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

Limiting shape of the LpL_{p}-Minkowski problem

Shi-Zhong Du The Department of Mathematics, Shantou University, Shantou, 515063, P. R. China. szdu@stu.edu.cn Xu-Jia Wang Institute for Theoretical Sciences, Westlake University, Hangzhou 310024, China. wangxujia@westlake.edu.cn  and  Bao-Cheng Zhu School of Mathematics and Statistics, Shaanxi Normal University, Xi’an, 710119, China bczhu@snnu.edu.cn
Abstract.

In [3], Andrews classified the limiting shape for isotropic curvature flow corresponding to the solutions of the LpL_{p}-Minkowski problem as pp\to-\infty in the planar case. In this paper, we use the group-invariant method to study the asymptotic shape of solutions to the LpL_{p}-Minkowski problem as pp\to-\infty in high dimensions. For any regular polytope TT, we establish the existence of a solution Ω(p){\Omega^{(p)}} to the LpL_{p}-Minkowski problem that converges to TT as pp\to-\infty, thereby revealing the intricate geometric structure underlying this limiting behavior. We also extend the result to the dual Minkowski problem.

Key words and phrases:
LpL_{p} Minkowski problem, limiting shape, polytope.
2010 Mathematics Subject Classification:
35J20; 35J60; 52A40; 53A15
The first author was supported by NSFC (12171299); the second author was supported by the start-up fund from Westlake University; the third author was supported by NSFC (12371060) and the Sydney Mathematical Research Institute at USYD

1. Introduction

The classical Minkowski problem looks for a convex body such that its surface measure matches a given Radon measure μ\mu on the unit sphere 𝕊n{\mathbb{S}}^{n} in n+1\mathbb{R}^{n+1}. This problem dates back to the early works of Minkowski, Aleksandrov and Fenchel-Jessen (see [33]). For a convex body Ω𝒦0\Omega\in{\mathcal{K}}_{0} (the set of convex bodies containing the origin) and two parameters p,qp,q\in{\mathbb{R}}, Lutwak,Yang and Zhang [30] introduced the measure

𝒞p,q(Ω,ω)=G1(ω)|y|qn1hp1(G(y))𝑑n,\mathcal{C}_{p,q}(\Omega,\omega)=\int_{G^{-1}(\omega)}\frac{|y|^{q-n-1}}{h^{p-1}(G(y))}d{\mathcal{H}}^{n},

for any measurable set ω𝕊n\omega\subset{\mathbb{S}}^{n}, and proposed the following dual Minkowski problem. Given a finite Radon measure μ\mu on 𝕊n{\mathbb{S}}^{n}, find a convex body Ω𝒦0\Omega\in{\mathcal{K}}_{0} such that d𝒞p,q(Ω,)=dμd\mathcal{C}_{p,q}(\Omega,\cdot)=d\mu. Here, G:Ω𝕊nG:\partial\Omega\rightarrow{\mathbb{S}}^{n} is the Gauss map, h:𝕊nh:{\mathbb{S}}^{n}\to{\mathbb{R}} is the support function of Ω𝒦0\Omega\in\mathcal{K}_{0}, and n\mathcal{H}^{n} is the nn-dimensional Hausdorff measure. In the smooth category, the measure μ\mu is given by dμ=fdσd\mu=fd\sigma for a positive function ff, where dσd\sigma is the spherical Lebesgue measure, the LpL_{p} dual Minkowski problem can then be formulated by the equation

(1.1) det(2h+hI)=fhp1(h2+|h|2)n+1q2,x𝕊n.\det(\nabla^{2}h+hI)=fh^{p-1}(h^{2}+|\nabla h|^{2})^{\frac{n+1-q}{2}},\ \ \forall\ x\in{\mathbb{S}}^{n}.

This is a fully nonlinear equation of Monge-Ampère type, which has received increasing attention in recent years (see e.g. [6, 11, 21, 22, 30, 35, 36]). When q=n+1q=n+1, equation (1.1) reduces to the LpL_{p}-Minkowski problem, introduced in [28], given by

(1.2) det(2h+hI)=fhp1,x𝕊n.\det(\nabla^{2}h+hI)=fh^{p-1},\ \ \forall\ x\in{\mathbb{S}}^{n}.

Here, \nabla denotes the standard Levi-Civita connection on the sphere 𝕊n{\mathbb{S}}^{n}, and II is the identity matrix.

The LpL_{p}-Minkowski problem has been extensively studied in recent years (see e.g. [1, 2, 4, 7, 8, 16, 19, 20, 23, 24, 27, 29, 31, 32, 37, 38]). According to the Blaschke-Santaló inequality, the problem is divided into three cases, the sub-critical case p>n1p>-n-1, the critical case p=n1p=-n-1, and the super-critical case p<n1p<-n-1. There are rich phenomena on the existence and multiplicity of solutions. In the sub-critical case p(n1,0]p\in(-n-1,0], the solution is usually not unique [25]. When p<np<-n, there may be infinitely many solutions [26]. In the critical case p=n1p=-n-1, due to a Kazdan-Warner type obstruction [13], there may be no solutions to the problem (1.2). In the super-critical case p<n1p<-n-1, surprisingly, Guang, Li and the second author [18] proved that for any p<n1p<-n-1 and any positive, smooth function ff, there is a solution to (1.2). An amazing result was obtained by Andrews [3]. He proved that in the planar case n=1n=1, for any regular polygon TkT_{k} (with kk-sides), there is a solution to (1.2) which converges to TkT_{k}, as pp tends to negative infinity. This raises a natural question of whether such limiting behavior holds in high dimensions:

Problem 1.1.

(i) Does there exist a solution h(p){h^{(p)}} to the LpL_{p}-Minkowski problem (1.2), such that the associated convex body Ω(p){\Omega^{(p)}} converges to a polytope TT, which is tangential to 𝕊n{\mathbb{S}}^{n}, as pp\to-\infty?
(ii) Moreover, for any regular polytope TT tangent to the unit sphere 𝕊n{\mathbb{S}}^{n}, does there exist a solution h(p){h^{(p)}} to (1.2) for negative large pp, such that the associated convex body Ω(p){\Omega^{(p)}} converges to the polytope TT as pp\to-\infty?

In fact, Andrews classified the limiting shape of the solutions to (1.2) in the planar case by solving an isotropic curvature flow. In this paper, we explore a new method, called the group-invariant method, to study the limiting shape of the solutions to (1.2) in high dimensions. To do so, we assume that, throughout the paper, ff satisfies the following conditions:

  • (i)

    ff is a measurable function satisfying

    (1.3) c1fc2c_{1}\leq f\leq c_{2}

    for positive constants c1,c2c_{1},c_{2}.

  • (ii)

    ff is invariant with respect to a group 𝒮{\mathcal{S}}, which is a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property (introduced in Section 2).

A particular case is when 𝒮\mathcal{S} is the discrete subgroup of O(n+1)O(n+1) generated by a regular polytope.

First, we obtain the asymptotic behavior of solutions to the LpL_{p}-Minkowski problem (1.2) for negative large pp in high dimensions, which solves part (i) of Problem 1.1.

Theorem 1.1.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfies the spanning property. Then for p<n1p<-n-1, there exists an 𝒮{\mathcal{S}}-invariant solution h(p){h^{(p)}} to the LpL_{p}-Minkowski problem (1.2), such that the associated convex body Ω(p){\Omega^{(p)}} sub-converges to a group invariant polytope TT, which is tangential to 𝕊n{\mathbb{S}}^{n}, as pp\to-\infty.

We say that a polytope TT is tangential to the unit sphere 𝕊n{\mathbb{S}}^{n} if every face of TT is tangential to 𝕊n{\mathbb{S}}^{n}. It is easy to see that if TT is tangential to 𝕊n{\mathbb{S}}^{n}, then TT is convex and TT contains the unit ball B1{B_{1}}.

By the regularity theory of Caffarelli [9, 10], the solution h(p)h^{(p)} is C1,αC^{1,\alpha} if ff is positive and bounded, and is C2,αC^{2,\alpha} if ff is also Hölder continuous. By the maximum principle, there is a unique solution h(p)h^{(p)} to (1.2) when p>n+1p>n+1 [13]. It is easy to see that the solution h(p)h^{(p)} tends to constant one when pp tends to infinity. In fact, at the maximum point of h(p)h^{(p)}, we have

hpn1det(2h+hI)/(fhn)1/f,h=h(p)h^{p-n-1}\leq\det(\nabla^{2}h+hI)/(fh^{n})\leq 1/f,\ \ h=h^{(p)}

by equation (1.2). Hence limp+suph(p)=1\lim_{p\to+\infty}\sup h^{(p)}=1. Similarly, one can show that limp+infh(p)=1\lim_{p\to+\infty}\inf h^{(p)}=1 by evaluating (1.2) at the minimum point of h(p)h^{(p)}. We conclude that the limiting shape of the solutions Ω(p)\Omega^{(p)} is the unit ball B1{B_{1}} as p+p\to+\infty.

Theorem 1.1 deals with the case p<n1p<-n-1. When f1f\equiv 1, there is a trivial solution h1h\equiv 1 to (1.2). Theorem 1.1 tells that there is a non-trivial solution h(p)h^{(p)} when pp is negative large. The solution h(p)h^{(p)} is a maximizer of the functional

(1.4) p(Ω)=V(Ω)(𝕊nfhp/𝕊nf)(n+1)/p\mathcal{F}_{p}(\Omega)=V(\Omega)\Big{(}\int_{{\mathbb{S}}^{n}}fh^{p}{\Big{/}}\int_{{\mathbb{S}}^{n}}f\Big{)}^{-(n+1)/p}

among all 𝒮\mathcal{S}-invariant convex bodies Ω\Omega containing the origin, where h=hΩh=h_{\Omega} is the support function of Ω\Omega,

V(Ω):=𝕊nhdet(2h+hI).V(\Omega):=\int_{{\mathbb{S}}^{n}}h\det(\nabla^{2}h+hI).

Here the value V(Ω)V(\Omega) is equal to (n+1)(n+1) times the volume |Ω||\Omega|. The spanning property of the group 𝒮\mathcal{S} guarantees the uniform estimate of group invariant solutions, and one can obtain the existence of solutions for all pn+1p\neq n+1. By property of the functional p\mathcal{F}_{p}, we show that the maximizer h(p)h^{(p)} is not a constant when pp is negative large, and converges to a limit as pp tends to negative infinity. Moreover, the limit is a polytope.

The next theorem shows that every regular polytope tangent to the unit sphere is a limit polytope, thereby resolving part (ii) of Problem 1.1.

Theorem 1.2.

For any regular polytope TT tangential to the unit sphere 𝕊n{\mathbb{S}}^{n}, there exists solution h(p){h^{(p)}} to (1.2) for negative large pp, such that the associated convex body Ω(p){\Omega^{(p)}} converges to the polytope TT as pp\to-\infty.

A regular polytope has strong symmetry. We prove Theorem 1.2 by considering maximizers of the functional p\mathcal{F}_{p} among convex bodies with the strong symmetry. Theorems 1.1 and 1.2 can be regarded as a higher dimensional extension of Andrews’s result in ([3], Theorem 1.5) in the planar case. Here we allow ff to be a measurable function.

Next we consider the dual Minkowski problem (1.1). Under the group invariant assumptions in Theorem 1.1, we have uniform estimate for the group invariant solution h(p,q)h^{(p,q)} to (1.1), which is a maximizer of the functional

(1.5) p,q(Ω)=Vq(Ω)(𝕊nfhp/𝕊nf)q/p\mathcal{F}_{p,q}(\Omega)=V_{q}(\Omega)\bigg{(}\int_{{\mathbb{S}}^{n}}fh^{p}{\Big{/}}\int_{{\mathbb{S}}^{n}}f\bigg{)}^{-q/p}

among all 𝒮\mathcal{S}-invariant convex bodies Ω𝒦0(𝒮)\Omega\subset\mathcal{K}_{0}(\mathcal{S}), where

Vq(Ω)=𝕊nrq(y)𝑑y,V_{q}(\Omega)=\int_{\mathbb{S}^{n}}r^{q}(y)dy,

and Vn+1(Ω)=V(Ω)V_{n+1}(\Omega)=V(\Omega).

Here we are interested in the asymptotic behaviour of the solution h(p,q)h^{(p,q)} as pp tends to negative infinity for fixed qq, and the asymptotic behaviour as qq tends to positive infinity for fixed pp.

Theorem 1.3.

Let 𝒮\mathcal{S} be as in Theorem 1.1. Let h(p,q)h^{(p,q)} be the maximizer of the functional p,q\mathcal{F}_{p,q} and Ω(p,q)\Omega^{(p,q)} be the associated convex body.

  • (1)

    For any given q>0q>0, Ω(p,q)\Omega^{(p,q)} sub-converges to a group invariant polytope TT, which contains the unit ball B1{B_{1}} and is tangential to 𝕊n\mathbb{S}^{n}, as pp\to-\infty.

  • (2)

    For any given q<0q<0, Ω(p,q)\Omega^{(p,q)} converges to the unit ball as pp\to-\infty.

  • (3)

    For any given p0p\neq 0, Ω(p,q)\Omega^{(p,q)} sub-converges to a group invariant polytope TT, which contains the unit ball B1{B_{1}} and is tangential to 𝕊n\mathbb{S}^{n}, as q+q\to+\infty.

The next theorem shows that every regular polytope tangential to the unit sphere is a limit polytope.

Theorem 1.4.

Let Tn+1T\subset{\mathbb{R}}^{n+1} be a regular polytope and let (p,q)2(p,q)\in{\mathbb{R}}^{2} satisfy pq0pq\not=0.

