Hostname: page-component-848d4c4894-2pzkn Total loading time: 0 Render date: 2024-06-08T08:37:32.096Z Has data issue: false hasContentIssue false

Anisotropic flow, entropy, and $L^p$-Minkowski problem

Published online by Cambridge University Press:  28 November 2023

Károly J. Böröczky
Affiliation:
Alfréd Rényi Institute of Mathematics, Budapest, Hungary e-mail: carlos@renyi.hu
Pengfei Guan*
Affiliation:
Department of Mathematics and Statistics, McGill University, Montreal, QC H3A 2K6, Canada
Rights & Permissions [Opens in a new window]

Abstract

We provide a natural simple argument using anistropic flows to prove the existence of weak solutions to Lutwak’s $L^p$-Minkowski problem on $S^n$ which were obtained by other methods.

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of The Canadian Mathematical Society

1 Introduction

For $\alpha>0$ and nonnegative $f\in L^1(\mathbb {S}^n)$ with positive integral, we are interested in finding a weak solution to the Monge–Ampére equation

(1.1) $$ \begin{align} u^{\frac1\alpha}\det(\bar{\nabla}^2_{ij} u+u\bar{g}_{ij})=f, \end{align} $$

or in other words, a weak solution to Lutwak’s $L^p$ -Minkowski problem on $S^n$ when $-n-1<p<1$ for $p=1-\frac 1\alpha $ where $\bar {\nabla }$ is the Levi-Civita connection of $\mathbb {S}^n$ , $\bar {g}_{ij}$ , with $\bar {g}$ being the induced round metric on the unit sphere. By a weak (Alexandrov) solution, we mean the following: Given a nontrivial finite Borel measure $\mu $ on $\mathbb {S}^n$ (for example, $d\mu =f\,d\theta $ for the Lebesgue measure $\theta $ on $S^n$ and the f in (1.1)), find a convex body $\Omega \subset \Bbb R^{n+1}$ with $o\in \Omega $ such that

(1.2) $$ \begin{align} d\mu=u^{\frac1\alpha}\,dS_\Omega, \end{align} $$

where $u(x)=\max _{z\in \Omega }\langle x,z\rangle $ is the support function and $S_\Omega $ is the surface area measure of $\Omega $ (see [Reference Schneider45]). If $\partial \Omega $ is $C^2_+$ , then

$$ \begin{align*}dS_\Omega=\det(\bar{\nabla}^2_{ij} u+u\bar{g}_{ij})d\theta=K^{-1}d\theta, \end{align*} $$

where $K(x)$ is the Gaussian curvature at the point of $\partial \Omega $ where $x\in S^n$ is the exterior unit normal (see [Reference Schneider45]). Concerning the regularity of the solution of (1.1), if $f\in C^{0,\beta }(S^n)$ and u are positive, then u is $C^{2,\beta }$ according to Caffarelli’s regularity theory in [Reference Caffarelli15, Reference Caffarelli16]. On the other hand, even if f is positive and continuous for $\alpha>\frac 1n$ , there might exist weak solution where $u(x)=0$ for some $x\in S^n$ and u is not even $C^1$ according to Example 4.2 in [Reference Bianchi, Böröczky and Colesanti7]. Moreover, even if $f\in C^{0,\beta }(S^n)$ is positive, it is possible that $u(x)=0$ for some $x\in S^n$ for $\alpha>\frac 1n$ , but Choi, Kim, and Lee [Reference Choi, Kim and Lee19] still managed to obtain some regularity results in this case.

The case $\alpha =\frac 1{n+2}$ of the Monge–Ampére equation (1.1) is the critical case when the left-hand side of (1.1) is invariant under linear transformations of $\Omega $ , and the case $\alpha =1$ is the so-called logarithmic Minkowski problem posed by Firey [Reference Firey23]. Setting $p=1-\frac 1\alpha <1$ , the Monge–Ampére equation (1.1) is Lutwak’s $L^p$ -Minkowski problem

(1.3) $$ \begin{align} u^{1-p}\det(\bar{\nabla}^2_{ij} u+u\bar{g}_{ij})=f. \end{align} $$

In this notation, (1.2) reads as

(1.4) $$ \begin{align} d\mu=u^{1-p}\,dS_\Omega; \end{align} $$

that equation makes sense for any $p\in \Bbb R$ . Within the rapidly developing $L^p$ -Brunn–Minkowski theory (where $p=1$ is the classical case originating from Minkowski’s oeuvre) initiated by Lutwak [Reference Lutwak39Reference Lutwak41], if $p>1$ and $p\neq n+1$ , then Hug, Lutwak, Yang, and Zhang [Reference Hug, Lutwak, Yang and Zhang30] (improving on Chou and Wang [Reference Chou and Wang20]) prove that (1.4) has an Alexandrov solution if and only if the $\mu $ is not concentrated onto any closed hemisphere, and the solution is unique. We note that there are examples in [Reference Guan and Lin25] (see also [Reference Hug, Lutwak, Yang and Zhang30]) and show that if $1<p<n+1$ , then it may happen that the density function f is a positive continuous in (1.3) and $o\in \partial K$ holds for the unique Alexandrov solution, and actually Bianchi, Böröczky, and Colesanti [Reference Bianchi, Böröczky and Colesanti7] exhibit an example that $o\in \partial K$ even if the density function f is a positive continuous in (1.3) assuming $-n-1<p<1$ .

In the case $p\in (0,1)$ (or equivalently, $\alpha>1$ ), if the measure $\mu $ is not concentrated onto any great subsphere of $S^n$ , then Chen, Li, and Zhu [Reference Chen, Li and Zhu17] prove that there exists an Alexandrov solution $K\in \mathcal {K}_o^n$ of (1.4) using a variational argument (see also [Reference Bianchi, Böröczky, Colesanti and Yang8]). We note that for $p\in (0,1)$ and $n\geq 2$ , no complete characterization of $L^p$ -surface area measures is known (see [Reference Böröczky and Trinh12] for the case $n=1$ , and [Reference Bianchi, Böröczky, Colesanti and Yang8, Reference Saroglou43] for partial results about the case when $n\geq 2$ and the support of $\mu $ is contained in a great subsphere of $S^n$ ).

Concerning the case $p=0$ (or equivalently, $\alpha =1$ ), the still open logarithmic Minkowski problem (1.3) or (1.4) was posed by Firey [Reference Firey23] in 1974. The paper [Reference Böröczky, Lutwak, Yang and Zhang11] characterized even measures $\mu $ such that (1.4) has an even solution for $p=0$ by the so-called subspace concentration condition (see (a) and (b) in Theorem 1.1). In general, Chen, Li, and Zhu [Reference Chen, Li and Zhu18] proved that if a nontrivial finite Borel measure $\mu $ on $S^{n-1}$ satisfies the same subspace concentration condition, then (1.4) has a solution for $p=0$ . On the other hand, Böröczky and Hegedus [Reference Böröczky and Hegedűs10] provide conditions on the restriction of the $\mu $ in (1.4) to a pair of antipodal points.

If $-n-1<p<0$ (or equivalently, $\frac 1{n+2}<\alpha <1$ ) and $f\in L_{\frac {n+1}{n+1+p}}(S^{n})$ in (1.3), then (1.3) has a solution according to [Reference Bianchi, Böröczky, Colesanti and Yang8]. For a rather special discrete measure $\mu $ satisfying that $\mu $ is not concentrated on any closed hemisphere and any n unit vectors in the support of $\mu $ are independent, Zhu [Reference Zhu47] solves the $L^p$ -Minkowski problem (1.4) for $p<0$ . The $p=-n-1$ (or equivalently, $\alpha =\frac 1{n+2}$ ) case of the $L^p$ -Minkowski problem is the critical case because its link with the $\mathrm {SL}(n)$ invariant centro-affine curvature whose reciprocal is $u^{n+2}\det (\bar {\nabla }^2_{ij} u+u\bar {g}_{ij})$ (see [Reference Hug29] or [Reference Ludwig38]). For positive results concerning the critical case $p=-n-1$ , see, for example, [Reference Guang, Li and Wang28, Reference Jian, Lu and Zhu34], and for obstructions for a solution, see, for example, [Reference Chou and Wang20, Reference Du22].

In the super-critical case $p<-n-1$ (or equivalently, $\alpha <\frac 1{n+2}$ ), there is a recent important work by Li, Guang, and Wang [Reference Guang, Li and Wang27] proving that for any positive $C^2$ function f, there exists a $C^4$ solution of (1.3). See also [Reference Du22] for non-existence examples.

The main contribution of this paper is to provide a very natural argument based on anisotropic flows developed by Andrews [Reference Andrews4] to handle the case $-n-1<p<1$ , or equivalently, the case $\frac 1{n+2}<\alpha <\infty $ .

Entropy functional. For any convex body $\Omega $ , a fixed positive function f on $\mathbb {S}^n$ and $\alpha \in (0, \infty )$ , define

(1.5) $$ \begin{align} \mathcal{E}_{\alpha, f} (\Omega) := \sup_{z\in\Omega}{\mathcal E}_{\alpha, f}(\Omega,z), \end{align} $$

where

(1.6) $$ \begin{align} {\mathcal E}_{\alpha, f}(\Omega,z) := \begin{cases} \frac{\alpha}{\alpha-1}\log\left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} u_{z}(x)^{1-\frac1\alpha}\,f(x) d\theta(x)\right),&\alpha\neq 1,\\ \hspace{3pt}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} \log(u_{z}(x))\, f(x) d\theta(x),&\alpha=1. \end{cases} \end{align} $$

Here, $u_{z}(x):=\sup _{y\in \Omega }\left \langle y-z,x\right \rangle $ is the support function of $\Omega $ in direction x with respect to $z_0$ and $\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n} h(x)\, d\theta (x)=\frac {1}{\omega _n} \int _{\mathbb {S}^n} h(x)$ with $\omega _n$ being the surface area of $\mathbb {S}^n$ and $\theta $ is the Lebesgue measure on $S^n$ . When $\alpha =1$ and $f(x)\equiv 1$ , then the above quantity agrees with the entropy in [Reference Guan and Ni26], first introduced by Firey [Reference Firey23] for the centrally symmetric $\Omega $ . General integral quantities were studied by Andrews in [Reference Andrews2, Reference Andrews4]. Here, we shall assume that $\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n} f(x)\, d\theta (x)=1$ , namely, $\frac 1{\omega _n}\,f(x)d\theta (x)$ is a probability measure. For the special case $f\equiv 1$ , $\mathcal {E}_{\alpha , f}(\Omega ) $ becomes the entropy $\mathcal {E}_\alpha (\Omega )$ in [Reference Andrews, Guan and Ni6].

