Weighted Berwald’s Inequality
Abstract.
The inequality of Berwald is a reverse-Hölder like inequality for the th average, of a non-negative, concave function over a convex body in We prove Berwald’s inequality for averages of functions with respect to measures that have some concavity conditions, e.g. -concave measures, 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; 28A751. Introduction
Let be the standard -dimensional real vector space with the Euclidean structure. We write for the -dimensional Lebesgue measure (volume) of a measurable set , where is the dimension of the minimal affine space containing . The volume of the unit ball is written as and its boundary, the unit sphere, will be denoted as usual A set is said to be convex if for every and We say is a convex body if it is a convex, compact set with non-empty interior; the set of all convex bodies in will be denoted by . The set of those convex bodies containing the origin will be denoted A convex body is centrally symmetric, or just symmetric, if . There exists an addition on the set of convex bodies: the Minkowski sum of and , and one has that
We recall that a non-negative function is said to be concave on if for every and one has
and that the support of a function is precisely 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 we shall assume it is not identically infinity, and so will have a finite maximum value, denoted . If is the support of a non-negative, concave function , then are the level sets of . Notice that the level sets are also convex. Additionally, if then for all If is even, then is symmetric and so too is each In any case, if is also bounded, then each (for each ).
We next recall that the classical Berwald inequality states that if is a non-negative, concave function supported on some convex set then, the function given by
(1) |
is decreasing for [4] with equality [22] if and only if the graph of is a certain cone with as its base. Here, the combinatorial coefficients are given by with the standard Gamma function, defined for except for when is negative integer. Usually written in the form for 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 has density if it has a locally integrable Radon-Nikodym derivative from to , i.e,
A Borel measure on is said to be -concave on a class of compact subsets of if there exists a continuous, (strictly) monotonic, invertible function such that, for every pair and every , one has
When this can be written as
and we say is -concave. When , we merely say the measure is concave. In the limit as , we obtain the case of log-concavity, which can also be obtained by taking :
The classical Brunn-Minkowski inequality (see for example [21]) asserts the -concavity of the Lebesgue measure on the class of all compact subsets of . From Borell’s classification on concave measures [7], a Radon measure (locally finite and regular Borel measure) is log-concave on Borel subsets of if, and only if, has a density that is log-concave, i.e. where and is convex. Similarly, a Radon measure is -concave on Borel subsets of , if, and only if, has a density that is -concave (if ) or -convex (if ), where However, all we will require is that a measure is -concave on a class of convex sets; we will discuss an important example below. Thus, our results in the case of -concave measures include measures beyond Borell’s classification.
We can now state our first main result, which is the Berwald inequality for -concave measures under different restrictions on the function . This result applies to a variety of measures, including -concave ones.
Theorem 1.1 (The Berwald Inequality for measures with concavity).
Let be a non-negative, concave function supported on . Let be a Borel measure such that and satisfies one of the below listed concavity assumptions on a collection of convex subsets of containing the level sets of . Then, for any we have
where
-
(1)
If is -concave, where is a continuous, increasing and invertible function:
There is equality if, and only if, for all the following formula holds
and for all must satisfy
-
(2)
If is -concave, where is a continuous, increasing and invertible function:
Equality is never obtained.
-
(3)
If is -concave, where is a continuous, decreasing and invertible function:
Equality is never obtained.
In all cases, is defined implicitly via where
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 being integrable implies or On the other hand, we will show that if there is equality, then would be finite. Similar logic holds for Case 3. However, the inequality is asymptotically sharp as is made arbitrarily large on its support.
We obtain the following corollary for -concave measures; the case where 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 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 -concave measures).
Let be a non-negative concave function supported on Let be a Borel measure finite on that is -concave, , on a collection of convex subsets of containing the level sets of . Then, for any we have
where
For we must restrict to for integrability.