  • (1)

    If TT is tangential to the unit sphere 𝕊n{\mathbb{S}}^{n}, there exists a local maximizer Ω(p,q)\Omega^{(p,q)} of the functional VqV_{q}, such that when q>0q>0, Ω(p,q)\Omega^{(p,q)} converges to TT as pp\to-\infty.

  • (2)

    When p0p\neq 0, there exists a local maximizer Ω(p,q)\Omega^{(p,q)} of VqV_{q}, which converges to a regular polytope similar to TT as q+q\to+\infty.

This paper is organized as follows. In Section 2, we introduce some notations and properties concerning subgroups 𝒮O(n+1)\mathcal{S}\subset O(n+1) and the corresponding group invariant polytopes. In Section 3, we prove the existence of non-trivial group invariant solutions to (1.2) by a variational method for negative large pp. By passing to the limit pp\to-\infty, we characterize the limiting convex body Ω()\Omega^{(-\infty)} in Section 4. In Section 5, we show that a regular polytope TT is a strictly local maximizer of the functional p\mathcal{F}_{p} and thus give a proof of Theorem 1.2. Finally, we consider the LpL_{p} dual Minkowski problem (1.1) and prove Theorems 1.3-1.4 in Sections 6-7.

2. Discrete subgroup and group invariant polytope

2.1. Notations

We will work in Euclidean space, where |x|=xx|x|=\sqrt{x\cdot x} for any xn+1x\in\mathbb{R}^{n+1}. A compact convex set with non-empty interior in n+1\mathbb{R}^{n+1} is called a convex body. Denote by 𝒦0{\mathcal{K}}_{0} the set of convex bodies in n+1{\mathbb{R}}^{n+1} containing the origin. Given a convex body Ω𝒦0\Omega\in\mathcal{K}_{0}, the support function h:𝕊nh:{\mathbb{S}}^{n}\to{\mathbb{R}} of Ω\Omega is defined by

h(x)=max{zx|zΩ},x𝕊n.h(x)=\max\big{\{}z\cdot x\ |\ z\in\Omega\big{\}},\ \ \forall x\in{\mathbb{S}}^{n}.

We denote by hij=ij2hh_{ij}=\nabla^{2}_{ij}h the Hessian matrix of hh with respect to an orthonormal frame {ei}i=1n\{e_{i}\}_{i=1}^{n} on the unit sphere 𝕊n{\mathbb{S}}^{n}, where δij\delta_{ij} is the Kronecker delta symbol. When Ω\Omega is uniformly convex, the matrix (Aij):=(hij+hδij)(A_{ij}):=(h_{ij}+h\delta_{ij}) is positive definite, with inverse (Aij)(A^{ij}). Let the radial distance function of Ω\partial\Omega be given by

r(y)=sup{λ>0|λyΩ},yn+1.r(y)=\sup\big{\{}\lambda>0\ |\ \lambda y\in\Omega\big{\}},\ \forall\ y\in{\mathbb{R}}^{n+1}.

There holds r(y)y=h(x)x+h(x)r(y)y=h(x)x+\nabla h(x), where xx is the unit outer normal of Ω\partial\Omega at the point r(y)yΩr(y)y\in\partial\Omega, such that G:r(y)yxG:\ r(y)y\to x is the Gauss map. Thus, Ω\Omega can be recovered from hh by

Ω={h(x)x+h(x)|x𝕊n}.\partial\Omega=\{h(x)x+\nabla h(x)\ |\ x\in{\mathbb{S}}^{n}\}.

The surface measure of Ω\partial\Omega is given by

dS=det(2h+hI)dσ,dS=\det(\nabla^{2}h+hI)d\sigma,

where dσd\sigma is the surface measure of the sphere 𝕊n{\mathbb{S}}^{n}.

2.2. Discrete subgroup and group invariant polytope

In this subsection, we introduce some notations and properties about the group invariant polytopes that will be used later. According to [34], a group is a set that is closed under multiplication, and satisfies the associative law, has a unit element, and has inverses. By definition, each discrete subgroup 𝒮{\mathcal{S}} of O(n+1)O(n+1) is a finite group. For example,

𝒮a={(cosθsinθsinθcosθ),θ=j×2π/a,j}O(2){\mathcal{S}}_{a}=\bigg{\{}\left(\begin{array}[]{cc}\cos\theta&\sin\theta\\ -\sin\theta&\cos\theta\end{array}\right),\ \ \theta=j\times 2\pi/a,\ \ j\in{\mathbb{Z}}\bigg{\}}\subset O(2)

is a discrete subgroup when aa is rational, and an infinite subgroup when aa is irrational.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1). We say that 𝒮{\mathcal{S}} satisfies the spanning property if for any a𝕊na\in{\mathbb{S}}^{n}, the set

Pa:=conv{ϕ(a),ϕ𝒮}P_{a}:=\text{conv}\{\phi(a),\phi\in{\mathcal{S}}\}

forms a non-degenerate, (n+1)(n+1)-dimensional polytope in n+1{\mathbb{R}}^{n+1}. Here we denote by conv(Ψ)\text{conv}(\Psi) the convex hull of the set Ψ\Psi. Moreover, we also define the bounded ratio condition by

(2.1) supa𝕊nγa<,γa:=max𝕊nha/min𝕊nha,\sup_{a\in{\mathbb{S}}^{n}}\gamma_{a}<\infty,\ \ \gamma_{a}:=\max_{{\mathbb{S}}^{n}}h_{a}/\min_{{\mathbb{S}}^{n}}h_{a},

for the support function hah_{a} of PaP_{a}. We have the following equivalence.

Proposition 2.1.

For a discrete subgroup 𝒮{\mathcal{S}} of O(n+1)O(n+1), the spanning property is equivalent to the bounded ratio condition (2.1).

Proof.

If the spanning property does not hold for some a𝕊na\in{\mathbb{S}}^{n}, then PaP_{a} is a degenerate polytope in (n+1)(n+1)-dimension. In this case, the bounded ratio condition fails.

Conversely, if the bounded ratio condition does not hold for a sequence aj𝕊na_{j}\in{\mathbb{S}}^{n} with aja_{j} tending to some a𝕊na\in{\mathbb{S}}^{n}, it is clear that PaP_{a} is degenerate. ∎

The spanning property is important in the sense that it guarantees the a-priori estimates for the maximizing sequence of our variational problem in Section 3. Let us point out that in the spanning property, the set Pa=conv{ϕ(a),ϕ𝒮}P_{a}=\text{conv}\{\phi(a),\phi\in{\mathcal{S}}\} is generated by only one arbitrary point aa.

For a given discrete subgroup 𝒮O(n+1){\mathcal{S}}\subset O(n+1) satisfying the spanning property, there may be more than one polytope corresponding to the subgroup 𝒮{\mathcal{S}}. For example, let 𝒮O(2)\mathcal{S}\subset O(2) be the discrete subgroup associated with the regular kk-polygon for k=5k=5, then all regular kk-polygons with k=5mk=5m are also 𝒮\mathcal{S}-invariant, for all integers m2m\geq 2.

Regular polytopes are of particular interest. According to [14, 15], we introduce the following definition for regular polytopes.

Definition 2.1.

A flag is a sequence of faces of a polytope TT, each contained in the next, with exactly one face from each dimension. More precisely, a flag ψ\psi of an nn-dimensional polytope TT is a set {F0,,Fn+1}\{F_{0},\cdots,F_{n+1}\}, where FiF_{i} for 0in+10\leq i\leq n+1 is the ii-dimensional face of TT, such that FiFi+1F_{i}\subset F_{i+1} for all ii.

Definition 2.2.

A regular polytope Tn+1T\subset{\mathbb{R}}^{n+1} is a polytope whose symmetry group 𝒮T{\mathcal{S}}_{T}, as a maximal subgroup of O(n+1)O(n+1), acts transitively on its flags in the sense for any flags x,yTx,y\subset T, there exists ϕ𝒮T\phi\in{\mathcal{S}}_{T} such that ϕ(x)=y\phi(x)=y.

Regular polytopes are the high dimensional analog of regular polygons (n=1n=1) and regular polyhedra (n=2n=2). A concise symbolic representation for regular polytopes was developed by Schläfli in the 19th century, and a slightly modified form has become standard nowadays [14]. Regular polytopes can be classified according to the symmetry group 𝒮O(n+1){\mathcal{S}}\subset O(n+1). Regular polytopes with kk vertices in dimension n1n\geq 1 are classified as follows,

(2.2) {When n=1,there exist k-regular polytopes for each k,k3.When n=2,there exist only k-regular polytopes for k=4,6,8,12,20.When n=3,there exist only k-regular polytopes for k=5,8,16,24,120,600.When n4,there exist only k-regular polytopes for k=n+2,2n+2,2n+1.\begin{cases}\mbox{When }n=1,\mbox{there exist }k\mbox{-regular polytopes for each }k\in{\mathbb{N}},k\geq 3.\\ \mbox{When }n=2,\mbox{there exist only }k\mbox{-regular polytopes for }k=4,6,8,12,20.\\ \mbox{When }n=3,\mbox{there exist only }k\mbox{-regular polytopes for }k=5,8,16,24,120,600.\\ \mbox{When }n\geq 4,\mbox{there exist only }k\mbox{-regular polytopes for }k=n+2,2n+2,2^{n+1}.\end{cases}
Example 2.1.

Let C=conv{(±1,±1,±1)}3C=\text{conv}\{(\pm 1,\pm 1,\pm 1)\}\subset{\mathbb{R}}^{3} be a standard cube. Let 𝒮O(3){\mathcal{S}}\subset O(3) be the associated symmetric group. The following statements hold.

  • (1)

    When a=(1,1,1)a=(1,1,1), conv{ϕ(a),ϕ𝒮}\text{conv}\{\phi(a),\phi\in{\mathcal{S}}\} is a standard cube in 3{\mathbb{R}}^{3}, which is a regular polytope.

  • (2)

    When a=(1,0,0)a=(1,0,0), conv{ϕ(a),ϕ𝒮}\text{conv}\{\phi(a),\phi\in{\mathcal{S}}\} is a cross-polytope in 3{\mathbb{R}}^{3}, which is also a regular polytope.

  • (3)

    When a=(1,1,r),r(0,1)a=(1,1,r),r\in(0,1), conv{ϕ(a),ϕ𝒮}\text{conv}\{\phi(a),\phi\in{\mathcal{S}}\} is a group invariant polytope formed by cutting eight solid angles from eight vertices of the cube, which is not a regular polytope.

3. Existence of group invariant solutions of (1.2)

Group invariant solutions have been studied in various geometric problems. Here we consider the discrete subgroup 𝒮O(n+1){\mathcal{S}}\subset O(n+1) that satisfies the spanning property. By the bounded ratio condition in Section 2, we can prove the following existence result.

Theorem 3.1.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property. For p<n1p<-n-1, equation (1.2) admits a group invariant solution h(p){{h^{(p)}}}, which is a maximizer of the global variational problem

(3.1) supΩ𝒦p(𝒮)V(Ω)\sup_{\Omega\in{\mathcal{K}}_{p}({\mathcal{S}})}V(\Omega)

over the group 𝒮{\mathcal{S}} invariant family

𝒦p(𝒮):={Ω𝒦0(𝒮)|𝕊nfhΩp=𝕊nf},{\mathcal{K}}_{p}({\mathcal{S}}):=\bigg{\{}\Omega\in{\mathcal{K}}_{0}({\mathcal{S}})\ \Big{|}\ \int_{{\mathbb{S}}^{n}}fh_{\Omega}^{p}=\int_{{\mathbb{S}}^{n}}f\bigg{\}},

where 𝒦0(𝒮){\mathcal{K}}_{0}({\mathcal{S}}) is the set of group 𝒮{\mathcal{S}} invariant convex bodies containing the origin, and hΩh_{\Omega} is the support function of Ω\Omega.

To prove Theorem 3.1, it suffices to prove

Lemma 3.1.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property. For each p<n1p<-n-1, the variational problem (3.1) admits a positive maximizer h(p){{h^{(p)}}} satisfying the equation (1.2) up to a constant multiplication.

To prove Lemma 3.1, we first prove the following lemmas.

Lemma 3.2.

For n1n\geq 1, p<0p<0, there exists a positive constant Cn(𝒮)C_{n}({\mathcal{S}}) depending only on nn and 𝒮{\mathcal{S}}, such that

(3.2) hΩ(x)Cn(𝒮),x𝕊n,h_{\Omega}(x)\leq C_{n}({\mathcal{S}}),\ \ \forall\ x\in{\mathbb{S}}^{n},

for each Ω𝒦p(𝒮)\Omega\in{\mathcal{K}}_{p}({\mathcal{S}}).

Proof.

For any Ω𝒦p(𝒮)\Omega\in{\mathcal{K}}_{p}({\mathcal{S}}), we can assume that ΩBR\Omega\subset B_{R} (the ball with radius RR centered at 0) and that RbΩRb\in\partial\Omega for some R>0R>0 and b𝕊nb\in{\mathbb{S}}^{n}. Noting that ΩPb=Rconv{ϕ(b)|ϕ𝒮}\Omega\supset P_{b}=R\,\text{conv}\{\phi(b)|\ \phi\in{\mathcal{S}}\} by convexity, we have

(3.3) max𝕊nhΩ/min𝕊nhΩγbγmax(𝒮),\max_{{\mathbb{S}}^{n}}h_{\Omega}/\min_{{\mathbb{S}}^{n}}h_{\Omega}\leq\gamma_{b}\leq\gamma_{max}({\mathcal{S}}),

where γmax(𝒮)=supa𝕊nγa\gamma_{max}({\mathcal{S}})=\sup_{a\in{\mathbb{S}}^{n}}\gamma_{a} and γa\gamma_{a} is the ratio introduced in (2.1). Thus, we obtain that

(3.4) 𝕊nf=𝕊nfhpmax𝕊nhp𝕊nf(max𝕊nhγmax(𝒮))p𝕊nf.{\begin{split}\int_{{\mathbb{S}}^{n}}f&=\int_{{\mathbb{S}}^{n}}fh^{p}\leq\max_{{\mathbb{S}}^{n}}h^{p}\int_{{\mathbb{S}}^{n}}f\\ &\leq\bigg{(}\frac{\max_{{\mathbb{S}}^{n}}h}{\gamma_{max}({\mathcal{S}})}\bigg{)}^{p}\int_{{\mathbb{S}}^{n}}f.\end{split}}

This gives exactly the desired estimate (3.2) for negative pp. ∎

Lemma 3.3.

