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

The LpL_{p}-Gaussian Minkowski problem

Jiaqian Liu School of Mathematics, Hunan University, Changsha, 410082, Hunan Province, China liujiaqian@hnu.edu.cn
Abstract.

In this paper, we extend the article that Minkowski problem in Gaussian probability space of Huang et al. to LpL_{p}-Gaussian Minkowski problem, and obtain the existence and uniqueness of oo-symmetry weak solution in case of p1p\geq 1.

Key words and phrases:
Minkowski problem, LpL_{p}-Gaussian Minkowski problem, Monge-Ampère equation, degree theory
2010 Mathematics Subject Classification:
52A40, 52A38, 35J96.
The author was partially supported by Tian Yuan Special Foundation(12026412) and Hunan Science and Technology Planning Project(2019RS3016).

1. Introduction

The Brunn-Minkowski theory mainly studies the geometric functional of convex bodies and their differential. The differentiable functional will generate some intrinsic geometric measures. In Euclidean space n\mathbb{R}^{n}, the most significant functional of convex body is volume denoted by VV, and Aleksandrov [1] gave the variational formula

limt0V(K+tL)V(K)t=Sn1hL(v)𝑑SK(v),\lim_{t\rightarrow 0}\frac{V(K+tL)-V(K)}{t}=\int_{S^{n-1}}h_{L}(v)dS_{K}(v), (1.1)

where K+tL={x+ty:xK and yL}K+tL=\{x+ty:\text{$x\in K$ and $y\in L$}\}, hL:Sn1h_{L}:S^{n-1}\rightarrow\mathbb{R} is the support function of convex body LL. It naturally generates the surface area measure SKS_{K} of convex body KK. The classical Minkowski problem characterizes the surface area measure, it reads : given a finite Borel measure μ\mu on Sn1S^{n-1}, what are the necessary and sufficient conditions on μ\mu does there exist a convex body KK in n\mathbb{R}^{n} such that surface area measure SK()=μ()S_{K}(\cdot\,)=\mu(\cdot\,)? If KK exists, is it unique? In the smooth case, it is equivalent to solve a certain fully nonlinear elliptic equation. The classical Minkowski problem was solved by Minkowski in [26] for polytope, and later solved by Aleksandrov [1], Fenchel and Jessen [16] in general.

In 1993, Lutwak developed the Brunn-Minkowski theory to LpL_{p}-Brunn-Minkowski theory. The LpL_{p}-Minkowski problem, characterizing LpL_{p}-surface area measure Sp(K)=hK1pSKS_{p}(K)={h_{K}}^{1-p}S_{K}. Clearly, when p=1p=1, the LpL_{p}-surface area measure is just the classical surface measure. Lutwak [39] testified that there exists a unique oo-symmetry convex body such that Sp(K)=μS_{p}(K)=\mu for p>1p>1. The LpL_{p}-Minkowski problem has been studied extensively in [3, 4, 11, 27, 24, 33, 43, 48, 47, 49] and so on. The uniqueness of the Minkowski problem is related to Brunn-Minkowski inequality. For p>1p>1, the LpL_{p}-Brunn-Minkowski inequality can be obtained easily. In case of 0<p<10<p<1, LpL_{p}-Brunn-Minkowski inequality and its condition of the equality holds have not been improved much until [8]. While the logarithmic Minkowski problem (p=0)(p=0) and the centro-affine Minkowski problem (p=n)(p=-n) are two special cases; see, e.g., [5], and [12]. The regularity of the LpL_{p}-Minkowski problem, for example [12, 28, 40].

In [38], the dual Brunn-Minkowski theory was developed in the 1970s. The most significant dual curvature measure and its associated Minkowski problem in the dual Brunn-Minkowski theory were studied in [29]. The work of [29] play a key role in promoting the development of partial differential equation and dual Brunn-Minkowski theory as [9, 31, 6, 41, 10, 45]. Recently, a number of meaningful and interesting work emerged on the Orlicz-Minkowski problem, such as [19, 20, 25, 34, 35, 46, 52].

In [30], the authors posed the Minkowski problem in Gaussian probability space. Gaussian volume functional denoted by γn\gamma_{n} plays a central role, which is given by

γn(K)=1(2π)nKe|x|22𝑑x.\gamma_{n}(K)=\frac{1}{(\sqrt{2\pi})^{n}}\int_{K}e^{\frac{-|x|^{2}}{2}}dx.

Using the variational formula from Huang et al. [29], the variational formula of γn\gamma_{n} is obtained as follows:

limt0γn(K+tL)γn(K)t=Sn1hL(v)𝑑Sγn,K(v),\lim_{t\rightarrow 0}\frac{\gamma_{n}(K+tL)-\gamma_{n}(K)}{t}=\int_{S^{n-1}}h_{L}(v)dS_{\gamma_{n},K}(v), (1.2)

where Gaussian surface area measure Sγn,KS_{\gamma_{n},K} is given by

Sγn,K(η)=1(2π)nνK1(η)e|x|22𝑑n1(x).S_{\gamma_{n},K}(\eta)=\frac{1}{(\sqrt{2\pi})^{n}}\int_{\nu_{K}^{-1}(\eta)}e^{\frac{-|x|^{2}}{2}}d\mathcal{H}^{n-1}(x).

About the mainly result of Gaussian Minkowski problem showed in [30] Theorem 1.4, stated as: Let μ\mu be a finite even Borel measure on Sn1S^{n-1} that is not concentrated on any closed hemisphere and |μ|<12π|\mu|<\frac{1}{\sqrt{2\pi}}. Then there exists a unique oo-symmetry convex body with γn(K)>12\gamma_{n}(K)>\frac{1}{2} such that Sγn,K=μ.S_{\gamma_{n},K}=\mu.

As remarked in [30], Gaussian Minkowski problem is different from the Minkowski problem in Lebesgue measure space in these aspects. At first, the Gaussian surface area of all convex set in n\mathbb{R}^{n} is not more than 4n144{n^{\frac{1}{4}}} which is shown detailed in Ball [2] and [18]. That is to say, the allowable μ\mu in Gaussian Minkowski problem can not have an arbitrarily big total mass. Secondly, Gaussian probability measure has neither translation invariance nor homogeneity, which makes our research more difficult. As a final remark, by the definition of Gaussian surface area, we can know that not only small convex bodies can have smaller Gaussian surface areas, but also large convex bodies. For instance, let BrB_{r} be a centered ball with radius r,r, the Gaussian surface area’s density of BrB_{r} is fr=1(2π)ner22rn1f_{r}=\frac{1}{(\sqrt{2\pi})^{n}}e^{\frac{-r^{2}}{2}}r^{n-1}. One can observe that 1(2π)ner22rn10\frac{1}{(\sqrt{2\pi})^{n}}e^{\frac{-r^{2}}{2}}r^{n-1}\rightarrow 0 as r0r\rightarrow 0 or r.r\rightarrow\infty.

Naturally, we develop the LpL_{p}-Brunn-Minkowski theory in Gaussian probability space. The purpose of this paper is to continue the study in [30] and to construct LpL_{p}-Gaussian surface area measures. A LpL_{p}-variational formula similar to Lemma 3.2 in Lutwak [39] is gained as below:

limt0γn([ht])γn(K)t=1pSn1f(v)p𝑑Sp,γn,K(v),\lim_{t\rightarrow 0}\frac{\gamma_{n}\left([h_{t}]\right)-\gamma_{n}\left(K\right)}{t}=\frac{1}{p}\int_{S^{n-1}}f(v)^{p}dS_{p,\gamma_{n},K}(v),

where ht(v)=(hK(v)p+tf(v)p)1ph_{t}(v)=(h_{K}(v)^{p}+tf(v)^{p})^{\frac{1}{p}} is LpL_{p}-combination for p0p\neq 0, and it leads to define the LpL_{p}-Gaussian surface area measures as follows

Sp,γn,K(η)=1(2π)nνK1(η)e|x|22(xνK(x))1p𝑑n1(x).S_{p,\gamma_{n},K}(\eta)=\frac{1}{(\sqrt{2\pi})^{n}}\int_{\nu_{K}^{-1}(\eta)}e^{\frac{-|x|^{2}}{2}}(x\cdot\nu_{K}(x))^{1-p}d\mathcal{H}^{n-1}(x).

The LpL_{p}-Gaussian Minkowski problem reads : given a finite Borel measure μ\mu on Sn1S^{n-1}, what are the necessary and sufficient conditions on μ\mu does there exist a convex body KK in n\mathbb{R}^{n} such that LpL_{p}-Gaussian surface area measure Sp,γn,K()=μ()S_{p,\gamma_{n},K}(\cdot\,)=\mu(\cdot\,)? If KK exist, is it unique?

The first assertion involves the uniqueness result, it is known to us that the Gaussian surface area’s density frf_{r} of BrB_{r} tends to 0 as r0r\rightarrow 0 or r,r\rightarrow\infty, this means that Gaussian surface area is may not unique. Our study in the restricted context that the Gaussian volume grater than or equal to 12\frac{1}{2}, then obtain the following uniqueness result. The same fact was firstly found in [30] in the situation of p=1p=1.

Theorem 1.1 (Uniqueness).

For p1p\geq 1, let K,L𝒦onK,L\in\mathcal{K}^{n}_{o} with the same LpL_{p}-Gaussian surface area, i.e.,

Sp,γn,K=Sp,γn,L.S_{p,\gamma_{n},K}=S_{p,\gamma_{n},L}.

If γn(K),γn(L)12,\gamma_{n}(K),\gamma_{n}(L)\geq\frac{1}{2}, then K=L.K=L.

Aleksandrov’s variational method implies the existence of solution with lagrange multipliers.

Theorem 1.2.

For p>0,p>0, let μ\mu be a nonzero finite Borel measure on Sn1S^{n-1} and be not concentrated in any closed hemisphere. Then there exist a K𝒦onK\in\mathcal{K}^{n}_{o} and a positive constant λ\lambda such that

μ=λpSp,γn,K.\mu=\frac{\lambda}{p}S_{p,\gamma_{n},K}.

Due to the lack of homogeneity of measure, just like all previous Orlicz-Minkowski problems, in the solving process, the lagrange multipliers produced by Aleksandrov’s variational method can not eliminated. Recently, the work of Huang et al. in [30], the first one overcame this obstruct. Inspired their work, we obtain the existence of smooth solution.

Theorem 1.3 (Existence of smooth solution).

For p1,p\geq 1, let even function fC2,α(Sn1)f\in C^{2,\alpha}(S^{n-1}) is positive and |f|L1<2πrpaea22|f|_{L_{1}}<\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}}, the rr and aa are chosen such that γn(rB)=γn(P)=12\gamma_{n}(rB)=\gamma_{n}(P)=\frac{1}{2}, symmetry strip P={xn:|x1|a}P=\left\{x\in\mathbb{R}^{n}:|x_{1}|\leq a\right\}. Then there exists a unique C4,αC^{4,\alpha} convex body K𝒦enK\in\mathcal{K}^{n}_{e} with γn(K)>12\gamma_{n}(K)>\frac{1}{2}, the support function hKh_{K} solves the equation

1(2π)nhK1pe(|hK|2+hK2)2det(2hK+hKI)=f.\frac{1}{(\sqrt{2\pi})^{n}}{h_{K}}^{1-p}e^{\frac{-(|\nabla h_{K}|^{2}+h_{K}^{2})}{2}}\det(\nabla^{2}h_{K}+h_{K}I)=f. (1.3)

Finally, the existence of weak solution of the LpL_{p}-Gaussian Minkowski problem follows by an approximation method with the help of a uniform estimate.

Theorem 1.4 (Existence of weak solution ).