If there is equality if, and only if, for all and
implying
If or 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 when the measure is -concave and -homogeneous, . Recall that a Borel measure is said to be -homogeneous for some if for all compact sets and so that If has density , then one can check using the Lebesgue differentiation theorem that this implies that is homogeneous.
We say a set with is star-shaped if every line passing through the origin crosses the boundary of exactly twice. We say is a star body if it is a compact, star-shaped set whose radial function given by is continuous. For the Minkowski functional of is defined to be The Minkowski functional of is a norm on if is symmetric. If and satisfy that is a star body, then the generalized radial function of at is defined by . Note that for every is a star body for every .
One gets the following formula for when is -homogeneous, , and is a star body in .
(2) |
Crucial to the statement of equality conditions, and our investigations henceforth, will be the roof function associated to a star body , which we define as for and, for In polar coordinates, becomes an affine function in for :
(3) |
Note that if then we can also write for and otherwise. Observe that, for a non-negative, concave function supported on some one obtains for and that
(4) |
we will make liberal use of this bound throughout this work. Functions of the form for some and will also be referred to roof functions, with height and vertex . The reason for this vocabulary will become clearer below.
Using (2), one can verify by hand that the function satisfies, for an -concave, -homogeneous measure, that
Therefore, yields equality in the Berwald inequality for -concave measures, Corollary 1.2, under the additional assumption that is -homogeneous. The next theorem shows this is the only such function.
Theorem 1.3.
(The Berwald Inequality for -concave, -homogeneous measures) Let be a non-negative, concave function supported on . Let be a Radon measure containing in its support that is -concave, -homogeneous for some . Then, for any we have
Suppose . Then, there is equality if, and only if, is an affine function in i.e. one has
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 Berwald was able to show the inequality is true for 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 is a certain cone with as a base, i.e. that 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 Specifically they constructed the roof function in the following way: given a convex set (which will become the base of a hypercone), let be the height of the hypercone, and let be the location of the projection of vertex of the hypercone. Then, the roof function with height and vertex is equivalently defined as the non-negative, concave function whose graph is given by
where conv denotes the convex hull. From this formulation, we obtain an interesting formula for the level sets of a roof function for one has that
Corollary 1.4 (The Classical Berwald Inequality).
Let be a non-negative, concave function supported on . Then, for any we have
There is equality if, and only if, is an affine function in up to translation i.e. if is the point in where the maximum of is obtained, one has
Proof.
The inequality follows immediately from Theorem 1.3, as do the equality conditions if the maximum of is obtained at the origin. If the maximum of is not obtained at the origin, let be the point in where obtains its maximum. Let and Then, is a concave function supported on with maximum at the origin, and, for every
Therefore, since there is equality in the inequality for the function and the convex body by hypothesis, there is equality in the inequality for the function and the convex body . Consequently, we have
Using that and yields the result. ∎
We next list two applications for the standard Gaussian measure on which we recall is given by From Borell’s classification, we see that the Gaussian measure is log-concave on 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 and Borel sets and in , we have
(5) |
i.e. is concave, where . 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 and convex set ; the general case for two Borel sets of the Ehrhard’s inequality was proven by Borell [10]. Since 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 concave on 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 be a non-negative, concave function supported on . Then, we have the following:
-
(1)
The function
is strictly decreasing on
-
(2)
The function
is strictly decreasing on where
-
(3)
and, if the maximum of is at the origin and then the function
is decreasing on
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 and that
(6) |
i.e. is -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 and 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) |
That is, every measure is -concave over the class of symmetric convex bodies. To show how rich this class is, 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 be a non-negative, concave, even function supported on a symmetric Let be a measure in containing in its support. Then, for any
We emphasize that the -concavity of the Gaussian measure on shown in [28] and the -concavity of and other measures from over the class of symmetric convex bodies falls strictly outside the classification of -concave measures by Borell. This paper is organized as follows. In Section 2, we prove a version of Berwald’s inequality for -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 . 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 and a Borel set with positive -measure, will denote the normalized probability on with respect to , that is for measurable Notice that for every non-negative, measurable function on and such that , one has the layer cake formula