For positive $f\in C^{\infty }(\mathbb S^n)$ , consider the anisotropic flow for convex hypersurfaces $\tilde X(\cdot , \tau ): M_{\tau }\to \mathbb {R}^{n+1}$ :

(1.7) $$ \begin{align} \frac{\partial}{\partial \tau}\tilde X(x, \tau)= -f^{\alpha} (\nu) \tilde K^{\alpha}(x, \tau)\, \nu(x, \tau), \end{align} $$

where $\nu (x, \tau )$ is the unit exterior normal at $\tilde X(x, \tau )$ of $\tilde M_\tau =\tilde X(M, \tau )$ , and $\tilde K(x,\tau )$ is the Gauss curvature of $\tilde M_\tau $ at $\tilde X(x,\tau )$ . Andrews [Reference Andrews4] proved that flow (1.7) contracts to a point under finite time if the initial hypersurface $M_0$ is strictly convex. Under a proper normalization, the normalized anisotropy flow of (1.7) is

(1.8) $$ \begin{align} \frac{\partial}{\partial t}X(x, t)= -\frac{f^\alpha(\nu) K^{\alpha}(x, t)}{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}f^\alpha K^{\alpha-1}}\, \nu(x, t) +X(x,t). \end{align} $$

The basic observation is that a critical point for entropy $\mathcal {E}_{\alpha , f} (\Omega )$ defined in (1.5) under volume normalization is a solution to equation (1.1). The entropy is monotone along flow (1.8). One may view (1.1) is an “optimal solution” to this variational problem as the flow (1.8) provides a natural path to reach it. This approach was devised in [Reference Andrews, Böröczky, Guan and Ni5] with the aim to obtain convergence of the normalized flow (1.8). The main arguments in [Reference Andrews, Böröczky, Guan and Ni5] follows those in [Reference Andrews, Guan and Ni6, Reference Guan and Ni26] where convergence of isotropic flows by power of Gauss curvature (i.e., $f=1$ ) was established. Unfortunately, the entropy point estimate in [Reference Andrews, Guan and Ni6, Reference Guan and Ni26] fails for general anisotropic flows except $\frac {1}{n+2}<\alpha \le \frac {1}n$ [Reference Andrews4]. The convergence was obtained in [Reference Andrews, Böröczky, Guan and Ni5] assuming $M_0$ and f are invariant under a subgroup G of $O(n+1)$ which has no fixed point. We note that an inverse Gauss curvature flow argument was considered by Bryan, Ivaki, and Scheuer [Reference Bryan, Ivaki and Scheuer14] to produce a origin-symmetric solution to (1.1).

Since we are only interested in finding a weak solution to (1.2), one only needs certain “weak” convergence of the flow (1.8). The key steps are to control diameter with entropy under appropriate conditions on measure $\mu =f d\theta $ on $\mathbb S^n$ and use monotonicity of entropy to produce a solution to (1.2). The following is our main result.

Theorem 1.1 For $\alpha>\frac 1{n+2}$ and finite nontrivial Borel measure $\mu $ on $\mathbb {S}^n$ , $n\geq 1$ , there exists a weak solution of (1.2) provided the following holds:

  1. (i) If $\alpha>1$ and $\mu $ is not concentrated onto any great subsphere $x^\bot \cap \mathbb {S}^n$ , $x\in \mathbb {S}^n$ .

  2. (ii) If $\alpha =1$ and $\mu $ satisfies that for any linear $\ell $ -subspace $L\subset \Bbb R^{n+1}$ with $1\leq \ell \leq n$ , we have

    1. (a) $\displaystyle \mu (L\cap \mathbb {S}^n)\leq \frac {\ell }{n+1}\cdot \mu (\mathbb {S}^n)$ ;

    2. (b) equality in (a) for a linear $\ell $ -subspace $L\subset \Bbb R^{n+1}$ with $1\leq d\leq n$ implies the existence of a complementary linear $(n+1-\ell )$ -subspace $\widetilde {L}\subset \Bbb R^{n+1}$ such that $\mathrm {supp}\,\mu \subset L\cup \widetilde {L}$ .

  3. (iii) If $\frac 1{n+2}<\alpha <1$ and $d\mu =f\,d\theta $ for nonnegative $f\in L^{\frac {n+1}{n+2-\frac 1\alpha }}( \mathbb {S}^n)$ with $\int _{\mathbb {S}^n}f>0$ .

Let us briefly discuss what is known about uniqueness of the solution of the $L^p$ -Minkowski problem (1.4). If $p>1$ and $p\neq n$ , then Hug, Lutwak, Yang, and Zhang [Reference Hug, Lutwak, Yang and Zhang30] proved that the Alexandrov solution of the $L^p$ -Minkowski problem (1.4) is unique. However, if $p<1$ , then the solution of the $L^p$ -Minkowski problem (1.3) may not be unique even if f is positive and continuous. Examples are provided by Chen, Li, and Zhu [Reference Chen, Li and Zhu17, Reference Chen, Li and Zhu18] if $p\in [0,1)$ , and Milman [Reference Milman42] shows that for any $C\in \mathcal {K}_{(0)}$ , one finds $q\in (-n,1)$ such that if $p<q$ , then there exist multiple solutions to the $L^p$ -Minkowski problem (1.4) with $\mu =S_{C,p}$ ; or in other words, there exists $K\in \mathcal {K}_{(0)}$ with $K\neq C$ and $S_{K,p}=S_{C,p}$ . In addition, Jian, Lu, and Wang [Reference Jian, Lu and Wang33] and Li, Liu, and Lu [Reference Li, Liu and Lu37] prove that for any $p<0$ , there exists positive even $C^\infty $ function f with rotational symmetry such that the $L^p$ -Minkowski problem (1.3) has multiple positive even $C^\infty $ solutions. We note that in the case of the centro-affine Minkowski problem $p=-n$ , Li [Reference Li36] even verified the possibility of existence of infinitely many solutions without affine equivalence, and Stancu [Reference Stancu46] related unique solution in the cases $p=0$ and $p=-n$ .

The case when f is a constant function in the $L^p$ -Minkowski problem (1.3) has received a special attention since [Reference Firey23]. When $p=-(n+1)$ , (1.3) is self-similar solution of affine curvature flow. It is proved by Andrews that all solutions are centered ellipsoids. If $n=2$ and $p=2$ , the uniqueness was proved by Andrews [Reference Andrews3]. For general n and $p>-(n+1)$ , through the work of Lutwak [Reference Lutwak40], Guan-Ni [Reference Guan and Ni26], and Andrews, Guan, and Ni [Reference Andrews, Guan and Ni6], Brendle, Choi, and Daskalopoulos [Reference Brendle, Choi and Daskalopoulos13] finally classified that the only solutions are centered balls. See also [Reference Crasta and Fragalá21, Reference Ivaki and Milman32, Reference Saroglou44] for other approaches. Stability versions of these results have been obtained by Ivaki [Reference Ivaki31], but still no stability version is known in the case $p\in [0,1)$ if we allow any solutions of (1.3) not only even ones.

Concerning recent versions of the $L^p$ -Minkowski problem, see [Reference Böröczky, Koldobsky and Volberg9].

The paper is structured as follows: The required diameter bounds are discussed in Section 2. Section 3 verifies the main properties of the Entropy, Section 4 proves our main result (Theorem 4.1) about flows, and finally Theorem 1.1 is proved in Section 5 via weak approximation.

2 Entropy and diameter estimates

For $\delta \in [0,1)$ and linear i-subspace L of $\Bbb R^{n+1}$ with $1\leq \mathrm {dim}\,L\leq n$ , we consider the collar

$$ \begin{align*}\Psi(L\cap \mathbb{S}^n,\delta)=\{x\in \mathbb{S}^n:\langle x,y\rangle\leq \delta\mbox{ for }y\in L^\bot\cap \mathbb{S}^n\}. \end{align*} $$

Let $B(1)\subset \Bbb R^{n+1}$ be the unit ball centered at the origin.

Theorem 2.1 Let $\alpha>\frac 1{n+2}$ , let $\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n}f=1$ for a bounded measurable function f on $\mathbb {S}^n$ with $\inf f>0$ , and let $\Omega \subset \Bbb R^{n+1}$ be a convex body such that $|\Omega |=|B(1)|$ and $\mathrm {diam}\, \Omega = D$ . For any $\delta ,\tau \in (0,1)$ , we have

  1. (i) if $\alpha>1$ , and $\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\Psi (z^\bot \cap \mathbb {S}^n,\delta )}f\leq 1-\tau $ for any $z\in S^n$ , then

    $$ \begin{align*}\exp\left(\frac{\alpha-1}{\alpha}\,\mathcal{E}_{\alpha, f} (\Omega)\right) \geq \gamma_1 \tau\delta^{1-\frac1\alpha}D^{1-\frac1\alpha}, \end{align*} $$
    where $\gamma _1>0$ depends on n and $\alpha $ ;
  2. (ii) if $\alpha =1$ , and

    $$ \begin{align*}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(L\cap \mathbb{S}^n,\delta)}f< \frac{(1-\tau)i}{n+1}, \end{align*} $$
    for any linear i-subspace L of $\Bbb R^{n+1}$ , $i=1,\ldots ,n$ , then
    $$ \begin{align*}\mathcal{E}_{1, f} (\Omega)\geq\tau\log D +\log\delta-4\log(n+1); \end{align*} $$
  3. (iii) if $\frac 1{n+2}<\alpha <1$ , $p=1-\frac 1\alpha $ (where $-n-1<p<0$ ), $\tau \leq \frac 12\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n}f\cdot u^{1-\frac 1\alpha }$ and

    (2.1) $$ \begin{align} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(z^\bot\cap \mathbb{S}^n,\delta)}f^{\frac{n+1}{n+1+p}}\leq \tau^{\frac{n+1}{n+1+p}}, \end{align} $$
    for any $z\in S^{n-1}$ , then
    $$ \begin{align*}\mbox{either}\ D\leq 16n^2/\delta^2, \ \ \mbox{or } D\leq \left(\frac12\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}f\cdot u^{1-\frac1\alpha}\right)^{\frac2{p}}. \end{align*} $$
    Moreover, if $\tau \leq \frac 12\exp \left (\frac {\alpha -1}{\alpha }\,\mathcal {E}_{\alpha , f} (\Omega )\right )$ , then
    $$ \begin{align*}\mbox{either}\ D\leq 16n^2/\delta^2, \ \ \mbox{or } D\leq \left(\frac12\exp\left(\frac{\alpha-1}{\alpha}\,\mathcal{E}_{\alpha, f} (\Omega)\right)\right)^{\frac2{p}}. \end{align*} $$

Remark 2.2 We note that for any $\alpha \ge 1$ , bounded f with $\inf f>0$ and $\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n}f=1$ , and $\tau \in (0,1)$ , there exists $\delta \in (0,1)$ such that conditions in (i) and (ii) hold. In the case of $1>\alpha >\frac 1{n+2}$ , (iii) holds if in addition that $\tau \leq \frac 12\exp \left (\frac {1-\alpha }{\alpha }\,\mathcal {E}_{\alpha , f} (\Omega )\right )$ for the convex body $\Omega \subset \Bbb R^{n+1}$ .

Proof Given $\alpha>\frac 1{n+2}$ , bounded f with $\inf f>0$ and $\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n}f=1$ , and $\tau \in (0,1)$ , the existence of suitable $\delta \in (0,1)$ follows from the fact that the Lebesgue measure is a Borel measure.