For n1n\geq 1, p<np<-n, and a group 𝒮{\mathcal{S}} invariant positive function ff on 𝕊n\mathbb{S}^{n}, there exists a positive constant CnC_{n} depending only on nn such that

(3.5) h(x)[(p)nCnγmax(𝒮)max𝕊nfmin𝕊nf]1n+p,x𝕊n,h(x)\geq\Big{[}(-p)^{n}\cdot C_{n}\cdot\gamma_{max}({\mathcal{S}})\cdot\frac{\max_{{\mathbb{S}}^{n}}f}{\min_{{\mathbb{S}}^{n}}f}\Big{]}^{\frac{1}{n+p}},\ \ \forall\ x\in{\mathbb{S}}^{n},

for any support function hh with the associated convex body in 𝒦p(𝒮){\mathcal{K}}_{p}({\mathcal{S}}).

Proof.

Denote M=max𝕊nhM=\max_{{\mathbb{S}}^{n}}h and m=min𝕊nhm=\min_{{\mathbb{S}}^{n}}h. Assume that m=h(x0)m=h(x_{0}) for the south polar x0x_{0}. By the convexity of hh, we have

h(x)m+Msinθ,x𝕊n,h(x)\leq m+M\sin\theta,\ \ \forall x\in{\mathbb{S}}^{n},

where θ=dist(x,x0)\theta=\text{dist}(x,x_{0}) is the geodesic distance from x0x_{0} to xx. Hence,

𝕊nfhp\displaystyle\int_{{\mathbb{S}}^{n}}fh^{p} \displaystyle\geq C1min𝕊nf0π/2θn1dθ(m+Mθ)p\displaystyle C_{1}\min_{{\mathbb{S}}^{n}}f\int^{\pi/2}_{0}\frac{\theta^{n-1}d\theta}{(m+M\theta)^{-p}}
=\displaystyle= C1Mnmp+nmin𝕊nf0Mπ/(2m)tn1(1+t)p𝑑t\displaystyle C_{1}M^{-n}m^{p+n}\min_{{\mathbb{S}}^{n}}f\int^{M\pi/(2m)}_{0}\frac{t^{n-1}}{(1+t)^{-p}}dt

for a positive constant C1C_{1} depending only on nn. By taking some large constant C2>0C_{2}>0, noting that

0Mπ/(2m)tn1(1+t)p𝑑t01/(pC2)tn1(1+t)p𝑑tC31(p)n\int^{M\pi/(2m)}_{0}\frac{t^{n-1}}{(1+t)^{-p}}dt\geq\int^{1/(-pC_{2})}_{0}\frac{t^{n-1}}{(1+t)^{-p}}dt\geq C_{3}^{-1}(-p)^{-n}

holds for another positive constant C3C_{3} depending only on nn, inequality (3.5) then follows from Lemma 3.2 when p<np<-n. ∎

As a corollary of Lemma 3.3, we have the following result.

Corollary 3.1.

Let Ω(p)\Omega^{(p)} be the solution obtained in Lemma 3.1 with the support function h(p)h^{(p)}, we have

(3.6) limpmin𝕊nh(p)=1.\lim_{p\to-\infty}\min_{{\mathbb{S}}^{n}}h^{(p)}=1.
Proof.

By Lemma 3.3, sending pp to negative infinity in (3.5), we conclude that

limpminx𝕊nh(p)(x)1.\lim_{p\to-\infty}\min_{x\in{\mathbb{S}}^{n}}{{h^{(p)}}}(x)\geq 1.

On the other hand, evaluating the solution h(p){{h^{(p)}}} at its minimal point on the sphere 𝕊n{\mathbb{S}}^{n}, it follows from (1.2) by the maximum principle that

minx𝕊nh(p)(x)maxx𝕊nf1/(np)(x).\min_{x\in{\mathbb{S}}^{n}}{{h^{(p)}}}(x)\leq\max_{x\in{\mathbb{S}}^{n}}f^{1/(n-p)}(x).

Sending pp to negative infinity, the desired identity (3.6) follows. ∎

When a-priori lower-upper bounds of functions h𝒦p(𝒮)h\in{\mathcal{K}}_{p}({\mathcal{S}}) have been derived in Lemmas 3.2 and 3.3, by Blaschke’s selection theorem, a maximizing sequence Ωj𝒦p(𝒮)\Omega_{j}\in{\mathcal{K}}_{p}({\mathcal{S}}) converges to a limiting convex body Ω(p)𝒦p(𝒮){{\Omega^{(p)}}}\in{\mathcal{K}}_{p}({\mathcal{S}}). Due to the uniform convergence of the support functions hjh_{j} to h(p){{h^{(p)}}}, we see that Ω(p){{\Omega^{(p)}}} is actually a maximizer of the variational problem (3.1). Consequently, we can prove Lemma 3.1.

Proof of Lemma 3.1.

To show Lemma 3.1, it remains to verify that h(p){{h^{(p)}}} satisfies the Euler-Lagrange equation (1.2). This can be achieved by adapting the argument in ([12], Section 3) (see also [13], Section 5) with slight modifications. First, by calculating the first variation along the Wulff shape, it follows that the Monge-Ampère measure associated to the solution is bounded from above. By the duality of the polar bodies, we have also that the Monge-Ampère measure is bounded from below. Then, using the work of Caffarelli in [9], we can demonstrate the strict convexity and C1,αC^{1,\alpha} regularity of the solution. Consequently, we can show that the maximizer satisfies that

(3.7) 𝕊n[det(2h+hI)fhp1]g=0\int_{{\mathbb{S}}^{n}}\big{[}\det(\nabla^{2}h+hI)-fh^{p-1}\big{]}g=0

for all group invariant function gg, we thus reach the conclusion that h=h(p)h=h^{(p)} satisfies the Euler-Lagrange equation (1.2). Actually, one needs only to consider group invariant Hilbert space

(𝒮){gL2(𝕊n)|g is group invairant}{\mathbb{H}}({\mathcal{S}})\equiv\{g\in L^{2}({\mathbb{S}}^{n})|\ g\mbox{ is group invairant}\}

and use the fact that the orthonormal subspace of (𝒮){\mathbb{H}}({\mathcal{S}}) itself must be identical to {0}\{0\}. Then, (3.7) means that the group invariant function f~:=det(2h+hI)fhp1\widetilde{f}:=\det(\nabla^{2}h+hI)-fh^{p-1} belongs to the orthonormal subspace of (𝒮){\mathbb{H}}({\mathcal{S}}), and hence must be identical to zero.

If ff is Hölder continuous, by [10] we conclude that h(p)h^{(p)} is C2,αC^{2,\alpha} smooth. Higher regularity can be obtained by iteration using Schauder’s estimate. This completes the proof of Lemma 3.1. ∎

4. Limiting extremal bodies

By Theorem 3.1 and its proof, we have

Theorem 4.1.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property. For each p<n1p<-n-1, there holds

(4.1) p(Ω)C(n,p,f,𝒮),Ω𝒦0(𝒮).\mathcal{F}_{p}(\Omega)\leq C(n,p,f,{\mathcal{S}}),\ \ \forall\ \Omega\in{\mathcal{K}}_{0}({\mathcal{S}}).

with the best constant

C(n,p,f,𝒮):=supΩ𝒦p(𝒮)V(Ω).C(n,p,f,{\mathcal{S}}):=\sup_{\Omega\in{\mathcal{K}}_{p}({\mathcal{S}})}V(\Omega).

Moreover, the inequality (4.1) becomes equality at some smooth group invariant solution h(p){{h^{(p)}}} of (1.2), corresponding to a group invariant convex body Ω(p)𝒦0(𝒮){{\Omega^{(p)}}}\in{\mathcal{K}}_{0}({\mathcal{S}}).

Proof.

For any convex body Ω𝒦0(𝒮)\Omega\in{\mathcal{K}}_{0}({\mathcal{S}}) with support function hΩh_{\Omega}, we normalize it to Ω~\widetilde{\Omega} with support function

h~:=hΩ/λ𝒦0(𝒮),λ=(𝕊nfhΩp/𝕊nf)1/p.\widetilde{h}:=h_{\Omega}/\lambda\in{\mathcal{K}}_{0}({\mathcal{S}}),\ \ \lambda=\left(\int_{{\mathbb{S}}^{n}}fh_{\Omega}^{p}{\Big{/}}\int_{{\mathbb{S}}^{n}}f\right)^{1/p}.

Then we have

p(Ω)=V(Ω~)C(n,p,f,𝒮):=supΩ𝒦p(𝒮)V(Ω).\mathcal{F}_{p}(\Omega)=V(\widetilde{\Omega})\leq C(n,p,f,{\mathcal{S}}):=\sup_{\Omega\in{\mathcal{K}}_{p}({\mathcal{S}})}V(\Omega).

Moreover, the equality holds at multiples of h(p)h^{(p)}. This completes the proof. ∎

It is worth noting that when the Santaló center of Ω\Omega is the origin and when f=1f=1, our functional n1()\mathcal{F}_{-n-1}(\cdot) is exactly Mahler’s volume [33]. By Lemmas 3.2 and 3.3, the sequence of maximizers h(p){{h^{(p)}}} of the functional is a-priori bounded in Lip(𝕊n)Lip({\mathbb{S}}^{n}) by convexity. Hence, along a subsequence, the convex body Ω(p){{\Omega^{(p)}}} converges to a limiting convex body Ω()𝒦0(𝒮)\Omega^{(-\infty)}\in{\mathcal{K}}_{0}({\mathcal{S}}). We have the following characterization of Ω()\Omega^{(-\infty)}.

Lemma 4.1.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property. The best constant C(n,p,f,𝒮)C(n,p,f,{\mathcal{S}}) of the functional p()\mathcal{F}_{p}(\cdot) converges to the best constant C(n,,𝒮)C(n,-\infty,{\mathcal{S}}) of the limiting functional

(Ω):=V(Ω)/min𝕊nhΩn+1,Ω𝒦0(𝒮).\mathcal{F}_{-\infty}(\Omega):=V(\Omega)/\min_{{\mathbb{S}}^{n}}h_{\Omega}^{n+1},\ \ \Omega\in{\mathcal{K}}_{0}({\mathcal{S}}).

Moreover, Ω()\Omega^{(-\infty)} is exactly a maximizer of the functional ()\mathcal{F}_{-\infty}(\cdot).

Proof.

For each p<np<-n, we have

(4.2) p(Ω)(Ω)C(n,p,f,𝒮)C(n,,𝒮).\mathcal{F}_{p}(\Omega)\leq\mathcal{F}_{-\infty}(\Omega)\Rightarrow C(n,p,f,{\mathcal{S}})\leq C(n,-\infty,{\mathcal{S}}).

Conversely, for each Ω𝒦0(𝒮)\Omega\in{\mathcal{K}}_{0}({\mathcal{S}}), by the convergence property of LpL^{p}-norm to LL^{\infty}-norm, we have

(4.3) (Ω)=limpp(Ω)C(n,,𝒮)limpC(n,p,f,𝒮).\mathcal{F}_{-\infty}(\Omega)=\lim_{p\to-\infty}\mathcal{F}_{p}(\Omega)\Rightarrow C(n,-\infty,{\mathcal{S}})\leq\lim_{p\to-\infty}C(n,p,f,{\mathcal{S}}).

Combining (4.3) with (4.2) yields

(4.4) C(n,,𝒮)=limpC(n,p,f,𝒮).C(n,-\infty,{\mathcal{S}})=\lim_{p\to-\infty}C(n,p,f,{\mathcal{S}}).

On the other hand, by the uniform convergence of 1/h(p)1/{{h^{(p)}}} to 1/h()1/h^{(-\infty)}, and the condition (1.3), we have

[1(n+1)ωn+1𝕊nf|1/h(p)1/h()|p]1/p\displaystyle\left[\frac{1}{(n+1)\omega_{n+1}}\int_{{\mathbb{S}}^{n}}f|1/{{h^{(p)}}}-1/h^{(-\infty)}|^{-p}\right]^{-1/p}
max𝕊nf1/p1/h(p)1/h()L=op(1),\displaystyle\hskip 45.0pt\leq\max_{{\mathbb{S}}^{n}}f^{-1/p}||1/{{h^{(p)}}}-1/h^{(-\infty)}||_{L^{\infty}}=o_{p}(1),

where the infinitesimal op(1)o_{p}(1) is small as long as pp is negative large. As a result,

(Ω())\displaystyle\mathcal{F}_{-\infty}(\Omega^{(-\infty)}) =p(Ω())+op(1)=p(Ω(p))+op(1)\displaystyle=\mathcal{F}_{p}(\Omega^{(-\infty)})+o_{p}(1)=\mathcal{F}_{p}({{\Omega^{(p)}}})+o_{p}(1)
=C(n,p,f,𝒮)+op(1)=C(n,,𝒮)+op(1)\displaystyle=C(n,p,f,{\mathcal{S}})+o_{p}(1)=C(n,-\infty,{\mathcal{S}})+o_{p}(1)

by (4.4). This implies that Ω()\Omega^{(-\infty)} is exactly a maxmizer of the functional ()\mathcal{F}_{-\infty}(\cdot). ∎

The next lemma rules out the trivial possibility of the limiting sphere 𝕊n{\mathbb{S}}^{n}.