from the following use of Fubini’s theorem:
Additionally, if is -concave, with increasing and invertible, on a class of convex sets, then for in the support of a concave function , one has that the function given by is -concave, where as long as the level sets of belong to Indeed, since is concave, one has, for and , that
Using the -concavity of this yields
Inserting the definition of and this is precisely
Similarly one can check that if is -concave, with decreasing and invertible, on a class of convex sets, then for in the support of a concave function , one then has that the function is -convex, where That is, is a convex function on its support, as long as the level sets of belong to
We next need the appropriate layer cake formula for when Notice that for every non-negative, measurable function on a Borel set and such that for a Borel measure , one has the layer cake formula
from the following use of Fubini’s theorem:
We now recall the analytic extension of the Gamma function. We start with the definition of when the real part of is greater than zero:
If the real part of is less than zero, then one uses analytic continuation to extend via the multiplicative property . Now, let us obtain the formula for when the real part of is in . From the multiplicative property one can write
(8) |
where, for the second equality, integration by parts was performed and was viewed as the derivative of to maintain integrability. The fact that the layer cake formula looks similar to the formula for when the real part of is between and inspires the analytic continuation of Theorem 1.1 to negative . We will use the Mellin transformation, which was extended to in [20] for -concave functions. We further generalize the Mellin transform here.
The Mellin transform of a function such that is the analytic function for given by
(9) |
Following [20], consider the function
(10) |
Then, for all one has where is the constant defined in Corollary 1.2, that is Berwald’s inequality for -concave measures; notice again that in the case when , for to be integrable, we must have that
Motivated by this example, we need to define a function whose Mellin transform is related to the constant from Theorem 1.1, and this definition will depend on the concavity of . Recall that a function is -concave for a monotonic function if is either concave (if is increasing) or convex (if is decreasing). Similarly, is -affine if is an affine function. We will have three different restrictions on the function , matching those in Theorem 1.1 (and the notation as well). First, fix some . Then, we will consider the case when where represents those functions that are continuous, increasing and invertible; represents those functions that continuous, increasing and invertible; and represents those functions that are continuous, decreasing and invertible. We next define
(11) |
Notice that, if then if , and this also holds for any such that is integrable.
We will now work towards the proof of Theorem 1.1. Let be a non-negative function such that Then, for set
(12) |
where and is defined implicitly by Next, set for
(13) |
and
Lemma 2.1 (The Mellin-Berwald Inequality).
Let be an integrable, -concave function, (elaborated above (11)). Suppose that is right differentiable at the origin. Next, set where is defined via (12). Then,
-
(1)
and if is non-increasing then
-
(2)
for every Thus, defined via (13), is well-defined and analytic on
-
(3)
is non-increasing on
-
(4)
If there exists such that then is constant on Furthermore, is constant on if, and only if, for some in which case
Proof.
From the fact that one immediately has that Notice that for If is non-increasing, then from (9) one obtains that as well. Thus, for all and thus
For the second statement, clearly for So, fix some such that Then, Define the function Notice that and, by performing a variable substitution, via (9) for every In particular, for From the definition of we then obtain that Thus, from (9), one obtains
Consequently, the function changes signs at least once. But actually, this function changes sign exactly once. Indeed, let be the smallest positive value such that Then, Now, is affine. If then is concave and the slope of is negative. Since one has that on From the concavity, we must then have that on Similarly, if then is convex and the slope of is positive. Hence, on and on Taking inverses, we obtain in either case that on and on
Next, define
Clearly, One has Thus, is non-increasing on and non-decreasing on Hence for all Next, pick From integration by parts, one obtains
Hence,
We deduce that
(14) |
for every Sending we obtain for every and thus for One immediately obtains that is well-defined and analytic on Finally, taking the th root of (14) yields for that
i.e. is non-increasing on Suppose there exists an such that Then, there is equality in (14). But this yields for almost all We take a moment to notice that this then yields for every Anyway, since for almost all , we have for almost all . Hence, the concave function equals the affine function for almost all and thus for all . Consequently, Conversely, suppose that for some Then, direct substitution yields on Notice that is also true for any Consequently, by picking any we repeat the above arguments and deduce again that
This time, however, Consequently, this immediately implies that
for every This establishes that is non-increasing on as well. The argument for the equality conditions is the same. ∎
Proof of Theorem 1.1.