For p1,p\geq 1, let μ\mu be a nonzero finite even Borel measure on Sn1S^{n-1} and be not concentrated in any closed hemisphere with |μ|<2πrpaea22,|\mu|<\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}}, the rr and aa are chosen such that γn(rB)=γn(P)=12\gamma_{n}(rB)=\gamma_{n}(P)=\frac{1}{2}, symmetry strip P={xn:|x1|a}P=\left\{x\in\mathbb{R}^{n}:|x_{1}|\leq a\right\}. Then there exists a unique K𝒦enK\in\mathcal{K}^{n}_{e} with γn(K)>12\gamma_{n}(K)>\frac{1}{2} such that

Sp,γn,K=μ.S_{p,\gamma_{n},K}=\mu.

We overcome the most important obstacle of the LpL_{p} Gaussian Minkowski problem about whether the solution degenerates or not, and obtain that it holds for asymmetric convex bodies see Theorem 3.3. But the limitation of this item is the lack of LpL_{p} Gaussian isoperimetric type inequality for general convex body such that it can only achieve the oo-symmetry convex body, see Theorem 4.7 and Theorem 5.5.

The paper is organized into five sections: we list some conceptual knowledge of convex bodies in Preliminaries part, and gain the key LpL_{p}-variational formula in Section 2. In Section 3, Aleksandrov’s variational method is applied to obtain the non-symmetry solution with lagrange multipliers for p>0p>0. The uniqueness of solution is proved in Section 4. Section 5 provides the existence result of oo-symmetry smooth solution for p1p\geq 1 by means of the degree theory of partial differential equation, then the oo-symmetry weak solution is obtained with the help of approximation argument.

2. LpL_{p}-Gaussian surface area and the LpL_{p}-Gaussian Minkowski problem

In this section, we make a brief review of some relevant notions about convex bodies. More detailed information of convex body theory can be found in the Schneider [44].

Let n\mathbb{R}^{n} represent the nn-dimensional Euclidean space, Sn1S^{n-1} be the unit sphere in n,\mathbb{R}^{n}, C+(Sn1)C^{+}(S^{n-1}) be positive functions defined on Sn1S^{n-1} and |μ||\mu| denote total measure of the Borel measure μ\mu on Sn1.S^{n-1}.

Convex body in n\mathbb{R}^{n} is compact convex set with nonempty interior denoted by 𝒦n\mathcal{K}^{n}. Let K\partial K denote the boundary of KK, 𝒦on\mathcal{K}^{n}_{o} be convex bodies that contain the origin in interior, and 𝒦en\mathcal{K}^{n}_{e} denote the oo-symmetry convex bodies.

The support function hKh_{K} of a compact convex subset KK in n\mathbb{R}^{n} is defined as

hK(y)=maxxKyxh_{K}(y)=\max_{x\in K}y\cdot x

for yny\in\mathbb{R}^{n}, where \cdot represents the inner product.

For K𝒦onK\in\mathcal{K}^{n}_{o}, the radial function ρK\rho_{K} is defined by

ρK(x)=max{λ:λxK}\rho_{K}(x)=\max\{\lambda:\lambda x\in K\}

for xn{0}x\in\mathbb{R}^{n}\setminus\{0\}. A simple observation allows us to obtain

K={ρK(u)u:uSn1}.\partial K=\{\rho_{K}(u)u:u\in S^{n-1}\}.

In fact, 𝒦n\mathcal{K}^{n} is the space of compact convex set in n\mathbb{R}^{n} and it can be equipped with Hausdorff metric which associated norm is

δ(K,L)=maxvSn1|hK(v)hL(v)|.\delta(K,L)=\max_{v\in S^{n-1}}|h_{K}(v)-h_{L}(v)|.

We can easily show that δ(K,L)\delta(K,L) is the distance between K,L𝒦nK,L\subset\mathcal{K}^{n}. Therefore, the convergence of compact convex set sequence can be characterized by the convergence of support function. That is to say, Ki𝒦nK𝒦nK_{i}\in\mathcal{K}^{n}\rightarrow K\in\mathcal{K}^{n} if

maxvSn1|hKi(v)hK(v)|0,\max_{v\in S^{n-1}}|h_{K_{i}}(v)-h_{K}(v)|\rightarrow 0,

as i.i\rightarrow\infty. Or if Ki𝒦onK_{i}\in\mathcal{K}^{n}_{o},

maxuSn1|ρKi(u)ρK(u)|0,\max_{u\in S^{n-1}}|\rho_{K_{i}}(u)-\rho_{K}(u)|\rightarrow 0,

as i,i\rightarrow\infty, then KiKK_{i}\rightarrow K.

Let hC+(Sn1)h\in C^{+}(S^{n-1}), the Wulff shape [h][h] generated by hh is a convex body defined by

[h]=vSn1{xn:xvh(v)}.[h]=\bigcap\limits_{v\in S^{n-1}}\left\{x\in\mathbb{R}^{n}:x\cdot v\leq h(v)\right\}.

Obviously, if K𝒦onK\in\mathcal{K}^{n}_{o},

[hK]=K.[h_{K}]=K.

Let ρC+(Sn1)\rho\in C^{+}(S^{n-1}), the convex hull ρ\langle\rho\rangle generated by ρ\rho is a convex body defined by

ρ=conv{ρ(u)u:uSn1}.\langle\rho\rangle=\operatorname{conv}\{\rho(u)u:u\in S^{n-1}\}.

Clearly, if K𝒦onK\in\mathcal{K}^{n}_{o},

ρK=K.\langle\rho_{K}\rangle=K.

For K,L𝒦nK,L\subset\mathcal{K}^{n} and real a,b>0a,b>0, the Minkowski combination, aK+bL𝒦naK+bL\subset\mathcal{K}^{n} defined as follow

aK+bL={ax+by:xK and yL},aK+bL=\{ax+by:\text{$x\in K$ and $y\in L$}\},

and support function is

haK+bL=ahK+bhL.h_{aK+bL}=ah_{K}+bh_{L}.

Firey was the first to define the LpL_{p}-combination of K,L𝒦onK,L\in\mathcal{K}^{n}_{o}, for a,b>0,a,b>0, p0,p\neq 0,

aK+pbL=vSn1{xn:xv(ahK(v)p+bhL(v)p)1p},a\cdot K+_{p}b\cdot L=\bigcap\limits_{v\in S^{n-1}}\left\{x\in\mathbb{R}^{n}:x\cdot v\leq\left(a{h_{K}(v)}^{p}+b{h_{L}(v)}^{p}\right)^{\frac{1}{p}}\right\},

and

aK+0bL=vSn1{xn:xvhK(v)ahL(v)b}.a\cdot K+_{0}b\cdot L=\bigcap\limits_{v\in S^{n-1}}\left\{x\in\mathbb{R}^{n}:x\cdot v\leq{h_{K}(v)}^{a}{h_{L}(v)}^{b}\right\}.

Here aK=a1pKa\cdot K=a^{\frac{1}{p}}K.

In case of p1,p\geq 1, (ahK(v)p+bhL(v)p)1p\left(a{h_{K}(v)}^{p}+b{h_{L}(v)}^{p}\right)^{\frac{1}{p}} is just the support function, but for 0<p<10<p<1 is not necessarily.

For K𝒦onK\in\mathcal{K}^{n}_{o}, its polar body K𝒦onK^{*}\in\mathcal{K}^{n}_{o} in n\mathbb{R}^{n} is defined by

K={xn:xy1,for all yK}.K^{*}=\{x\in\mathbb{R}^{n}:x\cdot y\leq 1,\ \text{for all }\ y\in K\}.

It is easy to show that, if K𝒦onK\in\mathcal{K}^{n}_{o},

ρK=1/hKandhK=1/ρK\rho_{K}=1/h_{K^{*}}\quad\text{and}\quad h_{K}=1/\rho_{K^{*}} (2.1)

for xn{0}x\in\mathbb{R}^{n}\setminus\{0\}. In addition, we can get (K)=K(K^{*})^{*}=K if K𝒦onK\in\mathcal{K}^{n}_{o}.

For K,Ki𝒦onK,K_{i}\in\mathcal{K}^{n}_{o}, by (2.1) and the previous characterization of convex body convergence, then KiKK_{i}\rightarrow K if and only if KiK.K_{i}^{*}\rightarrow K^{*}.

For convex body KnK\in\mathbb{R}^{n}, then its supporting hyperplane with outward unit normal vector vSn1v\in S^{n-1} represented by

HK(v)={xn:xv=hK(v)}.H_{K}(v)=\{x\in\mathbb{R}^{n}:x\cdot v=h_{K}(v)\}.

The boundary point of KK only has one supporting hyperplane called regularity point, otherwise, it is a singular point. The set of singular points denoted as σK\sigma_{K}, it is well known that σK\sigma_{K} has spherical Lebesgue measure 0 (see Schneider [44], p.84).

For xKσKx\in{\partial K\setminus\sigma_{K}}, its Gauss map νK\nu_{K} is represented by

νK(x)={vSn1:xv=hK(v)}.\nu_{K}(x)=\{v\in S^{n-1}:x\cdot v=h_{K}(v)\}.

Correspondingly, for Borel set ηSn1\eta\subset S^{n-1}, its reverse Gauss map denoted by νK1\nu_{K}^{-1},

νK1(η)={xK:νK(x)η}.\nu_{K}^{-1}(\eta)=\{x\in\partial K:\nu_{K}(x)\in\eta\}.

For Borel set ηSn1\eta\subset S^{n-1}, its surface area measure is defined as

SK(η)=n1(νK1(η)),S_{K}(\eta)=\mathcal{H}^{n-1}(\nu_{K}^{-1}(\eta)),

where n1\mathcal{H}^{n-1} is (n1)(n-1)-dimensional Hausdorff measure.

Then gC(Sn1),g\in C(S^{n-1}),

KσKg(νK(x))𝑑n1(x)=Sn1g(v)𝑑SK(v).\int_{\partial K\setminus\sigma_{K}}g(\nu_{K}(x))\,d\mathcal{H}^{n-1}(x)=\int_{S^{n-1}}g(v)\,dS_{K}(v). (2.2)

For Borel set ωSn1\omega\subset S^{n-1}, 𝜶K(ω)\boldsymbol{\alpha}_{K}(\omega) denotes its radial Gauss image and is defined as

𝜶K(ω)={vSn1:ρK(u)(uv)=hK(v)},\boldsymbol{\alpha}_{K}(\omega)=\{v\in S^{n-1}:\rho_{K}(u)(u\cdot v)=h_{K}(v)\},

for uωu\in\omega. When Borel set ω\omega have only one element uu, we will abbreviate 𝜶K({u})\boldsymbol{\alpha}_{K}(\{u\}) as 𝜶K(u)\boldsymbol{\alpha}_{K}(u). The subset of Sn1S^{n-1} which make 𝜶K(u)\boldsymbol{\alpha}_{K}(u) contain more than one element denoted by ωK\omega_{K} for each uωu\in\omega. The set ωK\omega_{K} has spherical Lebesgue measure 0 (see Schneider [44], Theorem 2.2.5).

The radial Guass map of KK is a map denoted by αK(u)\alpha_{K}(u), the only difference between αK(u)\alpha_{K}(u) and 𝜶K(u)\boldsymbol{\alpha}_{K}(u) is that the former is defined on Sn1ωKS^{n-1}\setminus\omega_{K} not on Sn1S^{n-1} which lead to 𝜶K(u)\boldsymbol{\alpha}_{K}(u) may have many elements but αK(u)\alpha_{K}(u) has only one. In other words, if 𝜶K(u)=({v})\boldsymbol{\alpha}_{K}(u)=(\{v\}), then 𝜶K(u)=αK(u)\boldsymbol{\alpha}_{K}(u)=\alpha_{K}(u).

For n1\mathcal{H}^{n-1}-integrable function g:Kg:\partial K\rightarrow\mathbb{R},

Kg(x)𝑑n1(x)=Sn1g(ρK(u)u)F(u)𝑑u,\int_{\partial K}g(x)d\mathcal{H}^{n-1}(x)=\int_{S^{n-1}}g(\rho_{K}(u)u)F(u)du, (2.3)

where FF is defined n1\mathcal{H}^{n-1}-a.e. on Sn1S^{n-1} by

F(u)=(ρK(u))nhK(αK(u)).F(u)=\frac{(\rho_{K}(u))^{n}}{h_{K}(\alpha_{K}(u))}.

