Convex Geometry
Convex Geometry
Convex
Geometry
Cetraro, Italy 2021
Andrea Colesanti · Monika Ludwig
Editors
Lecture Notes in Mathematics
Volume 2332
Editors-in-Chief
Jean-Michel Morel, Ecole Normale Supérieure Paris-Saclay, Paris, France
Bernard Teissier, IMJ-PRG, Paris, France
Series Editors
Karin Baur, University of Leeds, Leeds, UK
Michel Brion, UGA, Grenoble, France
Annette Huber, Albert Ludwig University, Freiburg, Germany
Davar Khoshnevisan, The University of Utah, Salt Lake City, UT, USA
Ioannis Kontoyiannis, University of Cambridge, Cambridge, UK
Angela Kunoth, University of Cologne, Cologne, Germany
Ariane Mézard, IMJ-PRG, Paris, France
Mark Podolskij, University of Luxembourg, Esch-sur-Alzette, Luxembourg
Mark Policott, Mathematics Institute, University of Warwick, Coventry, UK
Sylvia Serfaty, NYU Courant, New York, NY, USA
László Székelyhidi , Institute of Mathematics, Leipzig University, Leipzig,
Germany
Gabriele Vezzosi, UniFI, Florence, Italy
Anna Wienhard, Ruprecht Karl University, Heidelberg, Germany
2023
Paolo Salani
Daniele Angella
[email protected]
Shiri Artstein-Avidan • Gabriele Bianchi •
Andrea Colesanti • Paolo Gronchi • Daniel Hug •
Monika Ludwig • Fabian Mussnig
Convex Geometry
Cetraro, Italy 2021
Fabian Mussnig
Institut für Diskrete Mathematik und Geometrie
Technische Universität Wien
Wien, Austria
Editors
Andrea Colesanti Monika Ludwig
Dipartimento di Matematica e Informatica Institut für Diskrete Mathematik und Geometrie
“Ulisse Dini” Technische Universität Wien
University of Florence Wien, Austria
Firenze, Italy
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2023
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
This volume collects the lecture notes of the Summer School on Convex Geometry,
held in Cetraro, Italy, from August 30th to September 3rd, 2021.
Convex geometry is a very active area in mathematics with a solid tradition and
a promising future. Its main objects of study are convex bodies, that is, compact and
convex subsets of n-dimensional Euclidean space. The so-called Brunn–Minkowski
theory currently represents the central part of convex geometry.
The Summer School aimed to provide an introduction to various aspects of
convex geometry: The theory of valuations, including its recent developments
concerning valuations on function spaces; geometric and analytic inequalities,
including those which come from the .Lp Brunn–Minkowski theory; geometric and
analytic notions of duality, along with their interplay with mass transportation and
concentration phenomena; symmetrizations, which provide one of the main tools
to many variational problems (not only in convex geometry). Each of these parts is
represented by one of the courses given during the Summer School and corresponds
to one of the chapters of the present volume. The initial chapter contains some basic
notions in convex geometry, which form a common background for the subsequent
chapters.
The material of this book is essentially self-contained and, like the Summer
School, is addressed to PhD and post-doctoral students and to all researchers
approaching convex geometry for the first time.
We are deeply grateful to Fondazione CIME for giving us the opportunity to
carry out the Summer School and providing constant support for its organization.
v
Contents
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
vii
Contributors
ix
Chapter 1
Notation and Introductory Material
Andrea Colesanti
1.1 Introduction
This short note collects the preliminary material for the lecture notes of the courses
of the C.I.M.E. Summer School entitled “Convex Geometry”, held in Cetraro, Italy,
in the summer of 2021.
It contains basic notions in convex geometry, such as: the definition of a convex
body; the Minkowski addition; Hausdorff distance; support function; polar body,
gauge function; the Steiner formula; intrinsic and mixed volumes; the surface area
measure. These notions are supplied by examples and basic properties.
In general we do not include proofs. The most elementary statements are left as
exercises; for the rest we refer to [2] and [1], where the reader can find exhaustive
presentations of the material of this note.
The author is grateful to the other lecturers of the summer school, Shiri Artstein,
Gabriele Bianchi, Paolo Gronchi, Daniel Hug, Monika Ludwig and Fabian Mussnig,
for their advice, which contributed to improve this chapter.
A. Colesanti (O)
Dipartimento di Matematica e Informatica “U. Dini”, University of Florence, Florence, Italy
e-mail: andrea.colesanti@unifi.it
1.2 Notation
and
B n = {x ∈ Rn : |x| ≤ 1}.
.
For .k ∈ [0, n], .Hk denotes the k-dimensional Hausdorff measure in .Rn . In
particular .Hn is the Lebesgue measure in .Rn . For brevity, integration with respect
to the Lebesgue measure will be often denoted by “.dx”. We set
κn = Hn (B n ).
.
For reasons that will be clear once that the notion of intrinsic volumes will be
introduced, the Lebesgue measure in .Rn will be also denoted by .Vn (and referred to
as the volume).
A convex body is a non-empty compact and convex subset of .Rn ; the family of
convex bodies will be denoted by .Kn :
Examples
. The unit ball .B n .
. Let .p ≥ 1; the set
is the so-called .Lp -ball, and it is a convex body. The standard Euclidean ball is
retrieved for .p = 2.
. The unit cube:
Given .A ⊂ Rn , the convex hull of A, denoted by .conv A, is the set formed by the
points of the form
W
m
x=
. λi xi ,
i=1
W
m
(λ1 , . . . , λn ) ∈ [0, 1]n
. with λi = 1.
i=1
The convex hull of finitely many points in .Rn belongs to .Kn , and is called a
polytope. The family of polytopes will be denoted by .Pn .
The unit cube is an example of polytope, as it is the convex hull of the points
.x = (x1 , . . . , xn ), where each component .xi is either 0 or 1.
The aim of this part is to state some results concerning the existence of supporting
hyperplanes to a convex bodies. This question is connected to the notion of metric
projection.
Let .K ∈ Kn and let H be a hyperplane of .Rn . H is called a supporting hyperplane
to K if .K ∩H /= ∅, and K is contained in one of the two closed half-spaces bounded
by H . More precisely, let .u ∈ Sn−1 , and let H be a hyperplane orthogonal to u. Then
H can be written in the form
H = {x ∈ Rn : <x, u> = α}
.
We recall that given a (non-empty) subset A of .Rn and .x0 ∈ Rn , the distance of
.x0 from A is defined as
Exercise 1.2 Let .K ∈ Kn . For every .x ∈ Rn there exists a unique point .pK (x) ∈
∂K, such that
. dist(x, K) = |x − pK (x)|.
x − pK (x)
u=
.
|x − pK (x)|
K, passing through x.
Note that the supporting hyperplane through the boundary point of a convex body
may not be unique.
The previous result is complemented by the following one.
Proposition 1.6 Let .K ∈ Kn ; for every unit vector .u ∈ Sn−1 there exists a (unique)
supporting hyperplane H to K, with outer normal u.
Based on this result, we may give the following definition.
Definition 1.7 Let .K ∈ Kn and .u ∈ Sn−1 . We denote by .K(u) the subset of those
points .x ∈ ∂K such that u is an outer normal vector to K at x.
Exercise 1.8 Prove that, for every .K ∈ Kn and for every .u ∈ Sn−1 , .K(u) is a
non-empty closed convex set (a convex body), with .dim(K(u)) ≤ n − 1.
Let .K ∈ Kn and .x ∈ ∂K. We say that x is a regular point if there is only one
outer normal vector to K at x.
1 Notation and Introductory Material 5
Proposition 1.9 Let .K ∈ Kn ; then .∂K is .Hn−1 -measurable, and .Hn−1 (∂K) < ∞.
Moreover .Hn−1 -a.e. point of .∂K is regular.
K1 + K2 = {x + y : x ∈ K1 , y ∈ K2 }.
.
t K = {tx : x ∈ K}.