Let be the concavity of our measure Next, let Notice this is non-increasing, and thus from the statement of Lemma 2.1 is . Then, for
via the layer cake formula for ; similar computations yield the case for and follows from limits. Thus, we obtain from Lemma 2.1, Item 3, that the function
is non-increasing for Furthermore, , if, and only if,
We now insert the appropriate , starting with the case . This is precisely
(15) |
We then evaluate the above at to obtain
On the other hand, we also know that, for all we have
Inserting the formula for and the formula of from (15), we obtain
By performing a variable substitution in the denominator, we obtain that
Therefore, we have which means
Next, we show that equality never occurs for when , and the case is similar. From integrability, we have that or (where these are understood as limits from the left and the right, respectively). On the other hand, we have shown equality implies
Evaluating again at yields which would imply that ∎
Proof of Corollary 1.2.
We have that is -concave on the level sets of , 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 the case when is exactly the same (using the analytic continuation of the Beta function), and then the case follows from limits. Inserting yields
Focus on the function For this function to be integrable near zero, we require and, for the integrability near infinity, we require Thus, We will now manipulate to obtain a more familiar formula. Consider the variable substitution given by Writing as a function of this becomes
As and as We then obtain that
which equals our claim. ∎
Proof of Theorem 1.3.
From the assumptions on the measure , we obtain that for some -concave function Furthermore, is 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 is obtained at the origin. Equality conditions of Corollary 1.2 imply that
Using (2), this implies that
Using Fubini’s theorem, a variable substitution and the homogeneity of yields
One has from (4) that a concave function supported on whose maximum is at the origin satisfies
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 be a Radon measure that is -concave, -homogeneous, , and suppose that is given by (3) for some . Let be a concave function supported on , and suppose Then, one has
Proof.
Let be a concave perturbation of by , i.e. is picked small enough so that is concave with maximum at the origin for all and Next, consider the function given by, for
from Berwald’s inequality in Theorem 1.3, this function is greater than or equal to zero for all and equals zero when Hence, the derivative of this function is non-negative at . By taking the derivative of in the variable t, evaluating at , 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 behaves logarithmically, that is one can apply the second case of Theorem 1.1. Finally, for the third inequality, note that if is a concave function supported on some with maximum at the origin, then the level sets of are also in and thus one can apply the -concavity of the Gaussian measure over and use the first case of Corollary 1.2. ∎
Proof of Corollary 1.6.
Notice that if is an even, concave function supported on a symmetric , then the maximum of is at the origin (for every and so ) and the level sets of are all symmetric convex bodies. Thus, the result follows from the -concavity of measures in . ∎
2.1. Applications
We conclude this section by showing a few applications. The first example uses that the support of in Theorem 1.1 need not be compact.
Theorem 2.3.
Let . Denote and Denote
Then, for every Borel measure finite on with one of the following concavity conditions on subsets of :
-
(1)
If is -concave, where is an increasing and invertible function one has
for every where the integrals exist. In particular, if one obtains
-
(2)
If is -concave, where is an increasing and invertible function one has
for every where the integrals exist; the case for can be deduced. For the Gaussian measure especially, one can set and obtain
If one obtains for every that
-
(3)
If is -concave, where is a decreasing and invertible function one has
for every where the integrals exist; the case for can be deduced. In particular, if and one obtains
Finally, let be a Borel measure finite on some convex Suppose is either or concave, where the functions and are as given in Theorem 1.1. Next, consider a non-negative function so that is bounded and concave on for some . Inserting into Theorem 1.1 and picking appropriate choices of and we obtain that for every one has
(16) |
up to possible restrictions on admissible and so that all constants exist. In words, we have bounded the norm of a bounded, non-negative, -concave function by its norm when is either or -concave. Examples of interest are when is -concave. We obtain for a -concave measure and :
-
(1)
When :
-
(2)
When :
-
(3)
When and
We also highlight the following examples for the Gaussian measure.