For Borel set ηSn1\eta\subset S^{n-1}, its reverse radial Gauss image 𝜶K(η)\boldsymbol{\alpha}^{*}_{K}(\eta) is shown as

𝜶K(η)={uSn1:ρK(u)(uv)=hK(v)}\boldsymbol{\alpha}^{*}_{K}(\eta)=\{u\in S^{n-1}:\rho_{K}(u)(u\cdot v)=h_{K}(v)\}

for vηv\in\eta.

When Borel set η\eta have only one element vv, we will abbreviate 𝜶K({v})\boldsymbol{\alpha}^{*}_{K}(\{v\}) as 𝜶K(v)\boldsymbol{\alpha}^{*}_{K}(v). The subset of Sn1S^{n-1} which make 𝜶K(v)\boldsymbol{\alpha}^{*}_{K}(v) contain more than one element denoted by ηK\eta_{K} for each vηv\in\eta. The spherical Lebesgue measure of set ηK\eta_{K} is 0 (see Schneider [44], Theorem 2.2.11).

The reverse radial Gauss map of KK is a map denoted by αK(v)\alpha^{*}_{K}(v), if 𝜶K(v)=({u})\boldsymbol{\alpha}^{*}_{K}(v)=(\{u\}), then 𝜶K(v)=αK(v)\boldsymbol{\alpha}^{*}_{K}(v)=\alpha^{*}_{K}(v).

For this part, we give the definition of the LpL_{p}-Gaussian surface area measure as follows.

Definition 2.1.

Let K𝒦onK\in\mathcal{K}^{n}_{o}, for Borel set ηSn1\eta\in S^{n-1}, pp\in\mathbb{R}, define LpL_{p}-Gaussian surface area as below

Sp,γn,K(η)=1(2π)nνK1(η)e|x|22(xνK(x))1p𝑑n1(x),S_{p,\gamma_{n},K}(\eta)=\frac{1}{(\sqrt{2\pi})^{n}}\int_{\nu_{K}^{-1}(\eta)}e^{-\frac{|x|^{2}}{2}}(x\cdot\nu_{K}(x))^{1-p}d\mathcal{H}^{n-1}(x),

which is a Borel measure on Sn1S^{n-1}.

The above definition is deduced from the LpL_{p}-variational formula, and the proof of variational formula requires the following lemma. Our proof mainly relies on ideas developed in [29].

Lemma 2.2.

For p0,p\neq 0, let K𝒦onK\in\mathcal{K}^{n}_{o}, f:Sn1f:S^{n-1}\rightarrow\mathbb{R} be a continuous function. For enough small δ>0\delta>0, and each t(δ,δ)t\in(-\delta,\delta), define the continuous function ht:Sn1(0,)h_{t}:S^{n-1}\rightarrow(0,\infty) as

ht(v)=(hK(v)p+tf(v)p)1pvSn1.h_{t}(v)=(h_{K}(v)^{p}+tf(v)^{p})^{\frac{1}{p}}\quad v\in S^{n-1}.

Then,

limt0ρ[ht](u)ρK(u)t=f(αK(u))pphK(αK(u))pρK(u)\lim_{t\rightarrow 0}\frac{\rho_{[h_{t}]}(u)-\rho_{K}(u)}{t}=\frac{f(\alpha_{K}(u))^{p}}{ph_{K}(\alpha_{K}(u))^{p}}\rho_{K}(u)

holds for almost all uSn1u\in S^{n-1}. In addition, there exists M>0M>0 such that

|ρ[ht](u)ρK(u)|M|t|,|\rho_{[h_{t}]}(u)-\rho_{K}(u)|\leq M|t|,

for all uSn1u\in S^{n-1} and t(δ,δ)t\in(-\delta,\delta).

Proof.

Since

ht(v)=(hK(v)p+tf(v)p)1ph_{t}(v)=(h_{K}(v)^{p}+tf(v)^{p})^{\frac{1}{p}}

is equivalent to

loght(v)=loghK(v)+tf(v)pphK(v)p,\log h_{t}(v)=\log h_{K}(v)+\frac{tf(v)^{p}}{p{h_{K}(v)}^{p}},

the outcome is an easy application of the Lemmas 2.8 and 4.1 in [29]. ∎

The following LpL_{p}-variational formula, as we shall see in the next section, is the key to solve the LpL_{p}-Gaussian Minkowski problem.

Theorem 2.3 (LpL_{p}-variational formula in Gaussian space ).

For p0,p\neq 0, let K𝒦onK\in\mathcal{K}^{n}_{o}, f:Sn1f:S^{n-1}\rightarrow\mathbb{R} be continuous function. For sufficiently small δ>0\delta>0, and each t(δ,δ)t\in(-\delta,\delta), define the continuous function ht:Sn1(0,)h_{t}:S^{n-1}\rightarrow(0,\infty) by

ht(v)=(hK(v)p+tf(v)p)1pvSn1.h_{t}(v)=(h_{K}(v)^{p}+tf(v)^{p})^{\frac{1}{p}}\quad v\in S^{n-1}.

Then,

limt0γn([ht])γn(K)t=1pSn1f(v)p𝑑Sp,γn,K(v).\lim_{t\rightarrow 0}\frac{\gamma_{n}\left([h_{t}]\right)-\gamma_{n}\left(K\right)}{t}=\frac{1}{p}\int_{S^{n-1}}f(v)^{p}dS_{p,\gamma_{n},K}(v). (2.4)
Proof.

Applying polar coordinates, shows that

γn([ht])=1(2π)nSn10ρ[ht](u)er22rn1𝑑r𝑑u.\gamma_{n}([h_{t}])=\frac{1}{(\sqrt{2\pi})^{n}}\int_{S^{n-1}}\int_{0}^{\rho_{[h_{t}]}(u)}e^{\frac{-r^{2}}{2}}r^{n-1}drdu.

Since K𝒦onK\in\mathcal{K}^{n}_{o} and hthKh_{t}\rightarrow h_{K} uniformly as t0t\rightarrow 0, by Aleksandrov’s Convergence Lemma, we can get [ht][hK]=K[h_{t}]\rightarrow[h_{K}]=K, then radial function of [ht][h_{t}] converges to the radial function of KK, i.e., ρ[ht]ρK\rho_{[h_{t}]}\to\rho_{K}, and there exist m0,m1>0m_{0},m_{1}>0 such that ρ[ht],ρK[m0,m1]\rho_{[h_{t}]},\rho_{K}\in[m_{0},m_{1}].

For simplify, denote F(s)=0ser22rn1𝑑rF(s)=\int_{0}^{s}e^{\frac{-r^{2}}{2}}r^{n-1}dr. By mean value theorem, there exists θ[m0,m1]\theta\in[m_{0},m_{1}] such that

|F(ρ[ht](u))F(ρK(u))|=|F(θ)||ρ[ht](u)ρK(u)|M|F(θ)||t|,|F(\rho_{[h_{t}]}(u))-F(\rho_{K}(u))|=|F^{{}^{\prime}}(\theta)||\rho_{[h_{t}]}(u)-\rho_{K}(u)|\leq M|F^{{}^{\prime}}(\theta)||t|,

the MM is arisen as in Lemma 2.2. With the simple calculation,

|F(θ)|=|eθ22θn1|M1|F^{{}^{\prime}}(\theta)|=|e^{\frac{-\theta^{2}}{2}}\theta^{n-1}|\leq M_{1}

for some constant M1>0M_{1}>0 as θ[m0,m1]\theta\in[m_{0},m_{1}].

Therefore,

|F(ρ[ht](u))F(ρK(u))|MM1|t|.|F(\rho_{[h_{t}]}(u))-F(\rho_{K}(u))|\leq MM_{1}|t|.

Together with Lemma 2.2, employing the dominated convergence theorem, it follows that

limt0γn([ht])γn(K)t=1(2π)nSn1f(αK(u))pphK(αK(u))pρK(u)neρK(u)22𝑑u=1p(2π)nKf(νK(x))pe|x|22(xνK(x))1p𝑑n1(x)=1pSn1f(v)p𝑑Sp,γn,K(v).\begin{split}\lim_{t\rightarrow 0}\frac{\gamma_{n}\left([h_{t}]\right)-\gamma_{n}\left(K\right)}{t}&=\frac{1}{(\sqrt{2\pi})^{n}}\int_{S^{n-1}}\frac{f(\alpha_{K}(u))^{p}}{ph_{K}(\alpha_{K}(u))^{p}}\rho_{K}(u)^{n}e^{\frac{-\rho_{K}(u)^{2}}{2}}du\\ &=\frac{1}{p(\sqrt{2\pi})^{n}}\int_{\partial K}f(\nu_{K}(x))^{p}e^{\frac{-|x|^{2}}{2}}(x\cdot\nu_{K}(x))^{1-p}d\mathcal{H}^{n-1}(x)\\ &=\frac{1}{p}\int_{S^{n-1}}f(v)^{p}dS_{p,\gamma_{n},K}(v).\end{split}

By the definition of Sp,γn,KS_{p,\gamma_{n},K}, together with the conclusions that the Gauss surface area measure is weakly convergent and is absolutely continuous which were got in [30], it sufficient to reveal that Sp,γn,KS_{p,\gamma_{n},K} is weakly convergent with respect Hausdorff metric, and is absolutely continuous with respect to surface area measure.

3. The variational method to generate solution

This section aims to transform the LpL_{p}-Gaussian Minkowski problem into an optimization problem by employing variational method, and prove the optimizer is just the solution to the Minkowski problem of Sp,γn,KS_{p,\gamma_{n},K}.

3.1. An associated optimization problem

For any nonzero finite Borel measure μ\mu on Sn1S^{n-1}, p>0,p>0, define ϕ:𝒦on\phi:\mathcal{K}^{n}_{o}\rightarrow\mathbb{R} by

ϕ(Q)=Sn1hQ(v)p𝑑μ(v),\phi(Q)=\int_{S^{n-1}}h_{Q}(v)^{p}d\mu(v),

for each Q𝒦onQ\in\mathcal{K}^{n}_{o}. We take into account the following minimum problem

min{ϕ(Q):Q𝒦on ,γn(Q)=12}.\min\left\{\phi(Q):\text{$Q\in\mathcal{K}^{n}_{o}$ },\gamma_{n}(Q)=\frac{1}{2}\right\}. (3.1)
Lemma 3.1.

If K𝒦onK\in\mathcal{K}^{n}_{o} and satisfies

ϕ(K)=min{ϕ(Q):Q𝒦on ,γn(Q)=12},\phi(K)=\min\left\{\phi(Q):\text{$Q\in\mathcal{K}^{n}_{o}$ },\gamma_{n}(Q)=\frac{1}{2}\right\},

equivalent to

φ(hK)=min{φ(z):zC+(Sn1) ,γn([z])=12},\varphi(h_{K})=\min\left\{\varphi(z):\text{$z\in C^{+}(S^{n-1})$ },\gamma_{n}([z])=\frac{1}{2}\right\},

where φ:C+(Sn1)\varphi:C^{+}(S^{n-1})\rightarrow\mathbb{R} is defined by

φ(z)=Sn1z(v)p𝑑μ(v).\varphi(z)=\int_{S^{n-1}}z(v)^{p}d\mu(v).
Proof.

For the Wulff shape

[z]=vSn1{xn:xvz(v)},[z]=\bigcap\limits_{v\in S^{n-1}}\left\{x\in\mathbb{R}^{n}:x\cdot v\leq z(v)\right\},

it is easy to get h[z]zh_{[z]}\leq z and [h[z]]=[z][h_{[z]}]=[z], thus we have

φ(z)φ(h[z]).\varphi(z)\geq\varphi(h_{[z]}).