.
union over x (or vice versa). This may help to figure out how the sum of two convex
bodies looks like.
Exercise 1.10 Based on the previous definitions:
. draw a picture of the sum of a circular disk and a square in the plane;
. draw a picture of the sum of an equilateral triangle T and the rotation of T of an
angle .π ;
. determine the sum of two orthogonal segments in the plane, or of three pairwise
orthogonal segments in .R3 ;
. determine the sum of a two-dimensional circular disk in .R3 , and a segment
orthogonal to it.
Here is an important property of addition of convex bodies; the proof is left as an
exercise.
Proposition 1.11 For convex bodies .K1 , K2 ∈ Kn , and .t1 , t2 ≥ 0,
.t 1 K1 + t 2 K2
Kε = K + εB n
.
Kε = {x ∈ Rn : dist(x, K) ≤ ε}.
.
We have seen in the previous section that .Kn is endowed with two internal
operations. We will now equip .Kn with a metric, namely the one induced by the
Hausdorff distance.
Let .K1 and .K2 be convex bodies. Their Hausdorff distance is:
{ }
δ(K1 , K2 ) = max max dist(x, K2 ), max dist(y, K1 ) .
.
x∈K1 y∈K2
. lim δ(Ki , K) = 0
i→∞
With the metric induced by this distance, the space of convex bodies is a
complete metric space.
Besides completeness, .Kn has the following important compactness property.
Theorem 1.17 (Blaschke Selection Theorem) Let .Ki , .i ∈ N, be a bounded
sequence of elements of .Kn , i.e. there exists .r > 0 such that, for every i,
Ki ⊂ rB n .
.
. lim δ(Pj , K) = 0.
j →∞
. lim δ(Kj , K) = 0.
j →∞
We present here a powerful tool in the study of convex bodies, the support function.
Definition 1.20 Let .K ∈ Kn ; the support function of K, .hK : Rn → R, is defined
as
The support function of a convex body K may be also indicated by .h(K, ·).
The geometric meaning of the support function can be explained as follows. Let
K be a convex body in .Rn and assume that .o ∈ K. Fix a unit vector u and let
.Hu be the unique supporting hyperplane to K perpendicular to u and such that u
is the outer unit normal. The support function of K evaluated at u is the distance
of .Hu from the origin. The support function is then extended from .Sn−1 to .Rn as a
1-homogeneous function.
Exercise 1.21 Determine the support function of the following convex bodies.
. The ball centered at the origin with radius .r ≥ 0.
. A segment.
. The rectangle:
Let .K ∈ Kn ; its support function .hK , being the supremum of linear (and then
additive) functions, is sub-additive:
hK (u + v) ≤ hK (u) + hK (v).
.
(We remark that, here and throughout, by homogeneous we mean in fact positively
homogeneous). Hence .hK turns to be a convex function. We have thus proved the
following statement.
Proposition 1.22 For every .K ∈ Kn , .hK is a 1-homogeneous convex function.
Let .h : Rn → R be a 1-homogeneous convex function. Consider the set
n
.K = {x ∈ Rn : <x, u> ≤ h(u) for every u ∈ Sn−1 } = {x ∈ Rn : <x, u> ≤ h(u)}.
u∈Sn−1
hK = h.
.
We then see that to each convex body we can associate a unique 1-homogeneous
convex function, its support function; vice versa to each 1-homogeneous convex
function h we can associate a convex body having h as support function.
K⊂L
. if and only if hK ≤ hL .
. For .K ∈ Kn
o∈K
. if and only if hK ≥ 0.
One of the most important properties of the support function is its relation with the
operations on convex bodies that we have introduced before, as the following result
shows.
Proposition 1.28 For convex bodies .K, L ∈ Kn , and for .α, β ≥ 0
We now introduce another important notion in convex geometry: the polar body. For
simplicity, we restrict ourselves to polarity with respect to the origin.
Let
Kno = {K ∈ Kn : o ∈ K},
. Kn(o) = {K ∈ Kn : o is an interior point of K}.
Definition 1.30 Let .K ∈ Kn(o) . The polar body of K (with respect to the origin) is
defined as
1 n
K◦ =
. B .
r
Prove that .Q◦ is the convex hull of the four points .(±1, 0), .(0, ±1). .Q◦ is a cross
polytope, i.e. is the convex hull of the elements of an orthogonal basis and their
opposites. An analogous fact holds in general dimension .n ≥ 2.
Proposition 1.33 Let .K ∈ Kn(o) . Then
. .K ◦ ∈ Kn(o) ;
. .(K ◦ )◦ = K.
Let K be a convex body containing the origin in its interior, and assume moreover
that K is symmetric with respect to the origin. The function .||·||K : Rn → R defined
by:
|| · ||K = hK ◦ .
.
It is frequently said that convex geometry is based on the study of the relations
between volume and Minkowski addition. The Steiner formula is one of the simplest
and most significant examples of these relations.
The Steiner formula is also the first evidence of the following general phe-
nomenon: the volume has a polynomial behavior with respect to the addition of
convex bodies.
Let us recall that given a convex body .K ∈ Kn , for a fixed .ε > 0, the parallel set
of K is
Kε = K + εB n = {x ∈ Rn : dist(x, K) ≤ ε}.
.
Roughly speaking, from a geometric point of view .Kε is the union of K and a “shell”
of uniform width .ε surrounding K. The Steiner formula expresses the volume of the
parallel set of a convex body. We recall that .Vn denotes the Lebesgue measure on
.R .
n
Theorem 1.36 (Steiner Formula) Let .K ∈ Kn be a convex body in .Rn , .ε > 0 and
.Kε = K + εB be the parallel set of K. The volume of .Kε is a polynomial of degree
n
n in .ε. More precisely, there exist non-negative numbers .V0 (K), . . . , Vn−1 (K), such
that
W
n
Vn (Kε ) =
. Vi (K)κn−i εn−i ,
i=0
Remark 1.37 The Steiner formula can be presented in a different form, namely
n ( )
W n j
.Vn (Kε ) = ε Wj (K).
j
j =0
The coefficients appearing in this version, .Wi (K), .i = 0, . . . , n, are called the
quermassintegrals of K.
Some of the intrinsic volumes have a familiar geometric meaning.
. .V0 ≡ 1. In other words, this intrinsic volume is a constant independent of K.
Sometimes is useful to think about .V0 as the Euler characteristic, which is 1 for
every convex body.
. .V1 is proportional to the mean width. Given a convex body K, and .u ∈ Sn−1 , the
width .wK (u) of K in the direction u is the distance between the two supporting
hyperplanes to K, orthogonal to u. This quantity can be expressed in terms of the
support function of K, as follows:
The mean width .w(K) of K is the mean value of the width function:
f
1
.w(K) = wK (u) dHn−1 (u).
nκn Sn−1
. .Vn−1 (K) is, up to a factor, the so-called Minkowski content of K, defined as the
following limit
Vn (Kε ) − Vn (K)
. lim .
ε→0+ ε
1 n−1
Vn−1 (K) =
. H (∂K).
2
In general, given .K ∈ Kn , the quantity .2Vn−1 (K) will be called the surface area
of K.
. .Vn (K) is the volume of K.
1 Notation and Introductory Material 13
Intrinsic volumes are very well studied objects in convex geometry. We list here
some of their fundamental properties (see Chap. 3, Theorem 3.5, Theorem 3.14,
and the discussion after Theorem 3.16, for more details). For .j ∈ {0, . . . , n}, let us
consider .Vj as a functional from .Kn to .R.
Proposition 1.38 The following properties hold.
. .Vj (K) ≥ 0 for every .K ∈ Kn ; moreover, .Vj is monotone increasing:
K, L ∈ Kn , K ⊂ L
. ⇒ Vj (K) ≤ Vj (L).
Vj (λK) = λj Vj (K).
.
Vj (K ∪ L) + Vj (K ∩ L) = Vj (K) + Vj (L).
.