Now, we assume that the conditions in (i)–(iii) hold. We may assume that the centroid of $\Omega $ is the origin; thus, Kannan, Lovász, and Simonovics [Reference Kannan, Lovász and Simonovits35] yield the existence of an o-symmetric ellipsoid such that

(2.2) $$ \begin{align} E\subset\Omega\subset (n+1)E,\mbox{ and hence }-\Omega\subset (n+1)\Omega. \end{align} $$

Let u be the support function of $\Omega $ , and let $R=\max \{\|y\|:\,y\in \Omega \}\geq D/2$ and $z_0\in \mathbb {S}^n$ such that $Rz_0\in \partial \Omega $ . We observe that the definition of the entropy yields

$$ \begin{align*} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}fu^{1-\frac1\alpha}&\leq \exp\left(\frac{1-\alpha}{\alpha}\,\mathcal{E}_{\alpha, f} (\Omega)\right) \mbox{ if}\ \alpha>1;\\ \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}f\log u&\leq \mathcal{E}_{0, f} (\Omega);\\ \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}fu^{1-\frac1\alpha}&\geq \exp\left(\frac{1-\alpha}{\alpha}\,\mathcal{E}_{\alpha, f} (\Omega)\right) \mbox{ if}\ \frac1{n+2}<\alpha<1. \end{align*} $$

Case 1: $\alpha>1$ .

According to the condition in (i), we may choose $\zeta \in \{+1,-1\}$ such that

$$ \begin{align*}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Phi}f\geq \frac{\tau}2 \mbox{ for }\Phi=\{x\in \mathbb{S}^n:\langle x,\zeta z_0\rangle>\delta\}, \end{align*} $$

and hence $\frac {R\zeta z_0}{n+1}\in \Omega $ by (2.2). Since $u_\sigma (x)\geq \langle \frac {R\zeta z_0}{n+1},x\rangle \geq \frac {R\delta }{n+1}$ for $x\in \Phi $ , we have

$$ \begin{align*}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}fu^{1-\frac1\alpha}\geq\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Phi}f\left(\frac{R\delta}{n+1}\right)^{1-\frac1\alpha}\geq \frac{\tau}2\cdot\left(\frac{D\delta}{2(n+1)}\right)^{1-\frac1\alpha}. \end{align*} $$

Case 2: $\alpha =1$ .

To simplify notation, we consider the Borel probability measure $\mu (A)=\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _Af$ on $S^n$ . Let $e_1,\ldots ,e_{n+1}\in \mathbb {S}^n$ be the principal directions associated with the ellipsoid E in (2.2), and let $r_1,\ldots ,r_{n+1}>0$ be the half axes of E with $r_ie_i\in \partial E$ where we may assume that $r_1\leq \cdots \leq r_{n+1}$ . In particular, (2.2) yields that

(2.3) $$ \begin{align} (n+1)^{n+1}\prod_{i=1}^{n+1} r_i=\frac{|(n+1)E|}{|B(1)|}\geq \frac{|\Omega|}{|B(1)|}=1. \end{align} $$

We observe that for any $v\in \mathbb {S}^n$ , there exists $e_i$ such that $|\langle v,e_i\rangle |\geq \frac {1}{\sqrt {n+1}}> \frac {\delta }{n+1}$ . For $i=1,\ldots ,n+1$ , we define

$$ \begin{align*}B_i=\left\{v\in \mathbb{S}^n:|\langle v,e_i\rangle|\geq \frac{\delta}{n+1}\mbox{ and } |\langle v,e_j\rangle|<\frac{\delta}{n+1}\mbox{ for }j>i\right\}. \end{align*} $$

In particular, $B_i\subset \Psi (L_i\cap \mathbb {S}^n,\delta )$ for $i=1,\ldots ,n$ and $L_i=\mathrm {lin}\{e_1,\ldots ,e_i\}$ .

It follows that $\mathbb {S}^n$ is partitioned into the Borel sets $B_1,\ldots ,B_{n+1}$ , and as $B_i\subset \Psi (L_i\cap \mathbb {S}^n,\delta )$ for $i=1,\ldots ,n$ , we have

(2.4) $$ \begin{align} \mu(B_1)+\cdots+\mu(B_i)&\leq \frac{i(1-\tau)}{n+1} \mbox{ for}\ i=1,\ldots,n, \end{align} $$
(2.5) $$ \begin{align} \mu(B_1)+\cdots+\mu(B_{n+1})&= 1. \end{align} $$

For $\zeta =\frac {1-\tau }{n+1}$ , we have $0< \zeta <\frac 1{n+1}$ , and define

(2.6) $$ \begin{align} \beta_i&= \mu(B_i)-\zeta\mbox{ for}\ i=1,\ldots,n, \end{align} $$
(2.7) $$ \begin{align} \beta_{n+1}&= \mu(B_{n+1})-\zeta-\tau, \end{align} $$

where (2.4) and (2.5) yield

(2.8) $$ \begin{align} \beta_1+\cdots+\beta_i&\leq 0\mbox{ for}\ i=1,\ldots,m-1, \end{align} $$
(2.9) $$ \begin{align} \beta_1+\cdots+\beta_{n+1}&= 0. \end{align} $$

As $r_ie_i\in \Omega $ , it follows from the definition of $B_i$ that $u(x)\geq \langle x,r_ie_i\rangle \geq r_i\cdot \frac {\delta }{n+1}$ for $x\in B_i$ , $i=1,\ldots ,n+1$ . We deduce from applying (2.3), (2.5)–(2.9), $r_1\leq \cdots \leq r_{n+1}$ , and $\zeta <\frac 1{n+1}$ that

$$ \begin{align*} \int_{\mathbb{S}^n}\log u\,d\mu&= \sum_{i=1}^{n+1}\int_{B_i}\log u\,d\mu\\ &\geq \sum_{i=1}^{n+1}\mu(B_i)\log r_i+\sum_{i=1}^{n+1}\mu(B_i)\log \frac{\delta}{n+1} = \sum_{i=1}^{n+1}\mu(B_i)\log r_i+\log \frac{\delta}{n+1}\\ &= \sum_{i=1}^{n+1}\beta_i\log r_i+\sum_{i=1}^{n+1}\zeta \log r_i +\tau\log r_{n+1}+\log \frac{\delta}{n+1}\\ &\geq \sum_{i=1}^{n+1}\beta_i\log r_i+\zeta\log \frac{1}{(n+1)^{n+1}} +\tau\log r_{n+1}+\log \frac{\delta}{n+1}\\ &= (\beta_1+\cdots+\beta_{n+1})\log r_{n+1}+ \sum_{i=1}^{n}(\beta_1+\cdots+\beta_i)(\log r_i-\log r_{i+1})\\ & -(n+1)\zeta\log (n+1) +\tau\log r_{n+1}+\log \frac{\delta}{n+1}\\ &\geq -\log (n+1)+\tau\log r_{n+1}+\log \frac{\delta}{n+1}. \end{align*} $$

Now, $D\leq (n+1)\mathrm {diam}\,E=2(n+1)r_{n+1}\leq (n+1)^2r_{n+1}$ and $\tau <1$ , and hence

$$ \begin{align*} -\log (n+1)+\tau\log r_{n+1}+\log \frac{\delta}{n+1}&\geq -\log (n+1)+\tau\log \frac{D}{(n+1)^2} +\log \frac{\delta}{n+1}\\ &= \log\left(\delta D^\tau\right)-(2+2\tau)\log (n+1)\\ &\geq \tau\log D +\log\delta-4\log(n+1). \end{align*} $$

In particular, we conclude that

$$ \begin{align*}\mathcal{E}_{1, f} (\Omega)\geq \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}f\log u =\int_{\mathbb{S}^n}\log u\,d\mu \geq\tau\log D +\log\delta-4\log(n+1). \end{align*} $$

Case 3: $\frac 1{n+2}<\alpha <1$ .

In this case, $-(n+1)<1-\frac 1\alpha <0$ . We may assume that

$$ \begin{align*}D\geq 16n^2/\delta^2, \end{align*} $$

and we consider

$$ \begin{align*} \Phi_0&= \left\{x\in \mathbb{S}^n:\,u(x)> \sqrt{2R}\right\},\\ \Phi_1&= \left\{x\in \mathbb{S}^n:\,u(x)\leq \sqrt{2R}\right\}. \end{align*} $$

Concerning $\Phi _0$ , we have

(2.10) $$ \begin{align} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Phi_0}f\cdot u^{1-\frac1\alpha}\leq (2R)^{\frac12(1-\frac1\alpha)}\int_{\Phi_0}f\leq D^{\frac12(1-\frac1\alpha)}=D^{\frac{p}2}. \end{align} $$

On the other hand, we have $\pm \frac {R}{(n+1)}\,z_0\in \Omega $ by (2.2), thus any $x\in \Phi _1$ satisfies

$$ \begin{align*}\sqrt{2R}\geq u(x)\geq \left|\left\langle x,\frac{R}{n+1}\,z_0\right\rangle\right|, \end{align*} $$

and hence $|\langle x,z_0\rangle |\leq (n+1)\sqrt {\frac {2}{R}}\leq \frac {4n}{\sqrt {D}}\leq \delta $ ; or in other words,

$$ \begin{align*}\Phi_1\subset \Psi(z_{0}^\bot\cap \mathbb{S}^n,\delta). \end{align*} $$

It follows from $|\Omega |=|B(1)|$ and the Blaschke–Santaló inequality (cf. [Reference Schneider45]) that

$$ \begin{align*}\int_{\mathbb{S}^n} u^{-(n+1)}\leq (n+1)|B(1)|=\omega_{n},\mbox{ and hence } \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} u^{-(n+1)}\leq 1. \end{align*} $$

For $p=1-\frac 1\alpha \in (-n-1,0)$ , Hölder’s inequality and $\int _{\Phi _1}f^{\frac {n+1}{n+1+p}}< \tau ^{\frac {n+1}{n+1+p}}$ yield

$$ \begin{align*}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Phi_1}f\cdot u^{1-\frac1\alpha}\leq \left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Phi_1}f^{\frac{n+1}{n+1+p}}\right)^{\frac{n+1+p}{n+1}} \left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Phi_1} u_\sigma^{-(n+1)}\right)^{\frac{|p|}{n+1}}\leq \left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Phi_1}f^{\frac{n+1}{n+1+p}}\right)^{\frac{n+1+p}{n+1}}\leq \tau. \end{align*} $$

Finally, adding the last estimate to (2.10) yields

$$ \begin{align*}\exp\left(\frac{\alpha-1}{\alpha}\,\mathcal{E}_{\alpha, f} (\Omega)\right)\leq \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}f\cdot u^{1-\frac1\alpha}\leq D^{\frac{p}2}+\tau, \end{align*} $$

and hence the conditions either $\tau \leq \frac 12\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n}f\cdot u^{1-\frac 1\alpha }$ or $\tau \leq \frac 12\exp \left (\frac {1-\alpha }{\alpha }\,\mathcal {E}_{\alpha , f} (\Omega )\right )$ on $\tau $ implies (iii).