Clearly, K𝒦onK\in\mathcal{K}^{n}_{o} and satisfies

ϕ(K)=min{ϕ(Q):Q𝒦on ,γn(Q)=12}\phi(K)=\min\left\{\phi(Q):\text{$Q\in\mathcal{K}^{n}_{o}$ },\gamma_{n}(Q)=\frac{1}{2}\right\}

if and only if

φ(hK)=min{φ(z):zC+(Sn1) ,γn([z])=12}.\varphi(h_{K})=\min\left\{\varphi(z):\text{$z\in C^{+}(S^{n-1})$ },\gamma_{n}([z])=\frac{1}{2}\right\}.

Lemma 3.2.

For p>0,p>0, let μ\mu be a nonzero finite Borel measure on Sn1S^{n-1}. If K𝒦onK\in\mathcal{K}^{n}_{o} and satisfies

ϕ(K)=min{ϕ(Q):Q𝒦on ,γn(Q)=12},\phi(K)=\min\left\{\phi(Q):\text{$Q\in\mathcal{K}^{n}_{o}$ },\gamma_{n}(Q)=\frac{1}{2}\right\}, (3.2)

then there exists a constant λ>0\lambda>0 such that μ=λSp,γn,Kp\mu=\frac{\lambda S_{p,\gamma_{n},K}}{p}.

Proof.

By (3.2) and Lemma 3.1, we get

φ(hK)=min{φ(z):zC+(Sn1) ,γn([z])=12}.\varphi(h_{K})=\min\left\{\varphi(z):\text{$z\in C^{+}(S^{n-1})$ },\gamma_{n}([z])=\frac{1}{2}\right\}. (3.3)

For any fC(Sn1)f\in C(S^{n-1}) and t(δ,δ)t\in(-\delta,\delta) where δ>0\delta>0 is sufficiently small, let

ht(v)=(hK(v)p+tf(v)p)1p.h_{t}(v)=(h_{K}(v)^{p}+tf(v)^{p})^{\frac{1}{p}}.

(3.3) implies that hKh_{K} being a minimizer, there exists a constant λ>0\lambda>0 such that

ddt|t=0φ(ht)=λddt|t=0γn([ht]).\left.\frac{d}{dt}\right|_{t=0}\varphi(h_{t})=\lambda\left.\frac{d}{dt}\right|_{t=0}\gamma_{n}([h_{t}]).

Moreover, from the LpL_{p}-variational formula (2.4), we have

Sn1f(v)p𝑑μ(v)=λpSn1f(v)p𝑑Sp,γn,K(v).\int_{S^{n-1}}f(v)^{p}d\mu(v)=\frac{\lambda}{p}\int_{S^{n-1}}f(v)^{p}dS_{p,\gamma_{n},K}(v).

By the arbitrariness of ff, then μ=λpSp,γn,K\mu=\frac{\lambda}{p}S_{p,\gamma_{n},K}. ∎

3.2. Existence of an optimizer

Theorem 3.3.

For p>0,p>0, let μ\mu be a nonzero finite Borel measure on Sn1S^{n-1} and be not concentrated on any closed hemisphere. Then there exists a K𝒦onK\in\mathcal{K}^{n}_{o} such that

ϕ(K)=min{ϕ(Q):Q𝒦on ,γn(Q)=12}.\phi(K)=\min\left\{\phi(Q):\text{$Q\in\mathcal{K}^{n}_{o}$ },\gamma_{n}(Q)=\frac{1}{2}\right\}.
Proof.

Suppose Ql𝒦on{Q_{l}}\subset\mathcal{K}^{n}_{o} is a minimal sequence, i.e.,

limlϕ(Ql)=min{ϕ(Q):Q𝒦on ,γn(Q)=12}.\lim_{l\rightarrow\infty}\phi(Q_{l})=\min\left\{\phi(Q):\text{$Q\in\mathcal{K}^{n}_{o}$ },\gamma_{n}(Q)=\frac{1}{2}\right\}. (3.4)

We now claim that QlQ_{l} is uniformly bounded. If not, then there exists ulSn1u_{l}\in S^{n-1} such that ρQl(ul)\rho_{Q_{l}}(u_{l})\rightarrow\infty as ll\rightarrow\infty. By the definition of support function, ρQl(ul)(ulv)+hQl(v)\rho_{Q_{l}}(u_{l})(u_{l}\cdot v)_{+}\leq h_{Q_{l}}(v), then

ϕ(Ql)=Sn1hQl(v)p𝑑μ(v)Sn1ρQl(ul)p(ulv)+p𝑑μ(v)=ρQl(ul)pSn1(ulv)+p𝑑μ(v),\begin{split}\phi(Q_{l})&=\int_{S^{n-1}}h_{Q_{l}}(v)^{p}d\mu(v)\\ &\geq\int_{S^{n-1}}\rho_{Q_{l}}(u_{l})^{p}{(u_{l}\cdot v)_{+}}^{p}d\mu(v)\\ &=\rho_{Q_{l}}(u_{l})^{p}\int_{S^{n-1}}{(u_{l}\cdot v)_{+}}^{p}d\mu(v),\end{split}

where (t)+=max{t,0(t)_{+}=\max\{{t,0}} for any tt\in\mathbb{R}. Since μ\mu is not concentrated on any closed hemisphere, there exists a positive constant c0c_{0} such that

Sn1(ulv)+p𝑑μ(v)>c0.\int_{S^{n-1}}{(u_{l}\cdot v)_{+}}^{p}d\mu(v)>c_{0}.

Therefore,

ϕ(Ql)>ρQl(ul)pc0\phi(Q_{l})>\rho_{Q_{l}}(u_{l})^{p}c_{0}\rightarrow\infty

as l.l\rightarrow\infty. But this is contradicted to (3.4). Then we conclude QlQ_{l} is uniformly bounded. By Blaschke selection theorem, QlQ_{l} has convergent subsequence, still denoted by QlQ_{l}, converges to a compact convex set KK of n\mathbb{R}^{n}. By the continuity of Gaussian volume, we get

limlγn(Ql)=γn(K)=12.\lim_{l\rightarrow\infty}\gamma_{n}(Q_{l})=\gamma_{n}(K)=\frac{1}{2}.

Now we prove the uniform lower bound of support function by contradiction. For Ki𝒦onKK_{i}\in\mathcal{K}^{n}_{o}\rightarrow K with γn(Ki)=12\gamma_{n}(K_{i})=\frac{1}{2}. Suppose that there exists viSn1v_{i}\in S^{n-1} such that hKi(vi)0h_{K_{i}}(v_{i})\rightarrow 0. Then for any ϵ>0,\epsilon>0, viSn1v_{i}\in S^{n-1}, Ki{x|xviϵ}K_{i}\subset\left\{x|x\cdot v_{i}\geq-\epsilon\right\}. Combining the fact that hKi(vi)h_{K_{i}}(v_{i}) has upper bound, then KiBR{x|xviϵ}K_{i}\subset B_{R}\cap\left\{x|x\cdot v_{i}\geq-\epsilon\right\}. Let HH be halfspace, as computed in [2],

γn(n)=1(2π)nne|x|22𝑑x=1(2π)n×12×2π×(2π)n+1=1.\gamma_{n}(\mathbb{R}^{n})=\frac{1}{(\sqrt{2\pi})^{n}}\int_{\mathbb{R}^{n}}e^{\frac{-|x|^{2}}{2}}dx=\frac{1}{(\sqrt{2\pi})^{n}}\times\frac{1}{2}\times\sqrt{\frac{2}{\pi}}\times(\sqrt{2\pi})^{n+1}=1.

Then

12=1(2π)nHe|x|22𝑑x=1(2π)nBR2e|x|22𝑑x+1(2π)nHBR2e|x|22𝑑x,\frac{1}{2}=\frac{1}{(\sqrt{2\pi})^{n}}\int_{H}e^{\frac{-|x|^{2}}{2}}dx=\frac{1}{(\sqrt{2\pi})^{n}}\int_{B_{\frac{R}{2}}}e^{\frac{-|x|^{2}}{2}}dx+\frac{1}{(\sqrt{2\pi})^{n}}\int_{H\setminus B_{\frac{R}{2}}}e^{\frac{-|x|^{2}}{2}}dx,

which shows that for some t0>0t_{0}>0, 1(2π)nBR2e|x|22𝑑x+t0<12\frac{1}{(\sqrt{2\pi})^{n}}\int_{B_{\frac{R}{2}}}e^{\frac{-|x|^{2}}{2}}dx+t_{0}<\frac{1}{2}, and naturally for ϵ\epsilon small enough, there is γn(Ki)γn(BR{x|xviϵ})<12\gamma_{n}(K_{i})\leq\gamma_{n}(B_{R}\cap\left\{x|x\cdot v_{i}\geq-\epsilon\right\})<\frac{1}{2}, this contradicts to the condition γn(Ki)=12\gamma_{n}(K_{i})=\frac{1}{2}. Therefore, we conclude that KK is non-degenerate. Namely, KK is the desired convex body. ∎

Now we return to deal with the origin problem.

3.3. Existence of solution to the LpL_{p}-Gaussian Minkowski problem

Combining with Lemma 3.1, Lemma 3.2 and Theorem 3.3, we prove sufficiently the main existence theorem for the LpL_{p}-Gaussian Minkowski problem stated in the introduction.

Theorem 3.4.

For p>0,p>0, let μ\mu be a nonzero finite Borel measure on Sn1S^{n-1} and be not concentrated in any closed hemisphere. There exist K𝒦onK\in\mathcal{K}^{n}_{o} and constant λ>0\lambda>0 such that

μ=λpSp,γn,K.\mu=\frac{\lambda}{p}S_{p,\gamma_{n},K}.

The existence of a solution has been done as above, now we are devoted to solving the uniqueness.

4. Uniqueness of solution

In general, Brunn-Minkowski inequality implies the uniqueness of solution of Minkowski problem. There are many fruitful results on the inequality of Gaussian volume γn\gamma_{n}, such as [7, 15, 23]. In particular, Ehrhard inequality in [14] is one of the most significant Brunn-Minkowski type inequalities for Gaussian measure γn\gamma_{n}, stated as follows. But in Gaussian Minkowski problem, because there is no homogeneity, the uniqueness can not obtained from the Brunn-Minkowski inequality. Fortunately, by using Ehrhard inequality, we can get the uniqueness result when the Gaussian volume is more than or equal to half.

Theorem 4.1 (Ehrhard inequality).

Let K,LK,L be convex bodies in n\mathbb{R}^{n}, 0λ10\leq\lambda\leq 1, then

Φ1(γn(λL+(1λ)K))λΦ1(γn(L))+(1λ)Φ1(γn(K)),\Phi^{-1}(\gamma_{n}(\lambda L+(1-\lambda)K))\geq\lambda\Phi^{-1}(\gamma_{n}(L))+(1-\lambda)\Phi^{-1}(\gamma_{n}(K)),

with equality holds if and only if K=L.K=L. Where Φ(x)=12πxet22𝑑t\Phi(x)=\frac{1}{\sqrt{2\pi}}\int_{-\infty}^{x}e^{\frac{-t^{2}}{2}}dt.

In view of Ehrhard inequality, we obtain the log-concave property of γn\gamma_{n} as below.

Lemma 4.2.

Let K,LK,L be convex bodies in n\mathbb{R}^{n}, 0λ10\leq\lambda\leq 1, then

γn(λL+(1λ)K)γn(L)λγn(K)1λ\gamma_{n}(\lambda L+(1-\lambda)K)\geq{\gamma_{n}(L)}^{\lambda}{\gamma_{n}(K)}^{1-\lambda} (4.1)

with equality holds if and only if K=L.K=L.

