parallel to the longest direction of ellipsoid, that is the axis corresponding to λn+1\lambda_{n+1}. Moreover, Fubini’s Theorem gives

(4.14) 𝒲f~=proj{xn+1=0}𝒲γxn+1f~(x)𝑑x𝑑xn+1=|proj{xn+1=0}𝒲|=|W|.\int_{\mathcal{W}}\tilde{f}=\int_{\text{proj}_{\{x_{n+1}=0\}}\mathcal{W}}\int_{\gamma_{x_{n+1}}}\tilde{f}(x)dx^{\prime}dx_{n+1}=|\text{proj}_{\{x_{n+1}=0\}}\mathcal{W}|=|W^{*}|.

Thus, f~n,1\tilde{f}\in\mathcal{F}_{n,1} and satisfies (4.1), and we have completed the proof. ∎

For codimension m2m\geq 2 cases, we have following proposition.

Proposition 4.2.

If codimension m=2m=2 and 𝒲n+m\mathcal{W}\in\mathcal{E}_{n+m}, then there exists a sequence of functions fσn,mf_{\sigma}\in\mathcal{F}_{n,m} such that (4.2) holds. Hence when m=2m=2,

(4.15) supfn,2𝒲f=|W|.\sup_{f\in\mathcal{F}_{n,2}}\int_{\mathcal{W}}f=|W^{*}|.

If codimension m>2m>2 and 𝒲n+m\mathcal{W}\in\mathcal{E}_{n+m}, then there exists a sequence of functions fσn,mf_{\sigma}\in\mathcal{F}_{n,m} such that

(4.16) supfn,m𝒲flimσ1𝒲fσ=(n+m)|Bn+m|m|Bm||Bn||W|.\sup_{f\in\mathcal{F}_{n,m}}\int_{\mathcal{W}}f\geq\lim_{\sigma\to 1}\int_{\mathcal{W}}f_{\sigma}=\frac{(n+m)|B^{n+m}|}{m|B^{m}||B^{n}|}|W^{*}|.
Proof.

Denote 𝒲=Eλ1,λ2,,λn+m\mathcal{W}=E_{\lambda_{1},\lambda_{2},\cdots,\lambda_{n+m}} the (n+m)(n+m)-dimensional ellipsoid

(4.17) x12λ12+x22λ22++xn+m2λn+m21,\displaystyle\frac{x_{1}^{2}}{\lambda_{1}^{2}}+\frac{x_{2}^{2}}{\lambda_{2}^{2}}+\cdots+\frac{x_{n+m}^{2}}{\lambda_{n+m}^{2}}\leq 1,

where λ1λ2λn+m\lambda_{1}\leq\lambda_{2}\leq\cdots\leq\lambda_{n+m}, and Uσ=Eλ1,λ2,,λn+mEσλ1,σλ2,,σλn+mU_{\sigma}=E_{\lambda_{1},\lambda_{2},\cdots,\lambda_{n+m}}\setminus E_{\sigma\lambda_{1},\sigma\lambda_{2},\cdots,\sigma\lambda_{n+m}} for some 0<σ<10<\sigma<1.

Let P=𝒒+t1𝒓𝟏+t2𝒓𝟐++tm𝒓𝒎P=\boldsymbol{q}+t_{1}\boldsymbol{r_{1}}+t_{2}\boldsymbol{r_{2}}+\cdots+t_{m}\boldsymbol{r_{m}} be an mm-dimensional affine subspace which intersects with UσU_{\sigma}, where {𝒓𝒋}j=1m\{\boldsymbol{r_{j}}\}_{j=1}^{m} are orthonormal. The intersection region UσPU_{\sigma}\cap P is the shell of two mm-dimensional homothetic ellipsoids (if PP is very close to the boundary, it would be fully outside of Eσλ1,,σλn+mE_{\sigma\lambda_{1},\cdots,\sigma\lambda_{n+m}}, so in that case UσP=Eλ1,,λn+mPU_{\sigma}\cap P=E_{\lambda_{1},\cdots,\lambda_{n+m}}\cap P is just an ellipsoid). Denote

(4.18) Λ=(λ11λ21λn+m1).\displaystyle\Lambda=\begin{pmatrix}\lambda_{1}^{-1}&&&&\\ &\lambda_{2}^{-1}&&&\\ &&\ddots&&\\ &&&&\lambda_{n+m}^{-1}\\ \end{pmatrix}.

Note that Λ(Eλ1,,λn+m)=Bn+m\Lambda(E_{\lambda_{1},\cdots,\lambda_{n+m}})=B^{n+m}, so we can study the intersection between ellipsoid and lower dimensional plane by converting it to an intersection between ball and plane via Λ\Lambda, and converting everything back in the end by applying appropriate scaling. (For a detailed computation for m=1m=1 and n=2n=2, see [Kle12].)