3 Anisotropic flows and monotonicity of entropies

The following theorem was proved by Andrews in [Reference Andrews4] (see also for a discussion of contracting of non-homogeneous fully nonlinear anisotropic curvature flows in [Reference Guan, Huang and Liu24]).

Theorem 3.1 [Reference Andrews4]

For any $\alpha>0$ and positive $f\in C^{\infty }(\mathbb S^n)$ and any initial smooth, strictly convex hypersurface $\tilde M_0\subset \mathbb R^{n+1}$ , the hypersurfaces $\tilde M_{\tau }$ given by the solution of (1.7) exist for a finite time T and converge in Hausdorff distance to a point $p \in \mathbb R^{n+1}$ as $\tau $ approaches T.

Assuming

$$\begin{align*}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n}f=1, \quad |\Omega_0|=|B(1)|,\end{align*}$$

solution (1.7) yields a smooth convex solution to the normalized flow (1.8) with volume preserved.

Set

(3.1) $$ \begin{align} h_z(x,t)\doteqdot f(x) u_z^{-\frac{1}{\alpha}}(x,t)K(x,t), \quad d\sigma_t(x) =\frac{u_z(x,t)}{K(x,t)}d\theta(x).\end{align} $$

Note that $\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n} d\sigma _t(x) =\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n} d\theta (x) =1$ .

Since the un-normalized flow (1.7) shrinks to a point in finite time, we may assume that it is the origin. Then the support function $u(x,t)$ is positive for the normalized flow (1.8).

Lemma 3.2

  1. (a) The entropy $ \mathcal {E}_{\alpha , f}(\Omega _{t})$ defined in (1.5) is monotonically decreasing,

    (3.2) $$ \begin{align} \mathcal{E}_{\alpha, f}(\Omega_{t_2})\le \mathcal{E}_{\alpha, f}(\Omega_{t_1}), \quad \forall t_1\le t_2 \in [0, \infty).\end{align} $$
  2. (b) There is $D>0$ depending only on $\inf f, \sup f, \alpha , \Omega _0$ such that

    (3.3) $$ \begin{align} \mathrm{diam}\,\Omega_t=D(t)\le D, \ \forall t\ge 0.\end{align} $$
  3. (c) $\forall t_0\in [0, \infty )$ ,

    (3.4) $$ \begin{align} \mathcal{E}_{\alpha, f}(\Omega_{t_0}, 0)\ge \mathcal{E}_{\alpha, f, \infty} +\int_{t_0}^{\infty}\left(\frac{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h^{\alpha+1}(x,t)\, d\sigma_t} {\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h(x,t) \, d\sigma_t \cdot \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h^{\alpha}(x,t)\, d\sigma_t}-1\right)\, dt,\end{align} $$

    where

    $$\begin{align*}h(x,t)=h_0(x, t), \ \mathcal{E}_{\alpha, f, \infty}\doteqdot\lim_{t\to \infty} \mathcal{E}_{\alpha, f}(\Omega_{t}).\end{align*}$$

Proof

  1. (a) We follow argument in [Reference Guan and Ni26]. For each $T_0>$ fixed, pick $T> T_0$ . Let $a^{T}=(a^{T}_1,\ldots , a^{T}_{n+1})$ be an interior point of $\Omega _T$ . Set $u^T=u- e^{t-T}\sum _{i=1}^{n+1} a^T_ix_i$ ; it satisfies equation

    (3.5) $$ \begin{align} \frac{\partial}{\partial t}u^T(x, t)= -\frac{f^\alpha(x) K^{\alpha}(x, t)}{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}f^\alpha K^{\alpha-1}} +u^T(x,t).\end{align} $$

    Note that since $a^T$ is an interior point of $\Omega _T$ and $u(x,T)$ is the support function of $\Omega _T$ with respect to $a^T$ , $u^T(x, T)> 0, \forall x\in \mathbb S^n$ . We claim

    $$\begin{align*}u^T(x, t)>0, \ \forall t\in [0, T).\end{align*}$$
    Suppose $u^T(x_0, t')\le 0$ for some $0<t'<T, x_0\in \mathbb S^n$ , and equation (3.5) implies $u^T(x_0, t)<0$ for all $t>t'$ , which contradicts to $u^T(x, T)> 0$ .

    Set $a^T(t)=e^{t-T}a^T$ . By the claim, $a^T(t)$ is in the interior of $\Omega _t, \ \forall t\le T$ . Denote

    $$\begin{align*}d\sigma_{T,t}=u^T(x,t)K^{-1}(x,t)d\theta,\end{align*}$$
    we rewrite equation (3.3) as
    (3.6) $$ \begin{align} \frac{\partial}{\partial t}u_{a^T(t)}(x,t)= -\frac{f^\alpha(x) K^{\alpha}(x, t)}{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}h_{a^T(t)}^{\alpha}(x,t)\, d\sigma_{T,t}} +u_{a^T(t)}(x,t).\end{align} $$
    We have
    $$\begin{align*}\frac{\partial}{\partial t} \mathcal{E}_{\alpha, f}(\Omega_{t}, a^T(t))=\frac{-\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h_{a^T(t)}^{\alpha+1}(x,t)\, d\sigma_{T,t}} {\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h_{a^T(t)}(x,t) \, d\sigma_{T,t} \cdot \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h_{a^T(t)}^{\alpha}(x,t)\, d\sigma_{T,t}}+1.\end{align*}$$

    Thus, $\forall t<T$ ,

    (3.7) $$ \begin{align} &\mathcal{E}_{\alpha, f}(\Omega_{t}, a^T(t))-\mathcal{E}_{\alpha, f}(\Omega_T, a^T)\\ \nonumber &= \int_{t}^{T}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} \left(\frac{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h_{a^T(t)}^{\alpha+1}(x,t)\, d\sigma_{T,t}} {\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h_{a^T(t)}(x,t) \, d\sigma_{T,t} \cdot \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h_{a^T(t)}^{\alpha}(x,t)\, d\sigma_{T,t}}-1\right)\, dt\ge 0. \end{align} $$
    Therefore,
    $$\begin{align*}\mathcal{E}_{\alpha, f}(\Omega_{t})\ge\mathcal{E}_{\alpha, f}(\Omega_T, a^T), \ \forall t<T.\end{align*}$$
    Since $a^T$ is arbitrary, (3.2) is proved.
  2. (b) The boundedness of $D(t)$ follows from Theorem 2.1 combined with the estimate $\mathcal {E}_{\alpha , 1}(\Omega _{t})\leq \mathcal {E}_{\alpha , 1}(B(1))$ from (a) (see also [Reference Andrews, Guan and Ni6, Reference Guan and Ni26]). The only nontrivial case is when $\frac 1{n+2}<\alpha <1$ because we have to choose a $\tau $ independent of t. However, we may choose any $\tau \in (0,1)$ with $\tau \leq \frac 12\exp \left (\frac {1-\alpha }{\alpha }\,\mathcal {E}_{\alpha , f} (B(1))\right )$ according to $\mathcal {E}_{\alpha , 1}(\Omega _{t})\leq \mathcal {E}_{\alpha , 1}(B(1))$ .

  3. (c) $\forall \epsilon>0, \ \forall t_0$ fixed, pick $T>T_0>t_0$ . As $ \mathcal {E}_{\alpha , f}(\Omega _{T})$ is bounded by (a), $\exists a^T$ inside $\Omega _T$ such that $ \mathcal {E}_{\alpha , f}(\Omega _{T})\le \mathcal {E}_{\alpha , f}(\Omega _{T}, a^T)+\epsilon $ . By (3.7),

    $$ \begin{align*} &\mathcal{E}_{\alpha, f}(\Omega_{t_0}, a^T(t_0))-\mathcal{E}_{\alpha, f}(\Omega_{T})\\ &\ge \int_{t_0}^{T_0}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} \left(\frac{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h_{a^T(t)}^{\alpha+1}(x,t)\, d\sigma_{T,t}} {\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h_{a^T(t)}(x,t) \, d\sigma_{T,t} \cdot \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h_{a^T(t)}^{\alpha}(x,t)\, d\sigma_{T,t}}-1\right)\, dt-\epsilon. \end{align*} $$

    As $|a^T|\le D, \ \forall T$ , let $T\to \infty $ ,

    $$\begin{align*}a^T(t)\to 0, \ u^T(x,t)\to u(x,t), \ \ \mbox{ uniformly for}\ 0\le t\le T_0, x\in \mathbb S^n. \end{align*}$$

    We obtain $\forall t_0<T_0$ ,

    $$ \begin{align*} \mathcal{E}_{\alpha, f}(\Omega_{t_0}, 0)-\mathcal{E}_{\alpha, f, \infty}\ge \int_{t_0}^{T_0}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} \left(\frac{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h^{\alpha+1}(x,t)\, d\sigma_t} {\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h(x,t) \, d\sigma_t \cdot \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h^{\alpha}(x,t)\, d\sigma_t}-1\right)\, dt-\epsilon.\end{align*} $$

    Then let $T_0\to \infty $ , as $\epsilon>0$ is arbitrary, we obtain (3.4).

4 Weak convergence

The goal of this section is to prove the following statement.

Theorem 4.1 For a $C^\infty $ function $f:\mathbb {S}^n\to (0,\infty )$ and $\alpha>\frac 1{n+2}$ with $ \frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n}f=1$ , there exist $\lambda>0$ and a convex body $\Omega \subset \Bbb R^{n+1}$ with $o\in \Omega $ whose support function u is a (possibly weak) solution of the Monge–Ampère equation

(4.1) $$ \begin{align} u^{\frac1\alpha}\det(\bar{\nabla}^2_{ij} u+u\bar{g}_{ij})= f \end{align} $$

and $\Omega $ satisfies that

(4.2) $$ \begin{align} \mathcal{E}_{\alpha, f} (\lambda\Omega)\leq \mathcal{E}_{\alpha, f} (B(1)), \quad |\lambda\Omega|=|B(1)|, \end{align} $$

where $C^{-1}<\lambda <C$ for a $C>1$ depending only on the $\alpha , \tau , \delta $ in Theorem 2.1 such that f satisfies the conditions in Theorem 2.1.

From now on, we will assume that the f in Theorem 4.1 satisfies the corresponding condition in Theorem 2.1 and $\Omega _0=B(1)$ in (1.8). We note that for any $z\in B(1)$ , $v_z\leq 2$ for the support function $v_z$ of $B(1)$ at z, and hence if $\alpha>\frac 1{n+2}$ , then

(4.3) $$ \begin{align} \mathcal{E}_{\alpha, f_k} (B(1))\leq\left\{ \begin{array}{rl} \frac{\alpha}{\alpha-1}\cdot\log 2^{1-\frac1\alpha},&\mbox{ if }\alpha\neq 1,\\ \log 2,&\mbox{ if }\alpha=1. \end{array} \right. \end{align} $$

The following is a consequence of Theorem 2.1 and Lemma 3.2.