Indeed, this follows directly from the above Lemma and the fact that λL+p(1λ)KλL+p(1λ)K\lambda L+_{p}(1-\lambda)\cdot K\supset\lambda L+_{p^{\prime}}(1-\lambda)\cdot K for p>pp>p^{\prime}.

Lemma 4.3.

Let K,LK,L be convex bodies in n\mathbb{R}^{n}, for p1p\geq 1 and 0λ10\leq\lambda\leq 1, then

γn(λL+p(1λ)K)γn(L)λγn(K)1λ\gamma_{n}(\lambda L+_{p}(1-\lambda)\cdot K)\geq{\gamma_{n}(L)}^{\lambda}{\gamma_{n}(K)}^{1-\lambda} (4.2)

with equality holds if and only if K=L.K=L. And its differential equivalent form is

1pSn1hLp𝑑Sp,γn,K1pSn1hKp𝑑Sp,γn,K+γn(K)logγn(L)γn(K).\frac{1}{p}\int_{S^{n-1}}{h_{L}}^{p}dS_{p,\gamma_{n},K}\geq\frac{1}{p}\int_{S^{n-1}}{h_{K}}^{p}dS_{p,\gamma_{n},K}+\gamma_{n}(K)\log\frac{\gamma_{n}(L)}{\gamma_{n}(K)}. (4.3)

Naturally, it tells that

Lemma 4.4.

Let K,LK,L be convex bodies in n\mathbb{R}^{n}. For p1p\geq 1, if γn(K)=γn(L),\gamma_{n}(K)=\gamma_{n}(L), then

Sn1hLp𝑑Sp,γn,KSn1hKp𝑑Sp,γn,K\int_{S^{n-1}}{h_{L}}^{p}dS_{p,\gamma_{n},K}\geq\int_{S^{n-1}}{h_{K}}^{p}dS_{p,\gamma_{n},K}

with equality holds if and only if K=L.K=L.

Finally, we attempt to deal with the uniqueness. The main idea of proof is inspired by Lemma 5.1 in [30].

Lemma 4.5.

For p1p\geq 1, if K,L𝒦onK,L\in\mathcal{K}^{n}_{o} with the same LpL_{p}-Gaussian surface area, i.e.,

Sp,γn,K=Sp,γn,L.S_{p,\gamma_{n},K}=S_{p,\gamma_{n},L}.

If γn(K),γn(L)12,\gamma_{n}(K),\gamma_{n}(L)\geq\frac{1}{2}, then γn(K)=γn(L).\gamma_{n}(K)=\gamma_{n}(L).

Proof.

By the Ehrhard inequality,

Φ1(γn((1t)K+tL))(1t)Φ1(γn(K))+tΦ1(γn(L))\Phi^{-1}\left(\gamma_{n}((1-t)K+tL)\right)\geq(1-t)\Phi^{-1}(\gamma_{n}(K))+t\Phi^{-1}(\gamma_{n}(L))

with equality holds if and only if K=L.K=L. Where

Φ(x)=12πxet22𝑑t.\Phi(x)=\frac{1}{\sqrt{2\pi}}\int_{-\infty}^{x}e^{\frac{-t^{2}}{2}}dt.

For convenience, we write Ψ=Φ1\Psi=\Phi^{-1}, then

Ψ(γn((1t)K+tL))(1t)Ψ(γn(K))+tΨ(γn(L)).\Psi(\gamma_{n}((1-t)K+tL))\geq(1-t)\Psi(\gamma_{n}(K))+t\Psi(\gamma_{n}(L)).

For p1,p\geq 1, (1t)K+tL(1t)K+ptL(1-t)K+tL\subset(1-t)\cdot K+_{p}t\cdot L, together with the fact that Ψ\Psi is CC^{\infty} and strictly monotonically increasing, it follows that

Ψ(γn((1t)K+ptL))Ψ(γn((1t)K+tL))(1t)Ψ(γn(K))+tΨ(γn(L)).\Psi\left(\gamma_{n}\left((1-t)\cdot K+_{p}t\cdot L\right)\right)\geq\Psi(\gamma_{n}((1-t)K+tL))\geq(1-t)\Psi(\gamma_{n}(K))+t\Psi(\gamma_{n}(L)).

Then by applying the Theorem 2.3 we have,

Ψ(γn(K))pSn1hLphKpdSp,γn,KΨ(γn(L))Ψ(γn(K)).\frac{\Psi^{\prime}\left(\gamma_{n}(K)\right)}{p}\int_{S^{n-1}}h_{L}^{p}-h_{K}^{p}dS_{p,\gamma_{n},K}\geq\Psi(\gamma_{n}(L))-\Psi(\gamma_{n}(K)). (4.4)

Interchanging the position of KK and LL, we have

Ψ(γn(L))pSn1hKphLpdSp,γn,LΨ(γn(K))Ψ(γn(L)),\frac{\Psi^{\prime}\left(\gamma_{n}(L)\right)}{p}\int_{S^{n-1}}h_{K}^{p}-h_{L}^{p}dS_{p,\gamma_{n},L}\geq\Psi(\gamma_{n}(K))-\Psi(\gamma_{n}(L)),

or equivalently,

Ψ(γn(L))pSn1hLphKpdSp,γn,LΨ(γn(L))Ψ(γn(K)).\frac{\Psi^{\prime}\left(\gamma_{n}(L)\right)}{p}\int_{S^{n-1}}h_{L}^{p}-h_{K}^{p}dS_{p,\gamma_{n},L}\leq\Psi(\gamma_{n}(L))-\Psi(\gamma_{n}(K)). (4.5)

By the condition Sp,γn,K=Sp,γn,LS_{p,\gamma_{n},K}=S_{p,\gamma_{n},L}, Ψ(x)=2πeΨ(x)22>0\Psi^{\prime}(x)=\sqrt{2\pi}e^{\frac{\Psi(x)^{2}}{2}}>0, (4.4) and (4.5), we have

Ψ(γn(L))Ψ(γn(K))Ψ(γn(L))Ψ(γn(L))Ψ(γn(K))Ψ(γn(K)),\frac{\Psi(\gamma_{n}(L))-\Psi(\gamma_{n}(K))}{\Psi^{\prime}(\gamma_{n}(L))}\geq\frac{\Psi(\gamma_{n}(L))-\Psi(\gamma_{n}(K))}{\Psi^{\prime}(\gamma_{n}(K))},

i.e.,

(Ψ(γn(K))Ψ(γn(L)))(Ψ(γn(K))Ψ(γn(L)))0.\left(\Psi^{\prime}(\gamma_{n}(K))-\Psi^{\prime}(\gamma_{n}(L))\right)\left(\Psi(\gamma_{n}(K))-\Psi(\gamma_{n}(L))\right)\leq 0. (4.6)

On the one hand, Ψ′′(x)=2πeΨ(x)22Ψ(x)Ψ(x)>0\Psi^{\prime\prime}(x)=\sqrt{2\pi}e^{\frac{\Psi(x)^{2}}{2}}\Psi(x)\Psi^{\prime}(x)>0 on [12,][\frac{1}{2},\infty], Ψ\Psi^{\prime} is strictly increasing. On the other hand, Ψ\Psi is also strictly increasing, then we conclude

(Ψ(γn(K))Ψ(γn(L)))(Ψ(γn(K))Ψ(γn(L)))0\left(\Psi^{\prime}(\gamma_{n}(K))-\Psi^{\prime}(\gamma_{n}(L))\right)\left(\Psi(\gamma_{n}(K))-\Psi(\gamma_{n}(L))\right)\geq 0 (4.7)

with equality holds if and only if γn(K)=γn(L)\gamma_{n}(K)=\gamma_{n}(L). Combining (4.6) and (4.7), then γn(K)=γn(L).\gamma_{n}(K)=\gamma_{n}(L).

Theorem 4.6.

For p1p\geq 1, if K,L𝒦onK,L\in\mathcal{K}^{n}_{o} with the same LpL_{p}-Gaussian surface area, i.e.,

Sp,γn,K=Sp,γn,L.S_{p,\gamma_{n},K}=S_{p,\gamma_{n},L}.

If γn(K),γn(L)12,\gamma_{n}(K),\gamma_{n}(L)\geq\frac{1}{2}, then K=L.K=L.

Proof.

From Lemma 4.5, we get γn(K)=γn(L).\gamma_{n}(K)=\gamma_{n}(L). Combining the result of Lemma 4.4 with the condition Sp,γn,K=Sp,γn,LS_{p,\gamma_{n},K}=S_{p,\gamma_{n},L}, it sufficient to have

Sn1hLp𝑑Sp,γn,LSn1hKp𝑑Sp,γn,K,\int_{S^{n-1}}{h_{L}}^{p}dS_{p,\gamma_{n},L}\geq\int_{S^{n-1}}{h_{K}}^{p}dS_{p,\gamma_{n},K},

by changing the position of K,LK,L, the equality holds in Lemma 4.4, i.e., K=L.K=L.

The following isoperimetric type inequality can be derived from Ehrhard inequality, and will be used in the Section 5.

Theorem 4.7.

For p1,p\geq 1, let KK be an oo-symmetry convex body in n,\mathbb{R}^{n}, and symmetric strip P={|x1|a}P=\left\{|x_{1}|\leq a\right\} with γn(K)=γn(P)=12\gamma_{n}(K)=\gamma_{n}(P)=\frac{1}{2}. Then,

|Sp,γn,K|2πrpaea22,|S_{p,\gamma_{n},K}|\geq\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}},

where rr is chosen such that γn(rB)=12\gamma_{n}(rB)=\frac{1}{2}.

Proof.

For p1,p\geq 1, applying Ehrhard inequality, Φ1\Phi^{-1} is monotonically increasing, and (1t)K+tL(1t)K+ptL(1-t)K+tL\subset(1-t)\cdot K+_{p}t\cdot L, then,

Φ1(γn((1t)K+ptL))Φ1(γn((1t)K+tL))(1t)Φ1(γn(K))+tΦ1(γn(L)).\begin{split}\Phi^{-1}(\gamma_{n}((1-t)\cdot K+_{p}t\cdot L))&\geq\Phi^{-1}\left(\gamma_{n}((1-t)K+tL)\right)\\ &\geq(1-t)\Phi^{-1}(\gamma_{n}(K))+t\Phi^{-1}(\gamma_{n}(L)).\end{split}

Set L=rBL=rB such that γn(rB)=γn(K)=12\gamma_{n}(rB)=\gamma_{n}(K)=\frac{1}{2}, it follows that

|Sp,γn,K|rpSn1hK𝑑Sγn,K.|S_{p,\gamma_{n},K}|\geq r^{-p}\int_{S^{n-1}}h_{K}dS_{\gamma_{n},K}. (4.8)

By the known result in [36],

γn(K)=γn(P)ddt|t=1γn(tK)ddt|t=1γn(tP),\gamma_{n}(K)=\gamma_{n}(P)\Rightarrow\left.\frac{d}{dt}\right|_{t=1}\gamma_{n}(tK)\geq\left.\frac{d}{dt}\right|_{t=1}\gamma_{n}(tP),

i.e.,

Sn1hK𝑑Sγn,K2πaea22.\int_{S^{n-1}}h_{K}dS_{\gamma_{n},K}\geq\sqrt{\frac{2}{\pi}}ae^{\frac{-a^{2}}{2}}. (4.9)

Combining the (4.8) and (4.9), we have

|Sp,γn,K|2πrpaea22.|S_{p,\gamma_{n},K}|\geq\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}}.

5. Existence of weak solution

In this section, we do some prior estimates and apply degree theory to obtain the existence of oo-symmetry smooth solution, then with the help of approximation argument to get the existence of weak solution of LpL_{p}-Gaussian Minkowski problem for p1p\geq 1.

5.1. prior estimates

Lemma 5.1 (C0C^{0} estimate).