Specifically, the above computation yields that the length of the ii-th axis of the outer ellipsoid PEλ1,,λn+mP\cap E_{\lambda_{1},\cdots,\lambda_{n+m}} is

(4.19) 1d|Λ𝒓𝒊|,1im,\displaystyle\frac{\sqrt{1-d}}{|\Lambda\boldsymbol{r_{i}}|},\quad 1\leq i\leq m,

where d=|Λ𝒒|2Λ𝒒,Λ𝒓𝟏2|Λ𝒓𝟏|2Λ𝒒,Λ𝒓𝟐2|Λ𝒓𝟐|2Λ𝒒,Λ𝒓𝒎2|Λ𝒓𝒎|2d=|\Lambda\boldsymbol{q}|^{2}-\frac{\langle\Lambda\boldsymbol{q},\Lambda\boldsymbol{r_{1}}\rangle^{2}}{|\Lambda\boldsymbol{r_{1}}|^{2}}-\frac{\langle\Lambda\boldsymbol{q},\Lambda\boldsymbol{r_{2}}\rangle^{2}}{|\Lambda\boldsymbol{r_{2}}|^{2}}-\cdots-\frac{\langle\Lambda\boldsymbol{q},\Lambda\boldsymbol{r_{m}}\rangle^{2}}{|\Lambda\boldsymbol{r_{m}}|^{2}} is the squared distance between origin and ΛP\Lambda P. The length of ii-th axis of inner ellipsoid PEσλ1,,σλn+mP\cap E_{\sigma\lambda_{1},\cdots,\sigma\lambda_{n+m}} is

(4.20) σ2d|Λ𝒓𝒊|,1im.\displaystyle\frac{\sqrt{\sigma^{2}-d}}{|\Lambda\boldsymbol{r_{i}}|},\quad 1\leq i\leq m.

As discussed before, UσPU_{\sigma}\cap P is the shell formed by two homothetic ellipsoids only when PP cuts through Eσλ1,,σλn+mE_{\sigma\lambda_{1},\cdots,\sigma\lambda_{n+m}}, which is the same as ΛP\Lambda P cuts through Bn(σ)B^{n}(\sigma), or equivalently dσ2d\leq\sigma^{2}. Hence the intersection volume can be expressed as

(4.21) |UσP|=|Bm|[(1d)m/2(σ2d)+m/2]j=1m|Λ𝒓𝒋|1.|U_{\sigma}\cap P|=|B^{m}|[(1-d)^{m/2}-(\sigma^{2}-d)^{m/2}_{+}]\prod_{j=1}^{m}|\Lambda\boldsymbol{r_{j}}|^{-1}.

Since {𝒓𝒋}j=1m\{\boldsymbol{r_{j}}\}_{j=1}^{m} is an orthonormal basis, we have

(4.22) j=1m|Λ𝒓𝒋|1j=1mλn+m+1j.\prod_{j=1}^{m}|\Lambda\boldsymbol{r_{j}}|^{-1}\leq\prod_{j=1}^{m}\lambda_{n+m+1-j}.

Then it follows

(4.23) |UσP||Bm|m2(1σ2)j=1mλn+m+1j=:Cσ.|U_{\sigma}\cap P|\leq|B^{m}|\frac{m}{2}(1-\sigma^{2})\prod_{j=1}^{m}\lambda_{n+m+1-j}=:C_{\sigma}.

Now we consider

(4.24) fσ=1Cσ1Uσ,f_{\sigma}=\frac{1}{C_{\sigma}}1_{U_{\sigma}},

and it clearly satisfies supp(fσ)𝒲\text{supp}(f_{\sigma})\subset\mathcal{W} and fσ0f_{\sigma}\geq 0. For any mm-dimensional affine subspace PP in n+m\mathbb{R}^{n+m},

(4.25) Pfσ=|UσP|Cσ1.\int_{P}f_{\sigma}=\frac{|U_{\sigma}\cap P|}{C_{\sigma}}\leq 1.

On the other hand, the volume of the shell UσU_{\sigma} is given by the difference of two (n+m)(n+m)-dimensional ellipsoids, so

𝒲fσ\displaystyle\int_{\mathcal{W}}f_{\sigma} =|Uσ|Cσ=1Cσ|Bn+m|(1σn+m)j=1n+mλj\displaystyle=\frac{|U_{\sigma}|}{C_{\sigma}}=\frac{1}{C_{\sigma}}|B^{n+m}|(1-\sigma^{n+m})\prod_{j=1}^{n+m}\lambda_{j}
(4.26) =2(1σn+m)1σ2|Bn+m|m|Bm|j=1nλj.\displaystyle=\frac{2(1-\sigma^{n+m})}{1-\sigma^{2}}\frac{|B^{n+m}|}{m|B^{m}|}\prod_{j=1}^{n}\lambda_{j}.