-
(1)
-
(2)
If and the maximum of is obtained at the origin:
-
(3)
Let be a measure in . If is symmetric, and is even:
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 for the Lebesgue measure on .
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 finite on a Borel set in its support, the th mean of a non-negative is
(17) |
Jensen’s inequality states that for From continuity, one has and
Recall that for a star body , .
Definition 3.1.
Let be a Borel measure on and a convex body contained in the support of . Then, the -weighted th radial mean body of , denoted is the star body whose radial function is given, for and as
We will show why exists for later in this section. The usual radial mean bodies , first defined by Gardner and Zhang [22], are precisely the bodies in our notation. Sending to or , we obtain the limiting bodies and , given in terms of their radial functions by
where is the difference body of , given by
(18) |
A natural question is how behaves under linear transformations. We introduce the following notation: for a Borel measure and , we denote by the pushforward of by ; note that if has density then has density , and for a Borel set .
Proposition 3.2.
Let be a Borel measure finite on a convex body . Then, for and one has
Proof.
Suppose ; the case follows by continuity. Let be a star body in Then, one can verify from the definition (or see [21, page 20]) that
In particular, Then, observe that, by performing the variable substitution
∎
We will show that when is -concave, , then is a convex body for . But first, we must make a detour.
3.2. Weighted Projection Bodies
A convex body is uniquely determined by its support function . The dual body of is given by (notice this yields that ). For a Borel measure with density and a , the -measure of the boundary of , denoted , is
(19) |
where the second equality holds if the is a limit and there exists some canonical way to select how behaves on Here, is the -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 with density , as long as contains in its Lebesgue set (see Lemma 3.5 below).
Using weighted surface area measures, the centered, -weighted projection bodies of a convex body and a Borel measure with continuous density were defined as [30] the symmetric convex body whose support function is given by, for ,
(20) |
where is the Gauss map, which associates an element with its outer unit normal. Since the set is of measure zero, we will write integrals over involving as integrals over without any confusion.
As alluded to by the discussion after (19), the formula (20) is still well-defined when is not continuous but contains in its Lebesgue set; in which case, on is understood as, for ,
For such and ( density of ), the shift of with respect to is given by
where the second equality holds when is in Recall the notation that, if is a function, then there exists two non-negative functions, denoted and , such that . One can then write and obtain We define the -weighted projection body of to be the convex body defined via the support function, for every
(21) |
where the last integral is to emphasize that contains the origin in its interior. In the case when , one has where is the projection body of . This projection body is a fundamental tool in convex geometry; see e.g. [21]. It turns out that for :
(22) |
where the first equality is from (20), the orthogonal projection of onto a linear subspace is denoted by , 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 be a convex body in Then, for a Borel measure , the -covariogram of is the function given by
(23) |
The classical covariogram of is given by . 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 is Lipschitz on a bounded domain if, for every , one has for some .
Proposition 3.4 (The radial derivative of the covariogram, [30]).
Let . Suppose is a domain containing , and consider a Borel measure with density locally Lipschitz on . Then,
(24) |
We now briefly show that the assumption of Lipschitz density can be dropped. For a continuous function , the Wulff shape or Alexandrov body of is defined as
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 be a Borel measure on with locally integrable density . Let be a convex body such that , up to set of -dimensional Hausdorff measure zero, is in the Lebesgue set of . Then, for a continuous function on , one has that
Next, note that for any , . Also, for any convex body we have . Consequently,
Thus, the body is the Wulff shape of the function . Suppose we have a Borel measure with density , such that is in the Lebesgue set of . Then, observe that . Therefore, we obtain (24) from Lemma 3.5, with , and (21). We list this strengthened version as a separate result.
