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

Weighted Berwald’s Inequality

Dylan Langharst Dylan Langharst
Institut de Mathématiques de Jussieu, Sorbonne Université
Paris, 75252 France
dylan.langharst@imj-prg.fr
 and  Eli Putterman Eli Putterman
School of Mathematical Sciences, Tel Aviv University
Tel Aviv 66978, Israel
putterman@mail.tau.ac.il
Abstract.

The inequality of Berwald is a reverse-Hölder like inequality for the ppth average, p(1,),p\in(-1,\infty), of a non-negative, concave function over a convex body in n.\mathbb{R}^{n}. We prove Berwald’s inequality for averages of functions with respect to measures that have some concavity conditions, e.g. ss-concave measures, s.s\in\mathbb{R}. We also obtain equality conditions; in particular, this provides a new proof for the equality conditions of the classical inequality of Berwald. As applications, we generalize a number of classical bounds for the measure of the intersection of a convex body with a half-space and also the concept of radial means bodies and the projection body of a convex body.

Key words and phrases:
Berwald’s Inequality, Projection Bodies, Radial Mean Bodies, Zhang’s Inequality, Petty Projection Inequality.
1991 Mathematics Subject Classification:
52A39, 52A41; 28A75
The first named author was supported in part by the U.S. National Science Foundation Grant DMS-2000304 and the United States - Israel Binational Science Foundation (BSF) Grant 2018115. This work was completed while the first named author was a postdoctoral researcher funded by a Fondation Sciences Mathématiques de Paris fellowship. Both authors were supported during Fall 2022 by the National Science Foundation under Grant DMS-1929284 while in residence at the Institute for Computational and Experimental Research in Mathematics in Providence, RI, during “Harmonic Analysis and Convexity” program and during Spring 2024 by the Hausdorff Research Institute for Mathematics in Bonn, Germany, while in residence for the Dual Trimester Program: ”Synergies between modern probability, geometric analysis and stochastic geometry”.

1. Introduction

Let n\mathbb{R}^{n} be the standard nn-dimensional real vector space with the Euclidean structure. We write Volm(C)\text{\rm Vol}_{m}(C) for the mm-dimensional Lebesgue measure (volume) of a measurable set CnC\subset\mathbb{R}^{n}, where m=1,,nm=1,...,n is the dimension of the minimal affine space containing CC. The volume of the unit ball B2nB_{2}^{n} is written as κn,\kappa_{n}, and its boundary, the unit sphere, will be denoted as usual 𝕊n1.\mathbb{S}^{n-1}. A set KnK\subset\mathbb{R}^{n} is said to be convex if for every x,yKx,y\in K and λ[0,1],\lambda\in[0,1], (1λ)x+λyK.(1-\lambda)x+\lambda y\in K. We say KK is a convex body if it is a convex, compact set with non-empty interior; the set of all convex bodies in n\mathbb{R}^{n} will be denoted by 𝒦n\mathcal{K}^{n}. The set of those convex bodies containing the origin will be denoted 𝒦0n.\mathcal{K}^{n}_{0}. A convex body KK is centrally symmetric, or just symmetric, if K=KK=-K. There exists an addition on the set of convex bodies: the Minkowski sum of KK and LL, and one has that K+L={a+b:aK,bL}.K+L=\{a+b:a\in K,b\in L\}.

We recall that a non-negative function ff is said to be concave on n\mathbb{R}^{n} if for every x,ynx,y\in\mathbb{R}^{n} and λ[0,1]\lambda\in[0,1] one has

f((1λ)x+λy)(1λ)f(x)+λf(y),f((1-\lambda)x+\lambda y)\geq(1-\lambda)f(x)+\lambda f(y),

and that the support of a function is precisely supp(f)={xn:f(x)>0}¯.\text{supp}(f)=\overline{\{x\in\mathbb{R}^{n}:f(x)>0\}}. One can see that a non-negative, concave function will be supported on a convex set. It is easy to show if a non-negative, concave function takes the value infinity anywhere on its support, then the function is identically infinity on the interior of its support from convexity; therefore, throughout this paper, given a non-negative, concave function f,f, we shall assume it is not identically infinity, and so ff will have a finite maximum value, denoted f\|f\|_{\infty}. If KK is the support of a non-negative, concave function ff, then Kt={xn:f(x)t}={ft}K_{t}=\{x\in\mathbb{R}^{n}:f(x)\geq t\}=\{f\geq t\} are the level sets of ff. Notice that the level sets are also convex. Additionally, if f=f(0),\|f\|_{\infty}=f(0), then 0Kt0\in K_{t} for all tf.t\leq\|f\|_{\infty}. If ff is even, then KK is symmetric and so too is each Kt.K_{t}. In any case, if KK is also bounded, then each K,Kt𝒦nK,K_{t}\in\mathcal{K}^{n} (for each tft\leq\|f\|_{\infty}).

We next recall that the classical Berwald inequality states that if ff is a non-negative, concave function supported on some convex set Kn,K\subset\mathbb{R}^{n}, then, the function given by

(1) tf(p)=((n+pp)1Voln(K)Kfp(x)𝑑x)1/pt_{f}(p)=\left({\binom{n+p}{p}}\frac{1}{\text{\rm Vol}_{n}(K)}\int_{K}f^{p}(x)dx\right)^{1/p}

is decreasing for p(1,)p\in(-1,\infty) [4] with equality [22] if and only if the graph of ff is a certain cone with KK as its base. Here, the combinatorial coefficients are given by (mp)=Γ(m+1)Γ(p+1)Γ(mp+1),\binom{m}{p}=\frac{\Gamma(m+1)}{\Gamma(p+1)\Gamma(m-p+1)}, with Γ(z)\Gamma(z) the standard Gamma function, defined for zz\in\mathbb{C} except for when zz is negative integer. Usually written in the form tf(q)tf(p)t_{f}(q)\leq t_{f}(p) for 1<pq<,-1<p\leq q<\infty, Berwald’s inequality has several applications in the fields of convex geometry and probability theory, see for example [22, 39, 5, 24]. The first goal of this paper is to establish generalizations of Berwald’s inequality for measures with density under certain concavity assumptions. We will also analyze equality conditions; in particular, we obtain equality conditions for the classical Berwald inequality by a method independent of other proofs (e.g., [1, 22, 9]). To accomplish these tasks, we first prove a generalized Berwald’s inequality, Lemma 2.1.

We will say a Borel measure μ\mu has density if it has a locally integrable Radon-Nikodym derivative from n\mathbb{R}^{n} to +\mathbb{R}^{+}, i.e,

dμ(x)dx=ϕ(x), with ϕ:n+,ϕLloc1(n).\frac{d\mu(x)}{dx}=\phi(x),\text{ with }\phi\colon\mathbb{R}^{n}\to\mathbb{R}^{+},\phi\in L^{1}_{\text{loc}}(\mathbb{R}^{n}).

A Borel measure μ\mu on n\mathbb{R}^{n} is said to be FF-concave on a class 𝒞\mathcal{C} of compact subsets of n\mathbb{R}^{n} if there exists a continuous, (strictly) monotonic, invertible function F:(0,μ(n))(,)F:(0,\mu(\mathbb{R}^{n}))\to(-\infty,\infty) such that, for every pair A,B𝒞A,B\in\mathcal{C} and every t[0,1]t\in[0,1], one has

μ(tA+(1t)B)F1(tF(μ(A))+(1t)F(μ(B))).\mu(tA+(1-t)B)\geq F^{-1}\left(tF(\mu(A))+(1-t)F(\mu(B))\right).

When F(x)=xs,s>0F(x)=x^{s},s>0 this can be written as

μ(tA+(1t)B)stμ(A)s+(1t)μ(B)s,\mu(tA+(1-t)B)^{s}\geq t\mu(A)^{s}+(1-t)\mu(B)^{s},

and we say μ\mu is ss-concave. When s=1s=1, we merely say the measure is concave. In the limit as s0s\rightarrow 0, we obtain the case of log-concavity, which can also be obtained by taking F(x)=logxF(x)=\log x:

μ(tA+(1t)B)μ(A)tμ(B)1t.\mu(tA+(1-t)B)\geq\mu(A)^{t}\mu(B)^{1-t}.

The classical Brunn-Minkowski inequality (see for example [21]) asserts the 1/n1/n-concavity of the Lebesgue measure on the class of all compact subsets of n\mathbb{R}^{n}. From Borell’s classification on concave measures [7], a Radon measure (locally finite and regular Borel measure) is log-concave on Borel subsets of n\mathbb{R}^{n} if, and only if, μ\mu has a density ϕ(x)\phi(x) that is log-concave, i.e. ϕ(x)=Aeψ(x),\phi(x)=Ae^{-\psi(x)}, where A>0A>0 and ψ:n+\psi:\mathbb{R}^{n}\to\mathbb{R}^{+} is convex. Similarly, a Radon measure is ss-concave on Borel subsets of n\mathbb{R}^{n}, s(,0)(0,1/n),s\in(-\infty,0)\cup(0,1/n), if, and only if, μ\mu has a density ϕ(x)\phi(x) that is pp-concave (if s>0s>0) or pp-convex (if s<0s<0), where p=s/(1ns).p=s/(1-ns). However, all we will require is that a measure is ss-concave on a class of convex sets; we will discuss an important example below. Thus, our results in the case of ss-concave measures include measures beyond Borell’s classification.

We can now state our first main result, which is the Berwald inequality for FF-concave measures under different restrictions on the function FF. This result applies to a variety of measures, including ss-concave ones.

Theorem 1.1 (The Berwald Inequality for measures with concavity).

Let ff be a non-negative, concave function supported on KnK\subset\mathbb{R}^{n}. Let μ\mu be a Borel measure such that 0<μ(K)<0<\mu(K)<\infty and μ\mu satisfies one of the below listed concavity assumptions on a collection of convex subsets of KK containing the level sets of ff. Then, for any 1<pq<pmax-1<p\leq q<p_{\max} we have

C(p,μ,K)(1μ(K)Kf(x)p𝑑μ(x))1/pC(q,μ,K)(1μ(K)Kf(x)q𝑑μ(x))1/q,C(p,\mu,K)\left(\frac{1}{\mu(K)}\int_{K}f(x)^{p}d\mu(x)\right)^{1/p}\geq C(q,\mu,K)\left(\frac{1}{\mu(K)}\int_{K}f(x)^{q}d\mu(x)\right)^{1/q},

where

  1. (1)

    If μ\mu is FF-concave, where F:[0,μ(K)][0,)F:[0,\mu(K)]\to[0,\infty) is a continuous, increasing and invertible function: C(p,μ,K)=C(p,\mu,K)=

    {(pμ(K)01F1[F(μ(K))(1t)]tp1𝑑t)1pfor p>0(pμ(K)01tp1(F1[F(μ(K))(1t)]μ(K))𝑑t+1)1pfor p(1,0).\displaystyle\begin{cases}\left(\frac{p}{\mu(K)}\int_{0}^{1}F^{-1}\left[F(\mu(K))(1-t)\right]t^{p-1}dt\right)^{-\frac{1}{p}}&\text{for }p>0\\ \left(\frac{p}{\mu(K)}\int_{0}^{1}t^{p-1}(F^{-1}\left[F(\mu(K))(1-t)\right]-\mu(K))dt+1\right)^{-\frac{1}{p}}&\text{for }p\in(-1,0).\end{cases}

    There is equality if, and only if, F(0)=0,F(0)=0, for all t[0,f]t\in[0,\|f\|_{\infty}] the following formula holds

    μ({ft})=F1[F(μ(K))(1tf)],\mu(\{f\geq t\})=F^{-1}\left[F(\mu(K))\left(1-\frac{t}{\|f\|_{\infty}}\right)\right],

    and for all p(1,),p\in(-1,\infty), f\|f\|_{\infty} must satisfy

    f=C(p,μ,K)(1μ(K)Kf(x)p𝑑μ(x))1/p.\|f\|_{\infty}=C(p,\mu,K)\left(\frac{1}{\mu(K)}\int_{K}f(x)^{p}d\mu(x)\right)^{1/p}.
  2. (2)

    If μ\mu is QQ-concave, where Q:(0,μ(K)](,)Q:(0,\mu(K)]\to(-\infty,\infty) is a continuous, increasing and invertible function: C(p,μ,K)=C(p,\mu,K)=

    {(pμ(K)0Q1[Q(μ(K))t]tp1𝑑t)1pfor p>0(pμ(K)0tp1(Q1[Q(μ(K)t)]μ(K))𝑑t)1pfor p(1,0).\begin{cases}\left(\frac{p}{\mu(K)}\int_{0}^{\infty}Q^{-1}\left[Q(\mu(K))-t\right]t^{p-1}dt\right)^{-\frac{1}{p}}&\text{for }p>0\\ \left(\frac{p}{\mu(K)}\int_{0}^{\infty}t^{p-1}(Q^{-1}\left[Q(\mu(K)-t)\right]-\mu(K))dt\right)^{-\frac{1}{p}}&\text{for }p\in(-1,0).\end{cases}

    Equality is never obtained.

  3. (3)

    If μ\mu is RR-concave, where R:(0,μ(K)](0,)R:(0,\mu(K)]\to(0,\infty) is a continuous, decreasing and invertible function: C(p,μ,K)=C(p,\mu,K)=

    {(pμ(K)0R1[R(μ(K))(1+t)]tp1𝑑t)1pfor p>0(pμ(K)0tp1(R1[R(μ(K))(1+t)]μ(K))𝑑t)1pfor p(1,0).\begin{cases}\left(\frac{p}{\mu(K)}\int_{0}^{\infty}R^{-1}\left[R(\mu(K))(1+t)\right]t^{p-1}dt\right)^{-\frac{1}{p}}&\text{for }p>0\\ \left(\frac{p}{\mu(K)}\int_{0}^{\infty}t^{p-1}(R^{-1}\left[R(\mu(K))(1+t)\right]-\mu(K))dt\right)^{-\frac{1}{p}}&\text{for }p\in(-1,0).\end{cases}

    Equality is never obtained.

In all cases, pmaxp_{\max} is defined implicitly via pmax=sup{p>0:Tf(p)<},p_{\max}=\sup\{p>0:T_{f}(p)<\infty\}, where

Tf(p)=C(p,μ,K)(1μ(K)Kf(x)p𝑑μ(x))1/p.T_{f}(p)=C(p,\mu,K)\left(\frac{1}{\mu(K)}\int_{K}f(x)^{p}d\mu(x)\right)^{1/p}.

Tf(0)T_{f}(0) is defined via continuity.

We remark that cases 2 and 3 of Theorem 1.1 have a strict inequality due to the fact, for Case 2, that Q1[Q(μ(K))t]Q^{-1}[Q(\mu(K))-t] being integrable implies Q1()=0,Q^{-1}(-\infty)=0, or Q(0)=.Q(0)=-\infty. On the other hand, we will show that if there is equality, then |Q(0)||Q(0)| would be finite. Similar logic holds for Case 3. However, the inequality is asymptotically sharp as ff is made arbitrarily large on its support.

We obtain the following corollary for ss-concave measures; the case where s<0s<0 was previously done by Fradelizi, Guédon and Pajor [19], by modifying Borell’s proof [8] of the classical inequality of Berwald. Presented in [20] is a proof for all s,s\in\mathbb{R}, based on techniques from a work by Koldobsky, Pajor and Yaskin [27]. Both extensions do not mention equality conditions.

Corollary 1.2 (The Berwald Inequality for ss-concave measures).

Let ff be a non-negative concave function supported on Kn.K\subset\mathbb{R}^{n}. Let μ\mu be a Borel measure finite on KK that is ss-concave, ss\in\mathbb{R}, on a collection of convex subsets of KK containing the level sets of ff. Then, for any 1<pq<-1<p\leq q<\infty we have

(C(p,s)μ(K)Kf(x)p𝑑μ(x))1/p(C(q,s)μ(K)Kf(x)q𝑑μ(x))1/q,\left(\frac{C(p,s)}{\mu(K)}\int_{K}f(x)^{p}d\mu(x)\right)^{1/p}\geq\left(\frac{C(q,s)}{\mu(K)}\int_{K}f(x)^{q}d\mu(x)\right)^{1/q},

where

C(p,s)={(1s+pp)for s>0,Γ(p+1)1if s=0,s(p+1s)(1sp)for s<0.C(p,s)=\begin{cases}{{\frac{1}{s}+p}\choose p}&\text{for }s>0,\\ \Gamma(p+1)^{-1}&\text{if }s=0,\\ s\left(p+\frac{1}{s}\right){{-\frac{1}{s}}\choose p}&\text{for }s<0.\end{cases}

For s<0,s<0, we must restrict to p(1,1/s)p\in(-1,-1/s) for integrability.

If s>0,s>0, there is equality if, and only if, for all t[0,f]t\in[0,\|f\|_{\infty}] and p(1,):p\in(-1,\infty):

μ({xK:f(x)t})=μ(K)(1tf)1/s\mu(\{x\in K:f(x)\geq t\})=\mu(K)\left(1-\frac{t}{\|f\|_{\infty}}\right)^{1/s}

implying

fp=(1s+pp)1μ(K)Kf(x)p𝑑μ(x).\|f\|^{p}_{\infty}={{\frac{1}{s}+p}\choose p}\frac{1}{\mu(K)}\int_{K}f(x)^{p}d\mu(x).

If s=0s=0 or s<0,s<0, equality is never obtained.

The equality conditions to Corollary 1.2 may seem a bit strange; we are able to obtain an exact formula for the function ff when the measure μ\mu is ss-concave and 1/s1/s-homogeneous, s(0,1/n]s\in(0,1/n]. Recall that a Borel measure μ\mu is said to be α\alpha-homogeneous for some α>0\alpha>0 if μ(tA)=tαμ(A)\mu(tA)=t^{\alpha}\mu(A) for all compact sets AsuppμA\subset\operatorname{supp}\mu and t>0t>0 so that tAsuppμ.tA\subset\operatorname{supp}\mu. If μ\mu has density ϕ\phi, then one can check using the Lebesgue differentiation theorem that this implies that ϕ\phi is (αn)(\alpha-n)-homogeneous.

We say a set LL with 0int(L)0\in\text{int}(L) is star-shaped if every line passing through the origin crosses the boundary of LL exactly twice. We say LL is a star body if it is a compact, star-shaped set whose radial function ρL:n{0},\rho_{L}:\mathbb{R}^{n}\setminus\{0\}\to\mathbb{R}, given by ρL(y)=sup{λ:λyL},\rho_{L}(y)=\sup\{\lambda:\lambda y\in L\}, is continuous. For K𝒦0n,K\in\mathcal{K}^{n}_{0}, the Minkowski functional of KK is defined to be yK=ρK1(y)=inf{r>0:yrK}.\|y\|_{K}=\rho^{-1}_{K}(y)=\inf\{r>0:y\in rK\}. The Minkowski functional K\|\cdot\|_{K} of K𝒦0nK\in\mathcal{K}^{n}_{0} is a norm on n\mathbb{R}^{n} if KK is symmetric. If xnx\in\mathbb{R}^{n} and LnL\subset\mathbb{R}^{n} satisfy that LxL-x is a star body, then the generalized radial function of LL at xx is defined by ρL(x,y):=ρLx(y)\rho_{L}(x,y):=\rho_{L-x}(-y). Note that for every K𝒦n,K\in\mathcal{K}^{n}, KxK-x is a star body for every xint(K)x\in\text{int}(K).

One gets the following formula for μ(K)\mu(K) when μ\mu is α\alpha-homogeneous, α>0\alpha>0, and KK is a star body in n\mathbb{R}^{n}.

(2) μ(K)=𝕊n10ρK(θ)ϕ(rθ)rn1𝑑r𝑑θ=𝕊n1ϕ(θ)0ρK(θ)rα1𝑑r𝑑θ=1α𝕊n1ϕ(θ)ρKα(θ)𝑑θ.\begin{split}\mu(K)&=\int_{\mathbb{S}^{n-1}}\int_{0}^{\rho_{K}(\theta)}\phi(r\theta)r^{n-1}drd\theta\\ &=\int_{\mathbb{S}^{n-1}}\phi(\theta)\int_{0}^{\rho_{K}(\theta)}r^{\alpha-1}drd\theta=\frac{1}{\alpha}\int_{\mathbb{S}^{n-1}}\phi(\theta)\rho_{K}^{\alpha}(\theta)d\theta.\end{split}

Crucial to the statement of equality conditions, and our investigations henceforth, will be the roof function associated to a star body KK, which we define as K(0)=1,K(x)=0\ell_{K}(0)=1,\ell_{K}(x)=0 for xKx\neq K and, for xK{0},x\in K\setminus\{0\}, K(x)=(11ρK(x)).\ell_{K}(x)=\left(1-\frac{1}{\rho_{K}(x)}\right). In polar coordinates, K(rθ)\ell_{K}(r\theta) becomes an affine function in rr for r[0,ρK(θ)]r\in[0,\rho_{K}(\theta)]:

(3) K(rθ)=(1rρK(θ)).\ell_{K}(r\theta)=\left(1-\frac{r}{\rho_{K}(\theta)}\right).