Taking the limit as σ1\sigma\xrightarrow{}1, we obtain that

(4.27) limσ1𝒲fσ=(n+m)|Bn+m|m|Bm|j=1nλj=(n+m)|Bn+m|m|Bm||Bn||W|,\lim_{\sigma\to 1}\int_{\mathcal{W}}f_{\sigma}=\frac{(n+m)|B^{n+m}|}{m|B^{m}|}\prod_{j=1}^{n}\lambda_{j}=\frac{(n+m)|B^{n+m}|}{m|B^{m}||B^{n}|}|W^{*}|,

where we used |W|=|Bn|j=1nλj|W^{*}|=|B^{n}|\prod_{j=1}^{n}\lambda_{j} for the ellipsoid. In particular, when m=2m=2, the coefficient simplifies to 1, and we have completed the proof of (4.2). ∎

Next we will prove a lemma about the necessary condition discussed in Remark 1.8. We will use this lemma and the f~\tilde{f} and fσf_{\sigma} constructed above to verify (1.10) for general 𝒲n,m\mathcal{W}\in\mathcal{L}_{n,m} and m=1,2m=1,2.

Lemma 4.3.

Suppose 𝒲1\mathcal{W}_{1} and fσn,m𝒲1f_{\sigma}\in\mathcal{F}^{\mathcal{W}_{1}}_{n,m} satisfies

(4.28) supfn,m𝒲1𝒲1f=limσ1𝒲1fσ=|W1|,\sup_{f\in\mathcal{F}^{\mathcal{W}_{1}}_{n,m}}\int_{\mathcal{W}_{1}}f=\lim_{\sigma\to 1}\int_{\mathcal{W}_{1}}f_{\sigma}=|W_{1}^{*}|,

and let 𝒲2\mathcal{W}_{2} be generated from 𝒲1\mathcal{W}_{1} via gluing or cutting with respect to an nn-dimensional affine subspace PP such that projP𝒲1\text{proj}_{P}\mathcal{W}_{1} is an area-minimizing projection for 𝒲1\mathcal{W}_{1}. Then, projP𝒲2\text{proj}_{P}\mathcal{W}_{2} must also be an area-minimizing projection for 𝒲2\mathcal{W}_{2}, and

(4.29) supfn,m𝒲2𝒲2f=limσ1𝒲2fσ|𝒲2=|W2|.\sup_{f\in\mathcal{F}^{\mathcal{W}_{2}}_{n,m}}\int_{\mathcal{W}_{2}}f=\lim_{\sigma\to 1}\int_{\mathcal{W}_{2}}f_{\sigma}|_{\mathcal{W}_{2}}=|W_{2}^{*}|.
Proof.

Suppose 𝒲2𝒲1\mathcal{W}_{2}\supset\mathcal{W}_{1} is obtained via a gluing operation. Then by Definition 1.7, n,m𝒲1n,m𝒲2\mathcal{F}^{\mathcal{W}_{1}}_{n,m}\subset\mathcal{F}^{\mathcal{W}_{2}}_{n,m}. Moreover, we also have

(4.30) |W2||W1|=|projP𝒲1|=|projP𝒲2||W2|.\displaystyle|W_{2}^{*}|\geq|W_{1}^{*}|=|\text{proj}_{P}\mathcal{W}_{1}|=|\text{proj}_{P}\mathcal{W}_{2}|\geq|W_{2}^{*}|.

Thus, we conclude that |W1|=|W2||W_{1}^{*}|=|W_{2}^{*}|, and projP𝒲2\text{proj}_{P}\mathcal{W}_{2} is an area-minimizing projection for 𝒲2\mathcal{W}_{2}. By assumption, we can thus find a sequence of fσn,m𝒲1n,m𝒲2f_{\sigma}\in\mathcal{F}^{\mathcal{W}_{1}}_{n,m}\subset\mathcal{F}^{\mathcal{W}_{2}}_{n,m} such that

(4.31) limσ1𝒲2fσ=limσ1𝒲1fσ=|W1|=|W2|,\lim_{\sigma\to 1}\int_{\mathcal{W}_{2}}f_{\sigma}=\lim_{\sigma\to 1}\int_{\mathcal{W}_{1}}f_{\sigma}=|W_{1}^{*}|=|W_{2}^{*}|,

which finishes the proof for gluing operation.