Lemma 4.2 There exist $C_{\alpha , \tau , \delta }>0, D_{\alpha , \tau , \delta }>0$ , and $c_{\alpha , \tau , \delta }\in \mathbb R$ depending only on constants $\alpha , \tau , \delta $ in Theorem 2.1 such that, along (1.8), we have

(4.4) $$ \begin{align} D(t)\le D_{\alpha, \tau, \delta},\ \mathcal{E}_{\alpha, f}(\Omega_{t}, 0)\ge c_{\alpha, \tau, \delta}, \ \frac{1}{C_{\alpha, \tau, \delta}}\le \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} h(x,t)d\sigma_t\le C_{\alpha, \tau, \delta}.\end{align} $$

Proof For each $\alpha>\frac 1{n+2}$ fixed with condition on f as in Theorem 2.1, $\mathcal {E}_{\alpha , f}(\Omega _{t})$ is bounded from below in terms of the diameter $D(t)$ . Since $|\Omega _t|=|B(1)|$ , we have $D(t)\ge 2$ by the Isodiametric Inequality (cf. [Reference Schneider45]). By Theorem 2.1, $\mathcal {E}_{\alpha , f}(\Omega _{t})$ is bounded from below by a constant $c_{\alpha , \tau , \delta }>0$ , and hence $\mathcal {E}_{\alpha , f, \infty } \ge c_{\alpha , \tau , \delta }$ . It follows from Lemma 3.2 that $\mathcal {E}_{\alpha , f}(\Omega _{t})\le \mathcal {E}_{\alpha , f}(B(1))$ , and this estimate combined with (4.3) and Theorem 2.1 yields $D(t)\le D_{\alpha , \tau , \delta }$ where $D_{\alpha , \tau , \delta }$ depends only on constants in condition on f in Theorem 2.1. Finally, the inequalities follow from Lemma 3.2.

Set

(4.5) $$ \begin{align} \eta(t)=\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h(x,t) \, d\sigma_{t}. \end{align} $$

We note that $\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n} h(x,t) \, d\sigma _{t} $ is monotone and bounded from below and above by Lemma 4.2, and hence we have

(4.6) $$ \begin{align}C_{\alpha, \tau, \delta}\ge \lim_{t\to \infty} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h(x,t)=\eta\ge \frac1{C_{\alpha, \tau, \delta}}.\end{align} $$

By Lemma 3.2 and Corollary 4.2,

(4.7) $$ \begin{align} \int_0^{\infty} \left(\frac{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h^{\alpha+1}(x,t)\, d\sigma_t} {\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h(x,t) \, d\sigma_t \cdot \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h^{\alpha}(x,t)\, d\sigma_t}-1\right)\, dt<\infty.\end{align} $$

Since the integrand is nonnegative, $\exists t_k\to \infty $ such that

(4.8) $$ \begin{align} \lim_{k\to \infty} \left( \frac{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}h^{\alpha+1}(x,t_k)\, d\sigma_{t_k}} {\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h(x,t_k) \, d\sigma_{t_k} \cdot \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h^{\alpha}(x,t_k)\, d\sigma_{t_k}}-1\right)=0.\end{align} $$

This implies

(4.9) $$ \begin{align} \lim_{k\to \infty}\frac{\left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}h^{\alpha+1}(x,t_k)\, d\sigma_{t_k}\right)^{\frac{1}{1+\alpha}}}{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h(x,t_k) \, d\sigma_{t_k} }= \lim_{k\to \infty}\frac{\left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}h^{\alpha+1}(x,t_k)\, d\sigma_{t_k}\right)^{\frac{\alpha}{1+\alpha}}}{\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h^{\alpha}(x,t_k)\, d\sigma_{t_k}}= 1.\end{align} $$

After considering a subsequence, we may assume that

(4.10) $$ \begin{align} \Omega_{t_k}\to \Omega, \quad u(x,t_k)\to u(x),\end{align} $$

where u is the support function of $\Omega $ . In view of (4.9) and (4.6),

(4.11) $$ \begin{align} \lim_{k\to \infty} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}h^{\alpha+1}(x,t_k)\, d\sigma_{t_k}=\eta^{1+\alpha},\ \lim_{k\to \infty}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n} h^{\alpha}(x,t_k)\, d\sigma_{t_k}= \eta^{\alpha}.\end{align} $$

The following lemma is crucial for the weak convergence, which is a refined form of the classical Hölder inequality.Footnote 1

Lemma 4.3 Let $p,\ q\in \mathbb R^+$ with $\frac 1p+\frac 1q=1$ , and set $\beta =\min \{\frac 1p, \frac 1q\}$ . Let $(M,\mu )$ be a measurable space; $\forall F\in L^p, \ G\in L^q$ ,

(4.12) $$ \begin{align} \int_M |FG| d\mu \le \|F\|_{L^p}\|G\|_{L^q}\left(1-\beta\int_M \left(\frac{|F|^{\frac{p}2}}{(\int_M |F|^p d\mu)^{\frac12}}-\frac{|G|^{\frac{q}2}}{(\int_M |G|^qd\mu )^{\frac12}}\right)^2\right).\end{align} $$

Proof We first prove the following Claim. $\forall s, t\in \mathbb R$ ,

(4.13) $$ \begin{align} e^{\frac{s}{p}+\frac{t}{q}}\le \frac{e^s}{p}+\frac{e^t}{q}-\beta(e^{\frac{s}2}-e^{\frac{t}2})^2.\end{align} $$

We may assume $t\ge s$ , set $\tau =t-s$ , and (4.13) is equivalent to

(4.14) $$ \begin{align} e^{\frac{\tau}{q}}\le \frac{1}{p}+\frac{e^{\tau}}{q}-\beta(1-e^{\frac{\tau}2})^2, \ \forall \tau\ge 0.\end{align} $$

Set

$$\begin{align*}\xi(\tau)=\frac{1}{p}+\frac{e^{\tau}}{q}-\beta(1-e^{\frac{\tau}2})^2-e^{\frac{\tau}{q}}.\end{align*}$$

We have $\xi (0)=0$ ,

$$\begin{align*}\xi'(\tau)=\frac{e^{\frac{\tau}q}}{q}\rho, \ \mbox{where} \ \rho(\tau)=e^{\frac{\tau}{p}}(1-\beta q)+q\beta e^{\frac{\tau}2-\frac{\tau}q}-1.\end{align*}$$

If $\beta =\frac 1q$ , then $\frac 1q\le \frac 12$ ; since $\tau \ge 0$ ,

$$\begin{align*}\rho(\tau)=e^{\frac{\tau}{p}}(1-\beta q)+q\beta e^{\frac{\tau}2-\frac{\tau}q}-1=e^{\frac{\tau}2-\frac{\tau}q}-1\ge 0.\end{align*}$$

If $\beta =\frac 1p$ , then $\frac 1q\ge \frac 12$ ; we have

$$ \begin{align*} \rho'(\tau)&=e^{\frac{\tau}{p}}\left(\frac{1-\beta q}p+\beta q(\frac12-\frac1q)e^{\frac{\tau}2-\frac{\tau}q}\right)\\ & \ge e^{\frac{\tau}{p}}\left(\frac{1-\beta q}p+\beta q(\frac12-\frac1q)\right)\\ &\ge e^{\frac{\tau}{p}}\beta q(\frac12-\frac1p)\ge 0.\end{align*} $$

We conclude that

$$\begin{align*}\rho(\tau)\ge 0, \ \forall \tau\ge 0.\end{align*}$$

In turn,

$$\begin{align*}\xi'(\tau)\ge 0, \ \forall \tau\ge o.\end{align*}$$

This yields (4.14) and (4.13). The Claim is verified.

Back to the proof of the lemma. We may assume

$$\begin{align*}F\ge 0, \ g\ge 0, \ \int F^p>0, \ \int G^q>0.\end{align*}$$

Set

$$\begin{align*}e^s=\frac{F^p}{\int F^p}, \quad e^t=\frac{G^q}{\int G^q}.\end{align*}$$

Put them into (4.13) and integrate, as $\frac 1p+\frac 1q=1$ ,

$$\begin{align*}\frac{\int FG}{(\int F^p)^{\frac1{p}}(\int G^q)^{\frac1{q}}}\le \left(1-\beta\int (\frac{F^{\frac{p}2}}{(\int F^p)^{\frac12}}-\frac{G^{\frac{q}2}}{(\int G^q)^{\frac12}})^2\right).\\[-34pt] \end{align*}$$

We prove weak convergence.

Proposition 4.4 $\forall \alpha>\frac {1}{n+2}$ , suppose that (4.10) and (4.11) hold. Denote

$$\begin{align*}u_{k}=u(x, t_k), \ \sigma_{n,k}=\sigma_n(u_{ij}(x,t_k)+u(x,t_k)\delta_{ij}).\end{align*}$$

Then

(4.15) $$ \begin{align} \lim_{k\to \infty} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |u_k^{\frac{1}\alpha}\sigma_{n,k}-\frac{f}{\eta}| d\theta =0,\end{align} $$

where $\eta $ is defined in (4.5) which is bounded from below and above in (4.6). As a consequence, there is a convex body $\Omega \subset \mathbb R^{n+1}$ with $o\in \Omega $ ,

$$\begin{align*}|\Omega|=|B(1)|, \quad \mathcal{E}_{\alpha, f}(\Omega_{t})\le \mathcal{E}_{\alpha, f}(B(1)),\end{align*}$$

and its support function u satisfies

(4.16) $$ \begin{align} u^{\frac1{\alpha}} S_{\Omega}=\frac1{\eta} fd \theta. \end{align} $$

Proof We only need to verify (4.15). By (4.11), it is equivalent to prove

(4.17) $$ \begin{align} \lim_{k\to \infty} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |u_k^{\frac{1}\alpha}\sigma_{n,k}-f\eta^{-1}(t_k)| d\theta =0.\end{align} $$

Since $D(t_k)$ is bounded,

$$\begin{align*}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} u_k^{\frac{1}{\alpha^2}}\sigma_{n,k} d\theta\le (D(t_k))^{\frac1{\alpha^2}}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} u_k^{\frac{1}{\alpha^2}}\sigma_{n,k} d\theta\le (D(t_k))^{\frac1{\alpha^2}}|\partial \Omega_{t_k}|\le C.\end{align*}$$
(4.18) $$ \begin{align} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |u_k^{\frac{1}\alpha}\sigma_{n,k}-f\eta^{-1}(t_k)| d\theta &=\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\frac{f}{\eta(t_k)u_k^{\frac{1}\alpha}\sigma_{n,k}}-1 |u_k^{\frac{1}\alpha}\sigma_{n,k} d\theta\nonumber\\ &\le \left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\frac{f}{\eta(t_k)u_k^{\frac{1}\alpha}\sigma_{n,k}}-1 |^{1+\alpha} d\sigma_{t_k}\right)^{\frac{1}{1+\alpha}}\left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} u_k^{(\frac1{\alpha}-1)\frac{1+\alpha}{\alpha}}d\sigma_{t_k}\right)^{\frac{\alpha}{1+\alpha}}\nonumber \\ &= \left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\frac{f}{\eta(t_k)u_k^{\frac{1}\alpha}\sigma_{n,k}}-1 |^{1+\alpha} d\sigma_{t_k}\right)^{\frac{1}{1+\alpha}}\left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} u_k^{\frac{1}{\alpha^2}}\sigma_{n,k} d\theta\right)^{\frac{\alpha}{1+\alpha}}\nonumber \\ &\le C \left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |f\eta^{-1}(t_k)u_k^{-\frac{1}\alpha}\sigma^{-1}_{n,k}-1 |^{1+\alpha} d\sigma_{t_k}\right)^{\frac{1}{1+\alpha}}. \end{align} $$