Note that if K𝒦0n,K\in\mathcal{K}^{n}_{0}, then we can also write K(x)=1xK\ell_{K}(x)=1-\|x\|_{K} for xKx\in K and 0 otherwise. Observe that, for a non-negative, concave function supported on some K𝒦0nK\in\mathcal{K}^{n}_{0} one obtains for θ𝕊n1\theta\in\mathbb{S}^{n-1} and r[0,ρK(θ)]r\in[0,\rho_{K}(\theta)] that

(4) f(rθ)=f((rρK(θ)ρK(θ)+0(1rρK(θ)))θ)rρK(θ)f(ρK(θ)θ)+f(0)K(rθ)f(0)K(rθ);\begin{split}f(r\theta)&=f\left(\left(\frac{r}{\rho_{K}(\theta)}\rho_{K}(\theta)+0\left(1-\frac{r}{\rho_{K}(\theta)}\right)\right)\theta\right)\\ &\geq\frac{r}{\rho_{K}(\theta)}f(\rho_{K}(\theta)\theta)+f(0)\ell_{K}(r\theta)\geq f(0)\ell_{K}(r\theta);\end{split}

we will make liberal use of this bound throughout this work. Functions of the form f(x)=MKx0(xx0)f(x)=M\ell_{K-x_{0}}(x-x_{0}) for some M>0M>0 and x0Kx_{0}\in K will also be referred to roof functions, with height MM and vertex x0x_{0}. The reason for this vocabulary will become clearer below.

Using (2), one can verify by hand that the function K(x)\ell_{K}(x) satisfies, for μ\mu an ss-concave, 1/s1/s-homogeneous measure, that

KK(x)p𝑑μ(x)=(1s+p1s)1μ(K).\int_{K}\ell_{K}(x)^{p}d\mu(x)={{\frac{1}{s}+p}\choose\frac{1}{s}}^{-1}\mu(K).

Therefore, K(x)\ell_{K}(x) yields equality in the Berwald inequality for ss-concave measures, Corollary 1.2, under the additional assumption that μ\mu is 1/s1/s-homogeneous. The next theorem shows this is the only such function.

Theorem 1.3.

(The Berwald Inequality for ss-concave, 1/s1/s-homogeneous measures) Let ff be a non-negative, concave function supported on KnK\subset\mathbb{R}^{n}. Let μ\mu be a Radon measure containing KK in its support that is ss-concave, 1/s1/s-homogeneous for some s(0,1/n]s\in(0,1/n]. Then, for any 1<pq<-1<p\leq q<\infty we have

((1s+pp)1μ(K)Kf(x)p𝑑μ(x))1/p((1s+qq)1μ(K)Kf(x)q𝑑μ(x))1/q.\left({{\frac{1}{s}+p}\choose p}\frac{1}{\mu(K)}\int_{K}f(x)^{p}d\mu(x)\right)^{1/p}\geq\left({{\frac{1}{s}+q}\choose q}\frac{1}{\mu(K)}\int_{K}f(x)^{q}d\mu(x)\right)^{1/q}.

Suppose f=f(0)\|f\|_{\infty}=f(0). Then, there is equality if, and only if, f(rθ)f(r\theta) is an affine function in r.r. i.e. one has f(x)=fK(x).f(x)=\|f\|_{\infty}\ell_{K}(x).

In our applications below, we will always be considering functions whose maximum is obtained at the origin, and so the minor constraint on the equality conditions does not hinder us. We now prove the classical Berwald inequality with equality conditions. Favard first conjectured the inequality in one dimension, and Berwald verified the inequality for all dimensions [4], without equality conditions. In fact, when n=1,n=1, Berwald was able to show the inequality is true for 1<pq<,-1<p\leq q<\infty, and this was extended to all dimensions by Borell [9]. However, the generality of his technique makes it difficult to establish where equality occurs.

Gardner and Zhang [22] gave a different proof, which yields that equality is satisfied in the classical Berwald inequality precisely when the graph of ff is a certain cone with KK as a base, i.e. that ff is a roof function. In Corollary 1.4, we obtain a proof using Theorem 1.3, verifying that our techniques reduce to the known result. We must also mention that this result was also obtained in [1, Theorem 7.2] via a different technique. In that work, the roof function was defined via its graph in n+1.\mathbb{R}^{n+1}. Specifically they constructed the roof function in the following way: given a convex set KnK\subset\mathbb{R}^{n} (which will become the base of a hypercone), let M>0M>0 be the height of the hypercone, and let x0Kx_{0}\in K be the location of the projection of vertex of the hypercone. Then, the roof function with height MM and vertex x0x_{0} is equivalently defined as the non-negative, concave function ff whose graph is given by

{(x,t)K×:0tf(x)}=conv(K×{0},{(x0,M)}),\{(x,t)\in K\times\mathbb{R}:0\leq t\leq f(x)\}=\text{conv}\left(K\times\{0\},\left\{\left(x_{0},M\right)\right\}\right),

where conv denotes the convex hull. From this formulation, we obtain an interesting formula for the level sets of a roof function f:f: for 0tM,0\leq t\leq M, one has that Kt=tMx0+(1tM)K.K_{t}=\frac{t}{M}x_{0}+(1-\frac{t}{M})K.

Corollary 1.4 (The Classical Berwald Inequality).

Let ff be a non-negative, concave function supported on K𝒦nK\in\mathcal{K}^{n}. Then, for any 1<pq<-1<p\leq q<\infty we have

((n+pp)1Voln(K)Kf(x)p𝑑x)1/p((n+qq)1Voln(K)Kf(x)q𝑑x)1/q.\left({{n+p}\choose p}\frac{1}{\text{\rm Vol}_{n}(K)}\int_{K}f(x)^{p}dx\right)^{1/p}\geq\left({{n+q}\choose q}\frac{1}{\text{\rm Vol}_{n}(K)}\int_{K}f(x)^{q}dx\right)^{1/q}.

There is equality if, and only if, f(rθ)f(r\theta) is an affine function in rr up to translation i.e. if x0x_{0} is the point in KK where the maximum of ff is obtained, one has f(x)=fKx0(xx0).f(x)=\|f\|_{\infty}\ell_{K-x_{0}}(x-x_{0}).

Proof.

The inequality follows immediately from Theorem 1.3, as do the equality conditions if the maximum of ff is obtained at the origin. If the maximum of ff is not obtained at the origin, let x0x_{0} be the point in KK where ff obtains its maximum. Let g(x)=f(x+x0)g(x)=f(x+x_{0}) and K~=Kx0.\widetilde{K}=K-x_{0}. Then, g(x)g(x) is a concave function supported on K~\widetilde{K} with maximum at the origin, and, for every p(1,0)(0,)p\in(-1,0)\cup(0,\infty)

1Voln(K)Kf(x)p𝑑x=1Voln(K~)K~g(x)p𝑑x.\frac{1}{\text{\rm Vol}_{n}(K)}\int_{K}f(x)^{p}dx=\frac{1}{\text{\rm Vol}_{n}(\widetilde{K})}\int_{\widetilde{K}}g(x)^{p}dx.

Therefore, since there is equality in the inequality for the function ff and the convex body KK by hypothesis, there is equality in the inequality for the function gg and the convex body K~\widetilde{K}. Consequently, we have

g(x)=gK~(x).g(x)=\|g\|_{\infty}\ell_{\widetilde{K}}(x).

Using that f(x)=g(xx0)f(x)=g(x-x_{0}) and g=f\|g\|_{\infty}=\|f\|_{\infty} yields the result. ∎

We next list two applications for the standard Gaussian measure on n,\mathbb{R}^{n}, which we recall is given by dγn(x)=1(2π)n/2e|x|2/2dx.d\gamma_{n}(x)=\frac{1}{(2\pi)^{n/2}}e^{-|x|^{2}/2}dx. From Borell’s classification, we see that the Gaussian measure is log-concave on n\mathbb{R}^{n} over any collection of compact sets closed under Minkowski summation. Thus, we can apply the second case of Corollary 1.2 and obtain a Berwald-type inequality for the Gaussian measure in this case. However, the Ehrhard inequality shows one can improve on the log-concavity of the Gaussian measure: For 0<t<10<t<1 and Borel sets KK and LL in n\mathbb{R}^{n}, we have

(5) 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}\left(\gamma_{n}(K)\right)+t\Phi^{-1}\left(\gamma_{n}(L)\right),

i.e. Φ1γn\Phi^{-1}\circ\gamma_{n} is concave, where Φ(x)=γ1((,x))\Phi(x)=\gamma_{1}((-\infty,x)). The inequality (5) was first proven by Ehrhard for the case of two closed, convex sets [15, 14]. Latała [31] generalized Ehrhard’s result to the case of an arbitrary Borel set KK and convex set LL; the general case for two Borel sets of the Ehrhard’s inequality was proven by Borell [10]. Since Φ\Phi is log-concave, the log-concavity of the Gaussian measure is strictly weaker than the Ehrhard inequality. Additionally, Kolesnikov and Livshyts showed that the Gaussian measure is 12n\frac{1}{2n} concave on 𝒦0n,\mathcal{K}^{n}_{0}, the set of convex bodies containing the origin in their interior [28]. That is, by restricting the admissible sets in the concavity equation, the concavity can improve.

Corollary 1.5 (Berwald-type inequalities for the Gaussian Measure).

Let ff be a non-negative, concave function supported on KnK\subset\mathbb{R}^{n}. Then, we have the following:

  1. (1)

    The function

    g1(p)=1Γ(p+1)1/p(1γn(K)Kf(x)p𝑑γn(x))1/pg_{1}(p)=\frac{1}{\Gamma(p+1)^{1/p}}\left(\frac{1}{\gamma_{n}(K)}\int_{K}f(x)^{p}d\gamma_{n}(x)\right)^{1/p}

    is strictly decreasing on (1,);(-1,\infty);

  2. (2)

    The function

    g2(p)=C(p,K)(1γn(K)Kf(x)p𝑑γn(x))1/pg_{2}(p)=C(p,K)\left(\frac{1}{\gamma_{n}(K)}\int_{K}f(x)^{p}d\gamma_{n}(x)\right)^{1/p}

    is strictly decreasing on (1,),(-1,\infty), where C(p,K)=C(p,K)=

    {(pγn(K)0Φ[Φ1(γn(K))t]tp1𝑑t)1pfor p>0(pγn(K)0tp1(Φ[Φ1(γn(K)t)]γn(K))𝑑t)1pfor p(1,0);\begin{cases}\left(\frac{p}{\gamma_{n}(K)}\int_{0}^{\infty}\Phi\left[\Phi^{-1}(\gamma_{n}(K))-t\right]t^{p-1}dt\right)^{-\frac{1}{p}}&\text{for }p>0\\ \left(\frac{p}{\gamma_{n}(K)}\int_{0}^{\infty}t^{p-1}(\Phi\left[\Phi^{-1}(\gamma_{n}(K)-t)\right]-\gamma_{n}(K))dt\right)^{-\frac{1}{p}}&\text{for }p\in(-1,0);\end{cases}
  3. (3)

    and, if the maximum of ff is at the origin and K𝒦0n,K\in\mathcal{K}^{n}_{0}, then the function

    g3(p)=((2n+pp)1γn(K)Kf(x)p𝑑γn(x))1/pg_{3}(p)=\left({{2n+p}\choose p}\frac{1}{\gamma_{n}(K)}\int_{K}f(x)^{p}d\gamma_{n}(x)\right)^{1/p}

    is decreasing on (1,).(-1,\infty).

The equality condition for the third case of Corollary 1.5 can be deduced from Theorem 1.1, so we do not explicitly state it. If one further restricts the admissible sets, one can do even better. The Gardner-Zvavitch inequality states for symmetric K,L𝒦0nK,L\in\mathcal{K}^{n}_{0} and t[0,1]t\in[0,1] that

(6) γn((1t)K+tL)1/n(1t)γn(K)1/n+tγn(L)1/n,\gamma_{n}\left((1-t)K+tL\right)^{1/n}\geq(1-t)\gamma_{n}(K)^{1/n}+t\gamma_{n}(L)^{1/n},

i.e. γn\gamma_{n} is 1/n1/n-concave over the class of symmetric convex bodies. This inequality was first conjectured in [23] by Gardner and Zvavitch; a counterexample was shown in [40] when KK and LL are not symmetric. Important progress was made in [28], which lead to the proof of the inequality (6) by Eskenazis and Moschidis in [16] for symmetric convex bodies. Recently, Cordero-Erausquin and Rotem [12] extended this result to the class

(7) n={Borel measures μ on n:dμ(x)=ew(|x|)dx,w:[0,)(,] is an increasing function such that tw(et) is convex}.\begin{split}\mathcal{M}_{n}=\bigg{\{}&\text{Borel measures }\mu\text{ on }\mathbb{R}^{n}:d\mu(x)=e^{-w(|x|)}dx,w:[0,\infty)\to(-\infty,\infty]\\ &\text{ is an increasing function such that }t\to w(e^{t})\text{ is convex}\bigg{\}}.\end{split}

That is, every measure μn\mu\in\mathcal{M}_{n} is 1/n1/n-concave over the class of symmetric convex bodies. To show how rich this class is, n\mathcal{M}_{n} includes not only every rotationally invariant, log-concave measure (e.g. Gaussian), but also Cauchy-type measures. Combining these results, we obtain a Berwald-type inequality.

Corollary 1.6 (Berwald-type inequality for rotationally invariant log-concave measures).

Let ff be a non-negative, concave, even function supported on a symmetric K𝒦0n.K\in\mathcal{K}^{n}_{0}. Let μ\mu be a measure in n\mathcal{M}_{n} containing KK in its support. Then, for any 1<pq<:-1<p\leq q<\infty:

((n+pp)1μ(K)Kf(x)p𝑑μ(x))1/p((n+qq)1μ(K)Kf(x)q𝑑μ(x))1/q.\left({{n+p}\choose p}\frac{1}{\mu(K)}\int_{K}f(x)^{p}d\mu(x)\right)^{1/p}\geq\left({{n+q}\choose q}\frac{1}{\mu(K)}\int_{K}f(x)^{q}d\mu(x)\right)^{1/q}.

We emphasize that the (1/2n)(1/2n)-concavity of the Gaussian measure on 𝒦0n\mathcal{K}^{n}_{0} shown in [28] and the 1/n1/n-concavity of γn\gamma_{n} and other measures from n\mathcal{M}_{n} over the class of symmetric convex bodies falls strictly outside the classification of ss-concave measures by Borell. This paper is organized as follows. In Section 2, we prove a version of Berwald’s inequality for FF-concave measures. In Section 3, we discuss surface area measure, projection bodies, and radial mean bodies. Then, we apply our results to weighted generalizations of radial mean bodies. Along the way, we obtain more inequalities of Rogers and Shephard and of Zhang type. We would like to mention here that weighted extensions of concepts from the Brunn-Minkowski theory is a very rich field. This includes works on the surface area measure [3, 41, 33, 32, 36] and general measure extensions of the projection body of a convex body [35, 30]. Recently, it has been shown that these developments, in particular the concavities for the Gaussian measure and Borell’s classification, have led to a burgeoning weighted Brunn-Minkowski theory, see [29, 17, 18].

Acknowledgments We would like to thank Artem Zvavitch for the helpful feedback throughout this work, and we also thank Matthieu Fradelizi for the discussion concerning Theorem 1.1 when p(1,0)p\in(-1,0). We would also like to thank Michael Roysdon for the discussions concerning this work, in particular the suggestion of Corollary 2.2. This work began during a visit to the Laboratoire d’Analyse et de Mathématiques Appliquées at Université Gustave Eiffel, France, from October to December 2021 and was continued during a visit to Tel Aviv University, Israel, in March and April 2022 - heartfelt thanks are extended, respectively, to Matthieu Fradelizi and Semyon Alesker.

2. Generalizations of Berwald’s Inequality

In this section, we establish a generalization of Berwald’s inequality. In what follows, for a finite Borel measure μ\mu and a Borel set KK with positive μ\mu-measure, μK\mu_{K} will denote the normalized probability on KK with respect to μ\mu, that is for measurable An:A\subset\mathbb{R}^{n}: μK(A)=μ(KA)μ(K).\mu_{K}(A)=\frac{\mu(K\cap A)}{\mu(K)}. Notice that for every non-negative, measurable function ff on KK and p>0p>0 such that fLp(μ,K)f\in L^{p}(\mu,K), one has the layer cake formula

1μ(K)Kfp(x)𝑑μ(x)=p0μK({ft})tp1𝑑t\frac{1}{\mu(K)}\int_{K}f^{p}(x)d\mu(x)=p\int_{0}^{\infty}\mu_{K}(\left\{f\geq t\right\})t^{p-1}dt

from the following use of Fubini’s theorem:

1μ(K)Kfp(x)𝑑μ(x)\displaystyle\frac{1}{\mu(K)}\int_{K}f^{p}(x)d\mu(x) =pμ(K)K0f(x)tp1𝑑t𝑑μ(x)\displaystyle=\frac{p}{\mu(K)}\int_{K}\int_{0}^{f(x)}t^{p-1}dtd\mu(x)
=pμ(K)0μ({xK:f(x)t})tp1𝑑t.\displaystyle=\frac{p}{\mu(K)}\int_{0}^{\infty}\mu(\left\{x\in K:f(x)\geq t\right\})t^{p-1}dt.

Additionally, if μ\mu is FF-concave, with FF increasing and invertible, on a class 𝒞\mathcal{C} of convex sets, then for K𝒞K\in\mathcal{C} in the support of a concave function ff, one has that the function given by fμ(t)=μK({ft})f_{\mu}(t)=\mu_{K}(\left\{f\geq t\right\}) is F~\tilde{F}-concave, where F~(x)=F(μ(K)x),\tilde{F}(x)=F(\mu(K)x), as long as the level sets of ff belong to 𝒞.\mathcal{C}. Indeed, since ff is concave, one has, for λ[0,1]\lambda\in[0,1] and u,v0u,v\geq 0, that

{f(1λ)u+λv}(1λ){fu}+λ{fv}.\{f\geq(1-\lambda)u+\lambda v\}\supset(1-\lambda)\{f\geq u\}+\lambda\{f\geq v\}.

Using the FF-concavity of μ,\mu, this yields

F(μ({f(1λ)u+λv}))(1λ)F(μ({fu}))+λF(μ({fv})).F\left(\mu\left(\{f\geq(1-\lambda)u+\lambda v\}\right)\right)\geq(1-\lambda)F\left(\mu(\{f\geq u\})\right)+\lambda F\left(\mu(\{f\geq v\})\right).

Inserting the definition of F~\tilde{F} and fμ,f_{\mu}, this is precisely

F~fμ((1λ)u+λv)(1λ)F~fμ(u)+λF~fμ(v).\tilde{F}\circ f_{\mu}\left((1-\lambda)u+\lambda v\right)\geq(1-\lambda)\tilde{F}\circ f_{\mu}(u)+\lambda\tilde{F}\circ f_{\mu}(v).

Similarly one can check that if μ\mu is RR-concave, with RR decreasing and invertible, on a class 𝒞\mathcal{C} of convex sets, then for K𝒞K\in\mathcal{C} in the support of a concave function ff, one then has that the function fμf_{\mu} is R~\tilde{R}-convex, where R~(x)=R(μ(K)x).\tilde{R}(x)=R(\mu(K)x). That is, R~fμ\tilde{R}\circ f_{\mu} is a convex function on its support, as long as the level sets of ff belong to 𝒞.\mathcal{C}.

We next need the appropriate layer cake formula for when p<0.p<0. Notice that for every non-negative, measurable function ff on a Borel set KK and p<0p<0 such that fLp(μ,K)f\in L^{p}(\mu,K) for a Borel measure μ\mu, one has the layer cake formula

1μ(K)Kfp(x)𝑑μ(x)=p0tp1(μK({ft})1)𝑑t\frac{1}{\mu(K)}\int_{K}f^{p}(x)d\mu(x)=p\int_{0}^{\infty}t^{p-1}(\mu_{K}(\left\{f\geq t\right\})-1)dt

from the following use of Fubini’s theorem:

1μ(K)Kfp(x)𝑑μ(x)\displaystyle\frac{1}{\mu(K)}\int_{K}f^{p}(x)d\mu(x) =pμ(K)Kf(x)tp1𝑑t𝑑μ(x)\displaystyle=-\frac{p}{\mu(K)}\int_{K}\int_{f(x)}^{\infty}t^{p-1}dtd\mu(x)
=pμ(K)0tp1(μ({xK:f(x)t})μ(K))𝑑t.\displaystyle=\frac{p}{\mu(K)}\int_{0}^{\infty}t^{p-1}(\mu(\left\{x\in K:f(x)\geq t\right\})-\mu(K))dt.

We now recall the analytic extension of the Gamma function. We start with the definition of Γ(z)\Gamma(z) when the real part of zz is greater than zero:

Γ(z)=0tz1et𝑑t.\Gamma(z)=\int_{0}^{\infty}t^{z-1}e^{-t}dt.

If the real part of zz is less than zero, then one uses analytic continuation to extend Γ\Gamma via the multiplicative property Γ(z+1)=zΓ(z)\Gamma(z+1)=z\Gamma(z). Now, let us obtain the formula for Γ(z)\Gamma(z) when the real part of zz is in (1,0)(-1,0). From the multiplicative property one can write

(8) Γ(z)=1z0tzet𝑑t=0tz1(et1)𝑑t,\Gamma(z)=\frac{1}{z}\int_{0}^{\infty}t^{z}e^{-t}dt=\int_{0}^{\infty}t^{z-1}(e^{-t}-1)dt,

where, for the second equality, integration by parts was performed and ete^{-t} was viewed as the derivative of 1et,1-e^{-t}, to maintain integrability. The fact that the layer cake formula looks similar to the formula for Γ(z)\Gamma(z) when the real part of zz is between 1-1 and 0 inspires the analytic continuation of Theorem 1.1 to negative pp. We will use the Mellin transformation, which was extended to p(1,0)p\in(-1,0) in [20] for ss-concave functions. We further generalize the Mellin transform here.