The polynomiality of volume with respect to Minkowski addition goes far beyond
what the Steiner formula shows. For any natural number m, .K1 , . . . , Km ∈ Kn
convex bodies and .t1 , . . . , tm non-negative real numbers, we have
W
m
Vn (t1 K1 + · · · + tm Km ) =
. V (Ki1 , . . . , Kin )ti1 . . . tin .
i1 ,...,in =1
14 A. Colesanti
V (K1 , . . . , Kn )
.
As in the previous formula, the following notation will be used throughout: for
m ∈ {1, . . . , n}, .K, K1 , . . . , Kn−m ∈ Kn :
.
The following proposition gathers the basic properties of mixed volumes (see also
Theorem 3.14 in Chap. 3 for a more detailed presentation).
Proposition 1.41 The mixed volume functional .V : (Kn )n → R has the following
properties.
. For every .K ∈ Kn ,
V (K, . . . , K) = Vn (K).
.
V (K ∪ L, K2 , . . . , Kn ) + V (K ∩ L, K2 , . . . , Kn )
.
= V (K, K2 , . . . , Kn ) + V (L, K2 , . . . , Kn ).
as follows. For every regular point .x ∈ ∂K, we denote by .νK (x) the outer unit
normal to K at x. The function
is the Gauss map of K. For future use, we remark that more generally, given a subset
β of .∂K we can define the spherical image of K at .β as the set:
.
−1
It can be proved that if .ω is a Borel subset of .Sn−1 , then .νK (ω) is a .Hn−1
measurable subset of .∂K.
Definition 1.42 Let .K ∈ Kn ; the surface area measure .Sn−1 (K, ·) is a non-negative
Borel measure defined on .Sn−1 as follows: for every Borel subset .ω of .Sn−1
−1
Sn−1 (K, ω) = Hn−1 (νK
. (ω)).
Remark 1.43 The total mass of the surface area measure of .K ∈ Kn is just the
surface area of K
W
m
Sn−1 (P , ·) =
. Hn−1 (Fi )δνi (·).
i=1
Example Let .K ∈ Kn be of class .C 2,+ . In this case, the Gauss map .νK is a
diffeomorphism between .∂K and .Sn−1 . Moreover, the surface area measure of K is
absolutely continuous with respect to the .(n − 1)-dimensional Hausdorff measure
restricted to .Sn−1 , and
f
1
.Sn−1 (K, ω) =
−1
dHn−1 (x)
ω κK (νK (x))
for every Borel set .ω ⊂ Sn−1 . Here .κK : ∂K → R is the Gauss curvature.
The surface area measure has a continuity property with respect to the metric
induced on .Kn by the Hausdorff distance. We recall that a sequence of finite Borel
measures .μj , .j ∈ N, on .Sn−1 is said to converge weakly to the finite Borel measure
.μ on .S
n−1 , if
f f
. lim f (x) dμj (x) = f (x) dμ(x)
j →∞ Sn−1 Sn−1
. lim Kj = K.
j →∞
Then the sequence of measures .Sn−1 (Kj , ·) converges weakly to the measure
Sn−1 (K, ·).
.
We have seen that to each convex body .K ∈ Kn , we can associate its surface area
measure .Sn−1 (K, ·), which is a Borel measure on .Sn−1 . The Minkowski problem
asks to find a convex body which has a prescribed finite Borel measure on the unit
sphere as its surface area measure.
1 Notation and Introductory Material 17
Problem 1.45 Let .σ be a non-negative Borel measure on .Sn−1 ; find a convex body
K in .Kn such that
Let us conclude this section with a formula which shows that the area measure
is, roughly speaking, the first variation of the volume with respect to Minkowski
addition.
Theorem 1.48 Let .K, L ∈ Kn . Then
f
Vn (K + εL) − Vn (K)
. lim = hL (x) dSn−1 (K, x). (1.2)
ε→0+ ε Sn−1
References
1. D. Hug, W. Weil, Lectures on Convex Geometry, Graduate Texts in Mathematics, vol. 286
(Springer, Cham, 2020)
2. R. Schneider, Convex Bodies: the Brunn-Minkowski Theory, Encyclopedia of Mathematics and
its Applications, vol. 151, Second expanded edn. (Cambridge University Press, Cambridge,
2014)
Chapter 2
Valuations on Convex Bodies and
Functions
2.1 Introduction
In his Third Problem, Hilbert asked whether, given any two polytopes of equal
volume in .R3 , it is always possible to dissect the first into finitely many polytopes
which can be reassembled to yield the second. In 1900, it was known that the answer
to the corresponding question in .R2 is yes, but the question was open in higher
dimensions.
Let .Pn be the set of convex polytopes in .Rn . We say that .P ∈ Pn is dissected
into .P1 , . . . , Pm ∈ Pn and write .P = P1 u · · · u Pm , if .P = P1 ∪ · · · ∪ Pm and the
polytopes .P1 , . . . , Pm have pairwise disjoint interiors. So, Hilbert’s Third Problem
asks whether for any .P , Q ∈ Pn of equal volume there are dissections
P = P1 u · · · u Pm ,
. Q = Q1 u · · · u Q m ,
Pi = φi Qi
.
. Z(φP ) = Z(P )
for all .f, g ∈ X such that also their pointwise maximum .f ∨ g and pointwise
minimum .f ∧ g belong to X. Since we can embed spaces of convex bodies in
various function spaces in such a way that unions and intersections of convex bodies
correspond to pointwise minima and maxima of functions, this notion generalizes
the classical notion. We will discuss the results on valuations on convex functions.
E
. Z(P1 ∪ · · · ∪ Pm ) = (−1)|J |−1 Z(PJ ) (2.1)
∅/=J ⊂{1,...,m}
n
for .P1 , . . . , Pm ∈ S and .m ≥ 1 whenever .P1 ∪ · · · ∪ Pm ∈ S. Here .PJ := j ∈J Pj
and .|J | is the cardinality of the set J . The inclusion-exclusion principle holds for
every valuation on .Pn and every continuous valuation on .Kn (see [41, 72]). If .Z :
Pn → R is, in addition, simple, we have
for .P1 , . . . , Pm ∈ Pn .
For .K, L ∈ Kn , define the Minkowski sum by
K + L := {x + y : x ∈ K, y ∈ L}.
.
The following lemma describes a way to obtain new valuations from a given one.
Lemma 2.1 Let .Z : Kn → R be a valuation. If .C ∈ Kn is a fixed convex body and
K ∪ L + C = (K + C) ∪ (L + C).
. (2.3)
Thus .(K + C) ∩ (L + C) ⊂ (K ∩ L) + C.
Since it is easy to see that .(K ∩ L) + C ⊂ (K + C) ∩ (L + C), it follows that
(K + C) ∩ (L + C) = (K ∩ L) + C.
. (2.4)
22 M. Ludwig and F. Mussnig
Applying .Z to (2.3) and to (2.4) for convex bodies .C, K, L and adding, we obtain
the statement. u
n
For .p ∈ R, a functional .Z : Kn → R is called homogeneous of degree p (or
p-homogeneous), if
. Z(t K) = t p Z(K)
E
n
Vn (K + rB n ) =
. r n−j κn−j Vj (K), (2.5)
j =0
where .r ≥ 0 and .κj is the j -dimensional volume of the unit ball in .Rj (with the
convention that .κ0 = 1), that all intrinsic volumes .V0 , . . . , Vn are valuations on .Kn .
Recall that all intrinsic volumes are continuous and increasing functionals on .Kn
and that .V0 is the Euler characteristic and .V0 (K) = 1 for all .K ∈ Kn . Also, recall
that .Vj (K) is the j -dimensional volume of K if K is contained in a j -dimensional
plane and that .Vj is j -homogeneous.