Now suppose 𝒲2𝒲1\mathcal{W}_{2}\subset\mathcal{W}_{1} is obtained via a cutting operation. By Definition 1.7, projP𝒲1projP𝒲2\text{proj}_{P}\mathcal{W}_{1}\supset\text{proj}_{P}\mathcal{W}_{2}. For any xprojP𝒲1x\in\text{proj}_{P}\mathcal{W}_{1}, denote Qx=(projP)1xQ_{x}=(\text{proj}_{P})^{-1}x the mm-dimensional affine subspace orthogonal to projP𝒲1\text{proj}_{P}\mathcal{W}_{1} and passing through xx. By assumption, we can find a sequence fσn,m𝒲1f_{\sigma}\in\mathcal{F}^{\mathcal{W}_{1}}_{n,m} such that

(4.32) limσ1𝒲1fσ=|W1|.\lim_{\sigma\to 1}\int_{\mathcal{W}_{1}}f_{\sigma}=|W_{1}^{*}|.

Use the fact that projP𝒲1\text{proj}_{P}\mathcal{W}_{1} is an area-minimizing projection, we must have for a.e. xprojP𝒲1x\in\text{proj}_{P}\mathcal{W}_{1},

(4.33) limσ1Qx𝒲1fσ=1.\lim_{{\sigma\to 1}}\int_{Q_{x}\cap\mathcal{W}_{1}}f_{\sigma}=1.

On the other hand, for any xprojP𝒲2x\in\text{proj}_{P}\mathcal{W}_{2}, by definition of cutting

(4.34) 𝒲2=(projP)1(projP𝒲2)𝒲1(projP)1x𝒲1=Qx𝒲1\mathcal{W}_{2}=(\text{proj}_{P})^{-1}(\text{proj}_{P}\mathcal{W}_{2})\cap\mathcal{W}_{1}\supset(\text{proj}_{P})^{-1}x\cap\mathcal{W}_{1}=Q_{x}\cap\mathcal{W}_{1}

Certainly Qx𝒲2Qx𝒲1Q_{x}\cap\mathcal{W}_{2}\subset Q_{x}\cap\mathcal{W}_{1}, so Qx𝒲2=Qx𝒲1Q_{x}\cap\mathcal{W}_{2}=Q_{x}\cap\mathcal{W}_{1} for any xprojP𝒲2x\in\text{proj}_{P}\mathcal{W}_{2}. Therefore, fσ|𝒲2n,m𝒲2f_{\sigma}|_{\mathcal{W}_{2}}\subset\mathcal{F}^{\mathcal{W}_{2}}_{n,m} satisfies

(4.35) limσ1𝒲2fσ=limσ1projP𝒲2Qx𝒲2fσ=limσ1projP𝒲2Qx𝒲1fσ=|projP𝒲2|\lim_{{\sigma\to 1}}\int_{\mathcal{W}_{2}}f_{\sigma}=\lim_{{\sigma\to 1}}\int_{\text{proj}_{P}\mathcal{W}_{2}}\int_{Q_{x}\cap\mathcal{W}_{2}}f_{\sigma}=\lim_{{\sigma\to 1}}\int_{\text{proj}_{P}\mathcal{W}_{2}}\int_{Q_{x}\cap\mathcal{W}_{1}}f_{\sigma}=|\text{proj}_{P}\mathcal{W}_{2}|

Combining with (1.9), we obtain

(4.36) |W2||projP𝒲2|=limσ1𝒲2fσ|𝒲2supfn,m𝒲2𝒲2fσ|W2|.|W_{2}^{*}|\leq|\text{proj}_{P}\mathcal{W}_{2}|=\lim_{\sigma\to 1}\int_{\mathcal{W}_{2}}f_{\sigma}|_{\mathcal{W}_{2}}\leq\sup_{f\in\mathcal{F}^{\mathcal{W}_{2}}_{n,m}}\int_{\mathcal{W}_{2}}f_{\sigma}\leq|W_{2}^{*}|.

Hence projP𝒲2\text{proj}_{P}\mathcal{W}_{2} must be area-minimizing projection, which finishes the proof for cutting operation. ∎

Proposition 4.4.

Suppose 𝒲n,m\mathcal{W}\in\mathcal{L}_{n,m} is a long convex body, then there exists a function f~n,m\tilde{f}\in\mathcal{F}_{n,m} when m=1m=1, and a sequence of functions fσn,mf_{\sigma}\in\mathcal{F}_{n,m} when m=2m=2, such that (4.1) and (4.2) hold, respectively. Hence

(4.37) supfn,m𝒲f=|W|.\sup_{f\in\mathcal{F}_{n,m}}\int_{\mathcal{W}}f=|W^{*}|.
Proof.

By the discussion after Definition 1.7, any 𝒲n,m\mathcal{W}\in\mathcal{L}_{n,m} can be constructed through a finite number of gluing and cutting operations starting from an ellipsoid. It is also clear that Proposition 4.4 holds for ellipsoids by Proposition 4.1 and 4.2. Then for any long convex body, Proposition 4.4 directly follows from Lemma 4.3. In particular, let EE be the ellipsoid that generated 𝒲\mathcal{W}, then for m=1m=1 we can choose f~\tilde{f} to be the one in Proposition 4.1, and for m=2m=2 we can choose fσf_{\sigma} to be the sequence in Proposition 4.2, both restricted to 𝒲\mathcal{W}. ∎

Next, we establish the sharpness and rigidity properties when equality holds for 𝒲n,m\mathcal{W}\in\mathcal{L}_{n,m} and codimension m=1,2m=1,2.