The Mellin transform of a function ψ\psi such that supp(ψ)[0,B)\text{supp}(\psi)\subseteq[0,B) is the analytic function for p(1,0)(0,)p\in(-1,0)\cup(0,\infty) given by ψ(p)=\mathcal{M}_{\psi}(p)=

(9) {0Btp1(ψ(t)ψ(0))𝑑t+Bppψ(0)for p(1,0),0Btp1ψ(t)𝑑tfor p>0 such that tp1ψ(t)L1().\begin{cases}\int_{0}^{B}t^{p-1}(\psi(t)-\psi(0))dt+\frac{B^{p}}{p}\psi(0)&\text{for }p\in(-1,0),\\ \int_{0}^{B}t^{p-1}\psi(t)dt&\text{for }p>0\text{ such that }t^{p-1}\psi(t)\in L^{1}(\mathbb{R}).\end{cases}

Following [20], consider the function

(10) ψs(t)={(1t)1/sχ[0,1](t)for s>0,etχ(0,)(t)for s=0,(1+t)1/sχ(0,)(t)for s<0.\psi_{s}(t)=\begin{cases}(1-t)^{1/s}\chi_{[0,1]}(t)&\text{for }s>0,\\ e^{-t}\chi_{(0,\infty)}(t)&\text{for }s=0,\\ (1+t)^{1/s}\chi_{(0,\infty)}(t)&\text{for }s<0.\end{cases}

Then, for all p>1,p>-1, one has ψs(p)1=pC(p,s),\mathcal{M}_{\psi_{s}}(p)^{-1}=pC(p,s), where C(p,s)C(p,s) is the constant defined in Corollary 1.2, that is Berwald’s inequality for ss-concave measures; notice again that in the case when s<0s<0, for tp1(1+t)1/st^{p-1}(1+t)^{1/s} to be integrable, we must have that p<1/s.p<-1/s.

Motivated by this example, we need to define a function whose Mellin transform is related to the constant C(p,μ,K)C(p,\mu,K) from Theorem 1.1, and this definition will depend on the concavity of μ\mu. Recall that a function ψ\psi is ff-concave for a monotonic function ff if fψf\circ\psi is either concave (if ff is increasing) or convex (if ff is decreasing). Similarly, ψ\psi is ff-affine if fψf\circ\psi is an affine function. We will have three different restrictions on the function ff, matching those in Theorem 1.1 (and the notation as well). First, fix some A>0A>0. Then, we will consider the case when f{F,Q,R},f\in\{F,Q,R\}, where FF represents those functions F:[0,A][0,)F:[0,A]\to[0,\infty) that are continuous, increasing and invertible; QQ represents those functions Q:(0,A](,)Q:(0,A]\to(-\infty,\infty) that continuous, increasing and invertible; and RR represents those functions R:(0,A](0,)R:(0,A]\to(0,\infty) that are continuous, decreasing and invertible. We next define

(11) ψf,A(t)={F1(F(A)(1t))χ[0,1](t)if f=F,Q1(Q(A)t)χ(0,)(t)if f=Q,R1(R(A)(1+t))χ(0,)(t)iff=R.\psi_{f,A}(t)=\begin{cases}F^{-1}(F(A)(1-t))\chi_{[0,1]}(t)&\text{if }f=F,\\ Q^{-1}(Q(A)-t)\chi_{(0,\infty)}(t)&\text{if }f=Q,\\ R^{-1}(R(A)(1+t))\chi_{(0,\infty)}(t)&\text{if}f=R.\end{cases}

Notice that, if A=μ(K),A=\mu(K), then ψf,μ(K)(p)1=pμ(K)C(p,μ,K)p\mathcal{M}_{\psi_{f,\mu(K)}}(p)^{-1}=\frac{p}{\mu(K)}C(p,\mu,K)^{p} if p(1,0)p\in(-1,0), and this also holds for any p>0p>0 such that tp1ψf,μ(K)t^{p-1}\psi_{f,\mu(K)} is integrable.

We will now work towards the proof of Theorem 1.1. Let ψ\psi be a non-negative function such that ψ(0)=A>0.\psi(0)=A>0. Then, for p(1,0)(0,p1),p\in(-1,0)\cup(0,p_{1}), set

(12) Ωf,ψ(p)=ψ(p)ψf,A(p),\Omega_{f,\psi}(p)=\frac{\mathcal{M}_{\psi}(p)}{\mathcal{M}_{\psi_{f,A}}(p)},

where Ωf,ψ(0)=1\Omega_{f,\psi}(0)=1 and p1p_{1} is defined implicitly by p1=sup{p<:Ωf,ψ(p)<}.p_{1}=\sup\{p<\infty:\Omega_{f,\psi}(p)<\infty\}. Next, set for p(1,0)(0,p1)p\in(-1,0)\cup(0,p_{1})

(13) Gψ(p)=(Ωf,ψ(p))1/pG_{\psi}(p)=\left(\Omega_{f,\psi}(p)\right)^{1/p}

and Gψ(0)=exp(log(Ωf,ψ)(0)).G_{\psi}(0)=\exp(\log(\Omega_{f,\psi})^{\prime}(0)).

Lemma 2.1 (The Mellin-Berwald Inequality).

Let ψ:[0,)[0.)\psi:[0,\infty)\to[0.\infty) be an integrable, ff-concave function, f{F,Q,R}f\in\{F,Q,R\} (elaborated above (11)). Suppose that ψ\psi is right differentiable at the origin. Next, set p0=inf{p>1:Ωf,ψ(p)>0},p_{0}=\inf\{p>-1:\Omega_{f,\psi}(p)>0\}, where Ωf,ψ(p)\Omega_{f,\psi}(p) is defined via (12). Then,

  1. (1)

    p0[1,0)p_{0}\in[-1,0) and if ψ\psi is non-increasing then p0=1.p_{0}=-1.

  2. (2)

    Ωf,ψ(p)>0\Omega_{f,\psi}(p)>0 for every p(p0,p1).p\in(p_{0},p_{1}). Thus, Gψ(p),G_{\psi}(p), defined via (13), is well-defined and analytic on (p0,p1).(p_{0},p_{1}).

  3. (3)

    Gψ(p)G_{\psi}(p) is non-increasing on (p0,p1).(p_{0},p_{1}).

  4. (4)

    If there exists r,q(p0,p1)r,q\in(p_{0},p_{1}) such that Gψ(r)=Gψ(q),G_{\psi}(r)=G_{\psi}(q), then Gψ(p)G_{\psi}(p) is constant on (p0,p1).(p_{0},p_{1}). Furthermore, Gψ(p)G_{\psi}(p) is constant on (p0,p1)(p_{0},p_{1}) if, and only if, ψ(t)=ψf,A(tα)\psi(t)=\psi_{f,A}(\frac{t}{\alpha}) for some α>0,\alpha>0, in which case Gψ(p)=α.G_{\psi}(p)=\alpha.

Proof.

From the fact that Ωf,ψ(0)=ψ(0)=1>0,\Omega_{f,\psi}(0)=\psi(0)=1>0, one immediately has that p0[1,0).p_{0}\in[-1,0). Notice that ψf,A(p)<0\mathcal{M}_{\psi_{f,A}}(p)<0 for p(1,0).p\in(-1,0). If ψ\psi is non-increasing, then from (9) one obtains that ψ(p)<0\mathcal{M}_{\psi}(p)<0 as well. Thus, Ωf,ψ(p)=ψ(p)/ψf,A(p)>0\Omega_{f,\psi}(p)=\mathcal{M}_{\psi}(p)/\mathcal{M}_{\psi_{f,A}}(p)>0 for all p(1,0),p\in(-1,0), and thus p0=1.p_{0}=-1.

For the second statement, clearly Ωf,ψ(p)>0\Omega_{f,\psi}(p)>0 for p[0,p1].p\in[0,p_{1}]. So, fix some q(p0,0)q\in(p_{0},0) such that Ωf,ψ(q)>0.\Omega_{f,\psi}(q)>0. Then, Gψ(q)=(Ωf,ψ(q))1/q>0.G_{\psi}(q)=\left(\Omega_{f,\psi}(q)\right)^{1/q}>0. Define the function z(t)=ψf,A(t/Gψ(q)).z(t)=\psi_{f,A}(t/G_{\psi}(q)). Notice that z(0)=ψf,A(0)=Az(0)=\psi_{f,A}(0)=A and, by performing a variable substitution, z(p)=(Gψ(q))pψf,A(p)\mathcal{M}_{z}(p)=(G_{\psi}(q))^{p}\mathcal{M}_{\psi_{f,A}}(p) via (9) for every p(1,0)(0,p1).p\in(-1,0)\cup(0,p_{1}). In particular, for p=q.p=q. From the definition of Gψ(q),G_{\psi}(q), we then obtain that z(q)=(Gψ(q))qψf,A(q)=ψ(q).\mathcal{M}_{z}(q)=(G_{\psi}(q))^{q}\mathcal{M}_{\psi_{f,A}}(q)=\mathcal{M}_{\psi}(q). Thus, from (9), one obtains

0=ψ(q)z(q)=0tq1(ψ(t)z(t))𝑑t.0=\mathcal{M}_{\psi}(q)-\mathcal{M}_{z}(q)=\int_{0}^{\infty}t^{q-1}(\psi(t)-z(t))dt.

Consequently, the function ψ(t)z(t)\psi(t)-z(t) changes signs at least once. But actually, this function changes sign exactly once. Indeed, let t0t_{0} be the smallest positive value such that ψ(t0)=z(t0).\psi(t_{0})=z(t_{0}). Then, fψ(t0)=fz(t0).f\circ\psi(t_{0})=f\circ z(t_{0}). Now, fzf\circ z is affine. If f{F,Q},f\in\{F,Q\}, then fψf\circ\psi is concave and the slope of fzf\circ z is negative. Since ψ(0)=z(0)=A,\psi(0)=z(0)=A, one has that fψ(t)fz(t)f\circ\psi(t)\geq f\circ z(t) on [0,t0].[0,t_{0}]. From the concavity, we must then have that fψ(t)fz(t)f\circ\psi(t)\leq f\circ z(t) on [t0,).[t_{0},\infty). Similarly, if f=R,f=R, then fψf\circ\psi is convex and the slope of fzf\circ z is positive. Hence, fψ(t)fz(t)f\circ\psi(t)\leq f\circ z(t) on [0,t0][0,t_{0}] and fψ(t)fz(t)f\circ\psi(t)\geq f\circ z(t) on [t0,).[t_{0},\infty). Taking inverses, we obtain in either case that ψ(t)z(t)\psi(t)\geq z(t) on [0,t0][0,t_{0}] and ψ(t)z(t)\psi(t)\leq z(t) on [t0,).[t_{0},\infty).

Next, define

g(t)=tuq1(ψ(u)z(u))𝑑u.g(t)=\int_{t}^{\infty}u^{q-1}(\psi(u)-z(u))du.

Clearly, g(0)=g()=0.g(0)=g(\infty)=0. One has g(t)=tqt(ψ(t)z(t)).g^{\prime}(t)=-t^{q-t}(\psi(t)-z(t)). Thus, gg is non-increasing on [0,t0][0,t_{0}] and non-decreasing on [t0,).[t_{0},\infty). Hence g(t)0g(t)\leq 0 for all t[0,).t\in[0,\infty). Next, pick r(q,0).r\in(q,0). From integration by parts, one obtains

ψ(r)z(r)=0trqtq1(ψ(t)z(t))𝑑t=(rq)0trq1g(t)𝑑t0.\mathcal{M}_{\psi}(r)-\mathcal{M}_{z}(r)=\int_{0}^{\infty}t^{r-q}t^{q-1}(\psi(t)-z(t))dt=(r-q)\int_{0}^{\infty}t^{r-q-1}g(t)dt\leq 0.

Hence,

ψ(r)z(r)=(Gψ(q))rψf,A(r)<0.\mathcal{M}_{\psi}(r)\leq\mathcal{M}_{z}(r)=(G_{\psi}(q))^{r}\mathcal{M}_{\psi_{f,A}}(r)<0.

We deduce that

(14) Ωf,ψ(r)=ψ(r)ψf,A(r)(Gψ(q))r>0\Omega_{f,\psi}(r)=\frac{\mathcal{M}_{\psi}(r)}{\mathcal{M}_{\psi_{f,A}}(r)}\geq(G_{\psi}(q))^{r}>0

for every r(q,0).r\in(q,0). Sending qp0,q\to p_{0}, we obtain Ωf,ψ(p)>0\Omega_{f,\psi}(p)>0 for every p(p0,0)p\in(p_{0},0) and thus for p(p0,p1).p\in(p_{0},p_{1}). One immediately obtains that Gψ(p)G_{\psi}(p) is well-defined and analytic on (p0,p1).(p_{0},p_{1}). Finally, taking the rrth root of (14) yields for p0<q<r<0p_{0}<q<r<0 that

Gψ(r)=(Ωf,ψ(r))1/rGψ(q),G_{\psi}(r)=(\Omega_{f,\psi}(r))^{1/r}\leq G_{\psi}(q),

i.e. Gψ(p)G_{\psi}(p) is non-increasing on (p0,0).(p_{0},0). Suppose there exists an r(q,0)r\in(q,0) such that Gψ(q)=Gψ(r).G_{\psi}(q)=G_{\psi}(r). Then, there is equality in (14). But this yields g(t)=0g(t)=0 for almost all t.t. We take a moment to notice that this then yields Gψ(q)=Gψ(r)G_{\psi}(q)=G_{\psi}(r) for every q,r(p0,0).q,r\in(p_{0},0). Anyway, since g(t)=0g(t)=0 for almost all tt, we have ψ(t)=z(t)\psi(t)=z(t) for almost all tt. Hence, the concave function fψ(t)f\circ\psi(t) equals the affine function fz(t)f\circ z(t) for almost all tt and thus for all tt. Consequently, ψ(t)z(t)=ψf,A(t/Gψ(q)).\psi(t)\equiv z(t)=\psi_{f,A}(t/G_{\psi}(q)). Conversely, suppose that ψ(t)=ψf,A(t/α)\psi(t)=\psi_{f,A}(t/\alpha) for some α>0.\alpha>0. Then, direct substitution yields Gψ(p)=αG_{\psi}(p)=\alpha on (p0,0).(p_{0},0). Notice that z(q)=(Gψ(q))qψf,A(q)=ψ(q)\mathcal{M}_{z}(q)=(G_{\psi}(q))^{q}\mathcal{M}_{\psi_{f,A}}(q)=\mathcal{M}_{\psi}(q) is also true for any q(0,p1).q\in(0,p_{1}). Consequently, by picking any r(q,p1),r\in(q,p_{1}), we repeat the above arguments and deduce again that

ψ(r)z(r)=(Gψ(q))rψf,A(r).\mathcal{M}_{\psi}(r)\leq\mathcal{M}_{z}(r)=(G_{\psi}(q))^{r}\mathcal{M}_{\psi_{f,A}}(r).

This time, however, ψf,A(r)>0.\mathcal{M}_{\psi_{f,A}}(r)>0. Consequently, this immediately implies that

Gψ(r)=(Ωf,ψ(r))1/rGψ(q)G_{\psi}(r)=(\Omega_{f,\psi}(r))^{1/r}\leq G_{\psi}(q)

for every 0<qr<p1.0<q\leq r<p_{1}. This establishes that Gψ(p)G_{\psi}(p) is non-increasing on (0,p1)(0,p_{1}) as well. The argument for the equality conditions is the same. ∎

Proof of Theorem 1.1.

Let ww be the concavity of our measure μ.\mu. Next, let ψ(t)=μ({xK:f(x)t}).\psi(t)=\mu(\left\{x\in K:f(x)\geq t\right\}). Notice this ψ\psi is non-increasing, and thus p0p_{0} from the statement of Lemma 2.1 is 1-1. Then, for p(1,0):p\in(-1,0):

Ωw,μ(K),ψ(p)\displaystyle\Omega_{w,\mu(K),\psi}(p) =ψ(p)ψw,μ(K)(p)\displaystyle=\frac{\mathcal{M}_{\psi}(p)}{\mathcal{M}_{\psi_{w,\mu(K)}}(p)}
=Cp(p,μ,K)pμ(K)0tp1(μ({xK:f(x)t})μ(K))𝑑t\displaystyle=C^{p}(p,\mu,K)\frac{p}{\mu(K)}\int_{0}^{\infty}t^{p-1}(\mu(\left\{x\in K:f(x)\geq t\right\})-\mu(K))dt
=Cp(p,μ,K)1μ(K)Kfp(x)𝑑μ(x)\displaystyle=C^{p}(p,\mu,K)\frac{1}{\mu(K)}\int_{K}f^{p}(x)d\mu(x)

via the layer cake formula for p(1,0)p\in(-1,0); similar computations yield the case for p>0,p>0, and p=0p=0 follows from limits. Thus, we obtain from Lemma 2.1, Item 3, that the function

Gψ(p)=C(p,μ,K)(1μ(K)Kfp(x)𝑑μ(x))1/pG_{\psi}(p)=C(p,\mu,K)\left(\frac{1}{\mu(K)}\int_{K}f^{p}(x)d\mu(x)\right)^{1/p}

is non-increasing for p(1,pmax).p\in(-1,p_{\text{max}}). Furthermore, Gψ(p)α>0G_{\psi}(p)\equiv\alpha>0, if, and only if,

μ({xK:f(x)t})=ψ(t)=ψw,μ(K)(t/α).\mu(\left\{x\in K:f(x)\geq t\right\})=\psi(t)=\psi_{w,\mu(K)}(t/\alpha).

We now insert the appropriate ψw,μ(K)\psi_{w,\mu(K)}, starting with the case w=Fw=F. This is precisely

(15) αt=1F(μ({ft}))F(μ(K))μ({ft})=F1[F(μ(K))(1αt)].\alpha t=1-\frac{F(\mu(\{f\geq t\}))}{F(\mu(K))}\longleftrightarrow\mu(\{f\geq t\})=F^{-1}\left[F(\mu(K))\left(1-\alpha t\right)\right].

We then evaluate the above at t=f,t=\|f\|_{\infty}, to obtain

α=(1F(0)F(μ(K)))/f.\alpha=\left(1-\frac{F(0)}{F(\mu(K))}\right)/\|f\|_{\infty}.

On the other hand, we also know that, for all p(0,)p\in(0,\infty) we have

αp=01F1[F(μ(K))(1t)]tp1𝑑t0fμ({ft})tp1𝑑t.\displaystyle\alpha^{p}=\frac{\int_{0}^{1}F^{-1}\left[F(\mu(K))(1-t)\right]t^{p-1}dt}{\int_{0}^{\|f\|_{\infty}}\mu(\{f\geq t\})t^{p-1}dt}.

Inserting the formula for α\alpha and the formula of μ({ft})\mu(\{f\geq t\}) from (15), we obtain

(1F(0)F(μ(K)))pfp=01F1[F(μ(K))(1t)]tp1𝑑t0fF1[F(μ(K))(1(1F(0)F(μ(K)))ft)]tp1𝑑t.\frac{(1-\frac{F(0)}{F(\mu(K))})^{p}}{\|f\|_{\infty}^{p}}=\frac{\int_{0}^{1}F^{-1}\left[F(\mu(K))(1-t)\right]t^{p-1}dt}{\int_{0}^{\|f\|_{\infty}}F^{-1}\left[F(\mu(K))\left(1-\frac{(1-\frac{F(0)}{F(\mu(K))})}{\|f\|_{\infty}}t\right)\right]t^{p-1}dt}.

By performing a variable substitution in the denominator, we obtain that

1=01F1[F(μ(K))(1t)]tp1𝑑t0(1F(0)F(μ(K)))F1[F(μ(K))(1t)]tp1𝑑t.1=\frac{\int_{0}^{1}F^{-1}\left[F(\mu(K))(1-t)\right]t^{p-1}dt}{\int_{0}^{(1-\frac{F(0)}{F(\mu(K))})}F^{-1}\left[F(\mu(K))\left(1-t\right)\right]t^{p-1}dt}.

Therefore, we have (1F(0)F(μ(K)))=1,(1-\frac{F(0)}{F(\mu(K))})=1, which means F(0)=0.F(0)=0.

Next, we show that equality never occurs for when w=Qw=Q, and the case w=Rw=R is similar. From integrability, we have that Q1()=0,Q^{-1}(-\infty)=0, or Q(0)=Q(0)=-\infty (where these are understood as limits from the left and the right, respectively). On the other hand, we have shown equality implies

αt=Q(μ(K))Q(μ(K)fμ(t)).\alpha t=Q(\mu(K))-Q(\mu(K)f_{\mu}(t)).

Evaluating again at t=ft=\|f\|_{\infty} yields αf=Q(μ(K))Q(0),\alpha\|f\|_{\infty}=Q(\mu(K))-Q(0), which would imply that |Q(0)|<.|Q(0)|<\infty.

Proof of Corollary 1.2.