We will use the following notation. Let .e1 , . . . , en be the vectors of the canonical
basis of .Rn . For .x, y ∈ Rn , we write .<x, y> for the inner product and .|x| for the
Euclidean norm of x. The convex hull of subsets .A1 , . . . , Am ⊂ Rn is written as
.[A1 , . . . , Am ] and the convex hull of .x1 , . . . , xm ∈ R as .[x1 , . . . , xm ]. If .E ⊂ R is
n n
an affine plane in .R , then .K(E) and .P(E) are the sets of convex bodies and convex
n
Blaschke [14] obtained the first classification theorem of invariant valuations on .Kn .
Theorem 2.2 (Blaschke) A functional .Z : Kn → R is a continuous, translation
and .SL(n) invariant valuation if and only if there are constants .c0 , cn ∈ R such that
for every .K ∈ Kn .
In the next section, we will obtain a complete classification of translation invariant
valuations in the one-dimensional case, and in the following section, a complete
2 Valuations on Convex Bodies and Functions 23
ζ (x + y) = ζ (x) + ζ (y)
.
for every .x, y ∈ [0, x). Cauchy functions are well understood and can be
completely described (if we assume the axiom of choice) by their values on a Hamel
basis.
Proposition 2.3 A functional .Z : P1 → R is a translation invariant valuation if
and only if there are a constant .c0 ∈ R and a Cauchy function .ζ : [0, x) → R such
that
( )
. Z(P ) = c0 V0 (P ) + ζ V1 (P )
for every .P ∈ P1 .
Proof Set .c0 := Z({0}) and define .Z̃ : P1 → R by
Z̃(P ) := Z(P ) − c0 V0 (P ).
.
Note that .Z̃ is a simple, translation invariant valuation on .P1 . Define the function
.ζ : [0, x) → R by setting
for every .x, y ∈ [0, x). Hence .ζ is a Cauchy function. Using that .Z̃ is translation
invariant, we get .Z̃(P ) = ζ (V1 (P )) for .P ∈ P1 , which concludes the proof. u
n
Since every continuous Cauchy function is linear, we obtain the following result.
24 M. Ludwig and F. Mussnig
. Z(P ) = c0 V0 (P ) + c1 V1 (P )
for every .P ∈ P1 .
A corresponding classification result holds for upper semicontinuous and translation
invariant valuations on .P1 . Such a result also holds for Borel measurable and
translation invariant valuations on .P1 , since every Borel measurable Cauchy
function is linear.
( )
. Z(P ) = c0 V0 (P ) + ζ Vn (P )
for every .P ∈ Pn .
Proof Set .c0 := Z({0}) and define .Z̃ : Pn → R by
Z̃(P ) := Z(P ) − c0 V0 (P ).
.
Note that .Z̃ is a translation invariant valuation on .Pn that vanishes on singletons,
that is, sets of the form .{x} with .x ∈ Rn . We show that there is a Cauchy function
.ζ : [0, x) → R such that
Z̃(P ) = ζ (Vn (P ))
. (2.6)
for every .P ∈ Pn .
We use induction on the dimension n. By Proposition 2.3, the statement (2.6) is
true for .n = 1. Let .n ≥ 2. Assume that it is true for valuations on .Pn−1 . Hence
it is also true for valuations on .P(E) with E any hyperplane in .Rn . The induction
assumption implies that there is a Cauchy function .ζ̃ : [0, x) → R such that
Z̃(P ) = ζ̃ (Vn−1 (P ))
.
2 Valuations on Convex Bodies and Functions 25
for every .P ∈ P(E). Note that the invariance properties of .Z̃ imply that .ζ̃ does not
depend on E. Since .Z̃ vanishes on singletons, we have .ζ̃ (0) = 0. Let E be spanned
by the first .(n − 1) basis vectors .e1 , . . . , en−1 and define .φ ∈ SL(n) by setting
.φe1 = t e1 with .t > 0 and .φej = ej for .1 < j < n and .φen = en . Since .φE = E,
1
t
it follows from the .SL(n) invariance of .Z̃ that
for every .t > 0. This implies that .ζ̃ ≡ 0 and shows that .Z̃ is simple. Thus it suffices
to show that (2.6) holds for every simple, translation and .SL(n) invariant valuation
.Z̃ : P → R.
n
for every simplex .S ∈ Pn , as the valuation .Z̃ is simple, translation and .SL(n)
invariant and√every n-dimensional simplex is a translate of an .SL(n) image of
the simplex . n s n! [0, e1 , . . . , en ] for some .s > 0. For .0 < r < 1, we dissect
the n-dimensional simplex with vertices .v0 , . . . , vn ∈ Rn into the n-dimensional
simplices .T1 with vertices .v0 , r v0 + (1 − r)v1 , v2 , . . . , vn and .T2 with the vertices
.(1 − r)v0 + r v1 , v1 , v2 , . . . , vn (Fig. 2.1). Since .Z̃ is a simple valuation,
√
Choosing .v0 := 0 and .vj := n (s + t)n! ej for .j = 1, . . . , n as well as .r :=
s/(s + t), we obtain from (2.7) that
ζ (s + t) = ζ (s) + ζ (t)
.
for every .s, t ∈ (0, x). Hence, .ζ is a Cauchy function. By (2.2) and since we
can dissect every polytope into simplices, we conclude that (2.6) holds for every
.P ∈ P . n
u
n
. Z(P ) = c0 V0 (P ) + cn Vn (P )
for every .P ∈ Pn .
Corresponding statements hold for upper semicontinuous valuations and for Borel
measurable valuations.
We remark that classification results for .SL(n) invariant valuations are also
known without assuming translation invariance (see [57]). In particular, the follow-
ing result holds. Let .Pno be the space of convex polytopes containing the origin.
Theorem 2.7 A functional .Z : Pno → R is a continuous, .SL(n) invariant valuation
if and only if there are constants .c0 , cn ∈ R such that
. Z(P ) = c0 V0 (P ) + cn Vn (P )
It follows from (2.8) that .o vanishes on polytopes and is therefore not continuous.
The valuation property of .o on .Kn follows directly from (2.8).
Note that .o is translation invariant and that .Ω(K) = 0 if K is lower dimensional.
Hence we may assume in the following that the origin is an interior point of K.
Clearly, (2.8) can be rewritten as
f
1
o(K) =
. κ0 (K, x) n+1 dVK (x), (2.10)
∂K
where
κ(K, x)
κ0 (K, x) :=
.
<x, nK (x)>n+1
28 M. Ludwig and F. Mussnig
and
Here, .nK (x) is the unit outer normal vector of K at x, which is uniquely defined
almost everywhere on .∂K, and .<x, nK (x)> is the distance to the origin of the tangent
hyperplane to K at such x. In (2.10), it is easy to see that .o is .SL(n) invariant.
Indeed, for a Borel set .B ⊂ ∂K, using the fact that the volume of a cone is the
product of its height divided by n and the .(n − 1)-dimensional volume of its base,
we see that . n1 VK (B) is just the n-dimensional volume of the set .{t B : t ∈ [0, 1]}.
Consequently,
for every .φ ∈ SL(n) and every .x ∈ ∂K where .κ0 (K, x) > 0. This is a simple
consequence of the following geometric interpretation of .κ0 (K, x),
κn2
κ0 (K, x) =
. ,
Vn (EK (x))2
where .EK (x) is the unique centered ellipsoid that osculates K at x. We remark that
f
K |→
. ζ (κ0 (K, x)) dVK (x) (2.11)
∂K
is an .SL(n) invariant valuation on .Kn(o) , the set of convex bodies containing the origin
in their interiors when .ζ : [0, x) → [0, x) is a suitable continuous function. The
functionals defined in (2.11) are called Orlicz affine surface areas. If .ζ (t) := t p for
.t > 0 with .p > −n, the so-called .Lp affine surface area of K is obtained, which
was introduced by Lutwak [59]. Classification results for .SL(n) invariant valuations
on .Kn(o) were established in [38, 49, 56] and characterizations of .Lp and Orlicz affine
surface areas in [56].
The following result from [48, 55] strengthens Theorem 2.2 and establishes a
characterization of affine surface area.