Proposition 4.5.

Let codimension m=1,2m=1,2 and 𝒲n,m\mathcal{W}\in\mathcal{L}_{n,m}. Then the equality in (1.10) holds if and only if Σ\Sigma is homothetic to WW^{*}.

Proof.

It is clear from Remark 2.2 that if Σ\Sigma is homothetic to some WW^{*} then equality holds, so we only need to prove the converse.

First of all, if equality (1.10) holds for some Σ\Sigma, by (3.14) Σ\Sigma must be connected. Moreover, from Proposition 4.4 we know supn,m𝒲f=|W|\sup_{\mathcal{F}_{n,m}}\int_{\mathcal{W}}f=|W^{*}|, and Σ\Sigma achieves the equality of (1.10) if and only if Σ\Sigma achieves the equality of (1.7).

When m=1m=1, let f~n,m\tilde{f}\in\mathcal{F}_{n,m} be the density function found in Proposition 4.4 such that

(4.38) 𝒲f~=supn,m𝒲f=|W|.\int_{\mathcal{W}}\tilde{f}=\sup_{\mathcal{F}_{n,m}}\int_{\mathcal{W}}f=|W^{*}|.

For such f~\tilde{f}, the equalities in (3.12) must hold, meaning that

(4.39) |detDT(x,y)|=(PΦ(Σ)n|Σ|)nfor(x,y)Aa.e..\displaystyle|\det DT(x,y)|=\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}~{}\text{for}~{}(x,y)\in A~{}\text{a.e.}.

Moreover, we have |π(A)|=|Σ||\pi(A)|=|\Sigma| where π(A)={x:|{y:(x,y)A}|>0}\pi(A)=\{x:|\{y:(x,y)\in A\}|>0\} is the projection onto Σ\Sigma. Thus, the arithmetic-geometric mean inequality in Lemma 3.2 must be equality for a.e. points in AA, which only happens when DΣ2u(x)I​IΣ(x),y=PΦ(Σ)n|Σ|IdnD_{\Sigma}^{2}u(x)-\langle\textrm{I\!I}_{\Sigma}(x),y\rangle=\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\textrm{Id}_{n} is a constant multiple of identity matrix. Thus I​IΣ(x)=0\textrm{I\!I}_{\Sigma}(x)=0 a.e. in π(A)\pi(A), equivalent to a.e. in Σ\Sigma. Since Σ\Sigma is smooth, it is in some codimension mm affine subspace PP. By codimension zero Wulff inequality in Theorem 1.3, Σ\Sigma is homothetic to the Wulff shape corresponding to the restricted weight Φ|P\Phi|_{P}. Then by Lemma 2.1, Σ\Sigma must be homothetic to projP𝒲\text{proj}_{P}\mathcal{W}. By assumption Σ\Sigma achieves equality, so by Remark 2.2 we have

(4.40) PΦ(Σ)|Σ|n1n=PΦ(projP𝒲)|projP𝒲|n1n=n|projP𝒲|1nn|W|1n.\frac{P_{\Phi}(\Sigma)}{|\Sigma|^{\frac{n-1}{n}}}=\frac{P_{\Phi}(\text{proj}_{P}\mathcal{W})}{|\text{proj}_{P}\mathcal{W}|^{\frac{n-1}{n}}}=n|\text{proj}_{P}\mathcal{W}|^{\frac{1}{n}}\geq n|W^{*}|^{\frac{1}{n}}.

On the other hand, by (1.9) and the equality of (1.7), we also have

(4.41) PΦ(Σ)|Σ|n1nn|W|1n.\displaystyle\frac{P_{\Phi}(\Sigma)}{|\Sigma|^{\frac{n-1}{n}}}\leq n|W^{*}|^{\frac{1}{n}}.

Therefore, |projP𝒲|=|W||\text{proj}_{P}\mathcal{W}|=|W^{*}| and Σ\Sigma is homothetic to some area-minimizing projection WW^{*} in the codimension one case.

When m=2m=2, let the sequence fσn,mf_{\sigma}\in\mathcal{F}_{n,m} be the corresponding approximating density functions defined in Proposition 4.4, these functions satisfy

(4.42) limσ1𝒲fσ=supn,m𝒲f=|W|.\displaystyle\lim_{\sigma\to 1}\int_{\mathcal{W}}f_{\sigma}=\sup_{\mathcal{F}_{n,m}}\int_{\mathcal{W}}f=|W^{*}|.