We have that μ\mu is ss-concave on the level sets of ff, and thus the proof is a direct application of Theorem 1.1; in the first case, the coefficients become a beta function and in the second case they become a gamma function. As for the third case, a bit more work is required. We will show the case when p(0,1/s);p\in(0,-1/s); the case when p(1,0)p\in(-1,0) is exactly the same (using the analytic continuation of the Beta function), and then the case p=0p=0 follows from limits. Inserting R(x)=xs,s<0R(x)=x^{s},s<0 yields

C(p,s)=(p0(1+t)1/stp1𝑑t)1.C(p,s)=\left(p\int_{0}^{\infty}\left(1+t\right)^{1/s}t^{p-1}dt\right)^{-1}.

Focus on the function q(t)=(1+t)1/stp1.q(t)=\left(1+t\right)^{1/s}t^{p-1}. For this function to be integrable near zero, we require 1<p1,-1<p-1, and, for the integrability near infinity, we require 1s+p1<1.\frac{1}{s}+p-1<-1. Thus, p(0,1/s).p\in(0,-1/s). We will now manipulate C(p,s)C(p,s) to obtain a more familiar formula. Consider the variable substitution given by t=z1z.t=\frac{z}{1-z}. Writing zz as a function of t,t, this becomes

z=111+tz(t)=1(1+t)2.z=1-\frac{1}{1+t}\quad\longrightarrow\quad z^{\prime}(t)=\frac{1}{(1+t)^{2}}.

As t0+,z0+,t\to 0^{+},z\to 0^{+}, and as t,z1.t\to\infty,z\to 1^{-}. We then obtain that

C(p,s)\displaystyle C(p,s) =(p01(1z)(p+1/s)1zp1𝑑z)1=Γ(1s)pΓ(p)Γ(p1s)\displaystyle=\left(p\int_{0}^{1}\left(1-z\right)^{-(p+1/s)-1}z^{p-1}dz\right)^{-1}=\frac{\Gamma\left(-\frac{1}{s}\right)}{p\Gamma\left(p\right)\Gamma\left(-p-\frac{1}{s}\right)}
=s(p+1s)Γ(11s)Γ(1+p)Γ(1p1s),\displaystyle=s\left(p+\frac{1}{s}\right)\frac{\Gamma\left(1-\frac{1}{s}\right)}{\Gamma\left(1+p\right)\Gamma\left(1-p-\frac{1}{s}\right)},

which equals our claim. ∎

Proof of Theorem 1.3.

From the assumptions on the measure μ\mu, we obtain that dμ(x)=ϕ(x)dxd\mu(x)=\phi(x)dx for some p=s/(1ns)p=s/(1-ns)-concave function ϕ.\phi. Furthermore, ϕ\phi is (1/s)n(1/s)-n homogeneous. Observe that Corollary 1.2 yields the inequality; all that remains to show is the equality conditions. By hypothesis, the maximum of the function ff is obtained at the origin. Equality conditions of Corollary 1.2 imply that

f1/s=0fμK({ft})t1/s1𝑑t01(1t)1/st1/s1𝑑t.\|f\|^{1/s}_{\infty}=\frac{\int_{0}^{\|f\|_{\infty}}\mu_{K}(\{f\geq t\})t^{1/s-1}dt}{\int_{0}^{1}(1-t)^{1/s}t^{1/s-1}dt}.

Using (2), this implies that

Kf1/s(x)𝑑μ(x)\displaystyle\int_{K}f^{1/s}(x)d\mu(x) =μ(K)s01[f(1t)]1/s𝑑t\displaystyle=\frac{\mu(K)}{s}\int_{0}^{1}[\|f\|_{\infty}(1-t)]^{1/s}dt
=𝕊n1ϕ(θ)ρK(θ)1/s𝑑θ01[f(1t)]1/st1/s1𝑑t.\displaystyle=\int_{\mathbb{S}^{n-1}}\phi(\theta)\rho_{K}(\theta)^{1/s}d\theta\int_{0}^{1}[\|f\|_{\infty}(1-t)]^{1/s}t^{1/s-1}dt.

Using Fubini’s theorem, a variable substitution tt/ρK(θ)t\to t/\rho_{K}(\theta) and the homogeneity of ϕ\phi yields

Kf1/s(x)𝑑μ(x)\displaystyle\int_{K}f^{1/s}(x)d\mu(x) =𝕊n10ρK(θ)[f(1tρK(θ))]1/stn1ϕ(tθ)𝑑t𝑑θ\displaystyle=\int_{\mathbb{S}^{n-1}}\int_{0}^{\rho_{K}(\theta)}\left[\|f\|_{\infty}\left(1-\frac{t}{\rho_{K}(\theta)}\right)\right]^{1/s}t^{n-1}\phi(t\theta)dtd\theta
=K[f(11ρK(x))]1/s𝑑x.\displaystyle=\int_{K}\left[\|f\|_{\infty}\left(1-\frac{1}{\rho_{K}(x)}\right)\right]^{1/s}dx.

One has from (4) that a concave function ff supported on K𝒦0nK\in\mathcal{K}^{n}_{0} whose maximum is at the origin satisfies

f1/s(x)[f(11ρK(x))]1/s,xK{0}.f^{1/s}(x)\geq\left[\|f\|_{\infty}\left(1-\frac{1}{\rho_{K}(x)}\right)\right]^{1/s},\;x\in K\setminus\{0\}.

By the above integral, we have equality. ∎

We next obtain an interesting result by perturbing Theorem 1.3, inspired by the standard proof (see e.g. [21]) of Minkowski’s first inequality by perturbing the Brunn-Minkowski inequality.

Corollary 2.2.

Let μ\mu be a Radon measure that is ss-concave, 1/s1/s-homogeneous, s(0,1/n]s\in(0,1/n], and suppose that K\ell_{K} is given by (3) for some K𝒦nK\in\mathcal{K}^{n}. Let ψ\psi be a concave function supported on KK, and suppose 0<pq<.0<p\leq q<\infty. Then, one has

(1s+p1s)KKp(x)(ψ(x)K(x))𝑑μ(x)(1s+q1s)KKq(x)(ψ(x)K(x))𝑑μ(x).{{\frac{1}{s}+p}\choose\frac{1}{s}}\int_{K}\ell_{K}^{p}(x)\left(\frac{\psi(x)}{\ell_{K}(x)}\right)d\mu(x)\geq{{\frac{1}{s}+q}\choose\frac{1}{s}}\int_{K}\ell_{K}^{q}(x)\left(\frac{\psi(x)}{\ell_{K}(x)}\right)d\mu(x).
Proof.

Let zK(t,x)z_{K}(t,x) be a concave perturbation of K\ell_{K} by ψ\psi, i.e. δ>0\delta>0 is picked small enough so that zK(t,x)=K(x)+tψ(x)z_{K}(t,x)=\ell_{K}(x)+t\psi(x) is concave with maximum at the origin for all xKx\in K and |t|<δ.|t|<\delta. Next, consider the function given by, for 0<pq0<p\leq q

BK(t)=\displaystyle B_{K}(t)= ((1s+p1s)1μ(K)KzK(x,t)𝑑μ(x))1/p\displaystyle\left({{\frac{1}{s}+p}\choose\frac{1}{s}}\frac{1}{\mu(K)}\int_{K}z_{K}(x,t)d\mu(x)\right)^{1/p}
((1s+q1s)1μ(K)KzK(x,t)𝑑μ(x))1/q,\displaystyle-\left({{\frac{1}{s}+q}\choose\frac{1}{s}}\frac{1}{\mu(K)}\int_{K}z_{K}(x,t)d\mu(x)\right)^{1/q},

from Berwald’s inequality in Theorem 1.3, this function is greater than or equal to zero for all |t|<δ,|t|<\delta, and equals zero when t=0.t=0. Hence, the derivative of this function is non-negative at t=0t=0. By taking the derivative of BK(t)B_{K}(t) in the variable t, evaluating at t=0t=0, and setting this computation be greater than or equal to zero, one immediately obtains the result. ∎

We now prove the corollaries for the Gaussian measure and rotational invariant log-concave measures.

Proof of Corollary 1.5.

From Borell’s classification, the Gaussian measure is log-concave, and thus one can use the second case of Corollary 1.2 for the first inequality. For the second inequality, the function Φ1\Phi^{-1} behaves logarithmically, that is one can apply the second case of Theorem 1.1. Finally, for the third inequality, note that if ff is a concave function supported on some K𝒦0nK\in\mathcal{K}^{n}_{0} with maximum at the origin, then the level sets of ff are also in 𝒦0n,\mathcal{K}^{n}_{0}, and thus one can apply the 12n\frac{1}{2n}-concavity of the Gaussian measure over 𝒦0n\mathcal{K}^{n}_{0} and use the first case of Corollary 1.2. ∎

Proof of Corollary 1.6.

Notice that if ff is an even, concave function supported on a symmetric K𝒦0nK\in\mathcal{K}^{n}_{0}, then the maximum of ff is at the origin (for every xK,xKx\in K,-x\in K and so f(0)=f(12x+12(x))12f(x)+12f(x)=f(x)f(0)=f(\frac{1}{2}x+\frac{1}{2}(-x))\geq\frac{1}{2}f(x)+\frac{1}{2}f(-x)=f(x)) and the level sets of ff are all symmetric convex bodies. Thus, the result follows from the 1/n1/n-concavity of measures in n\mathcal{M}_{n}. ∎

2.1. Applications

We conclude this section by showing a few applications. The first example uses that the support of ff in Theorem 1.1 need not be compact.

Theorem 2.3.

Let θ𝕊n1\theta\in\mathbb{S}^{n-1}. Denote H=θH=\theta^{\perp} and H+={xn:x,θ>0}.H_{+}=\{x\in\mathbb{R}^{n}:\langle x,\theta\rangle>0\}. Denote

x,θ+=x,θχH+(x)={x,θifx,θ>0,0otherwise.\langle x,\theta\rangle_{+}=\langle x,\theta\rangle\chi_{H_{+}}(x)=\begin{cases}\langle x,\theta\rangle\;&\text{if}\;\langle x,\theta\rangle>0,\\ 0&\text{otherwise.}\end{cases}

Then, for every Borel measure μ\mu finite on H+H_{+} with one of the following concavity conditions on subsets of H+H_{+}:

  1. (1)

    If μ\mu is FF-concave, where F:[0,μ(H+)][0,)F:[0,\mu(H_{+})]\to[0,\infty) is an increasing and invertible function one has

    (nx,θ+q𝑑μ(x))1/q(nx,θ+p𝑑μ(x))1/p(q01(F1[F(μ(H+))(1t)]μ(H+))tq1𝑑t+μ(H+))1/q(p01(F1[F(μ(H+))(1t)]μ(H+))tp1𝑑t+μ(H+))1/p\frac{\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{q}d\mu(x)\right)^{1/q}}{\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{p}d\mu(x)\right)^{1/p}}\leq\frac{\left(q\int_{0}^{1}\left(F^{-1}\left[F(\mu(H_{+}))(1-t)\right]-\mu(H_{+})\right)t^{q-1}dt+\mu(H_{+})\right)^{1/q}}{\left(p\int_{0}^{1}\left(F^{-1}\left[F(\mu(H_{+}))(1-t)\right]-\mu(H_{+})\right)t^{p-1}dt+\mu(H_{+})\right)^{1/p}}

    for every 1<pq<-1<p\leq q<\infty where the integrals exist. In particular, if F(x)=xs,s>0,F(x)=x^{s},s>0, one obtains

    (nx,θ+q𝑑μ(x))1/qμ(H+)1q1p(1s+pp)1/p(1s+qq)1/q(nx,θ+p𝑑μ(x))1/p.\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{q}d\mu(x)\right)^{1/q}\leq\mu(H_{+})^{\frac{1}{q}-\frac{1}{p}}\frac{{{\frac{1}{s}+p}\choose p}^{1/p}}{{{\frac{1}{s}+q}\choose q}^{1/q}}\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{p}d\mu(x)\right)^{1/p}.
  2. (2)

    If μ\mu is QQ-concave, where Q:(0,μ(H+)](,)Q:(0,\mu(H_{+})]\to(-\infty,\infty) is an increasing and invertible function one has

    (nx,θ+q𝑑μ(x))1/q(nx,θ+p𝑑μ(x))1/p(q0Q1[Q(μ(H+))t]tq1𝑑t)1/q(p0Q1[Q(μ(H+))t]tp1𝑑t)1/p\frac{\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{q}d\mu(x)\right)^{1/q}}{\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{p}d\mu(x)\right)^{1/p}}\leq\frac{\left(q\int_{0}^{\infty}Q^{-1}\left[Q(\mu(H_{+}))-t\right]t^{q-1}dt\right)^{1/q}}{\left(p\int_{0}^{\infty}Q^{-1}\left[Q(\mu(H_{+}))-t\right]t^{p-1}dt\right)^{1/p}}

    for every 0<pq<0<p\leq q<\infty where the integrals exist; the case for 1<pq<-1<p\leq q<\infty can be deduced. For the Gaussian measure especially, one can set Q=Φ1Q=\Phi^{-1} and obtain

    (nx,θ+q𝑑γn(x))1/q(nx,θ+p𝑑γn(x))1/p(q0Φ[Φ1(γn(H+))t]tq1𝑑t)1/q(p0Φ[Φ1(γn(H+))t]tp1𝑑t)1/p.\frac{\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{q}d\gamma_{n}(x)\right)^{1/q}}{\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{p}d\gamma_{n}(x)\right)^{1/p}}\leq\frac{\left(q\int_{0}^{\infty}\Phi\left[\Phi^{-1}(\gamma_{n}(H_{+}))-t\right]t^{q-1}dt\right)^{1/q}}{\left(p\int_{0}^{\infty}\Phi\left[\Phi^{-1}(\gamma_{n}(H_{+}))-t\right]t^{p-1}dt\right)^{1/p}}.

    If Q(x)=log(x)Q(x)=\log(x) one obtains for every 1<pq<-1<p\leq q<\infty that

    (nx,θ+q𝑑μ(x))1/qμ(H+)1q1pΓ(q+1)1/qΓ(p+1)1/p(nx,θ+p𝑑μ(x))1/p.\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{q}d\mu(x)\right)^{1/q}\leq\mu(H_{+})^{\frac{1}{q}-\frac{1}{p}}\frac{{\Gamma\left(q+1\right)}^{1/q}}{{\Gamma\left(p+1\right)}^{1/p}}\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{p}d\mu(x)\right)^{1/p}.
  3. (3)

    If μ\mu is RR-concave, where R:(0,μ(H+)](0,)R:(0,\mu(H_{+})]\to(0,\infty) is a decreasing and invertible function one has

    (nx,θ+q𝑑μ(x))1/q(nx,θ+p𝑑μ(x))1/p(q0R1[R(μ(H+))(1+t)]tq1𝑑t)1/q(p0R1[R(μ(H+))(1+t)]tp1𝑑t)1/p\frac{\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{q}d\mu(x)\right)^{1/q}}{\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{p}d\mu(x)\right)^{1/p}}\leq\frac{\left(q\int_{0}^{\infty}R^{-1}\left[R(\mu(H_{+}))(1+t)\right]t^{q-1}dt\right)^{1/q}}{\left(p\int_{0}^{\infty}R^{-1}\left[R(\mu(H_{+}))(1+t)\right]t^{p-1}dt\right)^{1/p}}

    for every 0<pq<0<p\leq q<\infty where the integrals exist; the case for 1<pq<-1<p\leq q<\infty can be deduced. In particular, if R(x)=xs,s<0,R(x)=x^{s},s<0, and 1<pq<1/s,-1<p\leq q<-1/s, one obtains

    (nx,θ+q𝑑μ(x))1/qμ(H+)1q1p(s(p+1s)(1sp))1/p(s(q+1s)(1sq))1/q(nx,θ+p𝑑μ(x))1/p.\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{q}d\mu(x)\right)^{1/q}\leq\mu(H_{+})^{\frac{1}{q}-\frac{1}{p}}\frac{\left(s\left(p+\frac{1}{s}\right){{-\frac{1}{s}}\choose p}\right)^{1/p}}{\left(s\left(q+\frac{1}{s}\right){{-\frac{1}{s}}\choose q}\right)^{1/q}}\left(\int_{\mathbb{R}^{n}}\langle x,\theta\rangle_{+}^{p}d\mu(x)\right)^{1/p}.

Finally, let μ\mu be a Borel measure finite on some convex Kn.K\subset\mathbb{R}^{n}. Suppose μ\mu is either F,QF,Q or RR concave, where the functions F,QF,Q and RR are as given in Theorem 1.1. Next, consider a non-negative function ff so that fβf^{\beta} is bounded and concave on KK for some β>0\beta>0. Inserting fβ,f^{\beta}, into Theorem 1.1 and picking appropriate choices of pp and q,q, we obtain that for every q1q\geq 1 one has

(16) (Kf(x)q𝑑μ(x))1/qμ(K)1qq(C(1β,μ,K)C(qβ,μ,K))1βKf(x)𝑑μ(x),\left(\int_{K}f(x)^{q}d\mu(x)\right)^{1/q}\leq\mu(K)^{\frac{1-q}{q}}\left(\frac{C(\frac{1}{\beta},\mu,K)}{C(\frac{q}{\beta},\mu,K)}\right)^{\frac{1}{\beta}}\int_{K}f(x)d\mu(x),

up to possible restrictions on admissible β\beta and qq so that all constants exist. In words, we have bounded the Lq(K,μ)L^{q}(K,\mu) norm of a bounded, non-negative, β\beta-concave function ff by its L1(K,μ)L^{1}(K,\mu) norm when μ\mu is either F,QF,Q or RR-concave. Examples of interest are when μ\mu is ss-concave. We obtain for a ss-concave measure μ\mu and q1q\geq 1:

  1. (1)

    When s>0s>0:

    (Kf(x)q𝑑μ(x))1/q(1s+1β1β)μ(K)(μ(K)(1s+qβqβ))1/qKf(x)𝑑μ(x).\left(\int_{K}f(x)^{q}d\mu(x)\right)^{1/q}\leq\frac{{{\frac{1}{s}+\frac{1}{\beta}}\choose\frac{1}{\beta}}}{\mu(K)}\left(\frac{\mu(K)}{{{\frac{1}{s}+\frac{q}{\beta}}\choose\frac{q}{\beta}}}\right)^{1/q}\int_{K}f(x)d\mu(x).
  2. (2)

    When s=0s=0:

    (Kf(x)q𝑑μ(x))1/qΓ(1+1β)μ(K)(μ(K)Γ(1+qβ))1/qKf(x)𝑑μ(x).\left(\int_{K}f(x)^{q}d\mu(x)\right)^{1/q}\leq\frac{\Gamma(1+\frac{1}{\beta})}{\mu(K)}\left(\frac{\mu(K)}{\Gamma(1+\frac{q}{\beta})}\right)^{1/q}\int_{K}f(x)d\mu(x).
  3. (3)

    When s<0,s<0, β>s\beta>-s and q[1,βs):q\in[1,-\frac{\beta}{s}):

    (Kf(x)q𝑑μ(x))1/qs(q+1s)(1sq)μ(K)(μ(K)s(qβ+1s)(1sqβ))1/qKf(x)𝑑μ(x).\left(\int_{K}f(x)^{q}d\mu(x)\right)^{1/q}\leq\frac{s\left(q+\frac{1}{s}\right){{-\frac{1}{s}}\choose q}}{\mu(K)}\left(\frac{\mu(K)}{s\left(\frac{q}{\beta}+\frac{1}{s}\right){{-\frac{1}{s}}\choose\frac{q}{\beta}}}\right)^{1/q}\int_{K}f(x)d\mu(x).

We also highlight the following examples for the Gaussian measure.

  1. (1)
    (Kf(x)q𝑑γn(x))1/qβq1q(q0Φ[Φ1(γn(K))t]tq1𝑑t)1/q0Φ[Φ1(γn(K))t]tp1𝑑tKf(x)𝑑γn(x).\left(\int_{K}f(x)^{q}d\gamma_{n}(x)\right)^{1/q}\leq\beta^{\frac{q-1}{q}}\frac{\left(q\int_{0}^{\infty}\Phi\left[\Phi^{-1}(\gamma_{n}(K))-t\right]t^{q-1}dt\right)^{1/q}}{\int_{0}^{\infty}\Phi\left[\Phi^{-1}(\gamma_{n}(K))-t\right]t^{p-1}dt}\int_{K}f(x)d\gamma_{n}(x).
  2. (2)

    If K𝒦0nK\in\mathcal{K}^{n}_{0} and the maximum of fβf^{\beta} is obtained at the origin:

    (Kf(x)q𝑑γn(x))1/q(12n+1β1β)γn(K)(γn(K)(12n+qβqβ))1/qKf(x)𝑑γn(x).\left(\int_{K}f(x)^{q}d\gamma_{n}(x)\right)^{1/q}\leq\frac{{{\frac{1}{2n}+\frac{1}{\beta}}\choose\frac{1}{\beta}}}{\gamma_{n}(K)}\left(\frac{\gamma_{n}(K)}{{{\frac{1}{2n}+\frac{q}{\beta}}\choose\frac{q}{\beta}}}\right)^{1/q}\int_{K}f(x)d\gamma_{n}(x).
  3. (3)

    Let μ\mu be a measure in n\mathcal{M}_{n}. If K𝒦0nK\in\mathcal{K}^{n}_{0} is symmetric, and fβf^{\beta} is even:

    (Kf(x)q𝑑μ(x))1/q(1n+1β1β)μ(K)(μ(K)(1n+qβqβ))1/qKf(x)𝑑μ(x).\left(\int_{K}f(x)^{q}d\mu(x)\right)^{1/q}\leq\frac{{{\frac{1}{n}+\frac{1}{\beta}}\choose\frac{1}{\beta}}}{\mu(K)}\left(\frac{\mu(K)}{{{\frac{1}{n}+\frac{q}{\beta}}\choose\frac{q}{\beta}}}\right)^{1/q}\int_{K}f(x)d\mu(x).