Lemma 4.2.

Let PB1P\subset{B_{1}} be any group 𝒮{\mathcal{S}} invariant polytope, then

(4.5) (P)>(B1).\mathcal{F}_{-\infty}(P)>\mathcal{F}_{-\infty}(B_{1}).

Consequently, the limiting shape Ω()\Omega^{(-\infty)} of Ω(p){{\Omega^{(p)}}}, when pp\to-\infty, cannot be a sphere.

Proof.

Let PP be a group 𝒮{\mathcal{S}} invariant polytope. Setting m=minhPm=\min h_{P} to be the inner-radius of PP, we have mB1m{B_{1}} as a smaller ball contained inside and tangential to PP. Hence, the desired inequality (4.5) follows from

(P)(B1)=(P)(mB1)=V(P)/mn+1V(mB1)/mn+1>1.\frac{\mathcal{F}_{-\infty}(P)}{\mathcal{F}_{-\infty}({B_{1}})}=\frac{\mathcal{F}_{-\infty}(P)}{\mathcal{F}_{-\infty}(m{B_{1}})}=\frac{V(P)/m^{n+1}}{V(m{B_{1}})/m^{n+1}}>1.

By Lemma 4.1, the limiting shape Ω()\Omega^{(-\infty)} is a maximizer of the functional ()\mathcal{F}_{-\infty}(\cdot). Thus, it can not be a ball by (4.5). ∎

The next lemma implies that the maximizer of the limiting functional must be a polytope.

Lemma 4.3.

For any non-polytope group invariant convex body Ω𝒦0(𝒮)\Omega\in{\mathcal{K}}_{0}({\mathcal{S}}), there exists a group 𝒮{\mathcal{S}} invariant polytope PP containing Ω\Omega such that

(4.6) (Ω)<(P).\mathcal{F}_{-\infty}(\Omega)<\mathcal{F}_{-\infty}(P).
Proof.

Let x0Ωx_{0}\in\partial\Omega and BrΩΩB_{r_{\Omega}}\subset\Omega for the inner-radius rΩr_{\Omega} of Ω\Omega, we denote H(x0)H(x_{0}) to be the tangent hyperplane of Ω\partial\Omega at x0x_{0}, and let H(x0)H^{-}(x_{0}) be the half space of H(x0)H(x_{0}) containing Ω\Omega. Define

Px0=ϕ𝒮ϕ(H(x0)).P_{x_{0}}=\cap_{\phi\in{\mathcal{S}}}\phi(H^{-}(x_{0})).

Obviously, Px0P_{x_{0}} is a group 𝒮{\mathcal{S}} invariant polytope containing Ω\Omega, then

(Px0)(Ω)=V(Px0)V(Ω)>1,\displaystyle\frac{\mathcal{F}_{-\infty}(P_{x_{0}})}{\mathcal{F}_{-\infty}(\Omega)}=\frac{V(P_{x_{0}})}{V(\Omega)}>1,

as long as Ω\Omega is not a polytope. ∎

Combining with Theorems 3.1 and 4.1, along with Lemma 4.3, we obtain the following theorem, which characterizes the limiting extremal body as a maximal group-invariant polytope.

Theorem 4.2.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property. The maximizer Ω()\Omega^{(-\infty)} of the functional ()\mathcal{F}_{-\infty}(\cdot) must be a group invariant polytope, after scaling such that the volume |Ω()|>0|\Omega^{(-\infty)}|>0.

Proof of Theorem 1.1.

Theorem 1.1 follows from Theorem 4.2 and Corollary 3.1. ∎

5. Local maximizer of the functional and proof of Theorem 1.2

In this section, we use Tn+1T\subset\mathbb{R}^{n+1} to denote a regular polytope that is tangential to the unit sphere 𝕊n{\mathbb{S}}^{n}, and 𝒮TO(n+1)\mathcal{S}_{T}\subset O(n+1) to represent the symmetry group of TT. Denote

𝒩δ(T)={Ω𝒦0(𝒮T)|dist(Ω,T)<δ}{\mathcal{N}}_{\delta}(T)=\big{\{}\Omega\in{\mathcal{K}}_{0}({\mathcal{S}}_{T})\ \big{|}\ dist(\Omega,T)<\delta\big{\}}

the δ\delta-neighborhood of TT for a positive constant δ\delta and the Hausdorff distance dist(,)dist(\cdot,\cdot), and denote 𝒩δ(T)\partial{\mathcal{N}}_{\delta}(T) the boundary of 𝒩δ(T){\mathcal{N}}_{\delta}(T). Denote also the normalized group invariant family

𝒦min(𝒮T)={Ω𝒦0(𝒮T)|inf𝕊nhΩ=1},{\mathcal{K}}_{min}({\mathcal{S}}_{T})=\big{\{}\Omega\in{\mathcal{K}}_{0}({\mathcal{S}}_{T})\ \big{|}\ \inf_{{\mathbb{S}}^{n}}h_{\Omega}=1\big{\}},

and the normalized δ\delta-neighborhood

𝒦min(δ,𝒮T)=𝒦min(𝒮T)𝒩δ(T){\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})={\mathcal{K}}_{min}({\mathcal{S}}_{T})\cap{\mathcal{N}}_{\delta}(T)

of TT. Then, we can prove that TT has strictly maximal volume in the normalized family 𝒦min(δ,𝒮T){\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}) for some δ=δT>0\delta=\delta_{T}>0 as follows.

Lemma 5.1.

For any regular polytope Tn+1T\subset{\mathbb{R}}^{n+1} tangential to the unit sphere 𝕊n{\mathbb{S}}^{n}, there exists a small constant δ=δT>0\delta=\delta_{T}>0 such that

(5.1) V(T)>V(Ω),Ω𝒦min(δ,𝒮T){T}.V(T)>V(\Omega),\ \ \forall\ \Omega\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})\setminus\{T\}.
Proof.

Step 1. Subdivision of a regular polytope.

Let Tn+1T\subset{\mathbb{R}}^{n+1} be a regular polytope that is tangential to the unit sphere 𝕊n{\mathbb{S}}^{n}. By the high symmetry of TT, each κ\kappa-dimensional face is also a regular polytope, where κ[2,n]\kappa\in[2,n] and κ\kappa\in{\mathbb{N}}. We will use this high symmetry to decompose TT into as many congruent pieces as possible. Let On+1O_{n+1}, the origin of n+1{\mathbb{R}}^{n+1}, be the centroid of TT. For an nn-dimensional face FnF_{n} of TT, we denote by OnO_{n} the centroid of FnF_{n}. Let Fn1F_{n-1} be one of the (n1)(n-1)-dimensional face of FnF_{n}, we denote its centroid by On1O_{n-1}. The bootstrap procedure will produce a sequence of faces

T=Fn+1FnFn1F1F0T=F_{n+1}\rightarrow F_{n}\rightarrow F_{n-1}\rightarrow\cdots\rightarrow F_{1}\rightarrow F_{0}

with lower dimensions, and yield a sequence of corresponding centroids

On+1OnOn1O1O0.O_{n+1}\rightarrow O_{n}\rightarrow O_{n-1}\rightarrow\cdots\rightarrow O_{1}\rightarrow O_{0}.

Note that

(5.2) OjOj1span{Oi,i=0,1,,j1},j=n+1,n,,1.O_{j}O_{j-1}\perp\text{span}\{O_{i},i=0,1,\cdots,j-1\},\ \ \forall j=n+1,n,\cdots,1.

Since OjOj1Fj1,j=n+1,n,,1O_{j}O_{j-1}\perp F_{j-1},\forall j=n+1,n,\cdots,1, one may assume that OjOj1O_{j}O_{j-1} is parallell to the xjx_{j}-axis by rotation and set Rn+1=1,Rj=|OjOj1|,j=n,n1,,2,1R_{n+1}=1,R_{j}=|O_{j}O_{j-1}|,\forall j=n,n-1,\cdots,2,1. Define Ωj:=conv(i=0jOi),j=0,1,,n\Omega_{j}:=\text{conv}(\cup_{i=0}^{j}O_{i}),j=0,1,\cdots,n, we have the recursive formula

(5.3) Ω0=O0,Ωj=conv{Oj,Ωj1},j=1,2,,n.\Omega_{0}=O_{0},\ \ \Omega_{j}=\text{conv}\{O_{j},\Omega_{j-1}\},\ \ j=1,2,\cdots,n.

Then, FnF_{n} can be decomposed into small pieces that are congruent to Ωn\Omega_{n}111Sometimes, we may regard Ωl{zn+1|zn+1=Rn+1,zn=Rn,,zl+1=Rl+1,zj[0,Rj],j[1,l]}\Omega_{l}\subset\{z\in{\mathbb{R}}^{n+1}|z_{n+1}=R_{n+1},z_{n}=R_{n},\cdots,z_{l+1}=R_{l+1},z_{j}\in[0,R_{j}],\forall j\in[1,l]\} as a subdomain of l{\mathbb{R}}^{l} for l=1,2,,nl=1,2,\cdots,n by simply omitting the coordinates zk=Rkz_{k}=R_{k} for k=n+1,n,,l+1k=n+1,n,\cdots,l+1. For example, Ωn{z¯n}\Omega_{n}\subset\{\bar{z}\in\mathbb{R}^{n}\} for each z=(z¯,zn+1)n+1z=(\bar{z},z_{n+1})\in\mathbb{R}^{n+1}.. Let w0=(w¯0,1)Ωnn+1w_{0}=(\overline{w}_{0},1)\in\Omega_{n}\subset{\mathbb{R}}^{n+1} be the point of the piece Ωn\Omega_{n} defined above with projection

y0=w0/|w0|Dn𝕊n,y_{0}=w_{0}/|w_{0}|\in D_{n}\subset{\mathbb{S}}^{n},

satisfying h(y0)=1h^{\prime}(y_{0})=1 for support function hh^{\prime} of another normalized group invariant convex body Ω𝒦min(𝒮T)\Omega^{\prime}\in\mathcal{K}_{min}({\mathcal{S}}_{T}).

[Uncaptioned image]

[Uncaptioned image]

Figure: subdivision of a polyhedron tangential to the unit sphere 𝕊2{\mathbb{S}}^{2} in 3\mathbb{R}^{3}.

Step 2. Claim 1\sharp 1: The regular polytope TT has local strictly maximal volume on the normalized group invariant family 𝒦min(δ,𝒮T){\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}) if the following integral function

(5.4) V¯(w¯0)=Ωn(1|1+|w¯0|21+z¯,w¯0|n+1)𝑑z¯\overline{V}(\overline{w}_{0})=\int_{\Omega_{n}}\left(1-\bigg{|}\frac{\sqrt{1+|\overline{w}_{0}|^{2}}}{1+\langle\overline{z},\overline{w}_{0}\rangle}\bigg{|}^{n+1}\right)d\overline{z}

is positive for all w¯0Ωn{0}\overline{w}_{0}\in\Omega_{n}\setminus\{0\} closing to 0. (See footnote 1 for Ωn\Omega_{n}.)

Proof of claim 1\sharp 1.

At first, the hyperplane H1H_{1} tangential to the unit sphere 𝕊n{\mathbb{S}}^{n} at On=en+1O_{n}=e_{n+1} is given by the radial function ρ1(y)=1/yn+1\rho_{1}(y)=1/y_{n+1} for y=(y1,y2,,yn+1)𝕊ny=(y_{1},y_{2},\ldots,y_{n+1})\in{\mathbb{S}}^{n}. While, the hyperplane H2H_{2} tangential to the 𝕊n\mathbb{S}^{n} at y0y_{0} is given by the radial function ρ2(y)=1/y,y0\rho_{2}(y)=1/\langle y,y_{0}\rangle. Therefore, we obtain that

(5.5) VDn(H1)VDn(H2)=1n+1Dn(1yn+1n+11|y,y0|n+1)𝑑σ,V_{D_{n}}(H^{-}_{1})-V_{D_{n}}(H^{-}_{2})=\frac{1}{n+1}\int_{D_{n}}\left(\frac{1}{y_{n+1}^{n+1}}-\frac{1}{|\langle y,y_{0}\rangle|^{n+1}}\right)d\sigma,

where VDn(H)V_{D_{n}}(H^{-}) represents the volume of the solid enclosed by the hyperplane HH and the infinite cone formed by connecting the origin to DnD_{n}. Note that for each normalized group invariant convex body Ω𝒦min(δ,𝒮T)\Omega^{\prime}\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}), we also have

(5.6) VDn(T)=VDn(H1),VDn(Ω)VDn(H2).V_{D_{n}}(T)=V_{D_{n}}(H^{-}_{1}),\ \ V_{D_{n}}(\Omega^{\prime})\leq V_{D_{n}}(H^{-}_{2}).

Regarding the coordinates z¯\overline{z} in the hyperplane H1={z=(z¯,zn+1)n+1|zn+1=1}H_{1}=\{z=(\overline{z},z_{n+1})\in{\mathbb{R}}^{n+1}|\ z_{n+1}=1\} as the local coordinates of 𝕊n{\mathbb{S}}^{n} by projection y=z/|z|y=z/|z|, we have the metric

gij=yziyzj=δij|z|2zizj|z¯|2|z|6g_{ij}=\frac{\partial y}{\partial z_{i}}\cdot\frac{\partial y}{\partial z_{j}}=\frac{\delta_{ij}}{|z|^{2}}-\frac{z_{i}z_{j}|\overline{z}|^{2}}{|z|^{6}}

since yzi=(ei|z|z¯zi|z|3,zi|z|3)\frac{\partial y}{\partial z_{i}}=\left(\frac{e_{i}}{|z|}-\frac{\bar{z}z_{i}}{|z|^{3}},-\frac{z_{i}}{|z|^{3}}\right), and thus the volume element is

dσ=det(g)dz¯=(1+|z¯|2)n+12dz¯.d\sigma=\sqrt{\det(g)}d\overline{z}=(1+|\overline{z}|^{2})^{-\frac{n+1}{2}}d\overline{z}.