For pp\in\mathbb{R}, K𝒦onK\in\mathcal{K}^{n}_{o} and γn(K)12\gamma_{n}(K)\geq\frac{1}{2}, assume hKC2(Sn1)h_{K}\in C^{2}(S^{n-1}) is the solution of the following equation

1(2π)ne|h|2+h22h1pdet(2h+hI)=f.\frac{1}{(\sqrt{2\pi})^{n}}e^{-\frac{|\nabla h|^{2}+h^{2}}{2}}h^{1-p}\det(\nabla^{2}h+hI)=f.

If 1C1<f<C1\frac{1}{C_{1}}<f<C_{1} for some positive constant C1C_{1}, then there exists a constant C2>0C_{2}>0 such that 1C2<hK<C2\frac{1}{C_{2}}<h_{K}<C_{2}.

Proof.

Now we prove the upper bounded. There exists v0Sn1v_{0}\in S^{n-1} such that hK(v0)=maxvSn1hK(v)h_{K}(v_{0})=\max_{v\in S^{n-1}}h_{K}(v), hK|v0=0\nabla h_{K}|_{v_{0}}=0 and 2hK|v00\nabla^{2}h_{K}|_{v_{0}}\leq 0, then

f(v0)\displaystyle f(v_{0}) =1(2π)n[e|h|2+h22h1pdet(2h+hI)](v0)\displaystyle=\frac{1}{(\sqrt{2\pi})^{n}}[e^{-\frac{|\nabla h|^{2}+h^{2}}{2}}h^{1-p}\det(\nabla^{2}h+hI)](v_{0})
1(2π)nehK(v0)22hK(v0)np,\displaystyle\leq\frac{1}{(\sqrt{2\pi})^{n}}e^{\frac{-h_{K}(v_{0})^{2}}{2}}h_{K}(v_{0})^{n-p},

which implies hK(v0)C2.h_{K}(v_{0})\leq C_{2}.

The bound from blew of support function is guaranteed by the condition that γn(K)12\gamma_{n}(K)\geq\frac{1}{2}. The proof is same as established as in Theorem 3.3. ∎

Lemma 5.2.

For p1,p\geq 1, α(0,1)\alpha\in(0,1), assume function fC2,α(Sn1)f\in C^{2,\alpha}(S^{n-1}) and there exists constant C>0C>0 such that |f|C2,α<C|f|_{C_{2,\alpha}}<C and 1C<f<C\frac{1}{C}<f<C. For K𝒦onK\in\mathcal{K}^{n}_{o} with γn(K)>12\gamma_{n}(K)>\frac{1}{2}, if hKC4,α(Sn1)h_{K}\in C^{4,\alpha}(S^{n-1}) and satisfies

1(2π)nh1pe(|h|2+h2)2det(2h+hI)=f,\frac{1}{(\sqrt{2\pi})^{n}}h^{1-p}e^{\frac{-(|\nabla h|^{2}+h^{2})}{2}}\det(\nabla^{2}h+hI)=f, (5.1)

then there exists a positive constant CC^{\prime} which only depends on CC such that the following priori estimates hold:

(1) C1C^{1} estimate: |hK|<C.|\nabla h_{K}|<C^{\prime}.

(2) C2C^{2} estimate: 1CI<(2hK+hKI)<CI\frac{1}{C^{\prime}}I<(\nabla^{2}h_{K}+h_{K}I)<C^{\prime}I and higher estimate |hK|C4,α<C|h_{K}|_{C^{4,\alpha}}<C^{\prime}.

Proof.

Assume M(v0)=maxvSn1(|hK(v)|2+hK(v)2)M(v_{0})=\max_{v\in S^{n-1}}(|\nabla h_{K}(v)|^{2}+h_{K}(v)^{2}). It admits

2hKhKi+2hKlhKli=0,2h_{K}h_{Ki}+2h_{Kl}h_{Kli}=0,

at v0v_{0}, i.e.,

hKl(hKδil+hKli)=0.h_{Kl}(h_{K}\delta_{il}+h_{Kli})=0.

In view of (5.1), it tells that the element of the matrix hKδil+hKli>0,h_{K}\delta_{il}+h_{Kli}>0, we have hK(v0)=0\nabla h_{K}(v_{0})=0, and M(v0)=hK(v0)2<CM(v_{0})=h_{K}(v_{0})^{2}<C^{\prime}, the second inequality is immediately from Lemma 5.1 and CC^{\prime} is only depends on CC.

To get C2C^{2} estimate, the key is to show that the eigenvalues of matrix 2h+hI\nabla^{2}h+hI are bounded from above and below. In other words, the Monge-Ampere equation (5.1) is uniformly elliptic. Thus, an immediate consequence of the standard Evans-Krylov-Safonov theory [21] is |hK|C4,α<C|h_{K}|_{C^{4,\alpha}}<C^{\prime}.

On the one hand, in light of the fact that f,hKf,h_{K} have positive upper and lower bounds, then combined with |hK|<C|\nabla h_{K}|<C^{\prime}, we conclude

det(2hK+hKI)=(2π)nhKp1fe|hK|2+hK22\det(\nabla^{2}h_{K}+h_{K}I)=(\sqrt{2\pi})^{n}{h_{K}}^{p-1}fe^{\frac{|\nabla h_{K}|^{2}+{h_{K}}^{2}}{2}}

also has positive upper and lower bounds.

On the other hand, we end the proof by claiming that the trace of 2hK+hKI\nabla^{2}h_{K}+h_{K}I has an upper bound.

For convenience, denote

H=trace(2hK+hKI)=hK+(n1)hK.H=trace(\nabla^{2}h_{K}+h_{K}I)=\triangle h_{K}+(n-1)h_{K}.

Assume H(v1)=maxvSn1H(v)H(v_{1})=\max_{v\in S^{n-1}}H(v), then H(v1)=0\nabla H(v_{1})=0 and 2H(v1)0\nabla^{2}H(v_{1})\leq 0. We can make the Hessian of hKh_{K}, (hK)ij(h_{K})_{ij}, is diagonal by choosing the suitable local orthogonal frame ei(i=1,2,,n)e_{i}(i=1,2,\ldots,n). Denote ωij=(hK)ij+hKδij\omega_{ij}=(h_{K})_{ij}+h_{K}\delta_{ij}, setting ωij=ωij1\omega^{ij}=\omega_{ij}^{-1}, the inverse matrix of ωij\omega_{ij}.

Then, based on the above conclusions, at v1,v_{1}, we have

0ωijHij=ωiiHii=ωiiωii(n1)2+Hi=1n1ωiiωiiωii(n1)2,0\geq\omega^{ij}H_{ij}=\omega^{ii}H_{ii}=\omega^{ii}\triangle\omega_{ii}-(n-1)^{2}+H\sum_{i=1}^{n-1}\omega^{ii}\geq\omega^{ii}\triangle\omega_{ii}-(n-1)^{2}, (5.2)

the second equality of (5.2) is from the commutator identity [22]:

Hii=ωii(n1)ωii+H.H_{ii}=\triangle\omega_{ii}-(n-1)\omega_{ii}+H.

Next we are going to estimate ωiiωii.\omega^{ii}\triangle\omega_{ii}.

From the equation

1(2π)nhK1pe(|hK|2+hK2)2det(2hK+hKI)=f,\frac{1}{(\sqrt{2\pi})^{n}}h_{K}^{1-p}e^{\frac{-(|\nabla h_{K}|^{2}+h_{K}^{2})}{2}}\det(\nabla^{2}h_{K}+h_{K}I)=f,

i.e.,

logdet(2hK+hKI)=logf+|hK|2+hK22+(p1)loghK+n2log(2π).\log\det(\nabla^{2}h_{K}+h_{K}I)=\log f+\frac{|\nabla h_{K}|^{2}+h_{K}^{2}}{2}+(p-1)\log h_{K}+\frac{n}{2}\log(2\pi).

Carry out differential operation twice on the above formula, we obtain

ωijωijα=(logf)α+hK(hK)α+(hK)l(hK)lα+(p1)(hK)αhK,\omega^{ij}\omega_{ij\alpha}=(\log f)_{\alpha}+h_{K}(h_{K})_{\alpha}+(h_{K})_{l}(h_{K})_{l\alpha}+(p-1)\frac{(h_{K})_{\alpha}}{h_{K}},

and

ωijωijαα+(ωij)α(ωijα)=(logf)αα+hK(hK)αα+((hK)α)2+(hK)lα2+(hK)l(hK)lαα+(p1)hK(hK)αα(hK)α2hK2,\begin{split}&\omega^{ij}\omega_{ij\alpha\alpha}+(\omega^{ij})_{\alpha}(\omega_{ij\alpha})\\ &=(\log f)_{\alpha\alpha}+h_{K}(h_{K})_{\alpha\alpha}+((h_{K})_{\alpha})^{2}+(h_{K})_{l\alpha}^{2}+(h_{K})_{l}(h_{K})_{l\alpha\alpha}\\ &+(p-1)\frac{h_{K}(h_{K})_{\alpha\alpha}-(h_{K})_{\alpha}^{2}}{h_{K}^{2}},\end{split}

it illustrates that

ωijωijαα=(ωij)α(ωijα)+(logf)αα+hK(hK)αα+((hK)α)2+(hK)lα2+(hK)l(hK)lαα+(p1)hK(hK)αα(hK)α2hK2.\begin{split}&\omega^{ij}\omega_{ij\alpha\alpha}=-(\omega^{ij})_{\alpha}(\omega_{ij\alpha})+(\log f)_{\alpha\alpha}+h_{K}(h_{K})_{\alpha\alpha}+((h_{K})_{\alpha})^{2}+(h_{K})_{l\alpha}^{2}\\ &+(h_{K})_{l}(h_{K})_{l\alpha\alpha}+(p-1)\frac{h_{K}(h_{K})_{\alpha\alpha}-(h_{K})_{\alpha}^{2}}{h_{K}^{2}}.\end{split} (5.3)

We estimate the right side of the (5.3).

Since

(ωijωij)α=(1)α=0,(\omega^{ij}\omega_{ij})_{\alpha}=(1)_{\alpha}=0,

i.e.,

(ωij)αωij+ωijωijα=0,(\omega^{ij})_{\alpha}\omega_{ij}+\omega^{ij}\omega_{ij\alpha}=0,

then we have

(ωij)α=ωijωijωijα,(\omega^{ij})_{\alpha}=-\omega^{ij}\omega^{ij}\omega_{ij\alpha},

and

(ωij)α(ωijα)=(ωij)2(ωijα)20.(\omega^{ij})_{\alpha}(\omega_{ij\alpha})=-(\omega^{ij})^{2}(\omega_{ij\alpha})^{2}\leq 0.