To see how (16) yields results for the relative entropy of two measures with concavity, based on the work by Bobkov and Madiman [5] for Boltzmann-Shannon entropy, see [6].

3. Applications to Convex Geometry

3.1. Weighted Radial Mean Bodies

Throughout this section, we write λn\lambda_{n} for the Lebesgue measure on n\mathbb{R}^{n}.

One of our motivations for generalizing Berwald’s inequality is to study generalizations of the projection body and radial mean bodies of a convex body. We first introduce weighted radial mean bodies. For a Borel measure μ\mu finite on a Borel set KK in its support, the ppth mean of a non-negative fLp(K,μ)f\in L^{p}(K,\mu) is

(17) Mp,μf=(1μ(K)Kf(x)p𝑑μ(x))1p.M_{p,\mu}f=\left(\frac{1}{\mu(K)}\int_{K}f(x)^{p}d\mu(x)\right)^{\frac{1}{p}}.

Jensen’s inequality states that Mμ,pfMμ,qfM_{\mu,p}f\leq M_{\mu,q}f for pq.p\leq q. From continuity, one has limpMp,μf=esssupxKf(x),\lim_{p\to\infty}M_{p,\mu}f=\text{ess}\sup_{x\in K}f(x), and

limp0Mp,μf=exp(1μ(K)Klogf(x)𝑑μ(x)).\lim_{p\rightarrow 0}M_{p,\mu}f=\exp\left(\frac{1}{\mu(K)}\int_{K}\log f(x)d\mu(x)\right).

Recall that for a star body KK, ρK(x,θ):=ρKx(θ)=sup{λ:λθKx}\rho_{K}(x,\theta):=\rho_{K-x}(-\theta)=\sup\{\lambda:-\lambda\theta\in K-x\}.

Definition 3.1.

Let μ\mu be a Borel measure on n\mathbb{R}^{n} and KK a convex body contained in the support of μ\mu. Then, the μ\mu-weighted ppth radial mean body of KK, denoted Rp,μK,R_{p,\mu}K, is the star body whose radial function is given, for p(1,)p\in(-1,\infty) and θ𝕊n1,\theta\in\mathbb{S}^{n-1}, as

ρRp,μK(θ)=(1μ(K)KρK(x,θ)p𝑑μ(x))1p.\rho_{R_{p,\mu}K}(\theta)=\left(\frac{1}{\mu(K)}\int_{K}\rho_{K}(x,\theta)^{p}d\mu(x)\right)^{\frac{1}{p}}.

We will show why Rp,μKR_{p,\mu}K exists for p(1,0)p\in(-1,0) later in this section. The usual radial mean bodies RpKR_{p}K, first defined by Gardner and Zhang [22], are precisely the bodies Rp,λnKR_{p,\lambda_{n}}K in our notation. Sending pp to \infty or 0, we obtain the limiting bodies R,μKR_{\infty,\mu}K and R0,μKR_{0,\mu}K, given in terms of their radial functions by

ρR,μK(θ)\displaystyle\rho_{R_{\infty,\mu}K}(\theta) =maxxKρK(x,θ)=ρDK(θ),and\displaystyle=\max_{x\in K}\rho_{K}(x,\theta)=\rho_{DK}(\theta),\quad\text{and}
ρR0,μK(θ)\displaystyle\rho_{R_{0,\mu}K}(\theta) =exp(1μ(K)KlogρK(x,θ)𝑑μ(x)),\displaystyle=\exp\left(\frac{1}{\mu(K)}\int_{K}\log\rho_{K}(x,\theta)d\mu(x)\right),

where DKDK is the difference body of KK, given by

(18) DK={x:K(K+x)}=K+(K).DK=\{x:K\cap(K+x)\neq\emptyset\}=K+(-K).

A natural question is how Rp,μKR_{p,\mu}K behaves under linear transformations. We introduce the following notation: for a Borel measure μ\mu and TSLnT\in SL_{n}, we denote by μT\mu^{T} the pushforward of μ\mu by T1T^{-1}; note that if μ\mu has density ϕ\phi then μT\mu^{T} has density ϕT\phi\circ T, and μT(A)=μ(TA)\mu^{T}(A)=\mu(TA) for a Borel set AA.

Proposition 3.2.

Let μ\mu be a Borel measure finite on a convex body KsuppμK\subset\operatorname{supp}\mu. Then, for TSLnT\in SL_{n} and p>1,p>-1, one has

Rp,μTK=TRp,μTK.R_{p,\mu}TK=TR_{p,\mu^{T}}K.
Proof.

Suppose p(1,0)(0,)p\in(-1,0)\cup(0,\infty); the case p=0p=0 follows by continuity. Let LL be a star body in n.\mathbb{R}^{n}. Then, one can verify from the definition (or see [21, page 20]) that

ρTL(x,θ)=ρL(T1x,T1θ).\rho_{TL}(x,\theta)=\rho_{L}(T^{-1}x,T^{-1}\theta).

In particular, ρTL(θ)=ρL(T1θ).\rho_{TL}(\theta)=\rho_{L}(T^{-1}\theta). Then, observe that, by performing the variable substitution x=Tz,x=Tz,

ρRp,μTKp(θ)\displaystyle\rho^{p}_{R_{p,\mu}TK}(\theta) =1μ(TK)TKρTK(x,θ)p𝑑μ(x)=1μ(TK)TKρK(T1x,T1θ)p𝑑μ(x)\displaystyle=\frac{1}{\mu(TK)}\int_{TK}\rho_{TK}(x,\theta)^{p}d\mu(x)=\frac{1}{\mu(TK)}\int_{TK}\rho_{K}(T^{-1}x,T^{-1}\theta)^{p}d\mu(x)
=1μT(K)KρK(z,T1θ)p𝑑μT(z)=ρRp,μTKp(T1θ)=ρTRp,μTKp(θ).\displaystyle=\frac{1}{\mu^{T}(K)}\int_{K}\rho_{K}(z,T^{-1}\theta)^{p}d\mu^{T}(z)=\rho^{p}_{R_{p,\mu^{T}}K}(T^{-1}\theta)=\rho^{p}_{TR_{p,\mu^{T}}K}(\theta).

We will show that when μ\mu is ss-concave, s0s\geq 0, then Rp,μKR_{p,\mu}K is a convex body for p0p\geq 0. But first, we must make a detour.

3.2. Weighted Projection Bodies

A convex body K𝒦nK\in\mathcal{K}^{n} is uniquely determined by its support function hK(x)=sup{x,y:yK}h_{K}(x)=\sup\{\langle x,y\rangle\colon y\in K\}. The dual body of K𝒦0nK\in\mathcal{K}^{n}_{0} is given by K={xn:hK(x)1}K^{\circ}=\left\{x\in\mathbb{R}^{n}:h_{K}(x)\leq 1\right\} (notice this yields that hK(x)=xKh_{K}(x)=\|x\|_{K^{\circ}}). For a Borel measure μ\mu with density ϕ\phi and a K𝒦nK\in\mathcal{K}^{n}, the μ\mu-measure of the boundary of KK, denoted K\partial K, is

(19) μ+(K):=lim infϵ0μ(K+ϵB2n)μ(K)ϵ=Kϕ(y)𝑑n1(y),\mu^{+}(\partial K):=\liminf_{\epsilon\to 0}\frac{\mu\left(K+\epsilon B_{2}^{n}\right)-\mu(K)}{\epsilon}=\int_{\partial K}\phi(y)d\mathcal{H}^{n-1}(y),

where the second equality holds if the lim inf\liminf is a limit and there exists some canonical way to select how ϕ\phi behaves on K.\partial K. Here, n1\mathcal{H}^{n-1} is the (n1)(n-1)-dimensional Hausdorff measure. It was folklore for quite some time that this formula holds for measures with continuous density (see e.g. K. Ball’s work on the Gaussian measure [2]). This was proved rigorously by Livshyts [35]. More recently, it was shown by the first-named author and Kryvonos [29] that the formula holds for every Borel measure μ\mu with density ϕ\phi, as long as ϕ\phi contains K\partial K in its Lebesgue set (see Lemma 3.5 below).

Using weighted surface area measures, the centered, μ\mu-weighted projection bodies of a convex body KK and a Borel measure μ\mu with continuous density ϕ\phi were defined as [30] the symmetric convex body whose support function is given by, for θ𝕊n1\theta\in\mathbb{S}^{n-1},

(20) hΠ~μK(θ)=12K|θ,nK(y)|ϕ(y)𝑑n1(y),h_{\widetilde{\Pi}_{\mu}K}(\theta)=\frac{1}{2}\int_{\partial K}|\langle\theta,n_{K}(y)\rangle|\phi(y)d\mathcal{H}^{n-1}(y),

where nK:K𝕊n1n_{K}:\partial K\rightarrow\mathbb{S}^{n-1} is the Gauss map, which associates an element yKy\in\partial K with its outer unit normal. Since the set NK={xK:nK(x) is not well-defined}N_{K}=\{x\in\partial K:n_{K}(x)\text{ is not well-defined}\} is of measure zero, we will write integrals over KNK\partial K\setminus N_{K} involving nKn_{K} as integrals over K\partial K without any confusion.

As alluded to by the discussion after (19), the formula (20) is still well-defined when ϕ\phi is not continuous but contains K\partial K in its Lebesgue set; in which case, ϕ\phi on K\partial K is understood as, for yKy\in\partial K,

ϕ(y)=limϵ01Voln(y+ϵB2n)y+ϵB2nϕ(x)𝑑x.\phi(y)=\lim_{\epsilon\to 0}\frac{1}{\text{\rm Vol}_{n}(y+\epsilon B_{2}^{n})}\int_{y+\epsilon B_{2}^{n}}\phi(x)dx.

For such KK and ϕ\phi (== density of μ\mu), the shift of KK with respect to μ\mu is given by

ημ,K=12KnK(y)ϕ(y)𝑑n1(y)=12Kϕ(y)𝑑y,\eta_{\mu,K}=\frac{1}{2}\int_{\partial K}n_{K}(y)\phi(y)d\mathcal{H}^{n-1}(y)=\frac{1}{2}\int_{K}\nabla\phi(y)dy,

where the second equality holds when ϕ\phi is in C1(K).C^{1}(K). Recall the notation that, if ff is a function, then there exists two non-negative functions, denoted f+f_{+} and ff_{-}, such that f=f+ff=f_{+}-f_{-}. One can then write |f|=f++f|f|=f_{+}+f_{-} and obtain |f|f=2f.|f|-f=2f_{-}. We define the μ\mu-weighted projection body of KK to be the convex body ΠμK\Pi_{\mu}K defined via the support function, for every θ𝕊n1\theta\in\mathbb{S}^{n-1}

(21) hΠμK(θ):=hΠ~μK(θ)ημ,K,θ=12K|θ,nK(y)|ϕ(y)𝑑n1(y)12Kθ,nK(y)ϕ(y)𝑑n1(y)=Kθ,nK(y)ϕ(y)𝑑n1(y),\begin{split}&h_{\Pi_{\mu}K}(\theta):=h_{\widetilde{\Pi}_{\mu}K}(\theta)-\langle\eta_{\mu,K},\theta\rangle\\ &=\frac{1}{2}\int_{\partial K}|\langle\theta,n_{K}(y)\rangle|\phi(y)d\mathcal{H}^{n-1}(y)-\frac{1}{2}\int_{\partial K}\langle\theta,n_{K}(y)\rangle\phi(y)d\mathcal{H}^{n-1}(y)\\ &=\int_{\partial K}\langle\theta,n_{K}(y)\rangle_{-}\phi(y)d\mathcal{H}^{n-1}(y),\end{split}

where the last integral is to emphasize that ΠμK\Pi_{\mu}K contains the origin in its interior. In the case when μ=λn\mu=\lambda_{n}, one has ΠK:=Π~λnK=ΠλnK,\Pi K:=\widetilde{\Pi}_{\lambda^{n}}K=\Pi_{\lambda^{n}}K, where ΠK\Pi K is the projection body of KK. This projection body is a fundamental tool in convex geometry; see e.g. [21]. It turns out that for θ𝕊n1\theta\in\mathbb{S}^{n-1}:

(22) hΠK(θ)=12K|θ,nK(y)|𝑑n1(y)=Voln1(PθK),h_{\Pi K}(\theta)=\frac{1}{2}\int_{\partial K}|\langle\theta,n_{K}(y)\rangle|d\mathcal{H}^{n-1}(y)=\text{\rm Vol}_{n-1}\left(P_{\theta^{\perp}}K\right),

where the first equality is from (20), the orthogonal projection of KK onto a linear subspace HH is denoted by PHKP_{H}K, and the last equality is known as Minkowski’s projection formula. We next introduce the weighted covariogram of a convex body.

3.3. The Covariogram and Radial Mean Bodies

Definition 3.3.

Let KK be a convex body in n.\mathbb{R}^{n}. Then, for a Borel measure μ\mu, the μ\mu-covariogram of KK is the function given by

(23) gμ,K(x)=μ(K(K+x)).g_{\mu,K}(x)=\mu(K\cap(K+x)).

The classical covariogram of KK is given by gK:=gλn,Kg_{K}:=g_{\lambda_{n},K}. In [30], the following was proven, which extends the volume case first shown by Matheron [37]. Recall that a domain is an open, connected set with non-empty interior, and that a function q:Ωq:\Omega\to\mathbb{R} is Lipschitz on a bounded domain Ω\Omega if, for every x,yΩx,y\in\Omega, one has |q(x)q(y)|C|xy||q(x)-q(y)|\leq C|x-y| for some C>0C>0.

Proposition 3.4 (The radial derivative of the covariogram, [30]).

Let K𝒦nK\in\mathcal{K}^{n}. Suppose Ω\Omega is a domain containing KK, and consider a Borel measure μ\mu with density ϕ\phi locally Lipschitz on Ω\Omega. Then,

(24) \diffgμ,K(rθ)r|r=0=hΠμK(θ).\diff{g_{\mu,K}(r\theta)}{r}\bigg{|}_{r=0}=-h_{\Pi_{\mu}K}(\theta).

We now briefly show that the assumption of Lipschitz density can be dropped. For a continuous function h:𝕊n1(0,)h:\mathbb{S}^{n-1}\to(0,\infty), the Wulff shape or Alexandrov body of hh is defined as

[h]=uSn1{xn:x,uh(u)}.[h]=\bigcap_{u\in S^{n-1}}\{x\in\mathbb{R}^{n}:\langle x,u\rangle\leq h(u)\}.

In [29], the first-named author and Kryvonos established the following formula, generalizing the volume case and extending the partial case found in [35].

Lemma 3.5 (Aleksandrov’s variational formula for arbitrary measures, [29]).

Let μ\mu be a Borel measure on n\mathbb{R}^{n} with locally integrable density ϕ\phi. Let KK be a convex body such that K\partial K, up to set of (n1)(n-1)-dimensional Hausdorff measure zero, is in the Lebesgue set of ϕ\phi. Then, for a continuous function ff on 𝕊n1\mathbb{S}^{n-1}, one has that

limt0μ([hK+tf])μ(K)t=Kf(nK(y))𝑑n1(y).\lim_{t\rightarrow 0}\frac{\mu([h_{K}+tf])-\mu(K)}{t}=\int_{\partial K}f(n_{K}(y))d\mathcal{H}^{n-1}(y).

Next, note that for any θn\theta\in\mathbb{R}^{n}, hK+rθ(u)=hK(u)+ru,θh_{K+r\theta}(u)=h_{K}(u)+r\langle u,\theta\rangle. Also, for any convex body LL we have L=u𝕊n1{x:u,xhL(u)}L=\bigcap_{u\in\mathbb{S}^{n-1}}\{x:\langle u,x\rangle\leq h_{L}(u)\}. Consequently,

K(K+rθ)\displaystyle K\cap(K+r\theta) =u𝕊n1{x:u,xhK(u)}u𝕊n1{x:u,xhK+rθ(u)}\displaystyle=\bigcap_{u\in\mathbb{S}^{n-1}}\{x:\langle u,x\rangle\leq h_{K}(u)\}\cap\bigcap_{u\in\mathbb{S}^{n-1}}\{x:\langle u,x\rangle\leq h_{K+r\theta}(u)\}
=u𝕊n1({x:u,xhK(u)}{x:u,xhK+rθ(u)})\displaystyle=\bigcap_{u\in\mathbb{S}^{n-1}}(\{x:\langle u,x\rangle\leq h_{K}(u)\}\cap\{x:\langle u,x\rangle\leq h_{K+r\theta}(u)\})
=u𝕊n1{x:u,xmin(hK(u),hK(u)+rθ,u)}\displaystyle=\bigcap_{u\in\mathbb{S}^{n-1}}\{x:\langle u,x\rangle\leq\min(h_{K}(u),h_{K}(u)+r\langle\theta,u\rangle)\}
=u𝕊n1{x:u,xhK(u)+rmin(0,u,θ)}.\displaystyle=\bigcap_{u\in\mathbb{S}^{n-1}}\{x:\langle u,x\rangle\leq h_{K}(u)+r\min(0,\langle u,\theta\rangle)\}.

Thus, the body Kr(θ)=K(K+rθ)K_{r}(\theta)=K\cap(K+r\theta) is the Wulff shape of the function uhK(u)ru,θu\mapsto h_{K}(u)-r\langle u,\theta\rangle_{-}. Suppose we have a Borel measure μ\mu with density ϕ\phi, such that K\partial K is in the Lebesgue set of ϕ\phi. Then, observe that gμ,K(rθ)=μ(Kr(θ))g_{\mu,K}(r\theta)=\mu(K_{r}(\theta)). Therefore, we obtain (24) from Lemma 3.5, with f(u)=u,θf(u)=-\langle u,\theta\rangle_{-}, and (21). We list this strengthened version as a separate result.

Theorem 3.6.

Let KK be a convex body in n\mathbb{R}^{n} and μ\mu a Borel measure whose density ϕ:n+\phi\colon\mathbb{R}^{n}\to\mathbb{R}^{+} contains K\partial K in its Lebesgue set and KK in its support. Then, for every fixed direction θ𝕊n1\theta\in\mathbb{S}^{n-1}, one has

(25) \diffgμ,K(rθ)r|r=0=hΠμK(θ).\diff{g_{\mu,K}(r\theta)}{r}\bigg{|}_{r=0}=-h_{\Pi_{\mu}K}(\theta).

From the Brunn-Minkowski inequality, gKg_{K} is a 1/n1/n-concave function supported on DKDK. One can readily check that the μ\mu-covariogram inherits the concavity of the measure μ\mu in general.

Proposition 3.7 (Concavity of the covariogram).

Consider a class of convex bodies 𝒞𝒦n\mathcal{C}\subseteq\mathcal{K}^{n} with the property that K𝒞K(K+x)𝒞K\in\mathcal{C}\rightarrow K\cap(K+x)\in\mathcal{C} for every xDKx\in DK. Let μ\mu be a Borel measure finite on every K𝒞.K\in\mathcal{C}. Suppose FF is a continuous and invertible function such that μ\mu is FF-concave on 𝒞\mathcal{C}. Then, for K𝒞,K\in\mathcal{C}, gμ,Kg_{\mu,K} is also FF-concave, in the sense that, if FF is increasing, then Fgμ,KF\circ g_{\mu,K} is concave, and if FF is decreasing, then Fgμ,KF\circ g_{\mu,K} is convex.

Proof.

We first observe the following set inclusion: for x,ynx,y\in\mathbb{R}^{n} and λ[0,1]\lambda\in[0,1], we have from convexity that

K(K+(1λ)x+λy)\displaystyle K\cap(K+(1-\lambda)x+\lambda y) =K((1λ)(K+x)+λ(K+y))\displaystyle=K\cap((1-\lambda)(K+x)+\lambda(K+y))
(1λ)(K(K+x))+λ(K(K+y)).\displaystyle\supset(1-\lambda)(K\cap(K+x))+\lambda(K\cap(K+y)).

Using this set inclusion, we obtain that

gμ,K((1λ)x+λy)μ((1λ)(K(K+x))+λ(K(K+y))).g_{\mu,K}((1-\lambda)x+\lambda y)\geq\mu((1-\lambda)(K\cap(K+x))+\lambda(K\cap(K+y))).

From the fact that μ\mu is FF-concave, we obtain

gμ,K((1λ)x+λy)\displaystyle g_{\mu,K}((1-\lambda)x+\lambda y) F1((1λ)F(μ(K(K+x)))+λF(μ(K(K+y))))\displaystyle\geq F^{-1}\left((1-\lambda)F\left(\mu(K\cap(K+x))\right)+\lambda F\left(\mu(K\cap(K+y))\right)\right)
=F1((1λ)F(gμ,K(x))+λF(gμ,K(y))).\displaystyle=F^{-1}\left((1-\lambda)F(g_{\mu,K}(x))+\lambda F(g_{\mu,K}(y))\right).