Therefore, the volume difference function

V~(y0):=(n+1)(VDn(H1)VDn(H2))=Ωn(11|z,y0|n+1)𝑑z¯{\begin{split}\widetilde{V}(y_{0}):&=(n+1)(V_{D_{n}}(H^{-}_{1})-V_{D_{n}}(H^{-}_{2}))\\ &=\int_{\Omega_{n}}\left(1-\frac{1}{|\langle z,y_{0}\rangle|^{n+1}}\right)d\overline{z}\end{split}}

holds for y0Dny_{0}\in D_{n}. (See footnote 1 in page 14 for Ωn\Omega_{n}.) Letting y0,n+1y_{0,n+1} be the (n+1)(n+1)-th coordinates of y0y_{0} and setting w0=(w¯0,1)=y0y0,n+1Ωnn+1w_{0}=(\overline{w}_{0},1)=\frac{y_{0}}{y_{0,n+1}}\in\Omega_{n}\subset\mathbb{R}^{n+1}, there holds

(5.7) V~(y0)=V¯(w¯0):=Ωn(1|1+|w¯0|21+z¯,w¯0|n+1)𝑑z¯,w¯0Ωn,\widetilde{V}(y_{0})=\overline{V}(\overline{w}_{0}):=\int_{\Omega_{n}}\left(1-\bigg{|}\frac{\sqrt{1+|\overline{w}_{0}|^{2}}}{1+\langle\overline{z},\overline{w}_{0}\rangle}\bigg{|}^{n+1}\right)d\overline{z},\ \ \overline{w}_{0}\in\Omega_{n},

due to y0=(w¯0,1)/|(w¯0,1)|y_{0}=(\overline{w}_{0},1)/|(\overline{w}_{0},1)|. This, together with (5.6), completes the proof of claim 1\sharp 1. ∎

Step 3. Claim 2\sharp 2: The regular polytope TT has locally strictly maximal volume on the normalized group invariant family 𝒦min(δ,𝒮T){\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}) for some δ>0\delta>0. More precisely, there exists a positive constant δ=δT\delta=\delta_{T} such that for any Ω𝒦min(δ,𝒮T){T}\Omega^{\prime}\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})\setminus\{T\}, there holds

(5.8) V(Ω)<V(T).V(\Omega^{\prime})<V(T).
Proof of claim 2\sharp 2.

The proof of the claim is equivalent to verifying that: If w0=(w¯0,1)Ωnw_{0}=(\overline{w}_{0},1)\in\Omega_{n} approaches On=en+1O_{n}=e_{n+1} without equaling OnO_{n}, then the volume difference function V¯(w¯0)\overline{V}(\overline{w}_{0}) in (5.4) is positive. Along the direction segment I={On+raΩn|r0,a+n}I=\{O_{n}+ra\in\Omega_{n}|r\geq 0,a\in{\mathbb{R}}^{n}_{+}\} 222+j:={zj|Σi=1jzi>0andzi0for eachi=1,2,,j},j=1,2,,n+1.{\mathbb{R}}^{j}_{+}:=\{z\in{\mathbb{R}}^{j}|\Sigma_{i=1}^{j}z_{i}>0\ \text{and}\ z_{i}\geq 0\ \text{for each}\ i=1,2,\cdots,j\},\ \ j=1,2,\cdots,n+1., the volume difference function can be simplified by for w¯0=ra\overline{w}_{0}=ra,

V(r,a):=V¯(w¯0)=Ωn(1|1+r2|a|21+rz¯,a|n+1)𝑑z¯.V(r,a):=\overline{V}(\overline{w}_{0})=\int_{\Omega_{n}}\bigg{(}1-\bigg{|}\frac{\sqrt{1+r^{2}|a|^{2}}}{1+r\langle\overline{z},a\rangle}\bigg{|}^{n+1}\bigg{)}d\overline{z}.

Taking the first differentiation in rr and then evaluating at r=0r=0, we get

(5.9) ddr|r=0V(r,a)=(n+1)Ωnz¯,a𝑑z¯>0,a+n\frac{d}{dr}\bigg{|}_{r=0}V(r,a)=(n+1)\int_{\Omega_{n}}\langle\overline{z},a\rangle d\overline{z}>0,\ \ \forall a\in{\mathbb{R}}^{n}_{+}

using the fact that Ωn+n\Omega_{n}\subset{\mathbb{R}}^{n}_{+} (See footnote 1 in page 14 for Ωn\Omega_{n}). This, together with Claim 1\sharp 1, yields V(r,a)>V(0,a)=0V(r,a)>V(0,a)=0, and hence the desired inequality (5.8) holds. ∎

Then Lemma 5.1 follows directly from claim 2\sharp 2. ∎

For the given regular polytope Tn+1T\subset{\mathbb{R}}^{n+1} tangential to 𝕊n{\mathbb{S}}^{n}, define the group 𝒮T{\mathcal{S}}_{T} invariant family

𝒦p,f(𝒮T){Ω𝒦0(𝒮T)|𝕊nfhΩp=𝕊nfhTp},p<0.{\mathcal{K}}_{p,f}({\mathcal{S}}_{T})\equiv\Big{\{}\Omega\in{\mathcal{K}}_{0}({\mathcal{S}}_{T})\ \big{|}\ \int_{{\mathbb{S}}^{n}}fh_{\Omega}^{p}=\int_{{\mathbb{S}}^{n}}fh_{T}^{p}\Big{\}},\ \ p<0.

As shown in Lemma 5.1, TT is a local strict maximizer of the functional (){\mathcal{F}}_{-\infty}(\cdot) on the normalized family 𝒦min(δ,𝒮T){\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}) for some δ>0\delta>0. Hence, there exists a positive constant δ=δT\delta=\delta_{T} such that

(5.10) (Ω)<(T),Ω𝒦min(δ,𝒮T)T.{\mathcal{F}}_{-\infty}(\Omega)<{\mathcal{F}}_{-\infty}(T),\ \ \forall\ \Omega\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})\setminus T.

We now consider the local variational scheme

(5.11) supΩ𝒦p,f(δ,𝒮T)V(Ω),\sup_{\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V(\Omega),

where 𝒦p,f(δ,𝒮T):=𝒦p,f(𝒮T)𝒩δ(T)\mathcal{K}_{p,f}(\delta,{\mathcal{S}}_{T}):={\mathcal{K}}_{p,f}({\mathcal{S}}_{T})\cap{\mathcal{N}}_{\delta}(T) and 𝒦p,f(δ,𝒮T):=𝒦p,f(𝒮T)𝒩δ(T)\partial\mathcal{K}_{p,f}(\delta,{\mathcal{S}}_{T}):={\mathcal{K}}_{p,f}({\mathcal{S}}_{T})\cap\partial{\mathcal{N}}_{\delta}(T). Note that 𝒦min(δ,𝒮T)=p<n1p(,p)𝒦p,f(δ,𝒮T)\mathcal{K}_{min}(\delta,{\mathcal{S}}_{T})=\cap_{p^{\prime}<-n-1}\cup_{p\in(-\infty,p^{\prime})}\mathcal{K}_{p,f}(\delta,{\mathcal{S}}_{T}), i.e., 𝒦p,f(δ,𝒮T)𝒦min(δ,𝒮T)\mathcal{K}_{p,f}(\delta,{\mathcal{S}}_{T})\rightarrow\mathcal{K}_{min}(\delta,{\mathcal{S}}_{T}) as pp\rightarrow-\infty. In fact, the convergence of 𝒦p,f(δ,𝒮T){\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T}) to the limiting family

(5.12) 𝒦,f(δ,𝒮T)={Ω𝒦0(𝒮T)|inf𝕊nhΩ=inf𝕊nhT=1}𝒩δ(T)=𝒦min(δ,𝒮T){\mathcal{K}}_{-\infty,f}(\delta,{\mathcal{S}}_{T})=\Big{\{}\Omega\in{\mathcal{K}}_{0}({\mathcal{S}}_{T})\ \big{|}\ \inf_{{\mathbb{S}}^{n}}h_{\Omega}=\inf_{{\mathbb{S}}^{n}}h_{T}=1\Big{\}}\cap{\mathcal{N}}_{\delta}(T)={\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})

under the Hausdorff distance. To clarify this convergence, we mean that for each Ω()𝒦,f(δ,𝒮T)\Omega^{(-\infty)}\in{\mathcal{K}}_{-\infty,f}(\delta,{\mathcal{S}}_{T}), there exists a sequence of Ω(p)𝒦p,f(δ,𝒮T)\Omega^{(p)}\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T}) converges to Ω()\Omega^{(-\infty)} under Hausdorff distance. Vice versa, for any sequence of Ω(p)𝒦p,f(δ,𝒮T)\Omega^{(p)}\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T}), we have Ω(p)\Omega^{(p)} sub-converges to some Ω()𝒦,f(δ,𝒮T)\Omega^{(-\infty)}\in{\mathcal{K}}_{-\infty,f}(\delta,{\mathcal{S}}_{T}).

We now show the existence of maximizer Ω(p)𝒦p,f(δ,𝒮T)\Omega^{(p)}\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T}) of (5.11) for negative large pp, whose support function h(p)h^{(p)} satisfies the Euler-Lagrange equation up to a constant multiplication. The main ingredient of the proof is to show that the maximizer Ω(p)\Omega^{(p)} of (5.11) stays away from the boundary of 𝒩δ(T){\mathcal{N}}_{\delta}(T), that is, supΩ𝒦p,f(δ,𝒮T)V(Ω)<supΩ𝒦p,f(δ,𝒮T)V(Ω)=V(Ω(p))\sup_{\Omega\in\partial{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V(\Omega)<\sup_{\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V(\Omega)=V(\Omega^{(p)}) and Ω(p)𝒩δ(T)\Omega^{(p)}\not\in\partial{\mathcal{N}}_{\delta}(T).

Lemma 5.2.

Let Tn+1T\subset{\mathbb{R}}^{n+1} be a regular polytope tangential to the unit sphere 𝕊n{\mathbb{S}}^{n}. For each p<n1p<-n-1, the variational problem (5.11) admits a maximizer Ω(p)\Omega^{(p)} staying away from the boundary of 𝒩δ(T){\mathcal{N}}_{\delta}(T) for a small δ=δT>0\delta=\delta_{T}>0.

Proof.

For simplicity, let 𝒦(δ,𝒮T)\mathcal{K}(\delta,{\mathcal{S}}_{T}) denote the union of 𝒦p,f(δ,𝒮T){\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T}) for p<n1p<-n-1, that is, 𝒦(δ,𝒮T)=p<n1𝒦p,f(δ,𝒮T)\mathcal{K}(\delta,{\mathcal{S}}_{T})=\cup_{p<-n-1}{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T}). By a-priori Lipschitz bound of convex bodies Ω𝒦(δ,𝒮T)\Omega\in{\mathcal{K}}(\delta,{\mathcal{S}}_{T}), the boundary 𝒦(δ,𝒮T)\partial{\mathcal{K}}(\delta,{\mathcal{S}}_{T}) is a sequential compact set under the Hausdorff distance. Noting that the limiting functional (){\mathcal{F}}_{-\infty}(\cdot) is continuous under the Hausdorff distance, it follows from (5.10) that for some σ>0\sigma>0,

(5.13) supΩ𝒦min(δ,𝒮T)(Ω)(T)σ.\sup_{\Omega\in\partial{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})}{\mathcal{F}}_{-\infty}(\Omega)\leq{\mathcal{F}}_{-\infty}(T)-\sigma.

Noting that for fixed Ω\Omega, we have the convergence

(5.14) limpp(Ω)=(Ω).\lim_{p\to-\infty}{\mathcal{F}}_{p}(\Omega)={\mathcal{F}}_{-\infty}(\Omega).

We claim that the functionals p(){\mathcal{F}}_{p}(\cdot) are equi-continuous for negative large pp. Actually, by Minkowski’s inequality for the ff-weighted Lfq(𝕊n)L^{q}_{f}({\mathbb{S}}^{n})-space defined by the norm

hLfq(𝕊n)=(𝕊nfhq)1/q,hLfq(𝕊n)\|h\|_{L^{q}_{f}({\mathbb{S}}^{n})}=\left(\int_{{\mathbb{S}}^{n}}fh^{q}\right)^{1/q},\ \ \forall h\in L^{q}_{f}({\mathbb{S}}^{n})

and a positive constant q>1q>1, we have for some constant C>0C>0,

|(𝕊nfhΩp)1/p𝕊nfhΩp)1/p|\displaystyle\left|\left(\int_{{\mathbb{S}}^{n}}fh_{\Omega}^{p}\right)^{-1/p}-\int_{{\mathbb{S}}^{n}}fh_{\Omega^{\prime}}^{p}\bigg{)}^{-1/p}\right| =\displaystyle= |hΩ1Lf|p|(𝕊n)hΩ1Lf|p|(𝕊n)|\displaystyle\left|\|h^{-1}_{\Omega}\|_{L^{|p|}_{f}({\mathbb{S}}^{n})}-\|h^{-1}_{\Omega^{\prime}}\|_{L^{|p|}_{f}({\mathbb{S}}^{n})}\right|
\displaystyle\leq hΩ1hΩ1Lf|p|(𝕊n)\displaystyle\|h^{-1}_{\Omega}-h^{-1}_{\Omega^{\prime}}\|_{L^{|p|}_{f}({\mathbb{S}}^{n})}
\displaystyle\leq ChΩhΩLf|p|(𝕊n).\displaystyle C\|h_{\Omega}-h_{\Omega^{\prime}}\|_{L^{|p|}_{f}({\mathbb{S}}^{n})}.