From the equation (4.11) in Cheng-Yau[13], we have

i(hK)i(Δ(hK)i)=i[(hK)i(ΔhK)i+(hK)i(hK)i(hK)i(hK)αδiα]=hKH(n1)|hK|2.\begin{split}&\sum_{i}(h_{K})_{i}(\Delta(h_{K})_{i})\\ &=\sum_{i}[(h_{K})_{i}(\Delta h_{K})_{i}+(h_{K})_{i}(h_{K})_{i}-(h_{K})_{i}(h_{K})_{\alpha}\delta_{i\alpha}]\\ &=\nabla h_{K}\cdot\nabla H-(n-1)|\nabla h_{K}|^{2}.\end{split}

Due to H=ΔhK+(n1)hKH=\Delta h_{K}+(n-1)h_{K}, H(v1)=0\nabla H(v_{1})=0, α=1n1hαα2(α=1n1hαα)22=(H(n1)hK)22\sum_{\alpha=1}^{n-1}h_{\alpha\alpha}^{2}\geq\frac{(\sum_{\alpha=1}^{n-1}h_{\alpha\alpha})^{2}}{2}=\frac{(H-(n-1)h_{K})^{2}}{2}, thus, at v1v_{1},

α=1n1ωiiωiiααΔlogf+hK(H(n1)hK)+(H(n1)hK)22(n2)|hK|2+(p1)HhK(p1)(n1)+(1p)|hK|2hK2=H22+[(2n)hK+p1hK]H+Δlogf(n1)2hK22(n2)|hK|2(p1)(n1)(p1)|hK|2hK2.\begin{split}&\sum_{\alpha=1}^{n-1}\omega^{ii}\omega_{ii\alpha\alpha}\\ &\geq\Delta\log f+h_{K}(H-(n-1)h_{K})+\frac{(H-(n-1)h_{K})^{2}}{2}\\ &-(n-2)|\nabla h_{K}|^{2}+(p-1)\frac{H}{h_{K}}-(p-1)(n-1)+(1-p)\frac{|\nabla h_{K}|^{2}}{h_{K}^{2}}\\ &=\frac{H^{2}}{2}+\left[(2-n)h_{K}+\frac{p-1}{h_{K}}\right]H+\Delta\log f-\frac{(n-1)^{2}h_{K}^{2}}{2}\\ &-(n-2)|\nabla h_{K}|^{2}-(p-1)(n-1)-(p-1)\frac{|\nabla h_{K}|^{2}}{h_{K}^{2}}.\end{split} (5.4)

Put this in (5.2), then

0H22+[(2n)hK+p1hK]H+Δlogf(n1)2hK22(n2)|hK|2(p1)(n1)(p1)|hK|2hK2(n1)2.\begin{split}0&\geq\frac{H^{2}}{2}+\left[(2-n)h_{K}+\frac{p-1}{h_{K}}\right]H+\Delta\log f-\frac{(n-1)^{2}h_{K}^{2}}{2}\\ &-(n-2)|\nabla h_{K}|^{2}-(p-1)(n-1)-(p-1)\frac{|\nabla h_{K}|^{2}}{h_{K}^{2}}-(n-1)^{2}.\end{split} (5.5)

Since ff, |f|C2,α|f|_{C^{2,\alpha}}, hKh_{K}, |hK||\nabla h_{K}| are bounded, which implies HC.H\leq C^{\prime}.

Next we focus on obtaining the existence of smooth solution by the degree theory for second-order nonlinear elliptic operators, the reader can conference to the Li [37] for some details. ∎

Theorem 5.3 (Existence and uniqueness of smooth solution).

Let α(0,1)\alpha\in(0,1), for p1p\geq 1, fC2,α(Sn1)f\in C^{2,\alpha}(S^{n-1}) which is positive even function and satisfies |f|L1<2πrpaea22|f|_{L_{1}}<\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}}, the rr and aa are chosen such that γn(rB)=γn(P)=12\gamma_{n}(rB)=\gamma_{n}(P)=\frac{1}{2}, symmetry trip P={xn:|x1|a}P=\left\{x\in\mathbb{R}^{n}:|x_{1}|\leq a\right\}, then there exists a unique C4,αC^{4,\alpha} convex body K𝒦enK\in\mathcal{K}^{n}_{e} with γn(K)>12\gamma_{n}(K)>\frac{1}{2}, its support function hKh_{K} satisfies

1(2π)ne|hK|2+hK22hK1pdet(2hK+hKI)=f.\frac{1}{(\sqrt{2\pi})^{n}}e^{-\frac{|\nabla h_{K}|^{2}+h_{K}^{2}}{2}}h_{K}^{1-p}\det(\nabla^{2}h_{K}+h_{K}I)=f. (5.6)
Proof.

On the one hand, the uniqueness result is guaranteed by Theorem 4.6 .

On the other hand, define F(;t):C4,α(Sn1)C2,α(Sn1)F(;t):C^{4,\alpha}(S^{n-1})\rightarrow C^{2,\alpha}(S^{n-1}) as follows:

F(h;t)=det(2h+hI)(2π)ne|h|2+h22hp1ft,F(h;t)=\det(\nabla^{2}h+hI)-{(\sqrt{2\pi})^{n}}e^{\frac{|\nabla h|^{2}+h^{2}}{2}}h^{p-1}f_{t},

where ft=(1t)c0+tff_{t}=(1-t)c_{0}+tf.

Define OC4,α(Sn1)O\subset C^{4,\alpha}(S^{n-1}) by

O={h:1C<h<C,1CI<(2h+hI)<CI,F(h;t)=0,|h|C4,α<C,γn(h)>12}.\displaystyle O=\left\{h:\frac{1}{C^{\prime}}<h<C^{\prime},\frac{1}{C^{\prime}}I<(\nabla^{2}h+hI)<C^{\prime}I,\right.\left.F(h;t)=0,|h|_{C^{4,\alpha}}<C^{\prime},\gamma_{n}(h)>\frac{1}{2}\right\}.

For hOh\in O, the eigenvalues of its hessian are bounded from above and below, the operator F(;t)F(\cdot;t) is uniformly elliptic on OO for any t[0,1].t\in[0,1].

When f=c0>0(t=0)f=c_{0}>0(t=0) is small enough, applying mean value theorem, it reveals that there exists a unique constant solution hK=r0h_{K}=r_{0} such that γn(K)>12\gamma_{n}(K)>\frac{1}{2}. Since spherical Laplacian has a discrete spectrum, then we can select c0c_{0} suitably such that |c0|L1<2πrpaea22|c_{0}|_{L_{1}}<\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}}, and the operator Lϕ=ΔSn1ϕ+((np)r02)ϕL\phi=\Delta_{S^{n-1}}\phi+((n-p)-{r_{0}}^{2})\phi is invertible.

Now we claim that OO is an open bounded set under the norm ||C4,α|\cdot|_{C^{4,\alpha}}, that is, we need to prove that if hOh\in\partial O, then F(h;t)0F(h;t)\neq 0.

If F(h;t)=0F(h;t)=0, in other words, hh is the solution of

1(2π)ne|h|2+h22h1pdet(2h+hI)=ft.\frac{1}{(\sqrt{2\pi})^{n}}e^{-\frac{|\nabla h|^{2}+h^{2}}{2}}h^{1-p}\det(\nabla^{2}h+hI)=f_{t}.

Since hO,h\in\partial O, then γn([h])=12\gamma_{n}([h])=\frac{1}{2}, from Theorem 4.7, we have |Sp,γn,[h]|2πrpaea22|S_{p,\gamma_{n},[h]}|\geq\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}}, which is contracted to the condition |ft|L1<2πrpaea22|f_{t}|_{L_{1}}<\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}}.

By means of the Proposition 2.2 of Li[37], we conclude that

deg(F(;0),O,0)=deg(F(;1),O,0).deg(F(\cdot;0),O,0)=deg(F(\cdot;1),O,0).

It is clear that if deg(F(;1),O,0)0,deg(F(\cdot;1),O,0)\neq 0, then there exists hOh\in O such that F(h;1)=0F(h;1)=0. Subsequently, we need to claim deg(F(;0),O,0)0.deg(F(\cdot;0),O,0)\neq 0. The Proposition 2.3 and Proposition 2.4 of Li [37] told that if Lr0L_{r_{0}} the linearized operator of FF at r0r_{0} is invertible, then we have

deg(F(;0),O,0)=deg(Lr0(;0),O,0)0.deg(F(\cdot;0),O,0)=deg(L_{r_{0}}(\cdot;0),O,0)\neq 0.

Our final goal is to verify Lr0L_{r_{0}} :C4,α(Sn1)C2,α(Sn1)C^{4,\alpha}(S^{n-1})\rightarrow C^{2,\alpha}(S^{n-1}),

Lr0=r0n2Sn1ϕ+((np)r0n2r0n)ϕ=r0n2(Sn1ϕ+((np)r02)ϕ)\begin{split}L_{r_{0}}&={r_{0}}^{n-2}\triangle_{S^{n-1}}\phi+((n-p){r_{0}}^{n-2}-{r_{0}}^{n})\phi\\ &={r_{0}}^{n-2}(\triangle_{S^{n-1}}\phi+((n-p)-{r_{0}}^{2})\phi)\end{split}

is invertible.

The choice of c0c_{0} ensures the reversibility of Lr0L_{r_{0}}, then we completed the claim. ∎

We via approximation argument to obtain the existence of weak solution of LpL_{p}-Gaussian Minkowski problem as follows.

Lemma 5.4.

If μ\mu is not concentrated on any closed hemisphere, let μi=fidv\mu_{i}=f_{i}dv be a sequence of measure which converges to μ\mu weakly as i.i\rightarrow\infty. Then there exist ϵ0>0,δ0>0,\epsilon_{0}>0,\delta_{0}>0, and N0>0N_{0}>0 such that for any i>N0,eSn1i>N_{0},e\in S^{n-1},

Sn1{v|ve>δ0}fi𝑑v>ϵ0.\int_{S^{n-1}\cap\left\{v|v\cdot e>\delta_{0}\right\}}f_{i}dv>\epsilon_{0}.
Proof.

Argue via contradiction. Take ϵ0=1k,δ0=1k\epsilon_{0}=\frac{1}{k},\delta_{0}=\frac{1}{k}, ekSn1e_{k}\in S^{n-1}, a subsequence which converges to e,e, we conclude that

Sn1{v|vek>1k}fk𝑑v1k.\int_{S^{n-1}\cap\left\{v|v\cdot e_{k}>\frac{1}{k}\right\}}f_{k}dv\leq\frac{1}{k}.

Since fidvμf_{i}dv\rightharpoonup\mu weakly, fix any η>0,\eta>0, for kk large enough,

Sn1{v|ve>η}𝑑μ=0.\int_{S^{n-1}\cap\left\{v|v\cdot e>\eta\right\}}d\mu=0.

Let η0,\eta\rightarrow 0, we have Sn1{v|ve>0}𝑑μ=0\int_{S^{n-1}\cap\left\{v|v\cdot e>0\right\}}d\mu=0, which is contradicted to the condition that μ\mu is not concentrated on any closed hemisphere. ∎

At the end, we are ready to state the Theorem for the existence of unique symmetry weak solution as exhibited in the following.

Theorem 5.5.

For p1p\geq 1, let μ\mu be a finite even Borel measure on Sn1S^{n-1} and be not concentrated in any closed hemisphere with |μ|<2πrpaea22|\mu|<\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}}, the rr and aa are chosen such that γn(rB)=γn(P)=12\gamma_{n}(rB)=\gamma_{n}(P)=\frac{1}{2}, symmetry trip P={xn:|x1|a}P=\left\{x\in\mathbb{R}^{n}:|x_{1}|\leq a\right\}. Then there exists a unique K𝒦enK\in\mathcal{K}^{n}_{e} with γn(K)>12\gamma_{n}(K)>\frac{1}{2} such that

Sp,γn,K=μ.S_{p,\gamma_{n},K}=\mu.
Proof.

Let μi=fidv\mu_{i}=f_{i}dv be a sequence of even measure which converges to μ\mu weakly as i,i\rightarrow\infty, fiC2,α,fi>0f_{i}\in C^{2,\alpha},f_{i}>0 with 0<|fi|L1<2πrpaea220<|f_{i}|_{L_{1}}<\sqrt{\frac{2}{\pi}}r^{-p}ae^{\frac{-a^{2}}{2}}. Then combine Theorem 5.3, there are C4,αC^{4,\alpha} convex bodies Ki𝒦enK_{i}\in\mathcal{K}^{n}_{e} with γn(Ki)>12\gamma_{n}(K_{i})>\frac{1}{2}, its support function satisfies

1(2π)ne|hKi|2+hKi22hKi1pdet(2hKi+hKiI)=fi,\frac{1}{(\sqrt{2\pi})^{n}}e^{-\frac{|\nabla h_{K_{i}}|^{2}+{h_{K_{i}}}^{2}}{2}}{h_{K_{i}}}^{1-p}\det(\nabla^{2}h_{K_{i}}+h_{K_{i}}I)=f_{i},

i.e., Sp,γn,Ki=fidv.S_{p,\gamma_{n},K_{i}}=f_{i}dv.