By (4.8), (4.11), and Lemma 4.3, with $p=\alpha +1$ , $F^{\frac {1}{1+\alpha }}=h(x,t_k)$ , $G=1$ ,

(4.19) $$ \begin{align} \lim_{k\to \infty} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int \left((\frac{h(x,t_k)}{\eta(t_k)})^{\frac{1+\alpha}2}-1\right)^2d\sigma_{t_k}=0.\end{align} $$

For $t_k$ fixed, let

$$\begin{align*}\gamma_k(x)=f\eta^{-1}(t_k)u_k^{-\frac{1}\alpha}\sigma^{-1}_{n,k}=h(x,t_k)\eta^{-1}(t_k)\end{align*}$$

and set

$$\begin{align*}\Sigma_{k}=\left\{x\in \mathbb S^n \ | \ |\gamma_k(x)-1|\le \frac12\right\}.\end{align*}$$

It is straightforward to check that $\exists A_{\alpha }\ge 1$ depending only on $\alpha $ such that

$$ \begin{align*} A_{\alpha}|\gamma^{\frac{1+\alpha}2}_k(x)-1|&\ge |\gamma_k(x)-1|, \ \forall x\in \Sigma_k, \\ A_{\alpha} |\gamma^{\frac{1+\alpha}2}_k(x)-1|^{2}&\ge |\gamma_k(x)-1|^{1+\alpha}, \ \forall x\in \Sigma_{k}^c.\end{align*} $$

Since $ |\gamma ^{\frac {1+\alpha }2}_k(x)-1|\le 2^{1+\alpha }, \ \forall x\in \Sigma _k$ , let $\delta =\min \{1+\alpha , 2\}$ ,

$$ \begin{align*} \ \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\gamma_k(x)-1 |^{1+\alpha} d\sigma_{t_k}=&\frac{1}{\omega_n}\left(\int_{\Sigma_k} |\gamma_k(x)-1 |^{1+\alpha} d\sigma_{t_k}+\int_{\Sigma_k^c} |\gamma_k(x)-1 |^{1+\alpha} d\sigma_{t_k}\right)\\ \le & \frac{A^{1+\alpha}_{\alpha}}{\omega_n} \left(\int_{\Sigma_k} |\gamma_k^{\frac{1+\alpha}2}(x)-1 |^{1+\alpha} d\sigma_{t_k}+\int_{\Sigma_k^c} |\gamma_k^{\frac{1+\alpha}2}(x)-1 |^{2} d\sigma_{t_k}\right)\\ \le & \frac{(2A_{\alpha})^{1+\alpha}}{\omega_n} \left(\int_{\Sigma_k} |\gamma_k^{\frac{1+\alpha}2}(x)-1 |^{\delta} d\sigma_{t_k}+\int_{\Sigma_k^c} |\gamma_k^{\frac{1+\alpha}2}(x)-1 |^{2} d\sigma_{t_k}\right)\\ \le & (2A_{\alpha})^{1+\alpha} \left(\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\gamma_k^{\frac{1+\alpha}2}(x)-1 |^{\delta} d\sigma_{t_k}+\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\gamma_k^{\frac{1+\alpha}2}(x)-1 |^{2} d\sigma_{t_k}\right)\\ \le & (2A_{\alpha})^{1+\alpha} \left((\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\gamma_k^{\frac{1+\alpha}2}(x)-1 |^{2} d\sigma_{t_k})^{\frac{\delta}2}+\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\gamma_k^{\frac{1+\alpha}2}(x)-1 |^{2} d\sigma_{t_k}\right).\end{align*} $$

By (4.19),

$$\begin{align*}\lim_{k\to \infty}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\gamma_k^{\frac{1+\alpha}2}(x)-1 |^{2} d\sigma_{t_k}=0.\end{align*}$$

Hence,

(4.20) $$ \begin{align} \lim_{k\to \infty} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb S^n} |\gamma_k(x)-1 |^{1+\alpha} d\sigma_{t_k}=0.\end{align} $$

Now, (4.17) follows from (4.18)–(4.20).

Proof Proof of Theorem 4.1.

It follows from Proposition 4.4 after a proper rescaling as $\eta $ satisfies (4.6) and (4.16).

5 The general Monge–Ampère equations – proof of Theorem 1.1

In order to prove Theorem 1.1, we need weak approximation in the following sense.

Lemma 5.1 For $\delta ,\varepsilon \in (0,\frac 12)$ and a Borel probability measure $\mu $ on $\mathbb {S}^n$ , $n\geq 1$ , there exists a sequence $d\mu _k=\frac 1{\omega _n}\,f_k\,d\theta $ of Borel probability measures whose weak limit is $\mu $ and $f_k\in C^\infty ( \mathbb {S}^n)$ satisfies $f_k>0$ and the following properties:

  1. (i) If $\mu \left (\Psi (z^\bot \cap \mathbb {S}^n,2\delta )\right )\leq 1-\varepsilon $ for any $z\in S^{n-1}$ , then

    (5.1) $$ \begin{align} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(z^\bot\cap \mathbb{S}^n,\delta)}f_k\leq 1-\varepsilon \mbox{ for any}\ z\in S^{n-1}. \end{align} $$
  2. (ii) If $\mu (\Psi (L\cap \mathbb {S}^n,2\delta ))<(1-2\delta )\cdot \frac {\ell }{n+1}$ for any linear $\ell $ -subspace L of $\Bbb R^{n+1}$ , $\ell =1,\ldots ,n$ , then

    (5.2) $$ \begin{align} \mu_k\left(\Psi\left(L\cap \mathbb{S}^n,\delta\right)\right)<(1-\delta)\cdot \frac{\ell}{n+1}. \end{align} $$
  3. (iii) If $d\mu =\frac 1{\omega _n}\,f\,d\theta $ for $f\in L^{r}(\mathbb {S}^n)$ where $r>1$ , and

    (5.3) $$ \begin{align} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(z^\bot\cap \mathbb{S}^n,2\delta)}f^r\leq \varepsilon \end{align} $$

    for any $z\in S^{n-1}$ , then

    (5.4) $$ \begin{align} \int_{\Psi(z^\bot\cap \mathbb{S}^n,\delta)}f_k^r\leq 2^r\varepsilon \mbox{ for any}\ z\in S^{n-1}. \end{align} $$

Proof For $k\geq 1$ , let $\{B_{k,i}\}_{i=1,\ldots ,m(k)}$ be a partition of $S^n$ into spherically convex Borel measurable sets $B_{k,i}$ with $\mathrm {diam}B_{k,i}\leq \frac 1k$ and $\theta (B_{k,i})>0$ . For each $B_{k,i}$ , we choose a $C^\infty $ function $h_{k,i}:\mathbb {S}^n\to [0,\infty )$ such that for $M_{k,i}=\max h_{k,i}$ and the probability measure $d\tilde {\theta }=\frac 1{\omega _n}\,d\theta $ , we have:

  • $h_{k,i}=0$ if $x\not \in B_{k,i}$ ;

  • $M_{k,i}\leq (1+\frac 1k)\cdot \frac {\mu (B_{k,i})}{\tilde {\theta }(B_{k,i})}$ ;

  • $\theta \left (\left \{x\in B_{k,i}:h_{k,i}(x)<M_{k,i}\right \}\right )<\frac 1k\,\theta (B_{k,i})$ ;

  • $\int _{B_{k,i}}h_{k,i}\,d\tilde {\theta }=\mu (B_{k,i})$ .

We consider the positive $C^\infty $ function $\tilde {f}_k\hspace{-0.5pt}=\hspace{-0.5pt}\frac 1k\hspace{-0.5pt}+\hspace{-0.5pt}\sum _{i=1}^{m(k)}h_{k,i}$ , and hence $f_k\hspace{-0.5pt}=\hspace{-0.5pt}\left (\hspace{2pt} \frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n}\tilde {f}_k\right )^{-1}\tilde {f}$ satisfies that the probability measure $d\mu _k=f_k\,d\tilde {\theta }$ tends weakly to $\mu $ , and for large $k\geq 1/\delta $ , $\mu _k$ satisfies (i), and if (ii) holds, then $\mu _k$ also satisfies (5.2).

Turning to (iii), we assume that $d\mu =f\,d\tilde {\theta }$ for $f\in L^{r}(\mathbb {S}^n)$ where $r>1$ , and f satisfies (5.3). For any large k and $i=1,\ldots ,m(k)$ , we deduce from the Hölder inequality that

$$ \begin{align*} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{B_{k,i}} \tilde{f}_k^r&=\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{B_{k,i}} \left(h_{k,i}+\frac1k\right)^r\leq 2^{r-1}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{B_{k,i}} h_{k,i}^r+2^{r-1}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{B_{k,i}} \frac1{k^r}\\ &\leq 2^{r-1}\tilde{\theta}(B_{k,i})M_{k,i}^r+2^{r-1}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{B_{k,i}} \frac1{k^r}\\ &\leq 2^{r-1}\left(1+\frac1k\right)^r\tilde{\theta}(B_{k,i}) \left(\frac{\int_{B_{k,i}} f}{\tilde{\theta}(B_{k,i})}\right)^r+2^{r-1}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{B_{k,i}} \frac1{k^r}\\ &\leq 2^{r-1}\left(1+\frac1k\right)^r\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{B_{k,i}} f^r+2^{r-1}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{B_{k,i}}\frac1{k^r}. \end{align*} $$

Summing this estimate up for large k and all $B_{k,i}$ with $B_{k,i}\cap \Psi (z^\bot \cap \mathbb {S}^n,\delta )\neq \emptyset $ , and using that $\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n}\tilde {f}_k\geq 2^{-1/2}$ for large k, we deduce that

$$ \begin{align*}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(z^\bot\cap \mathbb{S}^n,\delta)} f_k^r\leq \sqrt{2} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(z^\bot\cap \mathbb{S}^n,\delta)} \tilde{f}_k^r\leq \sqrt{2}\cdot2^{r-1}\left(1+\frac1k\right)^r\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(z^\bot\cap \mathbb{S}^n,2\delta)} f^r+ \sqrt{2}\cdot\frac{2^{r-1}}{k^r}\leq 2^r\varepsilon. \end{align*} $$

For $\alpha>0$ and $p=1-\frac 1\alpha $ , the $L^p$ -surface area $dS_{\Omega ,p}=u^{1-p}dS_\Omega $ was introduced in the seminal works [Reference Lutwak39Reference Lutwak41] for a convex body $\Omega \subset \Bbb R^{n+1}$ with $o\in \Omega $ and support function u. Since the surface area measure is weakly continuous for $p<1$ , and if $K\subset \Bbb R^{n+1}$ is an at most n-dimensional compact convex set, then $S_{K,p}\equiv 0$ for $p<1$ , we have the following statement.