Thus, the equi-continuity of the functional p(){\mathcal{F}}_{p}(\cdot) follows from the definition (1.4).

Combining with the sequential compactness of 𝒦(δ,𝒮T)\partial{\mathcal{K}}(\delta,{\mathcal{S}}_{T}) and the equi-continuity of the functionals p(){\mathcal{F}}_{p}(\cdot) for negative large pp, we have

(5.15) limpsupΩ𝒦(δ,𝒮T)|p(Ω)(Ω)|=0,\lim_{p\to-\infty}\sup_{\Omega\in\partial{\mathcal{K}}(\delta,{\mathcal{S}}_{T})}|{\mathcal{F}}_{p}(\Omega)-{\mathcal{F}}_{-\infty}(\Omega)|=0,

in particular,

limp|p(T)(T)|=0.\lim_{p\to-\infty}|{\mathcal{F}}_{p}(T)-{\mathcal{F}}_{-\infty}(T)|=0.

For short, we write

C(n,p,f,𝒮T)=supΩ𝒦p,f(δ,𝒮T)V(Ω).C(n,p,f,{\mathcal{S}_{T}})=\sup_{\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V(\Omega).

By the compactness of 𝒦(δ,𝒮T){\mathcal{K}}(\delta,{\mathcal{S}}_{T}) and formula (5.12), we have the convergence

(5.16) limpC(n,p,f,𝒮T)=C(n,,f,𝒮T)=supΩ𝒦min(δ,𝒮T)V(Ω).\lim_{p\to-\infty}C(n,p,f,{\mathcal{S}_{T}})=C(n,-\infty,f,{\mathcal{S}_{T}})=\sup_{\Omega\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})}V(\Omega).

Noting that C(n,,f,𝒮T)=supΩ𝒦min(δ,𝒮T)V(Ω)=V(T)=(T)C(n,-\infty,f,{\mathcal{S}_{T}})=\sup_{\Omega\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})}V(\Omega)=V(T)=\mathcal{F}_{-\infty}(T) by Lemma 5.1, it follows from (5.13)-(5.16) that

supΩ𝒦p,f(δ,𝒮T)p(Ω)\displaystyle\sup_{\Omega\in\partial{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}{\mathcal{F}}_{p}(\Omega) \displaystyle\leq supΩ𝒦(δ,𝒮T)p(Ω)\displaystyle\sup_{\Omega\in\partial{\mathcal{K}}(\delta,{\mathcal{S}}_{T})}{\mathcal{F}}_{p}(\Omega)
=\displaystyle= supΩ𝒦(δ,𝒮T)(Ω)+ε(p)\displaystyle\sup_{\Omega\in\partial{\mathcal{K}}(\delta,{\mathcal{S}}_{T})}{\mathcal{F}}_{-\infty}(\Omega)+\varepsilon(p)
\displaystyle\leq (T)σ+ε(p)\displaystyle{\mathcal{F}}_{-\infty}(T)-\sigma+\varepsilon(p)
=\displaystyle= supΩ𝒦p,f(δ,𝒮T)V(Ω)σ+ε(p)\displaystyle\sup_{\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V(\Omega)-\sigma+\varepsilon(p)

hold for negative large pp and some infinitesimal ε(p)\varepsilon(p), where in the second inequality we have also used (5.13) and the convergence of 𝒦p,f(δ,𝒮T)\partial{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T}) to 𝒦min(δ,𝒮T)\partial{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}) under Hausdorff distance. Using again the relation

(5.17) p(Ω)=(Ω)+ε(p)=V(Ω)+ε(p),Ω𝒦p,f(δ,𝒮T){\mathcal{F}}_{p}(\Omega)={\mathcal{F}}_{-\infty}(\Omega)+\varepsilon(p)=V(\Omega)+\varepsilon(p),\ \ \forall\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})

for some infinitesimal ε(p)\varepsilon(p) and negative large pp, we conclude from (5) that

(5.18) supΩ𝒦p,f(δ,𝒮T)V(Ω)supΩ𝒦p,f(δ,𝒮T)V(Ω)σ+ε(p)\sup_{\Omega\in\partial{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V(\Omega)\leq\sup_{\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V(\Omega)-\sigma+\varepsilon(p)

holds for small δ>0\delta>0 and negative large pp. We thus arrive at the existence result of local maximizer of the functional p(){\mathcal{F}}_{p}(\cdot) for the negative large pp. ∎

Lemma 5.3.

Let Tn+1T\subset{\mathbb{R}}^{n+1} be a regular polytope tangential to the unit sphere 𝕊n{\mathbb{S}}^{n}. If δ=δT>0\delta=\delta_{T}>0 is small and pp is negative large, the variational problem (5.11) admits a maximizer Ω(p)𝒩δ(T)\Omega^{(p)}\not\in\partial{\mathcal{N}}_{\delta}(T). Moreover, the maximizer Ω(p)\Omega^{(p)} satisfies the equation (1.2), and converges to the given regular polytope TT as pp\rightarrow-\infty.

Proof.

The existence of maximizer Ω(p)\Omega^{(p)} follows from Lemma 5.2. Owing to the property that Ω(p)\Omega^{(p)} stays away from the boundary of 𝒩δ(T){\mathcal{N}}_{\delta}(T), we can prove that Ω(p)\Omega^{(p)} satisfies the Euler-Lagrange equation (1.2) as in the proof of global variational problem. Noting that C(n,p,f,𝒮T)C(n,p,f,{\mathcal{S}_{T}}) converges to C(n,,f,𝒮T)C(n,-\infty,f,{\mathcal{S}_{T}}) as pp\to-\infty, the maximizer Ω(p)\Omega^{(p)} converges to a maximizer Ω()\Omega^{(-\infty)} of the limiting functional. However, by Lemma 5.1, TT is the unique maximizer of the limiting functional on the family 𝒦min(δ,𝒮T){\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}). This, together with the observation that (Ω)=V(Ω)\mathcal{F}_{-\infty}(\Omega)=V(\Omega) for all Ω𝒦min(δ,𝒮T)\Omega\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}), implies that Ω()=T\Omega^{(-\infty)}=T. This completes the proof of the lemma. ∎

Proof of Theorem 1.2..

Theorem 1.2 is a direct consequence of Lemma 5.3. ∎

From our proof, a regular polytope is a local maximizer of the functional p\mathcal{F}_{p}. We would like to point out that it may not be a global maximizer of p\mathcal{F}_{p}.

6. The LpL_{p} dual Minkowski problem

In this section, we turn to study the LpL_{p} dual Minkowski problem (1.1) and prove Theorem 1.3. Parallel to Section 3, we consider a variational problem

(6.1) supΩ𝒦p(𝒮)Vq(Ω)withVq(Ω)𝕊nrq(y)𝑑y,q0\sup_{\Omega\in{\mathcal{K}}_{p}({\mathcal{S}})}V_{q}(\Omega)\ \ \text{with}\ \ V_{q}(\Omega)\equiv\int_{{\mathbb{S}}^{n}}r^{q}(y)dy,\ \ q\not=0

on the group 𝒮{\mathcal{S}} invariant family

𝒦p(𝒮):={Ω𝒦0(𝒮)|𝕊nfhΩp=𝕊nf},p0.{\mathcal{K}}_{p}({\mathcal{S}}):=\Bigg{\{}\Omega\in{\mathcal{K}}_{0}({\mathcal{S}})\Big{|}\ \int_{{\mathbb{S}}^{n}}fh_{\Omega}^{p}=\int_{{\mathbb{S}}^{n}}f\Bigg{\}},\ \ p\not=0.

A similar argument of Lemma 3.1 gives the following solvability result.

Proposition 6.1.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property. Then for q0q\not=0 and p<np<-n, the variational problem (6.1) admits a smooth positive maximizer h(p,q)h^{(p,q)} satisfying the Euler-Lagrange equation (1.1) up to a constant.

Recall the functional

p,q(Ω)=Vq(Ω)(𝕊nfhΩp/𝕊nf)q/p\mathcal{F}_{p,q}(\Omega)=V_{q}(\Omega)\left(\int_{{\mathbb{S}}^{n}}fh_{\Omega}^{p}{\Big{/}}\int_{{\mathbb{S}}^{n}}f\right)^{-q/p}

defined on 𝒮\mathcal{S}-invariant convex bodies Ω𝒦0(𝒮)\Omega\subset\mathcal{K}_{0}(\mathcal{S}). Similarly to Section 4, the maximizer Ωh(p,q)\Omega_{h}^{(p,q)} in Proposition 6.1 also serves as the extremal body of the following inequality.

Proposition 6.2.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property. Then for q0q\not=0 and p<np<-n, we have

(6.2) p,q(Ω)C(n,p,q,f,𝒮),Ω𝒦0(𝒮)\mathcal{F}_{p,q}(\Omega)\leq C(n,p,q,f,{\mathcal{S}}),\ \ \forall\ \Omega\in{\mathcal{K}}_{0}({\mathcal{S}})

with

C(n,p,q,f,𝒮):=supΩ𝒦p(𝒮)q(Ω).C(n,p,q,f,{\mathcal{S}}):=\sup_{\Omega\in{\mathcal{K}}_{p}({\mathcal{S}})}\mathcal{F}_{q}(\Omega).

Moreover, the inequality (6.2) becomes equality at some smooth group invariant solution h(p,q)h^{(p,q)} of (1.1), corresponding to the group invariant convex body Ω(p,q)𝒦0(𝒮)\Omega^{(p,q)}\in{\mathcal{K}}_{0}({\mathcal{S}}).

Remark 6.1.

Let Ω={xn+1:xy1foryΩ}\Omega^{*}=\{x\in\mathbb{R}^{n+1}:x\cdot y\leq 1\ \text{for}\ \forall\ y\in\Omega\} be the polar body of Ω\Omega. Then there exists a duality

(6.3) q,p(Ω)=p,q(Ω):=(𝕊nhΩp)(𝕊nfrΩq/𝕊nf)p/q.\mathcal{F}_{-q,-p}(\Omega)=\mathcal{F}^{*}_{p,q}(\Omega^{*}):=\left(\int_{{\mathbb{S}}^{n}}h_{\Omega^{*}}^{p}\right)\left(\int_{{\mathbb{S}}^{n}}fr_{\Omega^{*}}^{q}{\Big{/}}\int_{{\mathbb{S}}^{n}}f\right)^{-p/q}.

Thus, the critical point Ω\Omega of the functional q,p()\mathcal{F}_{-q,-p}(\cdot) corresponds to the critical point Ω\Omega^{*} of the functional p,q()\mathcal{F}^{*}_{p,q}(\cdot). Consequently, an analogous solvability result to Proposition 6.1 also holds for p0p\not=0 and q>nq>n.

By the group invariance of the solution Ω(p,q)\Omega^{(p,q)}, one has the uniform a-priori bound for Ω(p,q)\Omega^{(p,q)}. Passing to the limit pp\to-\infty, we obtain the limit Ω(,q)\Omega^{(-\infty,q)} which is the extremal body of the functional ,q\mathcal{F}_{-\infty,q}.

Lemma 6.1.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property. The best constant C(n,p,q,f,𝒮)C(n,p,q,f,{\mathcal{S}}) of the functional p,q()\mathcal{F}_{p,q}(\cdot) converges to the best constant C(n,,q,𝒮)C(n,-\infty,q,{\mathcal{S}}) of the limiting functional

,q(Ω):=Vq(Ω)/(min𝕊nhΩ)q,Ω𝒦0(𝒮).\mathcal{F}_{-\infty,q}(\Omega):=V_{q}(\Omega){\Big{/}}(\min_{{\mathbb{S}}^{n}}h_{\Omega})^{q},\ \ \Omega\in{\mathcal{K}}_{0}({\mathcal{S}}).

Moreover, Ω(,q)\Omega^{(-\infty,q)} is a maximizer of the functional ,q()\mathcal{F}_{-\infty,q}(\cdot).

The proof of Lemma 6.1 is similar to that of Lemma 4.1 and is omitted here. Next we show

Lemma 6.2.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property.

  • (1)

    The maximizer Ω(,q)\Omega^{(-\infty,q)} is a group invariant polytope if q>0q>0.

  • (2)

    The maximizer Ω(,q)\Omega^{(-\infty,q)} is a ball BRn+1B_{R}\subset{\mathbb{R}}^{n+1} for some R>0R>0, if q<0q<0.

Proof.

Part (1) can be proven as that in Theorem 4.2. For part (2), let Ω\Omega^{*} be the polar body of Ω\Omega. Then,

,q(Ω)=q(Ω):=(𝕊nhΩq)/(max𝕊nrΩ)q,Ω𝒦0(𝒮).\mathcal{F}_{-\infty,q}(\Omega)=\mathcal{F}^{*}_{-q}(\Omega^{*}):=\left(\int_{{\mathbb{S}}^{n}}h_{\Omega^{*}}^{-q}\right){\Big{/}}(\max_{{\mathbb{S}}^{n}}r_{\Omega^{*}})^{-q},\ \ \Omega\in{\mathcal{K}}_{0}({\mathcal{S}}).

Noting that the functional q()\mathcal{F}^{*}_{-q}(\cdot) is invariant under scaling, we may assume that max𝕊nrΩ=1\max_{{\mathbb{S}}^{n}}r_{\Omega^{*}}=1. In this case, we have

q(Ω)=𝕊nhΩq(n+1)ωn+1,\mathcal{F}^{*}_{-q}(\Omega^{*})=\int_{{\mathbb{S}}^{n}}h_{\Omega^{*}}^{-q}\leq(n+1)\omega_{n+1},

with equality if and only if Ω=B1\Omega^{*}={B_{1}}. This implies that the extremal body Ω\Omega^{*} maximizing the functional q()\mathcal{F}^{*}_{-q}(\cdot) must be a ball. Hence, we conclude that Ω(,q)\Omega^{(-\infty,q)} is also a ball by duality. The proof is completed. ∎

Lemma 6.3.