Theorem 2.8 A functional .Z : Kn → R is an upper semicontinuous, translation
and .SL(n) invariant valuation if and only if there are constants .c0 , cn ∈ R and
.c ≥ 0 such that
for every .K ∈ Kn .
2 Valuations on Convex Bodies and Functions 29
We present the proof of Theorem 2.8 in the case .n = 2 from [48]. We call a
closed triangle .T = T (x, y) a support triangle of .K ∈ K2 with endpoints x and
y, if .x, y ∈ ∂K and T is bounded by support lines (that is, 1-dimensional support
hyperplanes) to K at x and y and the chord connecting x and y (Fig. 2.2).
A cap of a convex body K is the intersection of a closed half-space and K. We
set .δs (K, L) := V2 (KAL) for .K, L ∈ K2 , where .KAL := (K ∪ L)\(K ∩ L)
is the symmetric difference of K and L. Note that the symmetric difference metric
.δs induces on full-dimensional convex bodies the same topology as the Hausdorff
metric.
We require the following lemma, whose proof is omitted as it is very similar to
the proof of Proposition 2.3.
Lemma 2.9 If . Z : K2 → R is an upper semicontinuous, rotation invariant valua-
tion that vanishes on polytopes, then
. Z(C) = c o(C)
for every .K ∈ K2 .
30 M. Ludwig and F. Mussnig
for every .K ∈ K2 and for every sequence .Ek ∈ E2 such that .Ek → K. To prove the
statement of the proposition, assume on the contrary that there is .K ∈ K2 such that
for all sequences .Ek with .Ek ∈ E2 and .Ek → K. By (2.9), the affine surface area of
E is uniformly bounded for all convex bodies E with .δs (K, E) < 1, say. Therefore,
by (2.12), for every .ε > 0 small enough, there is .0 < δ < 1 such that
E
k
(k)
. V2 (Ti ) < δ (2.14)
i=1
||
k ( ||
k )
(k) (k)
Ek :=
. Ei ∪ K\ Ti .
i=1 i=1
Note that .Ek ∈ E2 and that (2.14) implies that .δs (K, Ek ) < δ.
Since .Z and .o vanish on polytopes, (2.13) implies that
E
k
(k)
E
k
( (k) (k) )
. Z(K∩Ti ) = Z(K) ≥ Z(Ek )+ε o(Ek ) = Z(Ek ∩Ti )+ε o(Ek ∩Ti ) .
i=1 i=1
2 Valuations on Convex Bodies and Functions 31
we have
Let .lk be the largest integer such that there are rotations .ψ1 , . . . , ψlk with the
property that .ψ1 (T̃ (k) ), . . . , ψlk (T̃ (k) ) are non-overlapping support triangles of .B 2 .
Since for a sector of .B 2 with an angle .2 α at the origin, the area of a support triangle
to .B 2 is .sin2 α tan α, we have
( ) ( ) ( ) ( )
π π π π
. sin
2
tan ≤ V2 (T̃ (k) ) ≤ sin2 tan . (2.17)
lk + 1 lk + 1 lk lk
||
lk ( ||
lk )
K̃k :=
. ψi (C̃ (k) ) ∪ B 2 \ ψi (T̃ (k) ) .
i=1 i=1
for k sufficiently large. Since .δs (K̃k , B 2 ) ≤ lk V2 (T̃ (k) ), it follows from (2.17) that
K̃k → B 2
. (2.19)
This is a contradiction since .ε > 0 and .o(B 2 ) > 0, which concludes the proof of
the proposition. u
n
32 M. Ludwig and F. Mussnig
Note that we can apply Proposition 2.10 with .Z = o and obtain that
for every .K ∈ K2 .
Proof of Theorem 2.8 for .n = 2 Let .Z : K2 → R be an upper semicontinuous,
translation and .SL(2) invariant valuation. By Theorem 2.5 and since upper semi-
continuous Cauchy functions are linear, there are .c0 , c2 ∈ R such that
. Z(P ) = c0 V0 (P ) + c2 V2 (P )
and note that .Z̃ is an upper semicontinuous, translation and .SL(2) invariant valuation
that vanishes on polytopes. For every .K ∈ K2 , there is a sequence of polytopes .Pk
with .Pk → K. Hence, the upper semicontinuity of .Z̃ implies that
which shows that .Z̃ is non-negative. Using Lemma 2.9 and the translation and .SL(2)
invariance of .Z̃, we see that
. Z̃(C) = c o(C)
for every cap C of a unit ellipse, where .c = Z(B 2 )/ o(B 2 ). Since .Z vanishes on
polytopes, it is a simple valuation, and it follows from (2.2) that
Z̃(E) = c o(E)
. (2.21)
for .E ∈ E2 . Proposition 2.10 and (2.20) now complete the proof of the theorem. u
n
For .K ∈ K2 , Blaschke [13] gave the following definition of affine surface area.
Choose subdivision points .x1(k) , . . . , xk(k) , xk+1
(k)
= x1(k) on .∂K and support triangles
1 , . . . , Tk such that .Tj = T (xj , xj +1 ). Define
(k) (k) (k) (k) (k)
.T
k /
E (k)
õ(K) := lim
.
3
8 V2 (Tj ) (2.22)
j =1
2 Valuations on Convex Bodies and Functions 33
(k)
. max V2 (Ti ) → 0
i=1,...,k
as .k → x. For smooth convex bodies in .K2 , Blaschke showed that this limit always
exists and that .õ(K) = o(K).
If we choose a further subdivision point .y ∈ ∂K in a support triangle .T (x, z)
of .K ∈ K2 , we obtain support triangles .T (x, y) and .T (y, z) and the following
elementary anti-triangle inequality holds
/ / /
.
3
8 V2 (T (x, z)) ≥ 3 8 V2 (T (x, y)) + 3 8 V2 (T (y, z))
E /
(cf. [13, p. 38] or [20]). This implies that . kj =1 3 8 V2 (Tj(k) ) decreases as the
subdivision is refined. Consequently, the limit in (2.22) exists and is independent
of the sequence of subdivisions chosen and
k /
E (k)
.õ(K) = inf 3
8 V2 (Tj )
j =1
where the infimum is taken over all subdivisions of .∂K. Thus .õ is well defined
on .K2 and Leichtweiß [46] proved that .õ(K) = o(K) for every .K ∈ K2 . This
is also a simple consequence of Theorem 2.8 for .n = 2. Indeed, .õ : K2 → R
is equi-affine invariant and vanishes on lower dimensional sets. As an infimum of
continuous functionals, .õ is upper semicontinuous. So we have only to show that
i ) → 0 as .k → x, we have for every line
(k)
.õ is a valuation. Since .maxi=1,...,k V2 (T
H,
where .H + and .H − are the closed halfspaces bounded by H . It is not difficult to see
that this implies that .õ is a valuation. Thus Theorem 2.8 for .n = 2 shows that
õ(K) = c o(K)
.
example, [17]). Nevertheless, we will restrict our attention to convex polytopes and
convex bodies in .Rn .
{ E
n }
S = x0 +
. ri xi : 1 ≥ r1 ≥ . . . ≥ rn ≥ 0 . (2.23)
i=1
E
n
x=
. t i pi
i=0
E E
with .ti ≥ 0 and . ni=0 ti = 1. Setting .ri = nj=i tj , we have
E
n
x = x0 +
. ri xi .
i=1
for .1 ≤ k ≤ n − 1, we have
||
n
( )
. S= (1 − t) S k + t S n−k
k=0
Qk (t) := (1 − t) S k + t S n−k ,
.