Lemma 5.2 If convex bodies $\Omega _m\subset \Bbb R^{n+1}$ tend to a compact convex set $K\subset \Bbb R^{n+1}$ where $o\in \Omega _m,K$ , and $\liminf _{m\to \infty }S_{\Omega _m,p}>0$ , then $\mathrm {int}K\neq \emptyset $ and $S_{\Omega _m,p}$ tends weakly to $S_{K,p}$ .

For the reader’s sake, let us recall Theorem 1.1.

Theorem 5.3 For $\alpha>\frac 1{n+2}$ and finite nontrivial Borel measure $\mu $ on $\mathbb {S}^n$ , $n\geq 1$ , there exists a weak solution of (1.2) provided the following holds:

  1. (i) If $\alpha>1$ and $\mu $ is not concentrated onto any great subsphere $x^\bot \cap \mathbb {S}^n$ , $x\in \mathbb {S}^n$ .

  2. (ii) If $\alpha =1$ and $\mu $ satisfies that for any linear $\ell $ -subspace $L\subset \Bbb R^{n+1}$ with $1\leq \ell \leq n$ , we have:

    1. (a) $\displaystyle \mu (L\cap \mathbb {S}^n)\leq \frac {\ell }{n+1}\cdot \mu (\mathbb {S}^n)$ ;

    2. (b) equality in (a) for a linear $\ell $ -subspace $L\subset \Bbb R^{n+1}$ with $1\leq d\leq n$ implies the existence of a complementary linear $(n+1-\ell )$ -subspace $\widetilde {L}\subset \Bbb R^{n+1}$ such that $\mathrm {supp}\,\mu \subset L\cup \widetilde {L}$ .

  3. (iii) If $\frac 1{n+2}<\alpha <1$ , assume $d\mu =fd\theta $ for nonnegative $f\in L^{\frac {n+1}{n+2-\frac 1\alpha }}( \mathbb {S}^n)$ with $\int _{\mathbb {S}^n}f>0$ .

Proof Let $\alpha>\frac 1{n+2}$ . After rescaling, we may assume that the $\mu $ in (1.2) is a probability measure. We consider the sequence $d\mu _k=\frac 1{\omega _n}f_k\,d\theta $ of Lemma 5.1 of Borel probability measures whose weak limit is $\mu $ and $f_k\in C^\infty ( \mathbb {S}^n)$ satisfies $f_k>0$ . For each $f_k$ , let $\Omega _k\subset \Bbb R^{n+1}$ be the convex body with $o\in \Omega _k$ provided by Theorem 4.1 whose support function $u_k$ is the solution of the Monge–Ampère equation

(5.5) $$ \begin{align} u_k^{\frac1\alpha}\,dS_{\Omega_k}=f _k\,d\theta; \end{align} $$

$\exists \lambda _k>0$ under control, with $|\lambda _k\Omega |=|B(1)|$ , $\Omega _k$ satisfies that

(5.6) $$ \begin{align} \mathcal{E}_{\alpha, f_k} (\lambda_k\Omega_k)\leq \mathcal{E}_{\alpha, f_k} (B(1)). \end{align} $$

We also need the observations that

(5.7) $$ \begin{align} |\Omega_k|=\frac1{n+1}\int_{\mathbb{S}^n}u_k\,dS_{\Omega_k}, \end{align} $$

and if $p=1-\frac 1\alpha $ , then

(5.8) $$ \begin{align} S_{\Omega_k,p}(\mathbb{S}^n)=\int_{\mathbb{S}^n}u_k^{1-\frac1\alpha}\,dS_{\Omega_k} =\omega_n\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}f_k=\omega_n. \end{align} $$

We claim that if there exists $\Delta>0$ depending on n, $\alpha $ , and $\mu $ such that

(5.9) $$ \begin{align} \mathrm{diam}\Omega_k\leq \Delta,\ \mbox{then Theorem 5.3 holds.} \end{align} $$

To prove this claim, we note that (5.9) yields the existence of a subsequence of $\{\Omega _k\}$ tending to a compact convex set $\Omega $ with $o\in \Omega $ , which is a convex body by (5.8) and Lemma 5.2. Moreover, Lemma 5.2 also yields that $\Omega $ is an Alexandrov solution of (1.2), verifying the claim (5.9).

We divide the rest of the argument verifying Theorem 5.3 into three cases.

Case 1: $\alpha>1$ .

Since $\mu $ is not concentrated to any great subsphere, there exist $\delta \in (0,\frac 12)$ depending on $\mu $ such that $\mu \left (\Psi (z^\bot \cap \mathbb {S}^n,2\delta )\right )\leq 1-2\delta $ for any $z\in S^{n-1}$ . It follows from Lemma 5.1 that we may assume that

(5.10) $$ \begin{align} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(z^\bot\cap \mathbb{S}^n,\delta)}f_k\leq 1-\delta \mbox{ for any}\ z\in S^{n-1}. \end{align} $$

Now, Theorem 4.1 implies that $\lambda _k\geq c$ for a constant $c>0$ depending on n, $\delta $ , and $\alpha $ , and in turn Theorem 4.1, (4.3), and $\frac 1{\alpha }-1<0$ yield that

$$ \begin{align*}\mathcal{E}_{\alpha, f} (\Omega_k)=\frac{\alpha}{\alpha-1}\cdot\log\lambda_k^{\frac1{\alpha}-1} +\mathcal{E}_{\alpha, f} (\lambda_k\Omega_k)\leq \frac{\alpha}{\alpha-1}\cdot\log\lambda_k^{\frac1{\alpha}-1} +\mathcal{E}_{\alpha, f} (B(1))\leq C \end{align*} $$

for a constant $C>0$ depending on n, $\delta $ , and $\alpha $ . Therefore, Theorem 2.1 and (5.10) imply that the sequence $\{\Omega _k\}$ is bounded, and in turn the claim (5.9) implies Theorem 5.3 if $\alpha>1$ .

Case 2: $\alpha =1$ .

The argument is by induction on $n\geq 0$ where we do not put any restriction on the probability measure $\mu $ in the case $n=0$ . For the case $n=0$ , we observe that any finite measure $\mu $ on $S^0$ can be represented in the form $d\mu =u\,dS_\Omega $ for a suitable segment $\Omega \subset \Bbb R^1$ .

For the case $n\geq 1$ , assuming that we have verified Theorem 5.3(ii) in smaller dimensions, we consider a Borel measure probability $\mu $ on $S^n$ satisfying (a) and (b).

Case 2.1: There exists a linear $\ell $ -subspace $L\subset \Bbb R^{n+1}$ with $1\leq \ell \leq n$ and $\mu (L\cap \mathbb {S}^n)= \frac {\ell }{n+1}\cdot \mu (\mathbb {S}^n)$ .

Let $\widetilde {L}\subset \Bbb R^{n+1}$ be the complementary linear $(n+1-\ell )$ -subspace with $\mathrm {supp}\,\mu \subset L\cup \widetilde {L}$ , and hence $\mu (\widetilde {L}\cap \mathbb {S}^n)= \frac {n+1-\ell }{n+1}\cdot \mu (\mathbb {S}^n)$ . It follows by induction that there exist an $\ell $ -dimensional compact convex set $K'\subset L$ and an $(n+1-\ell )$ -dimensional compact convex set $\widetilde {K}'\subset \widetilde {L}$ such that and . Finally, for $K=\widetilde {L}^\bot \cap (K'+L^\bot )$ and $\widetilde {K}=L^\bot \cap (\widetilde {K}'+\widetilde {L}^\bot )$ , there exist $\alpha ,\tilde {\alpha }>0$ such that

$$ \begin{align*}\mu=(n+1)V_{\alpha K+\tilde{\alpha}\widetilde{K}}. \end{align*} $$

Case 2.2: $\mu (L\cap \mathbb {S}^n)< \frac {\ell }{n+1}\cdot \mu (\mathbb {S}^n)$ for any linear $\ell $ -subspace $L\subset \Bbb R^{n+1}$ with $1\leq \ell \leq n$ .

It follows by a compactness argument that there exists $\delta \in (0,\frac 12)$ depending on $\mu $ such that $\mu (\Psi (L\cap \mathbb {S}^n,2\delta ))<(1-2\delta )\cdot \frac {\ell }{n+1}$ for any linear $\ell $ -subspace L of $\Bbb R^{n+1}$ , $\ell =1,\ldots ,n$ . We consider the sequence of probability measures $d\mu _k=\frac 1{\omega _n}f_k\,d\theta $ of Lemma 5.1 tending weakly to $\mu $ such that $f_k>0$ , $f_k\in C^\infty (\mathbb {S}^n)$ , and

(5.11) $$ \begin{align} \mu_k\left(\Psi\left(L\cap \mathbb{S}^n,\delta\right)\right)<(1-\delta)\cdot \frac{\ell}{n+1} \end{align} $$

for any linear $\ell $ -subspace L of $\Bbb R^{n+1}$ , $\ell =1,\ldots ,n$ .

For each $f_k$ , let $\Omega _k\subset \Bbb R^{n+1}$ with $o\in \Omega _k$ be the convex body provided by Theorem 4.1 whose support function $u_k$ is the solution of the Monge–Ampère equation (4.1) and satisfies (4.2) with $f=f_k$ and $\lambda =\lambda _k$ where $|B(1)|=|\lambda _k\Omega _k|$ for $\lambda _k>0$ , and

$$ \begin{align*} |\Omega_k|&=\frac1{n+1}\,\int_{\mathbb{S}^n}u_k\det(\bar{\nabla}^2_{ij} u_k+u_k\bar{g}_{ij})\,d\theta= \frac{\omega_n}{n+1}\,\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}u_k\det(\bar{\nabla}^2_{ij} u_k+u_k\bar{g}_{ij})\\ &=|B(1)|\,\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}f_k=|B(1)|, \end{align*} $$

and hence $\lambda _k=1$ . In particular, (4.3) yields

$$ \begin{align*}\mathcal{E}_{1, f_k} (\lambda_k\Omega_k)\leq \mathcal{E}_{1, f_k} (B(1))\leq \log 2. \end{align*} $$

Since $\mathcal {E}_{1, f_k} (\Omega _k)$ is bounded, (5.11) and Theorem 2.1 imply that the sequence $\Omega _k$ stays bounded, as well. Therefore, the claim (5.9) yields Theorem 5.3 if $\alpha =1$ .

Case 3: $\frac 1{n+2}<\alpha <1$ .