Let 𝒮{\mathcal{S}} be a discrete subgroup of O(n+1)O(n+1) satisfying the spanning property. For each p0p\neq 0, the limiting shape Ω(p,+)\Omega^{(p,+\infty)} of Ω(p,q)\Omega^{(p,q)} as q+q\to+\infty is a group invariant polytope.

Proof.

When p<0p<0 and q>nq>n, the maximizer of p,q()\mathcal{F}_{p,q}(\cdot) is also the maximizer of the functional

p,q(Ω):=(𝕊nrΩq)p/q(𝕊nfhΩp/𝕊nf).\mathcal{F}^{*}_{p,q}(\Omega):=\left(\int_{{\mathbb{S}}^{n}}r_{\Omega}^{q}\right)^{-p/q}\left(\int_{{\mathbb{S}}^{n}}fh_{\Omega}^{p}{\Big{/}}\int_{{\mathbb{S}}^{n}}f\right).

As q+q\to+\infty, we conclude that Ω(p,+)\Omega^{(p,+\infty)} is the maximizer of the limiting functional

p,+(Ω):=(𝕊nfhΩp/𝕊nf)/(max𝕊nrΩ)p.\mathcal{F}^{*}_{p,+\infty}(\Omega):=\left(\int_{{\mathbb{S}}^{n}}fh_{\Omega}^{p}{\Big{/}}\int_{{\mathbb{S}}^{n}}f\right){\Big{/}}(\max_{{\mathbb{S}}^{n}}r_{\Omega})^{p}.

Noting that the function p,+()\mathcal{F}^{*}_{p,+\infty}(\cdot) is invariant under scaling, we may assume that maxrΩ=maxhΩ=h(x0)=1\max r_{\Omega}=\max h_{\Omega}=h(x_{0})=1 for some x0𝕊nx_{0}\in\mathbb{S}^{n}. Let Px0=conv{ϕ(x0):ϕ𝒮}B1P_{x_{0}}=\text{conv}\{\phi(x_{0}):\phi\in{\mathcal{S}}\}\subset{B_{1}} be a group invariant polytope. It is clear that Px0P_{x_{0}} is contained inside of Ω\Omega by convexity. Thus,

p,+(Ω)<p,+(Px0)\mathcal{F}^{*}_{p,+\infty}(\Omega)<\mathcal{F}^{*}_{p,+\infty}(P_{x_{0}})

as long as Ω\Omega is not a polytope. Therefore, the limiting shape Ω(p,+)\Omega^{(p,+\infty)} can only be a polytope for p<0p<0.

When p>0p>0 and q>nq>n, the maximizer of p,q()\mathcal{F}_{p,q}(\cdot) is also the minimizer of the functional p,q()\mathcal{F}^{*}_{p,q}(\cdot). Passing to the limit as q+q\to+\infty, we conclude that Ω(p,+)\Omega^{(p,+\infty)} is the minimizer of the limiting functional p,+()\mathcal{F}^{*}_{p,+\infty}(\cdot). Without loss of generality, we may also assume that maxrΩ=rΩ(a)=1\max r_{\Omega}=r_{\Omega}(a)=1 for some a𝕊na\in\mathbb{S}^{n}. Then, it is clear that the minimizer of the functional p,+()\mathcal{F}^{*}_{p,+\infty}(\cdot) is attained at PaP_{a} for p>0p>0. ∎

Proof of Theorem 1.3.

Let Ω(p,q)\Omega^{(p,q)} be the solution obtained in Proposition 6.1 with the support function h(p,q)h^{(p,q)}. Along the same lines as in Corollary 3.1, we have for q0q\neq 0

(6.4) limpmin𝕊nh(p,q)=1.\lim_{p\to-\infty}\min_{{\mathbb{S}}^{n}}h^{(p,q)}=1.

This, together with Lemma 6.2, yields the cases (1) and (2) of Theorem 1.3.

Let Ω(p,q)\Omega^{(p,q)} be the solution obtained in Proposition 6.1 with the support function h(p,q)h^{(p,q)} for p0p\not=0 and q>nq>n as stated in Remark 6.1. Along the same lines as in Corollary 3.1, we have for p0p\neq 0

limq+min𝕊nh(p,q)=1.\lim_{q\to+\infty}\min_{{\mathbb{S}}^{n}}h^{(p,q)}=1.

This, together with Lemma 6.3, yields the case (3) of Theorem 1.3. ∎

7. The proof of Theorem 1.4

This section aims to prove Theorem 1.4. Firstly, let us consider q>0q>0. In this case, we have the following result, analogous to Lemma 5.1.

Lemma 7.1.

For q>0q>0, a regular polytope TT is a strict local maximal of the functional VqV_{q} in the family 𝒦min(δ,𝒮T){\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}) for some δ>0\delta>0. Namely, there exists a positive constant δ=δq,T\delta=\delta_{q,T} such that for any Ω𝒦min(δ,𝒮T){T}\Omega^{\prime}\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})\setminus\{T\}, there holds

(7.1) Vq(Ω)<Vq(T).V_{q}(\Omega^{\prime})<V_{q}(T).
Proof.

Similar to the proof of Claim 1\sharp 1 in Lemma 5.1, we can check that for some δ>0\delta>0,

Vq(T)=supΩ𝒦min(δ,𝒮T)Vq(Ω)V_{q}(T)=\sup_{\Omega\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})}V_{q}(\Omega)

if and only if

(7.2) V¯q(w¯0)=Ωn(1+|z¯|2)qn12(1|1+|w¯0|21+z¯,w¯0|q)𝑑z¯>0,\displaystyle\overline{V}_{q}(\overline{w}_{0})=\int_{\Omega_{n}}(1+|\overline{z}|^{2})^{\frac{q-n-1}{2}}\bigg{(}1-\bigg{|}\frac{\sqrt{1+|\overline{w}_{0}|^{2}}}{1+\langle\overline{z},\overline{w}_{0}\rangle}\bigg{|}^{q}\bigg{)}d\overline{z}>0,

for all w¯0Ωn{0}\overline{w}_{0}\in\Omega_{n}\setminus\{0\} closing to 0.

Let w0=(w¯0,1)w_{0}=(\overline{w}_{0},1) approach On=en+1O_{n}=e_{n+1} but w0Onw_{0}\neq O_{n}. Along the direction segment I={On+raΩn|r0,a+n}I=\{O_{n}+ra\in\Omega_{n}|r\geq 0,a\in{\mathbb{R}}^{n}_{+}\}, the volume difference function can be written as

Vq(r,a):=V¯q(w¯0)=Ωn(1+|z¯|2)qn12(1|1+r2|a|21+rz¯,a|q)𝑑z¯.V_{q}(r,a):=\overline{V}_{q}(\overline{w}_{0})=\int_{\Omega_{n}}(1+|\overline{z}|^{2})^{\frac{q-n-1}{2}}\bigg{(}1-\bigg{|}\frac{\sqrt{1+r^{2}|a|^{2}}}{1+r\langle\overline{z},a\rangle}\bigg{|}^{q}\bigg{)}d\overline{z}.

Then by the fact that Ωn+n\Omega_{n}\subset{\mathbb{R}}^{n}_{+}, we have

ddr|r=0Vq(r,a)=qΩn(1+|z¯|2)qn12z¯,a𝑑z¯>0,a+n.\frac{d}{dr}\bigg{|}_{r=0}V_{q}(r,a)=q\int_{\Omega_{n}}(1+|\overline{z}|^{2})^{\frac{q-n-1}{2}}\langle\overline{z},a\rangle d\overline{z}>0,\ \ \forall a\in{\mathbb{R}}^{n}_{+}.

This, together with (7.2), gives the desired inequality (7.1). ∎

To complete the proof of Theorem 1.4 for the case q>0q>0, we shall consider the local variational scheme

(7.3) supΩ𝒦p,f(δ,𝒮T)Vq(Ω).\sup_{\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V_{q}(\Omega).

Then we will show the existence of maximizer Ω(p,q)𝒦p,f(δ,𝒮T)\Omega^{(p,q)}\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T}) of (7.3) for negative large pp, whose support function h(p,q)h^{(p,q)} satisfies the Euler-Lagrange equation (1.1) up to a constant.

Lemma 7.2.

Let q>0q>0 and Tn+1T\subset{\mathbb{R}}^{n+1} be a regular polytope tangential to the unit sphere 𝕊n{\mathbb{S}}^{n}. If δ=δq,T>0\delta=\delta_{q,T}>0 is small and pp is negative large, the variational scheme (7.3) admits a maximizer Ω(p,q)𝒩δ(T)\Omega^{(p,q)}\not\in\partial{\mathcal{N}}_{\delta}(T). Moreover, the maximizer Ω(p,q)\Omega^{(p,q)} satisfies the equation (1.1), and converges to the given regular polytope TT as pp\rightarrow-\infty.

Proof.

The existence of maximizer Ω(p,q)\Omega^{(p,q)} to the variational scheme (7.3) follows from a-priori estimates for group invariant polytopes.

Next, we show that the property that Ω(p,q)𝒩δ(T)\Omega^{(p,q)}\not\in\partial{\mathcal{N}}_{\delta}(T). In fact, by Lemma 7.1, we know that TT is a local strictly maximizer of the functional ,q(){\mathcal{F}}_{-\infty,q}(\cdot) on the family 𝒦min(δ,𝒮T){\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}) for some δ>0\delta>0. Hence, there exists a positive constant δ=δq,T\delta=\delta_{q,T} such that

(7.4) ,q(Ω)<,q(T),Ω𝒦min(δ,𝒮T)T.{\mathcal{F}}_{-\infty,q}(\Omega)<{\mathcal{F}}_{-\infty,q}(T),\ \ \forall\Omega\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})\setminus T.

Thus we have that for some σ>0\sigma>0,

(7.5) supΩ𝒦min(δ,𝒮T),q(Ω),q(T)σ,\sup_{\Omega\in\partial{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})}{\mathcal{F}}_{-\infty,q}(\Omega)\leq{\mathcal{F}}_{-\infty,q}(T)-\sigma,

and

(7.6) limpp,q(Ω)=,q(Ω).\lim_{p\rightarrow-\infty}{\mathcal{F}}_{p,q}(\Omega)={\mathcal{F}}_{-\infty,q}(\Omega).

Similar to the proof of Lemma 5.2, we have

supΩ𝒦p,f(δ,𝒮T)p,q(Ω)supΩ𝒦p,f(δ,𝒮T)Vq(Ω)σ+ε(p)\displaystyle\sup_{\Omega\in\partial{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}{\mathcal{F}}_{p,q}(\Omega)\leq\sup_{\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V_{q}(\Omega)-\sigma+\varepsilon(p)

holds for negative large pp and some infinitesimal ε(p)\varepsilon(p), and hence

(7.7) supΩ𝒦p,f(δ,𝒮T)Vq(Ω)supΩ𝒦p,f(δ,𝒮T)Vq(Ω)σ+ε(p)\sup_{\Omega\in\partial{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V_{q}(\Omega)\leq\sup_{\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V_{q}(\Omega)-\sigma+\varepsilon(p)

holds for small δ>0\delta>0 and negative large pp. This implies that the property that Ω(p,q)\Omega^{(p,q)} stays away from the boundary of 𝒩δ(T){\mathcal{N}}_{\delta}(T). Then we can prove that Ω(p,q)\Omega^{(p,q)} satisfies the equation (1.1) as in the proof of the variational problem (6.1). Since

limpsupΩ𝒦p,f(δ,𝒮T)Vq(Ω)=supΩ𝒦min(δ,𝒮T)Vq(Ω),\lim_{p\to-\infty}\sup_{\Omega\in{\mathcal{K}}_{p,f}(\delta,{\mathcal{S}}_{T})}V_{q}(\Omega)=\sup_{\Omega\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T})}V_{q}(\Omega),

by the compactness of 𝒦(δ,𝒮T){\mathcal{K}}(\delta,{\mathcal{S}}_{T}). Thus, the maximizer Ω(p,q)\Omega^{(p,q)} converges to a maximizer Ω(,q)\Omega^{(-\infty,q)} of the limiting functional. However, by Lemma 7.1, TT is the unique maximizer of the limiting functional on the family 𝒦min(δ,𝒮T){\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T}). We reach the conclusion that Ω(,q)=T\Omega^{(-\infty,q)}=T by the observation that ,q(Ω)=Vq(Ω){\mathcal{F}}_{-\infty,q}(\Omega)=V_{q}(\Omega) for Ω𝒦min(δ,𝒮T)\forall\ \Omega\in\mathcal{K}_{min}(\delta,\mathcal{S}_{T}). This completes the proof. ∎

Proof of (1) in Theorem 1.4..

Part (1) of Theorem 1.4 is a direct consequence of Lemma 7.2. ∎

To prove the second part of Theorem 1.4, we denote

Vp,f(Ω)=𝕊nfhΩp,p0,f>0,V_{p,f}(\Omega)=\int_{{\mathbb{S}}^{n}}fh_{\Omega}^{p},\ \ \ p\not=0,\ f>0,

and

𝒦max(𝒮T)={Ω𝒦0(𝒮T)|max𝕊nhΩ=1},𝒦max(δ,𝒮T)=𝒦max(𝒮T)𝒩δ(T).{\begin{split}{\mathcal{K}}_{max}({\mathcal{S}}_{T})&=\big{\{}\Omega\in{\mathcal{K}}_{0}({\mathcal{S}}_{T})\big{|}\max_{{\mathbb{S}}^{n}}h_{\Omega}=1\big{\}},\\ {\mathcal{K}}_{max}(\delta,{\mathcal{S}}_{T})&={\mathcal{K}}_{max}({\mathcal{S}}_{T})\cap{\mathcal{N}}_{\delta}(T).\end{split}}
Lemma 7.3.