Denote S=σsupp(fσ)S=\cap_{\sigma}\text{supp}(f_{\sigma}). If 𝒲\mathcal{W} is an ellipsoid we simply have S=𝒲S=\partial\mathcal{W}, and if 𝒲\mathcal{W} is a general long convex body, S=𝒲En+2S=\mathcal{W}\cap\partial E^{n+2}, where En+2E^{n+2} is the original ellipsoid that generates 𝒲\mathcal{W}. As shown in [Bre21, Section 3], we claim

Claim 4.6.

For all xΣx\in\Sigma and all yNxΣy\in N_{x}\Sigma satisfying T(x,y)ST(x,y)\in S, the equality DΣ2u(x)I​IΣ(x),y=PΦ(Σ)n|Σ|IdnD_{\Sigma}^{2}u(x)-\langle\textrm{I\!I}_{\Sigma}(x),y\rangle=\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\textrm{Id}_{n} holds.

Proof.

Suppose that claim fails at some x0Σx_{0}\in\Sigma and y0Nx0Σy_{0}\in N_{x_{0}}\Sigma satisfying T(x0,y0)ST(x_{0},y_{0})\in S and (x0,y0)A(x_{0},y_{0})\in A. Then we have DΣ2u(x0)I​IΣ(x0),y0PΦ(Σ)n|Σ|IdnD_{\Sigma}^{2}u(x_{0})-\langle\textrm{I\!I}_{\Sigma}(x_{0}),y_{0}\rangle\neq\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\textrm{Id}_{n}. The arithmetic-geometric inequality and (3.1) imply

(4.43) det(DΣ2u(x0)I​IΣ(x0),y0)<(PΦ(Σ)n|Σ|)n.\det(D^{2}_{\Sigma}u(x_{0})-\langle\textrm{I\!I}_{\Sigma}(x_{0}),y_{0}\rangle)<\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}.

By continuity, there exists ε(0,1)\varepsilon\in(0,1), and a neighborhood UU of (x0,y0)(x_{0},y_{0}) such that for all (x,y)UA(x,y)\in U\cap A, the following holds

(4.44) det(DΣ2u(x)I​IΣ(x),y)(1ε)(PΦ(Σ)n|Σ|)n.\displaystyle\det(D^{2}_{\Sigma}u(x)-\langle\textrm{I\!I}_{\Sigma}(x),y\rangle)\leq(1-\varepsilon)\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}.

Using Lemma 3.2, we deduce that on AA,

(4.45) 0detDT(x,y)(1ε1U(x,y))(PΦ(Σ)n|Σ|)n.\displaystyle 0\leq\det DT(x,y)\leq(1-\varepsilon 1_{U}(x,y))\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}.

Certainly in the case where (x0,y0)(x_{0},y_{0}) not in AA, there exists a neighborhood UU of (x0,y0)(x_{0},y_{0}) disjoint from AA, and (4.45) still holds. The inequality (3.12) becomes

𝒲fσ(ξ)𝑑ξ\displaystyle\int_{\mathcal{W}}f_{\sigma}(\xi)d\xi ΣNxΣfσ(T(x,y))|detDT(x,y)|1A(x,y)𝑑y𝑑x\displaystyle\leq\int_{\Sigma}\int_{N_{x}\Sigma}f_{\sigma}(T(x,y))|\det DT(x,y)|1_{A}(x,y)dydx
(PΦ(Σ)n|Σ|)nΣNxΣfσ(Σu(x)+y)𝑑y𝑑x\displaystyle\leq\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}\int_{\Sigma}\int_{N_{x}\Sigma}f_{\sigma}(\nabla^{\Sigma}u(x)+y)dydx
(4.46) ε(PΦ(Σ)n|Σ|)nUfσ(Σu(x)+y).\displaystyle-\varepsilon\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\right)^{n}\int_{U}f_{\sigma}(\nabla^{\Sigma}u(x)+y).

For all σ\sigma sufficiently close to 1, the integral Ufσ(Σu(x)+y)\int_{U}f_{\sigma}(\nabla^{\Sigma}u(x)+y) is bounded below by some positive constant independent on σ\sigma. Indeed, equality in (3.12) holds when taking the limit on both sides with respect to fσf_{\sigma}, which forces NxΣf(Σu(x)+y)𝑑y\int_{N_{x}\Sigma}f(\nabla^{\Sigma}u(x)+y)dy to converge to 1 for almost every xΣx\in\Sigma. Thus, the integral is bounded below by the fact that the support of fσf_{\sigma} concentrates around SS by construction. Now taking the limit as σ1\sigma\to 1 on both side of (4), we obtain