We see that the (μ\mu-)covariogram connects the (μ\mu-)projection body and difference body of KK. The covariogram is also related to the radial mean bodies, since, for p>0p>0:

KρK(x,θ)p𝑑μ(x)\displaystyle\int_{K}\rho_{K}(x,\theta)^{p}d\mu(x) =pK0ρK(x,θ)rp1𝑑r𝑑μ(x)\displaystyle=p\int_{K}\int_{0}^{\rho_{K}(x,\theta)}r^{p-1}dr\,d\mu(x)
=p0ρDK(θ)(K(K+rθ)𝑑μ(x))rp1𝑑r\displaystyle=p\int_{0}^{\rho_{DK}(\theta)}\left(\int_{K\cap(K+r\theta)}d\mu(x)\right)r^{p-1}dr
=p0ρDK(θ)gμ,K(rθ)rp1𝑑r=pgμ,K(rθ)(p),\displaystyle=p\int_{0}^{\rho_{DK}(\theta)}g_{\mu,K}(r\theta)r^{p-1}dr=p\mathcal{M}_{g_{\mu,K}(r\theta)}(p),

where, in the second step, we used Fubini’s theorem and the fact that xKx\in K and rθKx-r\theta\in K-x for all 0rρK(x,θ)0\leq r\leq\rho_{K}(x,\theta). Therefore, we can write, for p>0p>0, that

(26) ρRp,μK(θ)=(pμ(K)0ρDK(θ)gμ,K(rθ)rp1𝑑r)1p=(pμ(K))1pgμ,K(rθ)(p)1p.\rho_{R_{p,\mu}K}(\theta)\!=\!\left(\frac{p}{\mu(K)}\int_{0}^{\rho_{DK}(\theta)}\!g_{\mu,K}(r\theta)r^{p-1}dr\right)^{\frac{1}{p}}\!=\!\left(\frac{p}{\mu(K)}\right)^{\frac{1}{p}}\mathcal{M}_{g_{\mu,K}(r\theta)}(p)^{\frac{1}{p}}.

Additionally, this formulation implies that Rp,μKR_{p,\mu}K is a convex body if μ\mu is ss-concave in the Borell sense for some s0s\geq 0 (see [2, Theorem 5] for p1p\geq 1 and [22, Corollary 4.2]).

3.4. The Negative Regime for p

We first convince the reader that Rp,μKR_{p,\mu}K exists for p(1,0)p\in(-1,0) when μ\mu has a bounded, positive density ϕ\phi. Under these assumptions, let M=minxKϕ(x).M=\min_{x\in K}\phi(x). Then, for p(1,0)(0,):p\in(-1,0)\cup(0,\infty):

Mϕ1Voln(K)KρK(x,θ)p𝑑x\displaystyle\frac{M}{\|\phi\|_{\infty}}\frac{1}{\text{\rm Vol}_{n}(K)}\int_{K}\rho_{K}(x,\theta)^{p}dx 1μ(K)KρK(x,θ)p𝑑μ(x)\displaystyle\leq\frac{1}{\mu(K)}\int_{K}\rho_{K}(x,\theta)^{p}d\mu(x)
ϕM1Voln(K)KρK(x,θ)p𝑑x.\displaystyle\leq\frac{\|\phi\|_{\infty}}{M}\frac{1}{\text{\rm Vol}_{n}(K)}\int_{K}\rho_{K}(x,\theta)^{p}dx.

One then deduces that under these constraints, for p>0,p>0, (Mϕ)1pRpKRp,μK(ϕM)1pRpK,\left(\frac{M}{\|\phi\|_{\infty}}\right)^{\frac{1}{p}}R_{p}K\subseteq R_{p,\mu}K\subseteq\left(\frac{\|\phi\|_{\infty}}{M}\right)^{\frac{1}{p}}R_{p}K, and, for p(1,0),p\in(-1,0), one has (Mϕ)1pRpKRp,μK(ϕM)1pRpK\left(\frac{M}{\|\phi\|_{\infty}}\right)^{\frac{1}{p}}R_{p}K\supseteq R_{p,\mu}K\supseteq\left(\frac{\|\phi\|_{\infty}}{M}\right)^{\frac{1}{p}}R_{p}K. There is equality if, and only if, ϕ\phi is constant on K.K. Notice these inclusions show that Rp,μKR_{p,\mu}K is well-defined for p(1,0)p\in(-1,0). By sending p1p\to-1 we deduce that Rp,μK{0}R_{p,\mu}K\to\{0\} as p1.p\to-1.

For a general Borel measure μ\mu with density, we now obtain a formula for Rp,μKR_{p,\mu}K when p(1,0).p\in(-1,0). This also establishes existence. Notice that, in this instance,

KρK\displaystyle\int_{K}\rho_{K} (x,θ)pdμ(x)=pKρK(x,θ)rp1𝑑r𝑑μ(x)\displaystyle(x,\theta)^{p}d\mu(x)=-p\int_{K}\int_{\rho_{K}(x,\theta)}^{\infty}r^{p-1}drd\mu(x)
=p0ρDK(θ)(KK(K+rθ)𝑑μ(x))rp1𝑑rpKρDK(θ)rp1𝑑r𝑑μ(x).\displaystyle=-p\int_{0}^{\rho_{DK}(\theta)}\left(\int_{K\setminus{K\cap(K+r\theta)}}d\mu(x)\right)r^{p-1}dr-p\int_{K}\int_{\rho_{DK}(\theta)}^{\infty}r^{p-1}drd\mu(x).

Adding and subtracting integration over K(K+rθ),K\cap(K+r\theta), we obtain

KρK(x,θ)p𝑑μ(x)\displaystyle\int_{K}\rho_{K}(x,\theta)^{p}d\mu(x) =p0ρDK(θ)(gμ,K(rθ)μ(K))rp1𝑑r+ρDKp(θ)μ(K)\displaystyle=p\int_{0}^{\rho_{DK}(\theta)}(g_{\mu,K}(r\theta)-\mu(K))r^{p-1}dr+\rho^{p}_{DK}(\theta)\mu(K)
=pgμ,K(rθ)(p).\displaystyle=p\mathcal{M}_{g_{\mu,K}(r\theta)}(p).

Notice this formulation could have been established directly via the continuity of the Mellin transform. Hence, we can write, for p(1,0),p\in(-1,0), that

(27) ρRp,μK(θ)=(pμ(K)0ρDK(θ)(gμ,K(rθ)μ(K))rp1𝑑r+ρDKp(θ))1p=(pμ(K))1pgμ,K(rθ)(p)1p.\begin{split}\rho_{R_{p,\mu}K}(\theta)&=\left(\frac{p}{\mu(K)}\int_{0}^{\rho_{DK}(\theta)}(g_{\mu,K}(r\theta)-\mu(K))r^{p-1}dr+\rho^{p}_{DK}(\theta)\right)^{\frac{1}{p}}\\ &=\left(\frac{p}{\mu(K)}\right)^{\frac{1}{p}}\mathcal{M}_{g_{\mu,K}(r\theta)}(p)^{\frac{1}{p}}.\end{split}

The last equality is to emphasis that (27) is the analytic continuation of (26), as discussed in Section 2. Now that we have shown the existence of Rp,μKR_{p,\mu}K, we can use properties of ppth averages of functions, i.e. Jensen’s inequality, to immediately obtain the following.

Theorem 3.8.

Let μ\mu be a Borel measure finite on a convex body KK contained in its support. Then one has that, for 1<pq,-1<p\leq q\leq\infty,

Rp,μKRq,μKR,μK=DK.R_{p,\mu}K\subseteq R_{q,\mu}K\subseteq R_{\infty,\mu}K=DK.

We now take a moment to discuss how p(1,0)p\in(-1,0) was originally handled in the volume case. Gardner and Zhang defined another family of star bodies depending on K𝒦nK\in\mathcal{K}^{n}, the spectral pth mean bodies of K,K, denoted SpK.S_{p}K. However, to apply Jensen’s inequality, they had to assume additionally that Voln(K)=1.\text{\rm Vol}_{n}(K)=1. To avoid this assumption, we change the normalization and define SpKS_{p}K as the star body whose radial function is given by, for p[1,),p\in[-1,\infty),

ρSpK(θ)=(PθKXθK(y)p(XθK(y)dyVoln(K)))1/p,\rho_{S_{p}K}(\theta)=\left(\int_{P_{\theta^{\perp}}K}X_{\theta}K(y)^{p}\left(\frac{X_{\theta}K(y)dy}{\text{\rm Vol}_{n}(K)}\right)\right)^{1/p},

where XθK(y)=Vol1(K(y+θ))X_{\theta}K(y)=\text{\rm Vol}_{1}(K\cap(y+\theta\mathbb{R})) is the X-ray of KK in the direction θ𝕊n1\theta\in\mathbb{S}^{n-1} for yPθKy\in P_{\theta^{\perp}}K (see [21, Chapter 1] for more on the properties of XθKX_{\theta}K, and note that PθKXθK(y)dyVoln(K)=1\int_{P_{\theta^{\perp}}K}\frac{X_{\theta}K(y)dy}{\text{\rm Vol}_{n}(K)}=1), ρSK(θ)=maxyθXθK(y)=ρDK(θ),\rho_{S_{\infty}K}(\theta)=\max_{y\in\theta^{\perp}}X_{\theta}K(y)=\rho_{DK}(\theta), ρS0K(θ)=exp(PθKlog(XθK(y))XθK(y)dyVoln(K)),\rho_{S_{0}K}(\theta)=\exp\left(\int_{P_{\theta^{\perp}}K}\log(X_{\theta}K(y))\frac{X_{\theta}K(y)dy}{\text{\rm Vol}_{n}(K)}\right), and

ρS1K(θ)=Voln(K)Voln1(PθK)1=Voln(K)ρΠK(θ).\rho_{S_{-1}K}(\theta)=\text{\rm Vol}_{n}(K)\text{\rm Vol}_{n-1}(P_{\theta^{\perp}}K)^{-1}=\text{\rm Vol}_{n}(K)\rho_{\Pi^{\circ}K}(\theta).

Here, ΠK(ΠK)\Pi^{\circ}K\equiv(\Pi K)^{\circ} is the polar projection body of KK.

Therefore, from Jensen’s inequality, we obtain, for 1pq,-1\leq p\leq q\leq\infty,

(28) Voln(K)ΠK=S1KSpKSqKSK=DK.\text{\rm Vol}_{n}(K)\Pi^{\circ}K=S_{-1}K\subseteq S_{p}K\subseteq S_{q}K\subseteq S_{\infty}K=DK.

The fact that, for p>1,p>-1,

(29) 1p+1PθKXθK(y)p+1𝑑y=PθK0XθK(y)rp𝑑r𝑑y=KρK(x,θ)p𝑑x\frac{1}{p+1}\int_{P_{\theta^{\perp}}K}X_{\theta}K(y)^{p+1}dy=\int_{P_{\theta^{\perp}}K}\int_{0}^{X_{\theta}K(y)}r^{p}drdy=\int_{K}\rho_{K}(x,\theta)^{p}dx

shows S0K=eR0KS_{0}K=eR_{0}K, SpK=(p+1)1/pRpK,p>0,S_{p}K=(p+1)^{1/p}R_{p}K,\,p>0, and that we can analytically continue RpKR_{p}K to p(1,0)p\in(-1,0) by RpK:=(p+1)1/pSpK.R_{p}K:=(p+1)^{-1/p}S_{p}K. As observed in [22], the relation RpK=(p+1)1/pSpKR_{p}K=(p+1)^{-1/p}S_{p}K shows that RpK{0}R_{p}K\to\{0\} as p1,p\to-1, but the shape of RpKR_{p}K tends to that of S1K=Voln(K)ΠKS_{-1}K=\text{\rm Vol}_{n}(K)\Pi^{\circ}K (note that due to the alternate normalization of SpK,S_{p}K, these relations are expressed differently in [22, Theorem 2.2]; in both instances, it is unknown if RpKR_{p}K and SpKS_{p}K are convex for p(1,0)p\in(-1,0)).

Gardner and Zhang then obtained a chain of inequalities concerning RpKR_{p}K ([22, Theorem 5.5]), a reverse to the one from Jensen’s inequality: for 1<pq<-1<p\leq q<\infty

(30) DKcn,qRqKcn,pRpKnVoln(K)ΠK,DK\subseteq c_{n,q}R_{q}K\subseteq c_{n,p}R_{p}K\subseteq n\text{\rm Vol}_{n}(K)\Pi^{\circ}K,

where cn,pc_{n,p} are constants defined via continuity at p=0p=0, and, for p(1,0)(0,).p\in(-1,0)\cup(0,\infty). cn,p=(nB(p+1,n))1/p,c_{n,p}=(nB(p+1,n))^{-1/p}, with B(x,y)B(x,y) the standard Beta function. There is equality in each inclusion in (30) if, and only if, KK is a nn-dimensional simplex. One obtains from (29) that, indeed, cn,pRpKc_{n,p}R_{p}K tends to nVoln(K)ΠKn\text{\rm Vol}_{n}(K)\Pi^{\circ}K as p1+p\to-1^{+}, since cn,p(p+1)1pc_{n,p}(p+1)^{-\frac{1}{p}} tends to nn. When p=np=n, one obtains, since Voln(RnK)=Voln(K)\text{\rm Vol}_{n}(R_{n}K)=\text{\rm Vol}_{n}(K), the following special case:

(31) Voln(DK)Voln(K)(2nn)nnVoln(K)n1Voln(ΠK).\frac{\text{\rm Vol}_{n}(DK)}{\text{\rm Vol}_{n}(K)}\leq{2n\choose n}\leq n^{n}\text{\rm Vol}_{n}(K)^{n-1}\text{\rm Vol}_{n}(\Pi^{\circ}K).

The left-hand side is the inequality of Rogers-Shephard inequality [42], and the right-hand side is Zhang’s inequality [44].

The goal of the next subsection is to generalize (30). To determine the behaviour of Rp,μKR_{p,\mu}K as p1+p\to-1^{+}, we will not pass through spectral mean bodies. To explain why, we shall, for simplicity, focus on the Gaussian measure and a symmetric K𝒦0nK\in\mathcal{K}^{n}_{0}. Suppose we defined Gaussian spectral mean bodies Sp,γnKS_{p,\gamma_{n}}K as the star body whose radial function is given by, for p[1,),p\in[-1,\infty),

ρSp,γnK(θ)=(PθKγ1(K(y+θ))p+1(dγn1(y)γn(K)))1/p.\rho_{S_{p,\gamma_{n}}K}(\theta)=\left(\int_{P_{\theta^{\perp}}K}\gamma_{1}(K\cap(y+\theta\mathbb{R}))^{p+1}\left(\frac{d\gamma_{n-1}(y)}{\gamma_{n}(K)}\right)\right)^{1/p}.

Notice that an analogue of (29), which relates the radial functions of RpKR_{p}K and SpKS_{p}K when p>1,p>-1, does not hold. Consequently, we cannot determine the shape of Rp,γnKR_{p,\gamma_{n}}K as p1p\to-1 via Sp,γnKS_{p,\gamma_{n}}K. Perhaps then, the focus should be on Sp,γnKS_{p,\gamma_{n}}K and not Rp,γnKR_{p,\gamma_{n}}K. But notice that

ρS1,γn(K)(θ)=γn(K)γn1(PθK)1γn(K)ρΠγnK(θ)\rho_{S_{-1,\gamma_{n}}(K)}(\theta)=\gamma_{n}(K)\gamma_{n-1}(P_{\theta^{\perp}}K)^{-1}\neq\gamma_{n}(K)\rho_{\Pi^{\circ}_{\gamma_{n}}K}(\theta)

since one does not have an equivalent of Minkowski’s integral formula in the weighted case. Furthermore, it is not necessarily true that γn1(PθK)\gamma_{n-1}(P_{\theta^{\perp}}K) is convex as a function of θ.\theta. Hence, it is not necessarily the Minkowski functional of a convex body. Additionally, ρS,γnK(θ)=maxyθγ1(K(y+θ))ρDK(θ).\rho_{S_{\infty,\gamma_{n}}K}(\theta)=\max_{y\in\theta^{\perp}}\gamma_{1}(K\cap(y+\theta\mathbb{R}))\neq\rho_{DK}(\theta). To summarize, Sp,γnKS_{p,\gamma_{n}}K is not related to DKDK or ΠγnK,\Pi_{\gamma_{n}}^{\circ}K, and Rp,γnKR_{p,\gamma_{n}}K is not related to Sp,γnK.S_{p,\gamma_{n}}K. It is for these reasons we do not study weighted spectral mean bodies.

We must determine the shape of Rp,μKR_{p,\mu}K as p1.p\to-1. Applying integration by parts to both (26) and (27), we obtain that, for all p(1,0)(0,),p\in(-1,0)\cup(0,\infty), one has

(32) ρRp,μK(θ)p=0ρDK(θ)(gμ,K(rθ)μ(K))rp𝑑r,\rho_{R_{p,\mu}K}(\theta)^{p}=\int_{0}^{\rho_{DK}(\theta)}\left(\frac{-g_{\mu,K}(r\theta)^{\prime}}{\mu(K)}\right)r^{p}dr,

where we used Lebesgue’s theorem to obtain that gμ,K(rθ)g_{\mu,K}(r\theta) is differentiable almost everywhere on [0,ρDK(θ)],[0,\rho_{DK}(\theta)], as it is monotonically decreasing in the variable r.r. Taking the limit as p1,p\to-1, we see that Rp,μK{o}.R_{p,\mu}K\to\{o\}.

On the other-hand, recall the following lemma.

Lemma 3.9 (Lemma 4 in [26] / Lemma 8 in [25]).

If φ:[0,)[0,)\varphi:[0,\infty)\rightarrow[0,\infty) is a measurable function with limt0+φ(t)=\lim_{t\rightarrow 0^{+}}\varphi(t)= φ(0)\varphi(0) and such that 0ts0φ(t)dt<\int_{0}^{\infty}t^{-s_{0}}\varphi(t)\mathrm{d}t<\infty for some s0(0,1)s_{0}\in(0,1), then

lims1(1s)0tsφ(t)dt=φ(0).\lim_{s\rightarrow 1^{-}}(1-s)\int_{0}^{\infty}t^{-s}\varphi(t)\mathrm{d}t=\varphi(0).

Therefore, identifying p=sp=-s in Lemma 3.9, we obtain from Theorem 3.6 that, for μ\mu with locally integrable density and a convex body KK such that K\partial K is in the Lebesgue set of the density of μ\mu, one has

(33) limp1(p+1)1/pρRp,μK(θ)=μ(K)ρΠμK(θ),\lim_{p\to-1}(p+1)^{1/p}\rho_{R_{p,\mu}K}(\theta)=\mu(K)\rho_{\Pi^{\circ}_{\mu}K}(\theta),

establishing that the shape of Rp,μKR_{p,\mu}K approaches that of μ(K)ΠμK\mu(K)\Pi^{\circ}_{\mu}K as p1+.p\to-1^{+}.

3.5. Set Inclusions for Weighted Radial Mean Bodies

In this subsection and the next, we obtain the reverse of Theorem 3.8 via Berwald’s inequality. We will need the following facts about concave functions.

Lemma 3.10.

Let ff be a concave function that is supported on a convex body L𝒦0nL\in\mathcal{K}^{n}_{0} such that

h(θ):=\difff(rθ)r|r=0+<0for all θ𝕊n1.h(\theta):=\diff{f(r\theta)}{r}\bigg{|}_{r=0^{+}}<0\quad\text{for all }\theta\in\mathbb{S}^{n-1}.

Define z(θ)=(h(θ))1f(0),z(\theta)=-\left(h(\theta)\right)^{-1}f(0), then

(34) <f(rθ)f(0)[1(z(θ))1r]-\infty<f(r\theta)\leq f(0)\left[1-(z(\theta))^{-1}r\right]

whenever θ𝕊n1\theta\in\mathbb{S}^{n-1} and r[0,ρL(θ)]r\in[0,\rho_{L}(\theta)]. In particular, if ff is non-negative, then we have

0f(rθ)f(0)[1(z(θ))1r]and ρL(θ)z(θ).0\leq f(r\theta)\leq f(0)\left[1-(z(\theta))^{-1}r\right]\quad\mbox{and }\rho_{L}(\theta)\leq z(\theta).

One has f(rθ)=f(0)[1(z(θ))1r]f(r\theta)=f(0)\left[1-(z(\theta))^{-1}r\right] for r[0,ρL(θ)]r\in[0,\rho_{L}(\theta)] if, and only if, ρL(θ)=z(θ).\rho_{L}(\theta)=z(\theta).

Proof.

From the concavity of the function ff, one has

f(rθ)f(0)[1+h(θ)f(0)r].f(r\theta)\leq f(0)\left[1+\frac{h(\theta)}{f(0)}r\right].