{ ( E
k
) ( E
k E
n
)
Qk (t) = (1 − t) x0 +
. ri xi + t x0 + xi + si xi :
i=1 i=1 i=k+1
}
1 ≥ r1 ≥ · · · ≥ rk ≥ 0, 1 ≥ sk+1 ≥ · · · ≥ sn ≥ 0
{ E
n }
= x0 + ti xi : 1 ≥ t1 ≥ · · · ≥ tk ≥ t ≥ tk+1 ≥ · · · ≥ tn ≥ 0 .
i=1
For .x ∈ S, this implies that .x ∈ Qk (t) for a suitable k. We have to show that the sets
.Qk (t) for .1 ≤ k ≤ n − 1 have pairwise disjoint interiors. If .x ∈ Qi (t) ∩ Qk (t) for
.i < k, then .ti+1 = · · · = tk = t and therefore .rj = 0 and .sj = 1 for .i+1 ≤ j ≤ k. It
36 M. Ludwig and F. Mussnig
follows that .x ∈ ∂Qi (t) and .x ∈ ∂Qj (t). This completes the proof of the statement.
u
n
We say that a simplex .<x0 ; x1 , . . . , xn > is orthogonal if the vectors .x1 , . . . , xn are
pairwise orthogonal. The following result is due to Hadwiger [39, Section 1.3.4].
Lemma 2.13 Let .z ∈ Rn be given. If .P ∈ Pn is n-dimensional, then there are
orthogonal simplices .S1 , . . . , Sm , .S1' , . . . , Sm
' , each with a vertex at z, such that
'
'
||
m ||
m
Pu
. Si ∼ Sj' .
i=1 j =1
Proof The statement is easy to prove for .n = 1. Assume that it is true in .P(E) for
every .(n − 1)-dimensional hyperplane E and every .zE ∈ E.
It suffices to prove the statement for an n-dimensional simplex S. Let F be one
of its facets whose affine hull E does not contain z. Let .zE be the closest point to z
in E. We use the induction assumption for polytopes in E with .zE and obtain that
there are .(n − 1)-dimensional simplices .F1 , . . . , Fk , F1' , . . . , Fk '' , each with a vertex
at .zE such that
'
||
k ||
k
Fu
. Fi ∼ Fj' .
i=1 j =1
Setting .Si := [z, Fi ] for .1 ≤ i ≤ k and .Sj' := [z, Fj' ] for .1 ≤ j ≤ k ' , we obtain the
statement for S. u
n
The question of whether every polytope in .Pn can be dissected into finitely many
orthogonal simplices is open. Hadwiger conjectured that it is possible, and his
conjecture has been proved for .n ≤ 5 (see, for example, [19]).
Proof Let S be an n-dimensional orthogonal simplex in .Rn and .0 < t < 1. In the
canonical simplex decomposition,
||
n
( )
S=
. (1 − t)S k + t S n−k , (2.24)
k=0
the simplices .S k and .S n−k are orthogonal and lie in orthogonal subspaces for each
1 ≤ k ≤ n−1. Hence .(1−t)S k +t S n−k is an orthogonal cylinder for .1 ≤ k ≤ n−1.
.
for .1 ≤ k ≤ n − 1 and we also see that .Z is simple. By (2.2), it now follows from
(2.24) that
Let .r, s > 0. Setting .α(r) := Z(r S̃) with .S = (r + s)S̃ and .t = r/(r + s), we obtain
( ||
m ) m'
( || )
. Z(t P ) + Z t Si = Z t Sj'
i=1 j =1
and
' '
E
m
( ') E m
( ) E
m E
m
'
. Z(t P ) = Z t Sj − Z t Si = t Z(Sj ) − t Z(Si ) = t Z(P ),
j =1 i=1 j =1 i=1
38 M. Ludwig and F. Mussnig
where (2.2) and (2.25) are used. Hence, .Z is homogeneous of degree 1 on polytopes.
Since .Z is continuous, this concludes the proof. u
n
For .1 ≤ l ≤ n, we say that .P ∈ Pn is a convex .l-cylinder if there are subspaces
.E1 , . . . , El of .R which are pairwise orthogonal and at least one-dimensional and
n
||
n ||
mP =
. τ (Cl )
l=1 τ ∈Tl,m
( )
for every integer .m ≥ 1, where . Tl,m is a set of at most . ml translations.
Proof It suffices to prove the statement for P an n-dimensional simplex S. Let
.S = <x0 ; x1 , . . . , xn >. For .i < j , define the simplices .Sij := <xi ; xi+1 , . . . , xj >. For
.0 < t < 1 and .1 ≤ k < n, set
(E
k−1 )
Qk (t) := (1 − t) S0k + t
. xi + Skn .
i=0
mS = m Q0 ( m1 ) u · · · u m Qn ( m1 ).
.
m Ql ( m1 ) = (m − 1) S0l + Sln
.
for .1 ≤ (j1) < · · · < jl = n. We use induction to show that each .Tl (j1 , . . . , jl )
appears . ml times in the decomposition. The statement is trivial for .m = 1. So,
let .m > 1. The polytope .Tl (j1 , . . . , jl ) appears when decomposing .(m − 1)S and
when decomposing .(m − 1)S0jl−1 + Sjl−1 n . By the induction assumption, it appears
(m−1) ( )
.
l times in the first case and . m−1
l−1 times in the second case, which proves the
2 Valuations on Convex Bodies and Functions 39
claim. The .l-cylinder .Cl is obtained as union of translates (with pairwise disjoint
interiors) of the convex .l-cylinders .Tl (j1 , . . . , jl ) for .1 ≤ j1 < · · · < jl = n. n
u
. Z = Z0 + · · · + Zn
. Z = Z0 + · · · + Zn
E
n E
n
. Zj (k P ) mj = Z(k m P ) = Zj (P ) (k m)j .
j =1 j =1
Therefore,
. Zj (k P ) = k j Zj (P ),
40 M. Ludwig and F. Mussnig
that is, .Zj is homogeneous of degree j with respect to multiplication with positive
integers. To show that .Zj is a valuation, it suffices to show that
. Zj (P u Q) = Zj (P ) + Zj (Q).
This follows using (2.26) for .P uQ, P and Q and comparing coefficients of .mj . n
u
Proof of Theorem 2.16 for Simple Valuations First, we show that for non-negative
λ ∈ Q,
.
E
n
. Z(λP ) = Zj (P ) λj .
j =1
E
n E
n
( p )j
. Z( pq P ) = Zj ( q1 P ) pj = Zj (P ) q .
j =1 j =1
So far, the valuations .Zj are only defined on .Pn . Note that the system of equations,
E
n
. Z(m P ) = Zj (P ) mj
j =1
E
n
. Zj (P ) = αij Z(i P )
i=1
K |→ Z̄(K[j ], K1 , . . . , Km−j )
.
for every .K, L ∈ Kn . The special case .m = 1 in Theorem 2.18 leads to the following
result.
Corollary 2.19 If . Z : Kn → R is a continuous, translation invariant valuation that
is homogeneous of degree 1, then .Z is Minkowski additive.
Theorem 2.16 allows to reduce questions on continuous and translation invariant
valuations to questions on such valuations with a given degree of homogeneity
.j ∈ {0, . . . , n}. It is easy to see that every continuous, translation invariant, and
. Z(P ) = c Vn (P )
for every .P ∈ Pn .
42 M. Ludwig and F. Mussnig
|| Z || := sup{| Z(K)| : K ∈ Kn , K ⊆ B n }.
.
The following rigid motion invariant, simple valuations are called Dehn invariants.
For .P ∈ P3 and .ζ : [0, x) → [0, x) a Cauchy function with .ζ (π ) = 0, set
E
. Dζ (P ) := V1 (E) ζ (αP (E))
where the sum is taken over all edges E of P and .αP (E) is the dihedral angle of P
at E. It is not difficult to see that .Dζ is a rigid motion invariant, simple valuation on
.P and that .Dζ vanishes on cubes. The regular tetrahedron T in .R has the dihedral
3 3
angle .α := arccos(1/3) at every edge, and the ratio .α/π is irrational. Hence there
are Cauchy functions with .ζ (α) /= 0. Hence .Dζ (T ) /= 0 for every regular simplex
T . Since .Dζ is a rigid motion invariant, simple valuation, it follows from (2.2) that
T is not equi-dissectable to any cube. This shows that Hilbert’s Third Problem has
a negative answer (for an introduction to Hilbert’s Third Problem and the dissection
theory of polytopes, see [16]). In general, .Dζ is far from being continuous.