(4.47) supn,m𝒲f<(PΦ(Σ)n|Σ|n1n)n=|W|,\sup_{\mathcal{F}_{n,m}}\int_{\mathcal{W}}f<\left(\frac{P_{\Phi}(\Sigma)}{n|\Sigma|^{\frac{n-1}{n}}}\right)^{n}=|W^{*}|,

where the last equality follows from Remark 2.2 and the assumption that Σ\Sigma achieves equality of (1.10). This leads to a contradiction with the equality of (1.7). ∎

From Claim 4.6, we obtain

(4.48) DΣ2u(x)I​IΣ(x),y=PΦ(Σ)n|Σ|Idn\displaystyle D_{\Sigma}^{2}u(x)-\langle\textrm{I\!I}_{\Sigma}(x),y\rangle=\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\textrm{Id}_{n}

for all (x,y)NΣ(x,y)\in N\Sigma satisfying T(x,y)ST(x,y)\in S. Note that the intersection T(NxΣ)ST(N_{x}\Sigma)\cap S is a 11-dimensional ellipse in the 22-dimensional plane T(NxΣ)T(N_{x}\Sigma). Indeed this is clear when 𝒲\mathcal{W} is an ellipsoid; and when 𝒲\mathcal{W} is some other long convex body, any xx such that T(NxΣ)ST(N_{x}\Sigma)\cap S is not a full ellipse would result in NxΣf(Σu(x)+y)𝑑y\int_{N_{x}\Sigma}f(\nabla^{\Sigma}u(x)+y)dy being strictly smaller than 11. In this case, equality cannot be achieved when taking limit on both sides of (3.12), so such xx forms at most a measure-zero set. Thus, we conclude from above that I​IΣ(x)NxΣ\textrm{I\!I}_{\Sigma}(x)\perp N_{x}\Sigma, which implies I​IΣ(x)=0\textrm{I\!I}_{\Sigma}(x)=0 and DΣ2u(x)=PΦ(Σ)n|Σ|IdnD_{\Sigma}^{2}u(x)=\frac{P_{\Phi}(\Sigma)}{n|\Sigma|}\textrm{Id}_{n} for all xΣx\in\Sigma. Arguing as codimension m=1m=1 case, we conclude that Σ\Sigma is homothetic to some WW^{*}. This completes the proof of Proposition 4.5.

Proof of Theorem 1.9.

Theorem 1.9 directly follows from Theorem 1.5, the fact that n|W|=PΦ(W)n|W^{*}|=P_{\Phi}(W^{*}), Proposition 4.4 and Proposition 4.5. ∎

5. Proof of Theorem 1.11

In this section, we prove Theorem 1.11 and we use the following lemma about John’s ellipsoid.

Lemma 5.1 (John’s ellipsoid [Sch14]).

If KdK\subset\mathbb{R}^{d} is a centrally symmetric convex body with interior points, then there exists an ellipsoid EE such that

(5.1) EKdE.E\subset K\subset\sqrt{d}E.
Proof of Theorem 1.11.

Since 𝒲\mathcal{W} is centrally symmetric, by Lemma 5.1 there is John’s ellipsoid En+mE\subset\mathbb{R}^{n+m} such that

(5.2) E𝒲n+mE.E\subset\mathcal{W}\subset\sqrt{n+m}E.

Denote EE^{*} as the projection that achieves minimal projection area for EE, then

(5.3) |E||W|(n+m)n2|E|.|E^{*}|\leq|W^{*}|\leq(n+m)^{\frac{n}{2}}|E^{*}|.

Let n,mE\mathcal{F}_{n,m}^{E} and n,m𝒲\mathcal{F}_{n,m}^{\mathcal{W}} be the corresponding sets of functions from (1.8) for EE and 𝒲\mathcal{W}, respectively. We note that E𝒲E\subset\mathcal{W} implies n,mEn,m𝒲\mathcal{F}_{n,m}^{E}\subset\mathcal{F}_{n,m}^{\mathcal{W}}.

Define c~n,m\tilde{c}_{n,m} to be

(5.4) c~n,m={1m=1,2,((n+m)|Bn+m|m|Bm||Bn|)1nm3.\tilde{c}_{n,m}=\begin{cases}1&m=1,2,\\ \left(\frac{(n+m)|B^{n+m}|}{m|B^{m}||B^{n}|}\right)^{\frac{1}{n}}&m\geq 3.\end{cases}

By Proposition 4.1 and Proposition 4.2 we have

(5.5) supfn,mE𝒲fc~n,mn|E|.\sup_{f\in\mathcal{F}_{n,m}^{E}}\int_{\mathcal{W}}f\geq\tilde{c}_{n,m}^{n}|E^{*}|.