Theorem 3.6.
Let be a convex body in and a Borel measure whose density contains in its Lebesgue set and in its support. Then, for every fixed direction , one has
(25) |
From the Brunn-Minkowski inequality, is a -concave function supported on . One can readily check that the -covariogram inherits the concavity of the measure in general.
Proposition 3.7 (Concavity of the covariogram).
Consider a class of convex bodies with the property that for every . Let be a Borel measure finite on every Suppose is a continuous and invertible function such that is -concave on . Then, for is also -concave, in the sense that, if is increasing, then is concave, and if is decreasing, then is convex.
Proof.
We first observe the following set inclusion: for and , we have from convexity that
Using this set inclusion, we obtain that
From the fact that is -concave, we obtain
∎
We see that the (-)covariogram connects the (-)projection body and difference body of . The covariogram is also related to the radial mean bodies, since, for :
where, in the second step, we used Fubini’s theorem and the fact that and for all . Therefore, we can write, for , that
(26) |
Additionally, this formulation implies that is a convex body if is -concave in the Borell sense for some (see [2, Theorem 5] for and [22, Corollary 4.2]).
3.4. The Negative Regime for p
We first convince the reader that exists for when has a bounded, positive density . Under these assumptions, let Then, for
One then deduces that under these constraints, for and, for one has . There is equality if, and only if, is constant on Notice these inclusions show that is well-defined for . By sending we deduce that as
For a general Borel measure with density, we now obtain a formula for when This also establishes existence. Notice that, in this instance,
Adding and subtracting integration over we obtain
Notice this formulation could have been established directly via the continuity of the Mellin transform. Hence, we can write, for that
(27) |
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 , we can use properties of th averages of functions, i.e. Jensen’s inequality, to immediately obtain the following.
Theorem 3.8.
Let be a Borel measure finite on a convex body contained in its support. Then one has that, for
We now take a moment to discuss how was originally handled in the volume case. Gardner and Zhang defined another family of star bodies depending on , the spectral pth mean bodies of denoted However, to apply Jensen’s inequality, they had to assume additionally that To avoid this assumption, we change the normalization and define as the star body whose radial function is given by, for
where is the X-ray of in the direction for (see [21, Chapter 1] for more on the properties of , and note that ), and
Here, is the polar projection body of .
Therefore, from Jensen’s inequality, we obtain, for
(28) |
The fact that, for
(29) |
shows , and that we can analytically continue to by As observed in [22], the relation shows that as but the shape of tends to that of (note that due to the alternate normalization of these relations are expressed differently in [22, Theorem 2.2]; in both instances, it is unknown if and are convex for ).
Gardner and Zhang then obtained a chain of inequalities concerning ([22, Theorem 5.5]), a reverse to the one from Jensen’s inequality: for
(30) |
where are constants defined via continuity at , and, for with the standard Beta function. There is equality in each inclusion in (30) if, and only if, is a -dimensional simplex. One obtains from (29) that, indeed, tends to as , since tends to . When , one obtains, since , the following special case:
(31) |
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 as , we will not pass through spectral mean bodies. To explain why, we shall, for simplicity, focus on the Gaussian measure and a symmetric . Suppose we defined Gaussian spectral mean bodies as the star body whose radial function is given by, for
Notice that an analogue of (29), which relates the radial functions of and when does not hold. Consequently, we cannot determine the shape of as via . Perhaps then, the focus should be on and not . But notice that
since one does not have an equivalent of Minkowski’s integral formula in the weighted case. Furthermore, it is not necessarily true that is convex as a function of Hence, it is not necessarily the Minkowski functional of a convex body. Additionally, To summarize, is not related to or and is not related to It is for these reasons we do not study weighted spectral mean bodies.