A complete classification of rigid motion invariant and continuous valuations on
.K was obtained by Hadwiger [39, Section 6.1.10] in his celebrated classification
n
theorem.
Theorem 2.24 (Hadwiger) A functional .Z : Kn → R is a continuous, translation
and rotation invariant valuation if and only if there are constants .c0 , . . . , cn ∈ R
such that
for every .K ∈ Kn .
Hadwiger [39, Section 6.1.10] also obtained a complete classification of monotone
increasing, translation and rotation invariant valuations by showing that any such
valuation is a linear combination with non-negative coefficients of intrinsic vol-
umes. McMullen [64] showed that every monotone increasing, translation invariant
valuation is continuous. Hence the monotone version of Hadwiger’s theorem is a
simple consequence of Theorem 2.24.
We present a variation of Hadwiger’s original proof, which we got to know
through lecture notes by Ulrich Betke. The main step is to prove the following result
for simple valuations.
Proposition 2.25 A functional .Z : Kn → R is a continuous, translation and
rotation invariant, simple valuation if and only if there is a constant .c ∈ R such
44 M. Ludwig and F. Mussnig
that
. Z(K) = c Vn (K)
for every .K ∈ Kn .
We first show how to deduce the Hadwiger theorem from this proposition and then
describe its proof. An alternate proof of the Hadwiger theorem is due to Dan Klain
[40]. It can also be found in [41] and [72]. Klain also uses the simple argument in
the following subsection.
Proof of the Hadwiger Theorem Using Proposition 2.25 We use induction on the
dimension n and note that the statement is true for .n = 1 by Proposition 2.3. Assume
that the statement is true in dimension .n − 1 and let E be an .(n − 1)-dimensional
linear subspace of .Rn . The restriction of .Z to .K(E) is a continuous, translation
and rotation invariant valuation on .K(E). By the induction assumption, there are
constants .c0 , . . . , cn−1 ∈ R such that
E
n−1
. Z= cj Vj
j =0
E
n−1
Z̃ := Z −
. cj Vj
j =0
and note that .Z̃ is a continuous, translation and rotation invariant valuation on .Kn .
Moreover, .Z̃ is simple, as .Z̃ vanishes on .K(E) and hence, because of its translation
and rotation invariance, on all convex bodies contained in an affine hyperplane.
Using Proposition 2.25, we obtain that there is a constant .cn ∈ R such that
Z̃(K) = cn Vn (K)
.
and the width of K in direction u is .hK (u) + hK (−u). For the intrinsic volume .V1 ,
which is defined in (2.5), we have
2 Valuations on Convex Bodies and Functions 45
f
1
V1 (K) =
. (hK (u) + hK (−u)) dHn−1 (u),
2 κn−1 Sn−1
1( )
M=
. ϑ1 K + · · · + ϑm K .
m
We require the following result due to Hadwiger [39, Section 4.5.3].
Theorem 2.26 For each .K ∈ Kn , there exists a sequence of rotational Minkowski
means of K that converges to a centered ball.
We remark that
f
y |→
. hϑK (y) dϑ
SO(n)
. Z(K) = c V1 (K)
for every .K ∈ Kn .
Proof For .K ∈ Kn , Theorem 2.26 implies that there exists a sequence .(Kj ) of
rotational Minkowski means of K with .Kj → rB n , where .rB n is a centered ball of
radius r, where r depends on K. As every .Kj is of the form
1( )
. Kj = ϑ1 K + · · · + ϑm K
m
The first intrinsic volume, .V1 , is continuous, translation and rotation invariant, and
Minkowski additive. Hence we also have .V1 (K) = r V1 (B n ). Combined this gives
.Z(K) = c V1 (K) with .c = Z(B )/V1 (B ). u
n
n n
KE |→ Z(KE + KF )
.
for every .KE ∈ K(E). It follows from Lemma 2.1 that .c : K(F ) → R is a valuation,
which is easily seen to be simple, continuous, translation and rotation invariant.
Hence, by the induction assumption, there is .ck ∈ R such that
for every .KE ∈ K(E) and .KF ∈ K(F ). Since .Z is translation and rotation invariant,
(2.27) holds for convex bodies .KE and .KF in any orthogonal and complementary
subspaces E and F with .dim E = k. Evaluating on the unit cube, we obtain that
.c1 = · · · = cn−1 =: c.
Define .Z̃ : Kn → R by
Note that .Z̃ is a simple, continuous, translation and rotation invariant valuation
that vanishes on orthogonal cylinders. By Proposition 2.14, it is homogeneous of
degree 1, and by Corollary 2.19, it is Minkowski additive. Using Theorem 2.27, we
obtain that there is a constant .d ∈ R such that
Z̃(K) = d V1 (K)
.
For valuations invariant under the action of subgroups of the orthogonal group,
O(n), Alesker [1, 3] obtained the following result.
.
Theorem 2.28 (Alesker) For a compact subgroup G of .O(n), the linear space of
continuous, translation and G invariant valuations on .Kn is finite dimensional if
and only if G acts transitively on .Sn−1 .
As the classification of such subgroups G is known, it is a natural task (which
was already proposed in [1]) to find bases for spaces of continuous, translation
and G invariant valuations for all such subgroups (see [2, 10, 12] for some of the
contributions).
The following result is a special case of the principal kinematic formula, which is
due to Blaschke, Chern, Federer and Santaló (see [41, 73]). We use integration with
respect to the Haar measure on .SO(n) x Rn , and the normalization is chosen so that
on .SO(n) we have the Haar probability measure, and translations are identified with
.R with the standard Lebesgue measure.
n
E
n
. Z(K, L) = cj (L)Vj (K)
j =0
E
n
. Z(K, L) = cij Vi (K) Vj (L)
i,j =0
for every .K, L ∈ Kn . The constants .cij can be determined by evaluating this
formula for suitable convex bodies K and L. u
n
We will extend and generalize valuations from (subsets of) the space of convex
bodies to function spaces. Let .F (Rn ; R) denote the space of all real-valued functions
on .Rn .
2.6.1 Definition
One way to represent the set of convex bodies, .Kn , within .F (Rn ; R) is to assign to
each body .K ∈ Kn its characteristic function .χK ∈ F (Rn ; R), which is given by
{
1 for x ∈ K
χK (x) =
.
0 for x ∈
/ K.
Z̃(K) := Z(χK )
.
for every .K ∈ Kn .
We now ask which conditions .Z needs to satisfy so that .Z̃ is a valuation. By the
definition of .Z̃ we have
for every .K, L ∈ Kn such that also .K ∪ L ∈ Kn . Here, .f ∨ g and .f ∧ g denote the
pointwise maximum and minimum of .f, g ∈ F (Rn ; R), respectively. This motivates
the following definition.
Definition 2.30 Let .X ⊆ F (Rn ; R). A map .Z : X → R is a valuation if
We claim that .Z is a valuation. Let .f, g ∈ X. Since .Rn can be represented as the
disjoint union
Note that this valuation often plays the role of volume. For example, the Prékopa–
Leindler inequality is a functional version of the Brunn–Minkowski inequality,
where the usual n-dimensional volume on subsets of .Rn is replaced by the integral
of a function (see, for example, [36]).
Further valuations on suitable function spaces are given by the map .f |→ f (x̄)
for some fixed .x̄ ∈ Rn , the pth power of the .Lp norm, the moment matrix, the
Fisher information matrix or the LYZ body. We refer to [9, 50, 52, 77, 78, 80] for
more details.
Defining analytic analogs of geometric concepts is, of course, not a new problem
(see, for example, [7, Chapter 9] and [72, Sections 9.5 and 10.15]). This section
focuses on results where analogs of important valuations in geometry were found
on function spaces.