Therefore

(5.6) supfn,m𝒲𝒲fsupfn,mE𝒲fc~n,mn|E|(1n+mc~n,m)n|W|=cn,mn|W|,\sup_{f\in\mathcal{F}_{n,m}^{\mathcal{W}}}\int_{\mathcal{W}}f\geq\sup_{f\in\mathcal{F}_{n,m}^{E}}\int_{\mathcal{W}}f\geq\tilde{c}_{n,m}^{n}|E^{*}|\geq(\frac{1}{\sqrt{n+m}}\tilde{c}_{n,m})^{n}|W^{*}|=c_{n,m}^{n}|W^{*}|,

where the third inequality follows from (5.3). Then, by Theorem 1.5 and the fact that n|W|=PΦ(W)n|W^{*}|=P_{\Phi}(W^{*}) from (2.2), we conclude that

(5.7) PΦ(Σ)|Σ|n1nn(supn,m𝒲𝒲f)1ncn,mn|W|1n=cn,mPΦ(W)|W|n1n.\frac{P_{\Phi}({\Sigma})}{|\Sigma|^{\frac{n-1}{n}}}\geq n(\sup_{\mathcal{F}^{\mathcal{W}}_{n,m}}\int_{\mathcal{W}}f)^{\frac{1}{n}}\geq c_{n,m}\cdot n|W^{*}|^{\frac{1}{n}}=c_{n,m}\frac{P_{\Phi}({W^{*}})}{|W^{*}|^{\frac{n-1}{n}}}.

References

  • [AHO74] Herbert J. Alexander, David A. Hoffman, and Robert Osserman. Area estimates for submanifolds of euclidean space. In Symposia Mathematica, volume 14, pages 445–455, 1974.
  • [Alm86] Frederick J. Almgren. Optimal isoperimetric inequalities. Indiana University mathematics journal, 35(3):451–547, 1986.
  • [Bre21] Simon Brendle. The isoperimetric inequality for a minimal submanifold in Euclidean space. J. Amer. Math. Soc., 34(2):595–603, 2021.
  • [BZ13] Yurii D. Burago and Viktor A. Zalgaller. Geometric inequalities, volume 285. Springer Science & Business Media, 2013.
  • [Cab08] Xavier Cabré. Elliptic PDE’s in probability and geometry: Symmetry and regularity of solutions. Discrete Contin. Dyn. Syst., 20(3):425–457, 2008.
  • [Car21] Torsten Carleman. Zur Theorie der Minimalflächen. Math. Z., 9:154–160, 1921.
  • [Cha01] Isaac Chavel. Isoperimetric inequalities: differential geometric and analytic perspectives, volume 145. Cambridge University Press, 2001.
  • [CROS16] Xavier Cabré, Xavier Ros-Oton, and Joaquim Serra. Sharp isoperimetric inequalities via the abp method. Journal of the European Mathematical Society (EMS Publishing), 18(12), 2016.
  • [DPP] Guido De-Philippis and Alessandro Pigati. Michael-simon inequality for anisotropic varifolds. In prepration.
  • [DRKS20] Antonio De Rosa, Sławomir Kolasiński, and Mario Santilli. Uniqueness of critical points of the anisotropic isoperimetric problem for finite perimeter sets. Archive for Rational Mechanics and Analysis, 238(3):1157–1198, 2020.
  • [FMP10] Alessio Figalli, Francesco Maggi, and Aldo Pratelli. A mass transportation approach to quantitative isoperimetric inequalities. Inventiones mathematicae, 182(1):167–211, 2010.
  • [Kle12] Peter Paul Klein. On the ellipsoid and plane intersection equation. Applied Mathematics, 3(11):1634–1640, 2012.
  • [MG13] Robert J. McCann and Nestor Guillen. Five lectures on optimal transportation: geometry, regularity and applications. In Analysis and geometry of metric measure spaces. Lecture notes of the 50th Séminaire de Mathématiques Supérieures (SMS), Montréal, Canada, June 27 – July 8, 2011, pages 145–180. Providence, RI: American Mathematical Society (AMS), 2013.
  • [Oss78] Robert Osserman. The isoperimetric inequality. Bulletin of the American Mathematical Society, 84(6):1182–1238, 1978.
  • [Sch14] Rolf Schneider. Convex bodies: the Brunn-Minkowski theory, volume 151 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, expanded edition, 2014.
  • [Tay74] Jean E. Taylor. Existence and structure of solutions to a class of nonelliptic variational problems. In Symposia Mathematica, volume 14, pages 499–508, 1974.
  • [Tay75] Jean E. Taylor. Unique structure of solutions to a class of nonelliptic variational problems. Differ. Geom., Proc. Symp. Pure Math. 27, Part 1, Stanford 1973, 419-427 (1975), 1975.
  • [Tay78] Jean E. Taylor. Crystalline variational problems. Bulletin of the American Mathematical Society, 84(4):568–588, 1978.
  • [Wul01] Georg. Wulff. Zur frage der geschwindigkeit des wachstums und der auflosung der kristall achen. Z. Kristallogr, 34:449–530, 1901.