We must determine the shape of as Applying integration by parts to both (26) and (27), we obtain that, for all one has
(32) |
where we used Lebesgue’s theorem to obtain that is differentiable almost everywhere on as it is monotonically decreasing in the variable Taking the limit as we see that
On the other-hand, recall the following lemma.
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 be a concave function that is supported on a convex body such that
Define then
(34) |
whenever and . In particular, if is non-negative, then we have
One has for if, and only if,
Proof.
From the concavity of the function , one has
Then, the inequality (34) follows from the definition of . If is additionally non-negative, then one obtains that by using that for a fixed and and then setting For the equality conditions, suppose then, by definition of being the support of one obtains Conversely, suppose Then, from the concavity of , one has and so there is equality. ∎
Using Proposition 3.7, Lemma 3.10 and (25), we obtain for a Borel measure with density such that is -concave, is an increasing and differentiable function, that
(35) |
for every such that is in the Lebesgue set of the density of .
Theorem 3.11.
Fix a convex body in . Let be a finite Borel measure containing in its support, such that is -concave on convex subsets of , where is a continuous, increasing, and invertible function. Then, for , one has
where
and, for the last set inclusion, we additionally assume that has a locally integrable density containing in its Lebesgue set and that is differentiable at the value The equality conditions are the following:
-
(1)
For the first two set inclusions there is equality of sets if, and only if, and
-
(2)
For the last set inclusion, the sets are equal if, and only if,
Proof.
We now obtain a result for -concave measures, the promised generalization of (30).
Corollary 3.12.
Fix a convex body in . Let be a Borel measure containing in its support that is -concave, on convex subsets of . Then, for , one has
where the last inclusion holds if has locally integrable density containing in its Lebesgue set.
There is equality in any set inclusion if, and only if, . If is a -concave Radon measure, then and equality occurs if, and only if, is -dimensional simplex, the density of is constant on , and .
Proof.
Setting in Theorem 3.11 yields, in the case when
and similarly for The equality conditions from Theorem 3.11 yields that is an affine function along rays for . If is a -concave Radon measure, then one must have . For such -concave measures, 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:
-
(i).
is a simplex.
-
(ii).
For any , either is empty or it is homothetic to .
The equivalence between and 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 be a -concave Radon measure on with density , Borel sets of positive measure, and , and suppose that
Then up to -null sets, there exist , such that and such that for all .
Proposition 3.14.
Let , , and an -concave Radon measure, whose density contains in its support. Then, is an affine function in for for every and if, and only if, is -dimensional simplex, is a constant on , and .
Proof.
Let lie in the interior of , and let . The fact that is affine on the segment precisely means that for ,
(36) |
where . Examining the proof of the Proposition 3.7, we see that and equality can hold in (36) only if . In particular, we have
(37) |
By Lemma 3.13, that is homothetic to for all in the interior of , which implies that is a -dimensional simplex.
It remains to show that the density of is constant on . For this we use the second conclusion of Lemma 3.13: for each there exists such that for each , , where is the affine transformation which maps onto . But note that is a continuous map from the compact, convex set to itself, so it has a fixed point by Brouwer’s fixed point theorem. For such , we have , implying . (Note that as a convex function is continuous on the interior of its domain [43, Theorem 1.5.3], the density of is continuous on ; in particular, is well-defined pointwise on the interior of , not just up to sets of measure zero.)
Once one knows that for any homothety mapping to and any , one verifies by tedious but elementary arguments (e.g., by starting with the faces of , working by induction on the dimension) that any two points in can be mapped into each other by a chain of such ’s, which implies that is indeed constant on . Thus we are back to the case of Lebesgue measure, which we know is -affine on pairs of homothetic bodies. Conversely, one verifies that if is a -dimensional simplex and is constant on then 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, may tend to as and so will tend to the origin. Hence, we lose the first set inclusion:
Theorem 3.15 (Logarithmic Case).