We will set about claiming that KiK_{i} is uniformly bounded. Suppose that maxvSn1hKi(v)\max_{v\in S^{n-1}}h_{K_{i}}(v) is attained at viSn1v_{i}\in S^{n-1}, and the corresponding point on KK is xi.x_{i}. In fact, |xi|=hKi(vi)|x_{i}|=h_{K_{i}}(v_{i}). Then by Lemma 5.4, there exist ϵ0>0,δ0>0,\epsilon_{0}>0,\delta_{0}>0, and N0>0N_{0}>0 such that for any i>N0,i>N_{0},

Sn1{v|vvi>δ0}fi𝑑v>ϵ0.\int_{S^{n-1}\cap\left\{v|v\cdot v_{i}>\delta_{0}\right\}}f_{i}dv>\epsilon_{0}. (5.7)

Denote Di=Sn1{v|vvi>δ0}D_{i}=S^{n-1}\cap\left\{v|v\cdot v_{i}>\delta_{0}\right\}, the points xx on KiK_{i} corresponding to the vector on DiD_{i} satisfy |x|>hKi(vi)δ0|x|>h_{K_{i}}(v_{i})\delta_{0}. Hence,

Sp,γn,Di=1(2π)nνKi1(Di)e|x|22(xνKi(x))1p𝑑n1(x)1(2π)ne(hKi(vi)δ0)2(hKi(vi)δ0)1phKi(vi)n10,\begin{split}S_{p,\gamma_{n},D_{i}}&=\frac{1}{(\sqrt{2\pi})^{n}}\int_{\nu_{K_{i}}^{-1}(D_{i})}e^{-\frac{|x|^{2}}{2}}(x\cdot\nu_{K_{i}}(x))^{1-p}d\mathcal{H}^{n-1}(x)\\ &\leq\frac{1}{(\sqrt{2\pi})^{n}}e^{-(h_{K_{i}}(v_{i})\delta_{0})^{2}}\left(h_{K_{i}}(v_{i})\delta_{0}\right)^{1-p}{h_{K_{i}}(v_{i})}^{n-1}\rightarrow 0,\end{split}

as hKi(vi).h_{K_{i}}(v_{i})\rightarrow\infty. Which contradicts to (5.7).

In view of the fact that KiK_{i} is uniformly bounded, γn(Ki)>12\gamma_{n}(K_{i})>\frac{1}{2}, it follows that hKi>1Ch_{K_{i}}>\frac{1}{C}. In summary, we have 1C<hKi<C,\frac{1}{C}<h_{K_{i}}<C, KiK𝒦enK_{i}\rightarrow K\in\mathcal{K}^{n}_{e} as i.i\rightarrow\infty. KK is the desired convex body. ∎

Acknowledgement

I would like to thank my supervisor, professor Yong Huang, for his patient guidance and encouragement. I am also deeply indebted to professor Shibing Chen for providing the valuable advice of Lemma 5.1.

References

  • [1] A.D. Aleksandrov, On the theory of mixed volumes. III. Extensions of two theorems of Minkowski on convex polyhedra to arbitrary convex bodies, Mat. Sb. (N.S.) 3 (1938), 27–46.
  • [2] K. Ball, The reverse isoperimetric problem for Gaussian measure, Discrete Comput. Geom. 10 (1993), no. 4, 411–420.
  • [3] K. J. Böröczky, P. Hegedűs and G. Zhu, On the discrete logarithmic Minkowski problem, Int. Math. Res. Not. IMRN 2016, no. 6, 1807–1838.
  • [4] K. J. Böröczky et al., The log-Brunn-Minkowski inequality, Adv. Math. 231 (2012), no. 3-4, 1974–1997.
  • [5] K. J. Böröczky et al., The logarithmic Minkowski problem, J. Amer. Math. Soc. 26 (2013), no. 3, 831–852.
  • [6] K. J. Böröczky et al., The dual Minkowski problem for symmetric convex bodies, Adv. Math. 356 (2019), 106805, 30 pp.
  • [7] C. Borell, The Brunn-Minkowski inequality in Gauss space, Invent. Math. 30 (1975), no. 2, 207–216.
  • [8] S. Chen et al., The LpL_{p}-Brunn-Minkowski inequality for p<1p<1, Adv. Math. 368 (2020), 107166, 21 pp.
  • [9] C. Chen, Y. Huang and Y. Zhao, Smooth solutions to the LpL_{p} dual Minkowski problem, Math. Ann. 373 (2019), no. 3-4, 953–976.
  • [10] S. Chen and Q.-R. Li, On the planar dual Minkowski problem, Adv. Math. 333 (2018), 87–117.
  • [11] S. Chen, Q. Li and G. Zhu, On the LpL_{p} Monge-Ampère equation, J. Differential Equations 263 (2017), no. 8, 4997–5011.
  • [12] K.-S. Chou and X.-J. Wang, The LpL_{p}-Minkowski problem and the Minkowski problem in centroaffine geometry, Adv. Math. 205 (2006), no. 1, 33–83.
  • [13] S. Y. Cheng and S. T. Yau, On the regularity of the solution of the nn-dimensional Minkowski problem, Comm. Pure Appl. Math. 29 (1976), no. 5, 495–516.
  • [14] A. Ehrhard, Symétrisation dans l’espace de Gauss, Math. Scand. 53 (1983), no. 2, 281–301.
  • [15] A. Eskenazis and G. Moschidis, The dimensional Brunn-Minkowski inequality in Gauss space, J. Funct. Anal. 280 (2021), no. 6, 108914.
  • [16] W. Fenchel,B. Jessen, Mengenfunktionen und konvexe Korper. Danske Vid. Selskab. Mat.-Fys. Medd., 16 (1938), 1–31.
  • [17] W. J. Firey, Shapes of worn stones, Mathematika 21 (1974), 1–11.
  • [18] F. Nazarov, On the maximal perimeter of a convex set in n{\mathbb{R}}^{n} with respect to a Gaussian measure, in Geometric aspects of functional analysis, 169–187, Lecture Notes in Math., 1807, Springer, Berlin.
  • [19] R. J. Gardner et al., General volumes in the Orlicz-Brunn-Minkowski theory and a related Minkowski problem I, Calc. Var. Partial Differential Equations 58 (2019), no. 1, Paper No. 12, 35 pp.
  • [20] R. J. Gardner et al., General volumes in the Orlicz-Brunn-Minkowski theory and a related Minkowski problem II, Calc. Var. Partial Differential Equations 59 (2020), no. 1, Paper No. 15, 33 pp.
  • [21] D. Gilbarg and N. S. Trudinger, Elliptic partial differential equations of second order, reprint of the 1998 edition, Classics in Mathematics, Springer-Verlag, Berlin, 2001.
  • [22] P. Guan, X.-N. Ma and F. Zhou, The Christofel-Minkowski problem. III. Existence and convexity of admissible solutions, Comm. Pure Appl. Math. 59 (2006), no. 9, 1352–1376.
  • [23] R. J. Gardner and A. Zvavitch, Gaussian Brunn-Minkowski inequalities, Trans. Amer. Math. Soc. 362 (2010), no. 10, 5333–5353.
  • [24] D. Hug et al., On the LpL_{p} Minkowski problem for polytopes, Discrete Comput. Geom. 33 (2005), no. 4, 699–715.
  • [25] C. Haberl et al., The even Orlicz Minkowski problem, Adv. Math. 224 (2010), no. 6, 2485–2510.
  • [26] H. Minkowski, Volumen und Oberfläche, Math. Ann. 57 (1903), no. 4, 447–495.
  • [27] 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.
  • [28] Y. Huang and Q. Lu, On the regularity of the LpL_{p} Minkowski problem, Adv. in Appl. Math. 50 (2013), no. 2, 268–280.
  • [29] Y. Huang et al., Geometric measures in the dual Brunn-Minkowski theory and their associated Minkowski problems, Acta Math. 216 (2016), no. 2, 325–388.
  • [30] Y. Huang; D. Xi and Y. Zhao. The Minkowski problem in Gaussian probability space. Accpeted by Adv. Math. (2021).
  • [31] Y. Huang and Y. Zhao, On the LpL_{p} dual Minkowski problem, Adv. Math. 332 (2018), 57–84.
  • [32] J.Hosle; A.Kolensnikov and G.Livshyts. On the LpL_{p}-Brunn-Minkowski and dimensional Brunn-Minkowski conjectures for log-concave measures. Arxiv.org/abs/2003.05282,2020.
  • [33] H. Jian, J. Lu and X.-J. Wang, Nonuniqueness of solutions to the LpL_{p}-Minkowski problem, Adv. Math. 281 (2015), 845–856.
  • [34] H. Jian and J. Lu, Existence of solutions to the Orlicz-Minkowski problem, Adv. Math. 344 (2019), 262–288.
  • [35] Y. Liu and J. Lu, A flow method for the dual Orlicz-Minkowski problem, Trans. Amer. Math. Soc. 373 (2020), no. 8, 5833–5853.
  • [36] R. Latała, On some inequalities for Gaussian measures, in Proceedings of the International Congress of Mathematicians, Vol. II (Beijing, 2002), 813–822, Higher Ed. Press, Beijing.
  • [37] Y. Y. Li, Degree theory for second order nonlinear elliptic operators and its applications, Comm. Partial Differential Equations 14 (1989), no. 11, 1541–1578.
  • [38] E. Lutwak, Dual mixed volumes, Pacific J. Math. 58 (1975), no. 2, 531–538.
  • [39] E. Lutwak, The Brunn-Minkowski-Firey theory. I. Mixed volumes and the Minkowski problem, J. Differential Geom. 38 (1993), no. 1, 131–150.
  • [40] E. Lutwak and V. Oliker, On the regularity of solutions to a generalization of the Minkowski problem, J. Differential Geom. 41 (1995), no. 1, 227–246.
  • [41] Q.-R. Li, W. Sheng and X.-J. Wang, Flow by Gauss curvature to the Aleksandrov and dual Minkowski problems, J. Eur. Math. Soc. (JEMS) 22 (2020), no. 3, 893–923.
  • [42] X. Luo, D. Ye and B. Zhu, On the polar Orlicz-Minkowski problems and the pp-capacitary Orlicz-Petty bodies, Indiana Univ. Math. J. 69 (2020), no. 2, 385–420.
  • [43] A. Stancu, The discrete planar L0L_{0}-Minkowski problem, Adv. Math. 167 (2002), no. 1, 160–174.
  • [44] R. Schneider, Convex bodies: the Brunn-Minkowski theory, second expanded edition, Encyclopedia of Mathematics and its Applications, 151, Cambridge University Press, Cambridge, 2014.
  • [45] H. Wang, N. Fang and J. Zhou, Continuity of the solution to the dual Minkowski problem for negative indices, Proc. Amer. Math. Soc. 147 (2019), no. 3, 1299–1312.
  • [46] S. Xing and D. Ye, On the general dual Orlicz-Minkowski problem, Indiana Univ. Math. J. 69 (2020), no. 2, 621–655.
  • [47] G. Zhu, The logarithmic Minkowski problem for polytopes, Adv. Math. 262 (2014), 909–931.
  • [48] G. Zhu, The LpL_{p} Minkowski problem for polytopes for 0<p<10<p<1, J. Funct. Anal. 269 (2015), no. 4, 1070–1094.
  • [49] G. Zhu, Continuity of the solution to the LpL_{p} Minkowski problem, Proc. Amer. Math. Soc. 145 (2017), no. 1, 379–386.
  • [50] Y. Zhao, Existence of solutions to the even dual Minkowski problem, J. Differential Geom. 110 (2018), no. 3, 543–572.
  • [51] D. Zou and G. Xiong, The Orlicz Brunn-Minkowski inequality for the projection body, J. Geom. Anal. 30 (2020), no. 2, 2253–2272.
  • [52] B. Zhu, S. Xing and D. Ye, The dual Orlicz-Minkowski problem, J. Geom. Anal. 28 (2018), no. 4, 3829–3855.