We set $p=1-\frac 1\alpha \in (-n-1,0)$ and $r=\frac {n+1}{n+1+p}>1$ , and

(5.12) $$ \begin{align} \tau= \frac12\cdot 2^{-\frac{|p|(n+1)}{|p|+n}}, \end{align} $$

and choose $\delta \in (0,\frac 12)$ such that

$$ \begin{align*}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(z^\bot\cap \mathbb{S}^n,2\delta)}f^r\leq \frac{\tau^r}{2^r} \end{align*} $$

for any $z\in S^{n-1}$ . We deduce from Lemma 5.1 that if $z\in S^{n-1}$ , then

(5.13) $$ \begin{align} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\Psi(z^\bot\cap \mathbb{S}^n,\delta)}f_k^r\leq \tau^r. \end{align} $$

We deduce from (5.5), (5.7), and $|\lambda _k\Omega _k|=|B(1)|=\frac {\omega _n}{n+1}$ that

(5.14) $$ \begin{align} \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}u_k^{p}f_k=\frac{n+1}{\omega_n} \int_{\mathbb{S}^n}u_k\,dS_{\Omega_k}= \frac{n+1}{\omega_n}\,|\Omega_k|=\lambda_k^{-n-1}. \end{align} $$

In particular, (4.3) and the upper bound on the entropy yield that

(5.15) $$ \begin{align} \nonumber 2^{p}&\leq \exp\left(p\cdot \mathcal{E}_{\alpha, f_k} (B(1))\right) \leq \exp\left(p\cdot \mathcal{E}_{\alpha, f} (\lambda_k\Omega_k)\right) \leq \frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}(\lambda_ku_k)^{p}f_k\\ &=\lambda_k^{p}\int_{\mathbb{S}^n}u_k\,dS_{\Omega_k}= \lambda_k^{p-n}\cdot\frac{n+1}{\omega_n}\cdot|\lambda_k\Omega_k|=\lambda_k^{p-n}. \end{align} $$

It follows from (5.15) that $\lambda _k\leq 2^{\frac {|p|}{|p|+n}}$ , and in turn (5.14) yields that

$$ \begin{align*}\frac{\ \ }{\ \ }{\hskip -0.4cm}\int_{\mathbb{S}^n}u_k^{p}f_k\geq 2^{-\frac{|p|(n+1)}{|p|+n}}. \end{align*} $$

Therefore, $\tau \leq \frac 12\frac {\ \ }{\ \ }{\hskip -0.4cm}\int _{\mathbb {S}^n}u_k^{p}f_k$ (cf. (5.12)), (5.13), and Theorem 2.1 yield that the sequence $\{\Omega _k\}$ is bounded, and in turn the claim (5.9) implies Theorem 5.3 if $\frac 1{n+2}<\alpha <1$ .

Footnotes

Böröczky is supported by OTKA 132002

1 We would like to thank referee for pointing out that the lemma was proved as Theorem 2.2 in [Reference Aldaz1]. Here, we provide a proof for completeness.

References

Aldaz, M., A stability version of Hölder’s inequality . J. Math. Anal. Appl. 343(2008), 842852.CrossRefGoogle Scholar
Andrews, B., Monotone quantities and unique limits for evolving convex hypersurfaces . Int. Math. Res. Not. IMRN 20(1997), 10011031.CrossRefGoogle Scholar
Andrews, B., Gauss curvature flow: the fate of rolling stone . Invent. Math. 138(1999), 151161.CrossRefGoogle Scholar
Andrews, B., Motion of hypersurfaces by gauss curvature . Pacific J. Math. 195(2000), no. 1, 134.CrossRefGoogle Scholar
Andrews, B., Böröczky, K., Guan, P. and Ni, L., Entropy and anisotropic flow by power of Gauss curvature, preprint, 2018.Google Scholar
Andrews, B., Guan, P., and Ni, L., Flow by the power of the gauss curvature . Adv. Math. 299(2016), 174201.CrossRefGoogle Scholar
Bianchi, G., Böröczky, K. J., and Colesanti, A., Smoothness in the ${L}^p$ Minkowski problem for $p<1$ . J. Geom. Anal. 30(2020), 680705.CrossRefGoogle Scholar
Bianchi, G., Böröczky, K. J., Colesanti, A., and Yang, D., The ${L}^p$ -Minkowski problem for $-n<p<1$ according to Chou-Wang. Adv. Math. 341(2019), 493535.CrossRefGoogle Scholar
Böröczky, K. J., The logarithmic Minkowski conjecture and the Lp-Minkowski problem . In: Koldobsky, A. and Volberg, A. (eds.), Harmonic analysis and convexity, DeGryuter, preprint, 2022. arxiv:2210.00194 Google Scholar
Böröczky, K. J. and Hegedűs, P., The cone volume measure of antipodal points . Acta Math. Hungar. 146(2015), 449465.CrossRefGoogle Scholar
Böröczky, K. J., Lutwak, E., Yang, D., and Zhang, G., The logarithmic Minkowski problem . J. Amer. Math. Soc. 26(2013), no. 3, 831852.CrossRefGoogle Scholar
Böröczky, K. J. and Trinh, H. T., The planar ${L}^p$ -Minkowski problem for $0<p<1$ . Adv. Appl. Math. 87(2017), 5881.CrossRefGoogle Scholar
Brendle, S., Choi, K., and Daskalopoulos, P., Asymptotic behavior of flows by powers of the Gaussian curvature . Acta Math. 219(2017), 116.CrossRefGoogle Scholar
Bryan, P., Ivaki, M., and Scheuer, J., A unified flow approach to smooth, even Lp-Minkowski problems . Analysis & PDE 12(2019), 259280.CrossRefGoogle Scholar
Caffarelli, L., A localization property of viscosity solutions to the Monge–Ampère equation and their strict convexity . Ann. Math. 131(1990), no. 1, 129134.CrossRefGoogle Scholar
Caffarelli, L., Interior ${W}^{2,p}$ estimates for solutions of the Monge–Ampère equation. Ann. Math. 131(1990), no. 1, 135150.CrossRefGoogle Scholar
Chen, S., Li, Q.-R., and Zhu, G., On the ${L}^p$ Monge–Ampère equation. J. Differential Equations 263(2017), 49975011.CrossRefGoogle Scholar
Chen, S., Li, Q.-R., and Zhu, G., The logarithmic Minkowski problem for non-symmetric measures . Trans. Amer. Math. Soc. 371(2019), 26232641.CrossRefGoogle Scholar
Choi, K., Kim, M., and Lee, T., Curvature bound for ${L}^p$ Minkowski problem, preprint, 2023. arXiv:2304.11617 Google Scholar
Chou, K.-S. and Wang, X.-J., The ${L}^p$ -Minkowski problem and the Minkowski problem in centroaffine geometry . Adv. Math. 205(2006), no. 1, 3383.CrossRefGoogle Scholar
Crasta, G. and Fragalá, I., Variational worn stones, preprint, 2023. arXiv:2303.11764 Google Scholar
Du, S.-Z., On the planar ${L}^p$ -Minkowski problem, J. Differential Equations 287 (2021), 3777.CrossRefGoogle Scholar
Firey, W.-J., On the shapes of worn stones . Mathematika 21(1974), 111.CrossRefGoogle Scholar
Guan, P., Huang, J., and Liu, J., Non-homogeneous fully nonlinear contracting flows of convex hypersurfaces, to appear in Adv. Nonlinear Stud. (special issue in honour of Joel Spruck).Google Scholar
Guan, P. and Lin, C. S., On equation $\det \left({u}_{ij}+{\delta}_{ij}u\right)={u}^pf$ on ${S}^n$ , preprint, 1999.Google Scholar
Guan, P. and Ni, L., Entropy and a convergence theorem for gauss curvature flow in high dimension . J. Eur. Math. Soc. 19(2017), no. 12, 37353761.CrossRefGoogle Scholar
Guang, Q., Li, Q.-R., and Wang, X.-J., The ${L}^p$ -Minkowski problem with super-critical exponents, preprint, 2022. arXiv:2203.05099 Google Scholar
Guang, Q., Li, Q.-R., and Wang, X.-J., Existence of convex hypersurfaces with prescribed centroaffine curvature. https://person.zju.edu.cn/person/attachments/2022-02/01-1645171178-851572.pdf Google Scholar
Hug, D., Contributions to affine surface area . Manuscripta Math. 91(1996), 283301.CrossRefGoogle Scholar
Hug, D., Lutwak, E., Yang, D., and Zhang, G., On the ${L}_p$ -Minkowski problem for polytopes . Discrete Comput. Geom. 33(2005), 699715.CrossRefGoogle Scholar
Ivaki, M., On the stability of the ${L}^p$ -curvature . J. Funct. Anal. 283(2022), 109684.CrossRefGoogle Scholar
Ivaki, M. and Milman, E., Uniqueness of solutions to a class of isotropic curvature problems, preprint, 2023. arXiv:2304.12839 CrossRefGoogle Scholar
Jian, H., Lu, J., and Wang, X.-J., Nonuniqueness of solutions to the ${L}^p$ -Minkowski problem . Adv. Math. 281(2015), 845856.CrossRefGoogle Scholar
Jian, H., Lu, J., and Zhu, G., Mirror symmetric solutions to the centro-affine Minkowski problem . Calc. Var. 55(2016), Article no. 41, 22 pp.CrossRefGoogle Scholar
Kannan, R., Lovász, L., and Simonovits, M., Isoperimetric problems for convex bodies and a localization lemma . Discrete Comput. Geom. 13(1995), 541559.CrossRefGoogle Scholar
Li, Q.-R., Infinitely many solutions for the centro-affine Minkowski problem . Int. Math. Res. Not. IMRN 2019(2019), 55775596.CrossRefGoogle Scholar
Li, Q.-R., Liu, J., and Lu, J., Non-uniqueness of solutions to the dual ${L}^p$ -Minkowski problem . Int. Math. Res. Not. IMRN 2022(2022), 91149150.CrossRefGoogle Scholar
Ludwig, M., General affine surface areas . Adv. Math. 224(2010), 23462360.CrossRefGoogle Scholar
Lutwak, E., Selected affine isoperimetric inequalities . In: Handbook of convex geometry, North-Holland, Amsterdam, 1993, pp. 151176.CrossRefGoogle Scholar
Lutwak, E., The Brunn–Minkowski–Firey theory. I. Mixed volumes and the Minkowski problem . J. Differ. Geom. 38(1993), 131150.CrossRefGoogle Scholar
Lutwak, E., The Brunn–Minkowski–Firey theory. II. Affine and geominimal surface areas . Adv. Math. 118(1996), 244294.CrossRefGoogle Scholar
Milman, E., A sharp centro-affine isospectral inequality of Szegő–Weinberger type and the ${L}^p$ -Minkowski problem. J. Diff. Geom., preprint, 2022. arXiv:2103.02994 Google Scholar
Saroglou, C., A non-existence result for the ${L}^p$ -Minkowski problem, preprint, 2021. arXiv:2109.06545 Google Scholar
Saroglou, C., On a non-homogeneous version of a problem of Firey . Math. Ann. 382(2022), 10591090.CrossRefGoogle Scholar
Schneider, R., Convex bodies: the Brunn–Minkowski theory, Encyclopedia of Mathematics, 44, Cambridge University Press, Cambridge, 1993.CrossRefGoogle Scholar
Stancu, A., Prescribing centro-affine curvature from one convex body to another . Int. Math. Res. Not. IMRN 2022(2022), 10161044.CrossRefGoogle Scholar
Zhu, G., The ${L}^p$ -Minkowski problem for polytopes for $p<0$ . Indiana Univ. Math. J. 66(2017), 13331350.CrossRefGoogle Scholar