Suppose a Borel measure on is finite on some convex body and -concave, where is an increasing and invertible function. Then, for , one has
where
and, for the second set inclusion, we additionally assume that has locally integrable density containing in its Lebesgue set and that is differentiable at the value . In particular, if is log-concave:
where is interpreted via continuity.
Proof.
The first inclusion follows from the second case of Theorem 1.1. For the second inclusion, suppose Then, one has
Since may possibly be negative, we shall leave inside the integral:
and so which yields the result. The case for is similar. ∎
We list the Gaussian measure case as a corollary.
Corollary 3.16.
Let be a convex body. Then, for , one has
where is interpreted via continuity, and
where
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 with homogeneity , then there exists a radial mean body whose measure is “of the same order” as that of itself. First, define the -translated-average of with respect to as
(38) |
The last equality follows from Fubini’s theorem, and this definition has appeared in [30]. Next, we see that when is homogeneous of degree , we obtain a relation between and
Lemma 3.17.
Fix a convex body and a Borel measure that is -homogeneous with density and a Borel measure on . Then, one has .
Proof.
Theorem 3.18 (Rogers-Shephard type inequality for an -homogeneous and a -concave measure).
Fix a convex body . Consider a Borel measure that is -homogeneous and a Borel measure on that is -concave, , on convex subsets of , whose locally integrable density contains in its Lebesgue set and in its support. Then,
with equality if, and only if, . If is a -concave Radon measure, then and equality occurs if, and only if, is -dimensional simplex, the density of is constant on , and .
An upper bound for when is -concave was first shown by Borell, [7]. However, the bound was not sharp.
Corollary 3.19 (Zhang’s Inequality for an -homogeneous and a -concave measure).
Fix a convex body . Consider a Borel measure that is -homogeneous and a Borel measure on that is -concave, , on convex subsets of , whose locally integrable density contains in its Lebesgue set and in its support. Then, one has
with equality if, and only if, . If is a -concave Radon measure, then and equality occurs if, and only if, is a -dimensional simplex, the density of is constant on , and .
3.7. The Gardner-Zvavitch Inequality and Radial Mean Bodies
We would like to apply the -concavity of the Gaussian measure over symmetric convex bodies in (6), which is also true for every , to obtain that Corollary 3.12 holds for such measures. We run into an issue: even if is symmetric, then is not symmetric in general. Therefore, Proposition 3.7 does not apply, i.e. being concave does not imply that is concave. To remedy this, we take a cue from [30] and define the polarized covariogram as
As can be found in [30], the set is a symmetric convex body when is, and inherits any concavity of over symmetric convex bodies.
Under the assumption that is a Borel measure with even density and is symmetric, notice that . One also has, for every ,
(39) |
This was first shown in [30] under the additional assumption that the density of is Lipschitz; arguing similarly to Theorem 3.6 allows one to weaken the assumption to merely the density of contains in its Lebesgue set. In fact, one does not need the symmetry assumptions on and for this proof; one will obtain in general
(40) |
but is not necessarily .
In order to obtain Corollary 3.12 for , we define the polarized weighted mean bodies, , as the star bodies on whose radial function is given by, for and ,
(41) |
The bodies are symmetric convex bodies for . Once again, are interpreted via continuity, and for contained in the support of . We again have that
(42) |
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 in . Let be a Borel measure containing in its support. Then, for , one has
There is equality in any set inclusion if, and only if, .
Recently, it was shown by Livshyts [34] that for any even, log-concave probability measure on , is -concave over the class of symmetric convex bodies, with . One can then use this result to formulate Theorem 3.20 for such measures, with in the coefficients replaced by . Here, is a bounded sequence that goes to as .
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 is symmetric (since being -concave on convex subsets of implies it is -concave on symmetric convex subsets of , and the fact that will then inherit the concavity) and replaced by (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 ” 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 -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 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 -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 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 -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 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