Then, the inequality (34) follows from the definition of z(θ)z(\theta). If ff is additionally non-negative, then one obtains that ρL(θ)z(θ)\rho_{L}(\theta)\leq z(\theta) by using that 01(z(θ))1r0\leq 1-(z(\theta))^{-1}r for a fixed θ𝕊n1\theta\in\mathbb{S}^{n-1} and r[0,ρL(θ)],r\in[0,\rho_{L}(\theta)], and then setting r=ρL(θ).r=\rho_{L}(\theta). For the equality conditions, suppose f(rθ)=f(0)[1(z(θ))1r];f(r\theta)=f(0)\left[1-(z(\theta))^{-1}r\right]; then, by definition of LL being the support of f,f, one obtains z(θ)=ρL(θ).z(\theta)=\rho_{L}(\theta). Conversely, suppose ρL(θ)=z(θ).\rho_{L}(\theta)=z(\theta). Then, from the concavity of ff, one has f(0)[1(ρL(θ))1r]f(rθ)f(0)[1(ρL(θ))1r],f(0)\left[1-(\rho_{L}(\theta))^{-1}r\right]\leq f(r\theta)\leq f(0)\left[1-(\rho_{L}(\theta))^{-1}r\right], and so there is equality. ∎

Using Proposition 3.7, Lemma 3.10 and (25), we obtain for a Borel measure μ\mu with density such that μ\mu is FF-concave, F:++F:\mathbb{R}^{+}\to\mathbb{R}^{+} is an increasing and differentiable function, that

(35) DKF(μ(K))F(μ(K))ΠμKDK\subseteq\frac{F(\mu(K))}{F^{\prime}(\mu(K))}\Pi_{\mu}^{\circ}K

for every K𝒦0nK\in\mathcal{K}^{n}_{0} such that K\partial K is in the Lebesgue set of the density of μ\mu.

Theorem 3.11.

Fix a convex body KK in n\mathbb{R}^{n}. Let μ\mu be a finite Borel measure containing KK in its support, such that μ\mu is FF-concave on convex subsets of KK, where F:[0,μ(K))[0,)F:[0,\mu(K))\to[0,\infty) is a continuous, increasing, and invertible function. Then, for 1<pq<-1<p\leq q<\infty, one has

DKC(q,μ,K)Rq,μKC(p,μ,K)Rp,μKF(μ(K))F(μ(K))ΠμK,DK\subseteq C(q,\mu,K)R_{q,\mu}K\subseteq C(p,\mu,K)R_{p,\mu}K\subseteq\frac{F(\mu(K))}{F^{\prime}(\mu(K))}\Pi_{\mu}^{\circ}K,

where C(p,μ,K)=C(p,\mu,K)=

{(pμ(K)01F1[F(μ(K))(1t)]tp1𝑑t)1pfor p>0(pμ(K)01tp1(F1[F(μ(K))(1t)]μ(K))𝑑t+1)1pfor p(1,0),\begin{cases}\left(\frac{p}{\mu(K)}\int_{0}^{1}F^{-1}\left[F(\mu(K))(1-t)\right]t^{p-1}dt\right)^{-\frac{1}{p}}&\text{for }p>0\\ \left(\frac{p}{\mu(K)}\int_{0}^{1}t^{p-1}(F^{-1}\left[F(\mu(K))(1-t)\right]-\mu(K))dt+1\right)^{-\frac{1}{p}}&\text{for }p\in(-1,0),\end{cases}

and, for the last set inclusion, we additionally assume that μ\mu has a locally integrable density containing K\partial K in its Lebesgue set and that F(x)F(x) is differentiable at the value x=μ(K).x=\mu(K). The equality conditions are the following:

  1. (1)

    For the first two set inclusions there is equality of sets if, and only if, F(0)=0F(0)=0 and Fgμ,K(x)=F(μ(K))DK(x).F\circ g_{\mu,K}(x)=F(\mu(K))\ell_{DK}(x).

  2. (2)

    For the last set inclusion, the sets are equal if, and only if, Fgμ,K(x)=F(μ(K))C(x),C=F(μ(K))F(μ(K))ΠμK.F\circ g_{\mu,K}(x)=F(\mu(K))\ell_{C}(x),\;C=\frac{F(\mu(K))}{F^{\prime}(\mu(K))}\Pi^{\circ}_{\mu}K.

Proof.

Observe that

C(p,μ,K)ρRp,μK(θ)=Ggμ,K(rθ)(p)C(p,\mu,K)\rho_{R_{p,\mu}K}(\theta)=G_{g_{\mu,K}(r\theta)}(p)

from (13). Thus, from Lemma 2.1, this function is non-increasing in pp, which establishes the first three set inclusions. For the last set inclusion, we have not yet established the behaviour of limp1C(p,μ,K)ρRp,μK(p).\lim_{p\to-1}C(p,\mu,K)\rho_{R_{p,\mu}K}(p). We do so now.

First, begin by writing

Ggμ,K(rθ)(p)=C(p,μ,K)(p+1)1/p(p+1)1/pρRp,μK(θ).G_{g_{\mu,K}(r\theta)}(p)=\frac{C(p,\mu,K)}{(p+1)^{1/p}}(p+1)^{1/p}\rho_{R_{p,\mu}K}(\theta).

Therefore, from (33), it suffices to show that, as p1,p\to-1,

C(p,μ,K)(p+1)1/pF(μ(K))F(μ(K))μ(K).\frac{C(p,\mu,K)}{(p+1)^{1/p}}\to\frac{F(\mu(K))}{F^{\prime}(\mu(K))\mu(K)}.

Indeed, from integration by parts we can write, for all p(1,0)(0,),p\in(-1,0)\cup(0,\infty), that

C(p,μ,K)=(F(μ(K))μ(K))1p(01[F(F1[F(μ(K))(1t)])]1tp𝑑t)1p.C(p,\mu,K)=\left(\frac{F(\mu(K))}{\mu(K)}\right)^{-\frac{1}{p}}\left(\int_{0}^{1}\left[F^{\prime}\left(F^{-1}[F(\mu(K))(1-t)]\right)\right]^{-1}t^{p}dt\right)^{-\frac{1}{p}}.

Therefore, the result follows from Lemma 3.9. ∎

We now obtain a result for ss-concave measures, s>0,s>0, the promised generalization of (30).

Corollary 3.12.

Fix a convex body KK in n\mathbb{R}^{n}. Let μ\mu be a Borel measure containing KK in its support that is ss-concave, s>0,s>0, on convex subsets of KK. Then, for 1<pq<-1<p\leq q<\infty, one has

DK(1s+qq)1qRq,μK(1s+pp)1pRp,μK1sμ(K)ΠμK,DK\subseteq{{\frac{1}{s}+q}\choose q}^{\frac{1}{q}}R_{q,\mu}K\subseteq{{\frac{1}{s}+p}\choose p}^{\frac{1}{p}}R_{p,\mu}K\subseteq\frac{1}{s}\mu(K)\Pi_{\mu}^{\circ}K,

where the last inclusion holds if μ\mu has locally integrable density φ(x)\varphi(x) containing K\partial K in its Lebesgue set.

There is equality in any set inclusion if, and only if, gμ,Ks(x)=μ(K)sDK(x)g_{\mu,K}^{s}(x)=\mu(K)^{s}\ell_{DK}(x). If μ\mu is a ss-concave Radon measure, then s(0,1/n]s\in(0,1/n] and equality occurs if, and only if, KK is nn-dimensional simplex, the density φ\varphi of μ\mu is constant on KK, and s=1/ns=1/n.

Proof.

Setting F(x)=xsF(x)=x^{s} in Theorem 3.11 yields, in the case when p>0,p>0,

C(p,μ,K)=(p01(1u)1/sup1𝑑u)1p=(pΓ(1s+1)Γ(p)Γ(1s+p+1))1p,C(p,\mu,K)=\left(p\int_{0}^{1}(1-u)^{1/s}u^{p-1}du\right)^{-\frac{1}{p}}=\left(\frac{p\Gamma(\frac{1}{s}+1)\Gamma(p)}{\Gamma(\frac{1}{s}+p+1)}\right)^{-\frac{1}{p}},

and similarly for p(1,0).p\in(-1,0). The equality conditions from Theorem 3.11 yields that gμ,Ks(x)g_{\mu,K}^{s}(x) is an affine function along rays for xDKx\in DK. If μ\mu is a ss-concave Radon measure, then one must have s(0,1/n]s\in(0,1/n]. For such ss-concave measures, gμ,Ks(x)g_{\mu,K}^{s}(x) being an affine function along rays is equivalent to the stated equality conditions via Proposition 3.14 below. ∎

We first remark that the following are equivalent:

  1. (i).

    KK is a simplex.

  2. (ii).

    For any xnx\in\mathbb{R}^{n}, either K(K+x)K\cap(K+x) is empty or it is homothetic to KK.

The equivalence between (i)(i) and (ii)(ii) can be found in [13, Section 6], or [11, 42]. Next, we recall a result of Milman and Rotem [38, Corollary 2.16]:

Lemma 3.13.

Let μ\mu be a ss-concave Radon measure on n\mathbb{R}^{n} with density ϕ\phi, A,BnA,B\subset\mathbb{R}^{n} Borel sets of positive measure, and λ(0,1)\lambda\in(0,1), and suppose that

μ(λA+(1λ)B)s=λμ(A)s+(1λ)μ(B)s.\mu(\lambda A+(1-\lambda)B)^{s}=\lambda\mu(A)^{s}+(1-\lambda)\mu(B)^{s}.

Then up to μ\mu-null sets, there exist c,m>0c,m>0, bnb\in\mathbb{R}^{n} such that B=mA+bB=mA+b and such that ϕ(mx+b)=cϕ(x)\phi(mx+b)=c\cdot\phi(x) for all xAx\in A.

Proposition 3.14.

Let K𝒦nK\in\mathcal{K}^{n}, s(0,1/n]s\in(0,1/n], and μ\mu an ss-concave Radon measure, whose density φ\varphi contains KK in its support. Then, gμ,K(rθ)sg_{\mu,K}(r\theta)^{s} is an affine function in rr for for every θ𝕊n1\theta\in\mathbb{S}^{n-1} and r[0,ρDK(θ)]r\in[0,\rho_{DK}(\theta)] if, and only if, KK is nn-dimensional simplex, φ\varphi is a constant on KK, and s=1/ns=1/n.

Proof.

Let xx lie in the interior of supp(gμ,K())\operatorname{supp}(g_{\mu,K}(\cdot)), and let t,λ(0,1)t,\lambda\in(0,1). The fact that gμ,K()g_{\mu,K}(\cdot) is affine on the segment [0,x][0,x] precisely means that for λ(0,1)\lambda\in(0,1),

(36) μ(Kλ(0,x))s=λμ(K)s+(1λ)μ(K1(0,x))s,\mu(K^{\lambda}(0,x))^{s}=\lambda\mu(K)^{s}+(1-\lambda)\mu(K^{1}(0,x))^{s},

where Kλ(0,y)=K(K+λy)K^{\lambda}(0,y)=K\cap(K+\lambda y). Examining the proof of the Proposition 3.7, we see that Kλ(0,x)(1λ)K+λK1(0,x)K^{\lambda}(0,x)\subseteq(1-\lambda)K+\lambda K^{1}(0,x) and equality can hold in (36) only if Kλ(0,λx)=(1λ)K+λK1(0,x)K^{\lambda}(0,\lambda x)=(1-\lambda)K+\lambda K^{1}(0,x). In particular, we have

(37) μ((1λ)K+λK1(0,x))s=λμ(K)s+(1λ)μ(K1(0,x))s,\mu((1-\lambda)K+\lambda K^{1}(0,x))^{s}=\lambda\mu(K)^{s}+(1-\lambda)\mu(K^{1}(0,x))^{s},

By Lemma 3.13, that K(Kx)K\cap(K\cap x) is homothetic to KK for all xx in the interior of DKDK, which implies that KK is a nn-dimensional simplex.

It remains to show that the density of μ\mu is constant on KK. For this we use the second conclusion of Lemma 3.13: for each xint(DK)x\in\mathrm{int}(DK) there exists c(x)>0c(x)>0 such that for each yKy\in K, φ(Axy)=c(x)φ(y)\varphi(A_{x}y)=c(x)\varphi(y), where AxA_{x} is the affine transformation which maps KK onto K(K+x)K\cap(K+x). But note that AxA_{x} is a continuous map from the compact, convex set KK to itself, so it has a fixed point yy by Brouwer’s fixed point theorem. For such yy, we have φ(y)=φ(Axy)=c(x)φ(y)\varphi(y)=\varphi(A_{x}y)=c(x)\varphi(y), implying c(x)=1c(x)=1. (Note that as a convex function is continuous on the interior of its domain [43, Theorem 1.5.3], the density of μ\mu is continuous on int(suppμ)\mathrm{int}(\operatorname{supp}\mu); in particular, φ\varphi is well-defined pointwise on the interior of KK, not just up to sets of measure zero.)

Once one knows that φ(Axy)=φ(y)\varphi(A_{x}y)=\varphi(y) for any homothety AxA_{x} mapping KK to K(K+x)K\cap(K+x) and any yKy\in K, one verifies by tedious but elementary arguments (e.g., by starting with the faces of KK, working by induction on the dimension) that any two points in KK can be mapped into each other by a chain of such AxA_{x}’s, which implies that φ\varphi is indeed constant on KK. Thus we are back to the case of Lebesgue measure, which we know is 1/n1/n-affine on pairs of homothetic bodies. Conversely, one verifies that if KK is a nn-dimensional simplex and ϕ\phi is constant on KK then gμ,K()1/ng_{\mu,K}(\cdot)^{1/n} is affine on radial segments, as in the classical Zhang inequality. ∎

Most of the inclusions in Theorem 3.11 continue to hold when the concavity of the measures behaves logarithmically. Unfortunately, in this instance, C(p,μ,K)C(p,\mu,K) may tend to 0 as p,p\to\infty, and so C(p,μ,K)Rp,μKC(p,\mu,K)R_{p,\mu}K will tend to the origin. Hence, we lose the first set inclusion:

Theorem 3.15 (Logarithmic Case).

Suppose a Borel measure μ\mu on n\mathbb{R}^{n} is finite on some convex body KK and QQ-concave, where Q:(0,μ(K)](,)Q:(0,\mu(K)]\to(-\infty,\infty) is an increasing and invertible function. Then, for 1<pq<-1<p\leq q<\infty, one has

C(q,μ,K)Rq,μKC(p,μ,K)Rp,μK1Q(μ(K))ΠμK,C(q,\mu,K)R_{q,\mu}K\subset C(p,\mu,K)R_{p,\mu}K\subset\frac{1}{Q^{\prime}(\mu(K))}{\Pi_{\mu}^{\circ}K},

where C(p,μ,K)=C(p,\mu,K)=

{(pμ(K)0Q1[Q(μ(K))t]tp1𝑑t)1pfor p>0(pμ(K)0tp1(Q1[Q(μ(K)t)]μ(K))𝑑t)1pfor p(1,0),\begin{cases}\left(\frac{p}{\mu(K)}\int_{0}^{\infty}Q^{-1}\left[Q(\mu(K))-t\right]t^{p-1}dt\right)^{-\frac{1}{p}}&\text{for }p>0\\ \left(\frac{p}{\mu(K)}\int_{0}^{\infty}t^{p-1}(Q^{-1}\left[Q(\mu(K)-t)\right]-\mu(K))dt\right)^{-\frac{1}{p}}&\text{for }p\in(-1,0),\end{cases}

and, for the second set inclusion, we additionally assume that μ\mu has locally integrable density containing K\partial K in its Lebesgue set and that Q(x)Q(x) is differentiable at the value x=μ(K)x=\mu(K). In particular, if μ\mu is log-concave:

1Γ(1+q)1qRq,μK1Γ(1+p)1pRp,μKμ(K)ΠμK,\frac{1}{\Gamma\left(1+q\right)^{\frac{1}{q}}}R_{q,\mu}K\subset\frac{1}{\Gamma\left(1+p\right)^{\frac{1}{p}}}R_{p,\mu}K\subset\mu(K){\Pi_{\mu}^{\circ}K},

where limp01Γ(1+p)1pRp,μK\lim_{p\to 0}\frac{1}{\Gamma\left(1+p\right)^{\frac{1}{p}}}R_{p,\mu}K is interpreted via continuity.

Proof.

The first inclusion follows from the second case of Theorem 1.1. For the second inclusion, suppose p>0.p>0. Then, one has

0gμ,K(rθ)Q1[Q(μ(K))(1Q(μ(K))Q(μ(K))rρΠμK(θ))].0\leq g_{\mu,K}(r\theta)\leq Q^{-1}\left[Q(\mu(K))\left(1-\frac{Q^{\prime}(\mu(K))}{Q(\mu(K))}\frac{r}{\rho_{\Pi_{\mu}^{\circ}K}(\theta)}\right)\right].

Since Q(μ(K))Q(\mu(K)) may possibly be negative, we shall leave Q(μ(K))Q(\mu(K)) inside the integral:

ρRp,μKp(θ)=pμ(K)0ρDK(θ)gμ,K(rθ)rp1𝑑r\displaystyle\rho^{p}_{R_{p,\mu}K}(\theta)=\frac{p}{\mu(K)}\int_{0}^{\rho_{DK}(\theta)}g_{\mu,K}(r\theta)r^{p-1}dr
pμ(K)0ρDK(θ)Q1[Q(μ(K))(1Q(μ(K))Q(μ(K))rρΠμK(θ))]rp1𝑑r.\displaystyle\leq\frac{p}{\mu(K)}\int_{0}^{\rho_{DK}(\theta)}Q^{-1}\left[Q(\mu(K))\left(1-\frac{Q^{\prime}(\mu(K))}{Q(\mu(K))}\frac{r}{\rho_{\Pi_{\mu}^{\circ}K}(\theta)}\right)\right]r^{p-1}dr.
=(ρΠμK(θ)Q(μ(K)))ppμ(K)\displaystyle=\left(\frac{\rho_{\Pi_{\mu}^{\circ}K}(\theta)}{Q^{\prime}(\mu(K))}\right)^{p}\frac{p}{\mu(K)}
×0Q(μ(K))ρDK(θ)ρΠμK(θ)Q1[Q(μ(K))u]up1du.\displaystyle\quad\quad\quad\quad\quad\times\int_{0}^{Q^{\prime}(\mu(K))\frac{\rho_{DK}(\theta)}{\rho_{\Pi_{\mu}^{\circ}K(\theta)}}}Q^{-1}\left[Q(\mu(K))-u\right]u^{p-1}du.

and so C(p,μ,K)ρRp,μK(θ)<1Q(μ(K))ρΠμK(θ),C(p,\mu,K)\rho_{R_{p,\mu}K}(\theta)<\frac{1}{Q^{\prime}(\mu(K))}\rho_{\Pi_{\mu}^{\circ}K}(\theta), which yields the result. The case for p(1,0)p\in(-1,0) is similar. ∎

We list the Gaussian measure case as a corollary.

Corollary 3.16.

Let KK be a convex body. Then, for 1<pq<-1<p\leq q<\infty, one has

1Γ(1+q)1qRq,γnK1Γ(1+p)1pRp,γnKγn(K)ΠγnK,\frac{1}{\Gamma\left(1+q\right)^{\frac{1}{q}}}R_{q,\gamma_{n}}K\subset\frac{1}{\Gamma\left(1+p\right)^{\frac{1}{p}}}R_{p,\gamma_{n}}K\subset\gamma_{n}(K){\Pi_{\gamma_{n}}^{\circ}K},

where limp01Γ(1+p)1pRp,γnK\lim_{p\to 0}\frac{1}{\Gamma\left(1+p\right)^{\frac{1}{p}}}R_{p,\gamma_{n}}K is interpreted via continuity, and

C(q,γn,K)Rq,γnKC(p,γn,K)Rp,μK2πeΦ1(γn(K))22ΠγnK,C(q,\gamma_{n},K)R_{q,\gamma_{n}}K\subset C(p,\gamma_{n},K)R_{p,\mu}K\subset\sqrt{\frac{2}{\pi}}e^{-\frac{\Phi^{-1}(\gamma_{n}(K))^{2}}{2}}{\Pi_{\gamma_{n}}^{\circ}K},

where C(p,γn,K)=C(p,\gamma_{n},K)=

{(pγn(K)0Φ[Φ1(γn(K))t]tp1𝑑t)1pfor p>0(pγn(K)0tp1(Φ[Φ1(γn(K)t)]γn(K))𝑑t)1pfor p(1,0).\begin{cases}\left(\frac{p}{\gamma_{n}(K)}\int_{0}^{\infty}\Phi\left[\Phi^{-1}(\gamma_{n}(K))-t\right]t^{p-1}dt\right)^{-\frac{1}{p}}&\text{for }p>0\\ \left(\frac{p}{\gamma_{n}(K)}\int_{0}^{\infty}t^{p-1}(\Phi\left[\Phi^{-1}(\gamma_{n}(K)-t)\right]-\gamma_{n}(K))dt\right)^{-\frac{1}{p}}&\text{for }p\in(-1,0).\end{cases}

3.6. Inequalities for Weighted Radial Mean Bodies

We next show an application of Corollary 3.12. In particular, if the set inclusions are applied to a measure ν\nu with homogeneity α\alpha, then there exists a radial mean body whose ν\nu measure is “of the same order” as that of KK itself. First, define the ν\nu-translated-average of KK with respect to μ\mu as

(38) νμ(K)=1μ(K)Kν(yK)𝑑μ(y)=1μ(K)DKgμ,K(x)𝑑ν(x).\nu_{\mu}(K)=\frac{1}{\mu(K)}\int_{K}\nu(y-K)d\mu(y)=\frac{1}{\mu(K)}\int_{DK}g_{\mu,K}(x)d\nu(x).