For any regular polytope Tn+1T\subset{\mathbb{R}}^{n+1} with outer radius one and positive group invariant function ff on 𝕊n{\mathbb{S}}^{n}, there exists a small constant δ=δp,f,T>0\delta=\delta_{p,f,T}>0 such that

(7.8) {Vp,f(T)<Vp,f(Ω),p>0,Ω𝒦max(δ,𝒮T){T},Vp,f(T)>Vp,f(Ω),p<0,Ω𝒦max(δ,𝒮T){T}.\begin{cases}V_{p,f}(T)<V_{p,f}(\Omega),&p>0,\ \ \forall\Omega\in{\mathcal{K}}_{max}(\delta,{\mathcal{S}}_{T})\setminus\{T\},\\ V_{p,f}(T)>V_{p,f}(\Omega),&p<0,\ \ \forall\Omega\in{\mathcal{K}}_{max}(\delta,{\mathcal{S}}_{T})\setminus\{T\}.\end{cases}
Proof.

Firstly, let p>0p>0 and assume ff is positive group invariant function. For any Ω𝒦max(δ,𝒮T)\Omega\in{\mathcal{K}}_{max}(\delta,{\mathcal{S}}_{T}), there exists at least one a𝕊na\in{\mathbb{S}}^{n} such that

hΩ(a)=max𝕊nhΩ=1.h_{\Omega}(a)=\max_{{\mathbb{S}}^{n}}h_{\Omega}=1.

Set Ωa=conv(ϕ𝒮Tϕ(a))\Omega_{a}=\text{conv}(\cup_{\phi\in{\mathcal{S}}_{T}}\phi(a)), it is clearly that

(7.9) Vp,f(Ω)Vp,f(Ωa)V_{p,f}(\Omega)\geq V_{p,f}(\Omega_{a})

since ΩaΩ\Omega_{a}\subset\Omega. We remain to show that

(7.10) Vp,f(Ωa)>Vp,f(T)V_{p,f}(\Omega_{a})>V_{p,f}(T)

if Ωa\Omega_{a} approaches TT without being equal to TT. Noting that Ωa,T𝒦max(δ,𝒮T)\Omega_{a},T\in{\mathcal{K}}_{max}(\delta,{\mathcal{S}}_{T}), if one takes their dual bodies Ωb\Omega_{b} (for some b𝕊nb\in\mathbb{S}^{n}), TT^{*} respectively, we have Ωb,T𝒦min(δ,𝒮T)\Omega_{b},T^{*}\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T^{*}}) are closing from each other under the Hausdorff distance. So, to prove (7.10) is equivalent to proving

(7.11) Vp,f(Ωb)>Vp,f(T),V^{*}_{-p,f}(\Omega_{b})>V^{*}_{-p,f}(T^{*}),

where

Vq,f(Ω):=𝕊nf(y)rΩq(y)𝑑y.V^{*}_{q,f}(\Omega):=\int_{{\mathbb{S}}^{n}}f(y)r_{\Omega}^{q}(y)dy.

The proof of (7.11) is similar to the proof of (5.8). Actually, after subdivision of TT^{*} into congruent pieces similar to one polytope Ωn+n\Omega_{n}\subset{\mathbb{R}}^{n}_{+}, the comparison of Vp,f(Ωb)V^{*}_{-p,f}(\Omega_{b}) with Vp,f(T)V^{*}_{-p,f}(T^{*}) can be reduced to the verification of the negativity of the function

V¯p,f(w¯0)=Ωnf(1+|z¯|2)pn12(1|1+|w¯0|21+z¯,w¯0|p)𝑑z¯\displaystyle\overline{V}_{-p,f}(\overline{w}_{0})=\int_{\Omega_{n}}f(1+|\overline{z}|^{2})^{\frac{-p-n-1}{2}}\bigg{(}1-\bigg{|}\frac{\sqrt{1+|\overline{w}_{0}|^{2}}}{1+\langle\overline{z},\overline{w}_{0}\rangle}\bigg{|}^{-p}\bigg{)}d\overline{z}

for w¯00\overline{w}_{0}\not=0 approaching 0. Along the direction segment I={On+raΩn|r0,a+n}I=\{O_{n}+ra\in\Omega_{n}|r\geq 0,a\in{\mathbb{R}}^{n}_{+}\}, we have

Vp,f(r,a):=V¯p,f(w¯0)=Ωnf(1+|z¯|2)pn12(1|1+r2|a|21+rz¯,a|p)𝑑z¯.V_{-p,f}(r,a):=\overline{V}_{-p,f}(\overline{w}_{0})=\int_{\Omega_{n}}f(1+|\overline{z}|^{2})^{\frac{-p-n-1}{2}}\bigg{(}1-\bigg{|}\frac{\sqrt{1+r^{2}|a|^{2}}}{1+r\langle\overline{z},a\rangle}\bigg{|}^{-p}\bigg{)}d\overline{z}.

Therefore, it follows from

(7.12) ddr|r=0Vp,f(r,a)=pΩnf(1+|z¯|2)pn12z¯,a𝑑z¯<0\frac{d}{dr}\bigg{|}_{r=0}V_{-p,f}(r,a)=-p\int_{\Omega_{n}}f(1+|\overline{z}|^{2})^{\frac{-p-n-1}{2}}\langle\overline{z},a\rangle d\overline{z}<0

for a0a\not=0 that (7.11) holds true.

After modifying the proof of (7.11) slightly, we can also show that: If p<0p<0, for Ωb𝒦min(δ,𝒮T),b𝕊n\Omega_{b}\in{\mathcal{K}}_{min}(\delta,{\mathcal{S}}_{T^{*}}),b\in{\mathbb{S}}^{n} approaches TT^{*} without being equal to TT^{*}, there holds

Vp,f(Ωb)<Vp,f(T),V^{*}_{-p,f}(\Omega_{b})<V^{*}_{-p,f}(T^{*}),

which is equivalent to

Vp,f(Ωa)<Vp,f(T).V_{-p,f}(\Omega_{a})<V_{-p,f}(T).

This, together with the fact that

(7.13) Vp,f(Ω)Vp,f(Ωa),Ω𝒦max(δ,𝒮T),V_{p,f}(\Omega)\leq V_{p,f}(\Omega_{a}),\ \ \forall\ \Omega\in{\mathcal{K}}_{max}(\delta,{\mathcal{S}}_{T}),

gives the desired result for case p<0p<0. ∎

With the help of Lemma 7.3, we know that the regular polytope TT is of local strictly extremal of Vp,fV_{p,f} for p0p\not=0 and f>0f>0. Considering the same local variational scheme (7.3) and modifying the argument in Section 5 slightly, we reach the following result.

Lemma 7.4.

Let p0p\not=0 and Tn+1T\subset{\mathbb{R}}^{n+1} be a regular polytope tangential with outer radius one, if δ=δp,T>0\delta=\delta_{p,T}>0 is small and qq is positive large, the variational scheme (7.3) admits a maximizer Ω(p,q)𝒩δ(T)\Omega^{(p,q)}\not\in{\partial\mathcal{N}}_{\delta}(T). Moreover, the maximizer Ω(p,q)\Omega^{(p,q)} satisfies the equation (1.1), and converges to a regular polytope similar to TT as q+q\rightarrow+\infty.

Proof of (2) in Theorem 1.4..

Part (2) of Theorem 1.4 is a direct consequence of Lemma 7.4. ∎

Acknowledgement. The authors would like to thank Xudong Wang for many helpful comments.

References

  • [1] B. Andrews, Gauss curvature flow: the fate of the rolling stones, Invent. Math., 138 (1999), 151-161.
  • [2] B. Andrews, Motion of hypersurfaces by gauss curvature, Pacific J. Math., 195 (2000), 1-34.
  • [3] B. Andrews, Classification of limiting shapes for isotropic curve flows, J. Amer. Math. Soc., 16 (2003), 443-459.
  • [4] G. Bianchi, K. Böröczky, A. Colesanti and D. Yang, The LpL_{p}-Minkowski problem for n<p<1-n<p<1, Adv. Math., 341 (2019), 493-535.
  • [5] K. Böröczky, The logarithmic Minkowski conjecture and the LpL_{p}-Minkowski problem, Adv. Anal. Geom., 9 (2023), 83-118.
  • [6] K. Böröczky and F. Fodor, The LpL_{p} dual Minkowski problem for p>1p>1 and q>0q>0, J. Diff. Equa., 266 (2019), 7980-8033.
  • [7] K. Böröczky, E. Lutwak, D. Yang and G. Zhang, The logarithmic Minkowski problem, J. Amer. Math. Soc., 26 (2013), 831-852.
  • [8] P. Bryan, M. Ivaki and J. Scheuer, A unified flow approach to smooth, even LpL_{p}-Minkowski problems, Analysis & PDE, 12 (2019), 259-280.
  • [9] L.A. Caffarelli, A localization property of viscosity solutions to the Monge-Ampère equation and their strict convexity, Ann. of Math., 131 (1990), 129-134.
  • [10] L.A. Caffarelli, Interior w2,pw^{2,p} estimates for solutions of the Monge-Ampère equation, Ann. of Math., 131 (1990), 135-150.
  • [11] C. Chen, Y. Huang and Y. Zhao, Smooth solutions to the LpL_{p} dual Minkowski problem, Math. Ann., 373 (2019), 953-976.
  • [12] H. Chen, S. Chen and Q. Li, Variations of a class of Monge-Ampère-type functionals and their applications, Analysis & PDE, 14 (2021), 689-716.
  • [13] K.S. Chou and X.-J. Wang, The LpL_{p}-Minkowski problem and the Minkowski problem in centroaffine geometry, Adv. Math., 205 (2006), 33-83.
  • [14] H.S.M. Coxeter, Regular polytopes, Courier Corporation, 1973.
  • [15] P.R. Cromwell, Polyhedra, Cambridge University Press, 1997.
  • [16] S. Du, On the planar LpL_{p}-Minkowski problem, J. Diff. Equa., 287 (2021), 37-77.
  • [17] P.M. Gruber, Convex and discrete geometry, Springer-Verlag, Berlin Heidelberg, 2007.
  • [18] Q. Guang, Q. Li and X.-J. Wang, The LpL_{p}-Minkowski problem with super-critical exponents, preprint.
  • [19] Y. He, Q. Li and X.-J. Wang, Multiple solutions of the LpL_{p}-Minkowski problem, Calc. Var. PDEs, 55 (2016), 13 pp.
  • [20] Y. Huang, J. Liu and L. Xu, On the uniqueness of LpL_{p}-Minkowski problems: the constant pp-curvature case in 3{\mathbb{R}}^{3}, Adv. Math., 281 (2015), 906-927.
  • [21] Y. Huang, E. Lutwak, D. Yang and G. Zhang, Geometric measures in the dual Brunn-Minkowski theory and their associated Minkowski problems, Acta Math., 216 (2016), 325-388.
  • [22] Y. Huang and Y. Zhao, On the LpL_{p} dual Minkowski problem, Adv. Math., 332 (2018), 57-84.
  • [23] D. Hug, E. Lutwak, D. Yang and G. Zhang, On the LpL_{p}-Minkowski problem for polytopes, Discrete &\& Comput. Geom., 33 (2005), 699-715.
  • [24] M. Ivaki and E. Milman, LpL_{p}-Minkowski problem under curvature pinching, Int. Math. Res. Not., IMRN (2024), 8638-8652.
  • [25] H. Jian, J. Lu and X.-J. Wang, Nonuniqueness of solutions to the LpL_{p}-Minkowski problem, Adv. Math., 281 (2015), 845-856.
  • [26] Q. Li, Infinitely many solutions for centro-affine Minkowski problem, Int. Math. Res. Not., 18 (2019), 5577-5596.
  • [27] J. Lu and X.-J. Wang, Rotationally symmetric solutions to the LpL_{p}-Minkowski problem, J. Diff. Equa., 254 (2013), 983-1005.
  • [28] E. Lutwak, The Brunn-Minkowski-Firey theory I. Mixed volumes and the Minkowski problem, J. Diff. Geom., 38 (1993), 131-150.
  • [29] E. Lutwak, D. Yang and G. Zhang, On the LpL_{p}-Minkowski problem, Trans. Am. Math. Soc., 356 (2004), 4359-4370.
  • [30] E. Lutwak, D. Yang and G. Zhang, LpL_{p} dual curvature measures, Adv. Math., 329 (2018), 85-132.
  • [31] E. Milman, Centro-affine differential geometry and the log-Minkowski problem, J. Eur. Math. Soc., to appear.
  • [32] E. Milman, A sharp centro-affine isospectral inequality of Szegö-Weinberger type and the LpL_{p}-Minkowski problem, J. Diff. Geom., 127 (2024), 373-408.
  • [33] R. Schneider, Convex bodies: the Brunn-Minkowski theory, Encyclopedia of Mathematics and its Applications, 151, Cambridge University Press, Cambridge, 2014.
  • [34] L. Serge, Algebra, Grad. Texts in Math., 211, Springer-Verlag, New York, 2002, xvi+914 pp.
  • [35] Y. Zhao, The dual Minkowski problem for negative indices, Calc. Var. PDEs, 56 (2017), 16 pp.
  • [36] Y. Zhao, Existence of solutions to the even dual Minkowski problem, J. Diff. Geom., 110 (2018), 543-572.
  • [37] G. Zhu, The LpL_{p}-Minkowski problem for polytopes for 0<p<10<p<1, J. Funct. Anal., 269 (2015), 1070-1094.
  • [38] G. Zhu, The LpL_{p}-Minkowski problem for polytopes for negative pp, Indiana Univ. Math. J., 66 (2017), 1333-1350.