The last equality follows from Fubini’s theorem, and this definition has appeared in [30]. Next, we see that when ν\nu is homogeneous of degree α\alpha, we obtain a relation between ν(Rα,μK)\nu(R_{\alpha,\mu}K) and νμ(K).\nu_{\mu}(K).

Lemma 3.17.

Fix a convex body KK and a Borel measure ν\nu that is α\alpha-homogeneous with density and a Borel measure μ\mu on n\mathbb{R}^{n}. Then, one has ν(Rα,μK)=νμ(K)\nu(R_{\alpha,\mu}K)=\nu_{\mu}(K).

Proof.

Let φ\varphi be the density of ν\nu. Using (2) and Fubini’s theorem, we obtain:

ν(Rα,μK)\displaystyle\nu(R_{\alpha,\mu}K) =1α𝕊n1ρRα,μKα(θ)φ(θ)𝑑θ=1α1μ(K)𝕊n1KρK(x,θ)α𝑑μ(x)φ(θ)𝑑θ\displaystyle=\frac{1}{\alpha}\int_{\mathbb{S}^{n-1}}\rho^{\alpha}_{R_{\alpha,\mu}K}(\theta)\varphi(\theta)d\theta=\frac{1}{\alpha}\frac{1}{\mu(K)}\int_{\mathbb{S}^{n-1}}\int_{K}\rho_{K}(x,\theta)^{\alpha}d\mu(x)\varphi(\theta)d\theta
=1α1μ(K)K𝕊n1ρK(x,θ)αφ(θ)𝑑θ𝑑μ(x)\displaystyle=\frac{1}{\alpha}\frac{1}{\mu(K)}\int_{K}\int_{\mathbb{S}^{n-1}}\rho_{K}(x,\theta)^{\alpha}\varphi(\theta)d\theta d\mu(x)
=1α1μ(K)K𝕊n1ρxK(θ)αφ(θ)𝑑θ𝑑μ(x),\displaystyle=\frac{1}{\alpha}\frac{1}{\mu(K)}\int_{K}\int_{\mathbb{S}^{n-1}}\rho_{x-K}(\theta)^{\alpha}\varphi(\theta)d\theta d\mu(x),

where the last equality follows from the fact that ρK(x,θ)=ρKx(θ)=ρxK(θ).\rho_{K}(x,\theta)=\rho_{K-x}(-\theta)=\rho_{x-K}(\theta). Using (2) again yields the result. ∎

Theorem 3.18 (Rogers-Shephard type inequality for an α\alpha-homogeneous and a ss-concave measure).

Fix a convex body KK. Consider a Borel measure ν\nu that is α\alpha-homogeneous and a Borel measure μ\mu on n\mathbb{R}^{n} that is ss-concave, s>0s>0, on convex subsets of KK, whose locally integrable density contains K\partial K in its Lebesgue set and KK in its support. Then,

ν(DK)(1s+αα)min{νμ(K),νμ(K)},\nu(DK)\leq{{\frac{1}{s}+\alpha}\choose\alpha}\min\{\nu_{\mu}(K),\nu_{\mu}(-K)\},

with equality if, and only if, gμ,Ks(x)=μ(K)sDK(x)g_{\mu,K}^{s}(x)=\mu(K)^{s}\ell_{DK}(x). If μ\mu is a ss-concave Radon measure, then s(0,1/n]s\in(0,1/n] and equality occurs if, and only if, KK is nn-dimensional simplex, the density φ\varphi of μ\mu is constant on KK, and s=1/ns=1/n.

Proof.

From Corollary 3.12 with p=αp=\alpha one obtains

ν(DK)ν((1s+αα)1αRμ,αK)=(1s+αα)ν(Rμ,αK).\nu(DK)\leq\nu\left({{\frac{1}{s}+\alpha}\choose\alpha}^{\frac{1}{\alpha}}R_{\mu,\alpha}K\right)={{\frac{1}{s}+\alpha}\choose\alpha}\nu(R_{\mu,\alpha}K).

Using Lemma 3.17 and that DK=D(K)DK=D(-K) completes the proof. ∎

An upper bound for μ(DK)/μ(K)\mu(DK)/\mu(K) when μ\mu is ss-concave was first shown by Borell, [7]. However, the bound was not sharp.

Corollary 3.19 (Zhang’s Inequality for an α\alpha-homogeneous and a ss-concave measure).

Fix a convex body KK. Consider a Borel measure ν\nu that is α\alpha-homogeneous and a Borel measure μ\mu on n\mathbb{R}^{n} that is ss-concave, s>0s>0, on convex subsets of KK, whose locally integrable density contains K\partial K in its Lebesgue set and KK in its support. Then, one has

sα(1s+αα)μ(K)ανμ(K)ν(ΠμK),s^{\alpha}{{\frac{1}{s}+\alpha}\choose\alpha}\leq\frac{\mu(K)^{\alpha}}{\nu_{\mu}(K)}\nu\left(\Pi_{\mu}^{\circ}K\right),

with equality if, and only if, gμ,Ks(x)=μ(K)sΠμK(x)g_{\mu,K}^{s}(x)=\mu(K)^{s}\ell_{\Pi_{\mu}^{\circ}K}(x). If μ\mu is a ss-concave Radon measure, then s(0,1/n]s\in(0,1/n] and equality occurs if, and only if, KK is a nn-dimensional simplex, the density φ\varphi of μ\mu is constant on KK, and s=1/ns=1/n.

Proof.

From Lemma 3.17 and Corollary 3.12 with p=α,p=\alpha, one obtains

(1s+αα)νμ(K)\displaystyle{{\frac{1}{s}+\alpha}\choose\alpha}\nu_{\mu}(K) =(1s+αα)ν(Rμ,αK)=ν((1s+αα)1αRμ,αK)\displaystyle={{\frac{1}{s}+\alpha}\choose\alpha}\nu(R_{\mu,\alpha}K)=\nu\left({{\frac{1}{s}+\alpha}\choose\alpha}^{\frac{1}{\alpha}}R_{\mu,\alpha}K\right)
ν(1sμ(K)ΠμK).\displaystyle\leq\nu\left(\frac{1}{s}\mu(K)\Pi_{\mu}^{\circ}K\right).

3.7. The Gardner-Zvavitch Inequality and Radial Mean Bodies

We would like to apply the 1/n1/n-concavity of the Gaussian measure over symmetric convex bodies in (6), which is also true for every μn\mu\in\mathcal{M}_{n}, to obtain that Corollary 3.12 holds for such measures. We run into an issue: even if KK is symmetric, then K(K+x)K\cap(K+x) is not symmetric in general. Therefore, Proposition 3.7 does not apply, i.e. μ\mu being 1/n1/n concave does not imply that gμ,K1ng_{\mu,K}^{\frac{1}{n}} is concave. To remedy this, we take a cue from [30] and define the polarized covariogram as

rμ,K(x)=μ((K+x2)(Kx2)).r_{\mu,K}(x)=\mu\left((K+\frac{x}{2})\cap(K-\frac{x}{2})\right).

As can be found in [30], the set (K+x2)(Kx2)(K+\frac{x}{2})\cap(K-\frac{x}{2}) is a symmetric convex body when KK is, and rμ,Kr_{\mu,K} inherits any concavity of μ\mu over symmetric convex bodies.

Under the assumption that μ\mu is a Borel measure with even density and K𝒦nK\in\mathcal{K}^{n} is symmetric, notice that ΠμK=Π~μK\Pi_{\mu}K=\widetilde{\Pi}_{\mu}K. One also has, for every θ𝕊n1\theta\in\mathbb{S}^{n-1},

(39) \diffrμ,K(rθ)r|r=0+=hΠμK(θ).\diff{r_{\mu,K}(r\theta)}{r}\bigg{|}_{r=0^{+}}=-h_{\Pi_{\mu}K}(\theta).

This was first shown in [30] under the additional assumption that the density of μ\mu is Lipschitz; arguing similarly to Theorem 3.6 allows one to weaken the assumption to merely the density of μ\mu contains K\partial K in its Lebesgue set. In fact, one does not need the symmetry assumptions on μ\mu and KK for this proof; one will obtain in general

(40) \diffrμ,K(rθ)r|r=0+=hΠ~μK(θ),\diff{r_{\mu,K}(r\theta)}{r}\bigg{|}_{r=0^{+}}=-h_{\widetilde{\Pi}_{\mu}K}(\theta),

but Π~μK\widetilde{\Pi}_{\mu}K is not necessarily ΠμK\Pi_{\mu}K.

In order to obtain Corollary 3.12 for μn\mu\in\mathcal{M}_{n}, we define the polarized weighted mean bodies, Pp,μKP_{p,\mu}K, as the star bodies on n\mathbb{R}^{n} whose radial function is given by, for p(1,)p\in(-1,\infty) and θ𝕊n1\theta\in\mathbb{S}^{n-1},

(41) ρPp,μK(θ)p=(pμ(K))rμ,K(rθ)(p)=0ρDK(θ)(rμ,K(rθ)μ(K))rp𝑑r.\rho_{P_{p,\mu}K}(\theta)^{p}=\left(\frac{p}{\mu(K)}\right)\mathcal{M}_{r_{\mu,K}(r\theta)}(p)=\int_{0}^{\rho_{DK}(\theta)}\left(\frac{-r_{\mu,K}(r\theta)^{\prime}}{\mu(K)}\right)r^{p}dr.

The bodies Pp,μKP_{p,\mu}K are symmetric convex bodies for p0p\geq 0. Once again, p=0,p=0,\infty are interpreted via continuity, and R,μK=DKR_{\infty,\mu}K=DK for KK contained in the support of μ\mu. We again have that

(42) limp1(p+1)1/pρPp,μK(θ)=μ(K)ρΠ~μK(θ).\lim_{p\to-1}(p+1)^{1/p}\rho_{P_{p,\mu}K}(\theta)=\mu(K)\rho_{\widetilde{\Pi}^{\circ}_{\mu}K}(\theta).

Consequently, the proof of the following theorem is verbatim the same as Theorem 3.11 and Corollary 3.12.

Theorem 3.20.

Fix a symmetric convex body KK in n\mathbb{R}^{n}. Let μn\mu\in\mathcal{M}_{n} be a Borel measure containing KK in its support. Then, for 1<pq<-1<p\leq q<\infty, one has

DK(n+qq)1qPq,μK(n+pp)1pPp,μKnμ(K)ΠμK.DK\subseteq{{n+q}\choose q}^{\frac{1}{q}}P_{q,\mu}K\subseteq{{n+p}\choose p}^{\frac{1}{p}}P_{p,\mu}K\subseteq n\mu(K)\Pi_{\mu}^{\circ}K.

There is equality in any set inclusion if, and only if, rμ,K(x)=μ(K)DK(x)nr_{\mu,K}(x)=\mu(K)\ell_{DK}(x)^{n}.

Recently, it was shown by Livshyts [34] that for any even, log-concave probability measure μ\mu on n\mathbb{R}^{n}, μ\mu is ss-concave over the class of symmetric convex bodies, with s=n4sns=n^{-4-s_{n}}. One can then use this result to formulate Theorem 3.20 for such measures, with nn in the coefficients replaced by n4+snn^{4+s_{n}}. Here, {sn}\{s_{n}\} is a bounded sequence that goes to 0 as nn\to\infty.

It is also manifest that Theorem 3.11 holds with the weighted radial mean bodies replaced by the polarized weighted mean bodies, the additional assumption that KK is symmetric (since μ\mu being FF-concave on convex subsets of KK implies it is FF-concave on symmetric convex subsets of KK, and the fact that rμ,Kr_{\mu,K} will then inherit the concavity) and ΠμK\Pi^{\circ}_{\mu}K replaced by Π~μK\widetilde{\Pi}^{\circ}_{\mu}K (due to (40) and (42)). We avoid the unnecessary formal statement of this slightly different theorem.

References

  • [1] David Alonso-Gutiérrez et al. “Rogers-Shephard and local Loomis-Whitney type inequalities” In Math. Ann. 374.3-4, 2019, pp. 1719–1771 DOI: 10.1007/s00208-019-01834-3
  • [2] Keith Ball “Logarithmically concave functions and sections of convex sets in 𝐑n{\bf R}^{n} In Studia Math. 88.1, 1988, pp. 69–84 DOI: 10.4064/sm-88-1-69-84
  • [3] Keith Ball “The reverse isoperimetric problem for Gaussian measure” In Discrete Comput. Geom. 10.4, 1993, pp. 411–420 DOI: 10.1007/BF02573986
  • [4] L. Berwald “Verallgemeinerung eines Mittelwertsatzes von J. Favard für positive konkave Funktionen” In Acta Math. 79, 1947, pp. 17–37 DOI: 10.1007/BF02404692
  • [5] Sergey Bobkov and Mokshay Madiman “Reverse Brunn-Minkowski and reverse entropy power inequalities for convex measures” In J. Funct. Anal. 262.7, 2012, pp. 3309–3339 DOI: 10.1016/j.jfa.2012.01.011
  • [6] Sergey G. Bobkov et al. “When can one invert Hölder’s inequality? (and why one may want to)”, To appear
  • [7] C. Borell “Convex set functions in dd-space” In Period. Math. Hungar. 6.2, 1975, pp. 111–136 DOI: 10.1007/BF02018814
  • [8] Christer Borell “Complements of Lyapunov’s inequality” In Math. Ann. 205, 1973, pp. 323–331 DOI: 10.1007/BF01362702
  • [9] Christer Borell “Integral inequalities for generalized concave or convex functions” In J. Math. Anal. Appl. 43, 1973, pp. 419–440 DOI: 10.1016/0022-247X(73)90083-8
  • [10] Christer Borell “The Ehrhard inequality” In C. R. Math. Acad. Sci. Paris 337.10, 2003, pp. 663–666 DOI: 10.1016/j.crma.2003.09.031
  • [11] Gustave Choquet “Existence des représentations intégrales dans les cônes convexes” In C. R. Acad. Sci. Paris 243, 1956, pp. 736–737
  • [12] Dario Cordero-Erausquin and Liran Rotem “Improved log-concavity for rotationally invariant measures of symmetric convex sets” In Annals of Probability, 2023
  • [13] H.. Eggleston, Branko Grünbaum and Victor Klee “Some semicontinuity theorems for convex polytopes and cell-complexes” In Comment. Math. Helv. 39, 1964, pp. 165–188 DOI: 10.1007/BF02566949
  • [14] Antoine Ehrhard “Élements extrémaux pours les inégalités de Brunn-Minkowski gaussienes” In Annals of the Institute Henri Poincaré Probability and Statistics 22, 1986, pp. 149–168 URL: www.numdam.org/item/AIHPB_1986__22_2_149_0/
  • [15] Antoine Ehrhard “Symétrisation dans l’espace de Gauss” In Mathematica Scandinavica 53, 1983, pp. 281–301 URL: http://eudml.org/doc/166873
  • [16] Alexandros Eskenazis and Georgios Moschidis “The dimensional Brunn-Minkowski inequality in Gauss space” In J. Funct. Anal. 280.6, 2021, pp. Paper No. 108914\bibrangessep19 DOI: 10.1016/j.jfa.2020.108914
  • [17] M. Fradelizi, D. Langharst, M. Madiman and A. Zvavitch In J. Math. Anal. Appl. 529, 2024, pp. 127519 DOI: https://doi.org/10.1016/j.jmaa.2023.127519
  • [18] M. Fradelizi, D. Langharst, M. Madiman and A. Zvavitch “Weighted Brunn-Minkowski Theory II: On Inequalities for Mixed Measures” In Preprint, 2023
  • [19] Matthieu Fradelizi, Olivier Guédon and Alain Pajor “Thin-shell concentration for convex measures” In Studia Math. 223.2, 2014, pp. 123–148 DOI: 10.4064/sm223-2-2
  • [20] Matthieu Fradelizi, Jiange Li and Mokshay Madiman “Concentration of information content for convex measures” In Electron. J. Probab. 25, 2020, pp. Paper No. 20\bibrangessep22 DOI: 10.1214/20-ejp416
  • [21] Richard J. Gardner “Geometric Tomography” 58, Encyclopedia of Mathematics and its Applications Cambridge, UK: Cambridge University Press, 2006
  • [22] Richard J. Gardner and Gaoyong Zhang “Affine inequalities and radial mean bodies” In Amer. J. Math. 120.3, 1998, pp. 505–528
  • [23] Richard J. Gardner and Artem Zvavitch “Gaussian Brunn-Minkowski inequalities” In Trans. Amer. Math. Soc. 362.10, 2010, pp. 5333–5353 DOI: 10.1090/S0002-9947-2010-04891-3
  • [24] A. Giannopoulos and M. Papadimitrakis “Isotropic surface area measures” In Mathematika 46.1, 1999, pp. 1–13 DOI: 10.1112/S0025579300007518
  • [25] Julián Haddad and Monika Ludwig “Affine fractional LpL^{p} Sobolev inequalities” In Math. Ann. 388.1, 2024, pp. 1091–1115 DOI: 10.1007/s00208-022-02540-3
  • [26] Julián E. Haddad and Monika Ludwig “Affine Fractional Sobolev and Isoperimetric Inequalities” In Preprint, arxiv: 2207.06375 arXiv, 2022 DOI: 10.48550/ARXIV.2207.06375
  • [27] A. Koldobsky, A. Pajor and V. Yaskin “Inequalities of the Kahane-Khinchin type and sections of LpL_{p}-balls” In Studia Math. 184.3, 2008, pp. 217–231 DOI: 10.4064/sm184-3-2
  • [28] Alexander V. Kolesnikov and Galyna V. Livshyts “On the Gardner-Zvavitch conjecture: symmetry in inequalities of Brunn-Minkowski type” In Adv. Math. 384, 2021, pp. Paper No. 107689\bibrangessep23 DOI: 10.1016/j.aim.2021.107689
  • [29] Liudmyla Kryvonos and Dylan Langharst “Weighted Minkowski’s Existence Theorem
    and Projection Bodies”
    In Trans. Amer. Math. Soc. 376, 2023, pp. 8447–8493 DOI: https://doi.org/10.1090/tran/8992
  • [30] Dylan Langharst, Michael Roysdon and Artem Zvavitch “General measure extensions of projection bodies” In Proc. Lond. Math. Soc. (3) 125.5, 2022, pp. 1083–1129 DOI: 10.1112/plms.12477
  • [31] Rafał Latała “A note on the Ehrhard inequality” In Studia Math. 118.2, 1996, pp. 169–174 DOI: 10.4064/sm-118-2-169-174
  • [32] Galyna Livshyts “Maximal surface area of a convex set in n\mathbb{R}^{n} with respect to exponential rotation invariant measures” In J. Math. Anal. Appl. 404.2, 2013, pp. 231–238 DOI: 10.1016/j.jmaa.2013.03.014
  • [33] Galyna Livshyts “Maximal surface area of polytopes with respect to log-concave rotation invariant measures” In Adv. in Appl. Math. 70, 2015, pp. 54–69 DOI: 10.1016/j.aam.2015.06.004
  • [34] Galyna V. Livshyts “A universal bound in the dimensional Brunn-Minkowski inequality for log-concave measures” In Trans. Amer. Math. Soc. 376.9, 2023, pp. 6663–6680 DOI: 10.1090/tran/8976
  • [35] Galyna V. Livshyts “An extension of Minkowski’s theorem and its applications to questions about projections for measures” In Adv. Math. 356, 2019, pp. 106803\bibrangessep40 DOI: 10.1016/j.aim.2019.106803
  • [36] Galyna V. Livshyts “Some remarks about the maximal perimeter of convex sets with respect to probability measures” In Commun. Contemp. Math. 23.5, 2021, pp. Paper No. 2050037\bibrangessep19 DOI: 10.1142/S0219199720500376
  • [37] Georges Matheron “Random Sets and Integral Geometry”, Wiley Series in Probability and Statistics Michigan: Wiley, 1975, pp. 86
  • [38] Emanuel Milman and Liran Rotem “Complemented Brunn-Minkowski inequalities and isoperimetry for homogeneous and non-homogeneous measures” In Adv. Math. 262, 2014, pp. 867–908 DOI: 10.1016/j.aim.2014.05.023
  • [39] V.. Milman and A. Pajor “Isotropic position and inertia ellipsoids and zonoids of the unit ball of a normed nn-dimensional space” In Geometric aspects of functional analysis (1987–88) 1376, Lecture Notes in Math. Springer, Berlin, 1989, pp. 64–104 DOI: 10.1007/BFb0090049
  • [40] Piotr Nayar and Tomasz Tkocz “A note on a Brunn-Minkowski inequality for the Gaussian measure” In Proc. Amer. Math. Soc. 141.11, 2013, pp. 4027–4030 DOI: 10.1090/S0002-9939-2013-11609-6
  • [41] Fedor 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 1807, Lecture Notes in Math. Springer, Berlin, 2003, pp. 169–187 DOI: 10.1007/978-3-540-36428-3“˙15
  • [42] C.. Rogers and G.. Shephard “The difference body of a convex body” In Arch. Math. (Basel) 8, 1957, pp. 220–233 DOI: 10.1007/BF01899997
  • [43] Rolf Schneider “Convex Bodies: the Brunn-Minkowski Theory” 151, Encyclopedia of Mathematics and its Applications Cambridge, UK: Cambridge University Press, Cambridge, 2014
  • [44] Gaoyong Zhang “Restricted chord projection and affine inequalities” In Geom. Dedicata 39.2, 1991, pp. 213–222 DOI: 10.1007/BF00182294