The survey [51] describes some of the first results on valuations on function
spaces. Among them are Tsang’s characterization of valuations on .Lp spaces and
.Lp stars [77, 78], as well as the characterization of the moment matrix [53],
Fisher information matrix [50], the LYZ body and its projection body [52]. In the
following, we will give a brief overview of some of the results not included in [51].
2 Valuations on Convex Bodies and Functions 51
Quasi-Concave Functions
{f ≥ t} := {x ∈ Rn : f (x) ≥ t},
.
are either empty or convex bodies for every .t > 0. A natural approach to extending
intrinsic volumes from convex bodies to quasi-concave functions is to integrate
intrinsic volumes of the level sets of a given function with respect to suitable
measures. The following result was proved in [22].
Theorem 2.31 A map .Z is a rigid motion invariant, continuous and increasing
valuation on the space of quasi-concave functions on .Rn if and only if there are
measures .νi ∈ Ni for .0 ≤ i ≤ n such that
n f
E
. Z(f ) = Vi ({f ≥ t}) dνi (t)
i=0 [0,x)
. Z(f ◦ φ −1 ) = Z(f )
Convex Functions
Recently, the first results on valuations on various spaces of convex functions were
obtained. Valuations on the space of coercive convex functions, .Convcoe (Rn ) (see
Sect. 2.7.1 for the definition), were first classified in [21]. A characterization of
analogs of the Euler characteristic and the n-dimensional volume as .SL(n) invariant
valuations was established in [25], and we will prove a special case of this result in
the next section. See also [66, 67].
52 M. Ludwig and F. Mussnig
. Z(u) = D[ζ ◦ u]
f x
h([ζ ◦ u], y) :=
. h({ζ ◦ u ≥ t}, y) dt,
0
and .DK := K + (−K) denotes the difference body of the convex body .K ∈ Kn .
For the definition of continuity of operators on .Convcoe (Rn ), we refer to Sect. 2.7.1.
There are various results on valuations on spaces of real-valued functions on the unit
sphere, .Sn−1 . Such functions are particularly interesting since they appear as radial
functions of star bodies or more general star-shaped sets. Results on valuations on
the set, .C(Sn−1 )+ , of positive, continuous functions on .Sn−1 are equivalent to results
on valuations on star bodies. The following result was established in [75, 79] as a
result for valuations on star bodies.
Theorem 2.33 A map .Z : C(Sn−1 )+ → R is a continuous and rotation invariant
valuation if and only if there is a continuous function .ζ : [0, x) → R such that
f
. Z(f ) = ζ (f (x)) dHn−1 (x)
Sn−1
. Z(f ◦ φ −1 ) = Z(f )
Lp Spaces
Definable Functions
We denote by
Observe that a function .u ∈ Conv(Rn ) is coercive if and only if there exist .a > 0
and .b ∈ R such that
.u(x) ≥ a|x| + b
for every .x ∈ Rn .
If .u ∈ Convcoe (Rn ), then its sublevel sets
{u ≤ t} := {x ∈ Rn : u(x) ≤ t}
.
are compact, convex subsets of .Rn for every .t ∈ R. We have .{u ≤ t} = ∅ for every
.t < minx∈Rn u(x) and .{u ≤ t} ∈ K for every .t ≥ minx∈Rn u(x). Note that u attains
n
its minimum since it is lower semicontinuous and coercive. We will see that many
operators can be generalized to .Convcoe (Rn ) using sublevel sets.
It is easy to see that
{u ∨ v ≤ t} = {u ≤ t} ∩ {v ≤ t} and
. {u ∧ v ≤ t} = {u ≤ t} ∪ {v ≤ t} (2.29)
for every .u, v ∈ Convcoe (Rn ) such that .u ∧ v ∈ Convcoe (Rn ) and .t ∈ R. Also,
observe that while the domain of u,
{uk ≤ t0 } = ∅
.
uk (x) := I(
.
) (x)
1− k1 B n
uk (x) ≥ a|x| + b
. and u(x) ≥ a|x| + b (2.30)
t ∨ +x = +x and
. t ∧ +x = t
. Z(u ◦ τ −1 ) = Z(u)
. Z(u ◦ ϑ −1 ) = Z(u)
{u ◦ τ −1 ≤ t} = τ {u ≤ t} and
. {u ◦ ϑ −1 ≤ t} = ϑ{u ≤ t} (2.31)
u |→ e− minx∈Rn u(x)
. (2.32)
and
f
. u |→ e−u(x) dx (2.33)
Rn
{u ≤ t} ⊆ {x : a|x| ≤ t} = at B n
.
2 Valuations on Convex Bodies and Functions 57
In the general case there exist .a > 0 and .b ∈ R such that .u(x) ≥ a|x| + b for
x ∈ Rn and thus, since .u(x) − b ≥ a|x| for .x ∈ Rn , we have
.
f f
n!
0≤
. e−u(x) dx = e−b e−(u(x)−b) dx ≤ κn e−b . (2.35)
Rn Rn an
The fact that (2.33) defines a valuation follows from (2.34) and (2.29) together
with the fact that the n-dimensional volume, .Vn , is a valuation on convex bodies.
Similarly, using (2.31), it is easy to obtain .SL(n) and translation invariance. The
valuation property can be proved analogous to (2.28).
It remains to show continuity. Let .uk be a sequence in .Convcoe (Rn ) that epi-
converges to .u ∈ Convcoe (Rn ). By Lemma 2.35, there exist .a > 0 and .b ∈ R such
that (2.30) holds. Thus, similar to (2.35), we obtain
f
n!
0≤
. e−uk (x) dx ≤ κn e−b
Rn an
for every .u ∈ Convcoe (Rn ). Thus, it is easy to see that the same arguments as above
can be applied. u
n
58 M. Ludwig and F. Mussnig
The proof above demonstrates the simple strategy for finding a functional analog
of an operator .Z : Kn → R by considering the map
f +x
u |→
. Z({u ≤ t}) e−t dt
−x
on .Convcoe (Rn ), where we set .Z(∅) := 0. Indeed, in many cases, this will define
an operator on . Convcoe (Rn ) with similar properties as the original operator .Z.
More generally, one can often replace .e−t dt by a suitable measure .dμ(t). For some
examples, see [15, 65].
for every .u ∈ Convcoe (Rn ) and .t ∈ R, if and only if there are constants .c0 , cn ∈ R
such that
f
− minx∈Rn u(x)
. Z(u) = c0 e + cn e−u(x) dx (2.37)
Rn
The properties of valuations on .LCcoe (Rn ) are defined analogously to the corre-
sponding properties for valuations on .Convcoe (Rn ). The following result, which is
equivalent to Theorem 2.37, is a consequence of [68, Theorem 4].
2 Valuations on Convex Bodies and Functions 59
. Z(sf ) = s Z(f )
for every .f ∈ LCcoe (Rn ) and .s > 0, if and only if there are constants .c0 , cn ∈ R
such that
f
. Z(f ) = c0 max f (x) + cn f (x) dx
n
x∈R Rn
{gK ≤ t} = tK
.
for every .K, L ∈ Kno such that .K ∪ L ∈ Kno and convergence of .Kj to K on .Kno
implies
gKj → gK ,
.
Z̃(K) := Z(gK )
.
60 M. Ludwig and F. Mussnig
such that
for every .K ∈ Kno . Note that this also implies that .Z̃ is translation invariant, which
is not evident from its definition. Next, observe that
and
f f x
1 e−t
. e−(gK (x)+t) dx = Vn (sK)e−s ds = e−t Vn (K)
n! Rn n! 0
for every .K ∈ Kno and .t ∈ R, where we used a similar computation as in (2.34). The
result now follows by setting .c0 := c̃0 and .cn := c̃n /n!, since
. Z1 (u) = Z2 (u)
Proof (for .n = 1) Since .Z1 and .Z2 are continuous, it is enough to consider the case
where .u ∈ Convcoe (Rn ) is such that
A
m
u=
. wi
i=1
with affine functions .wi : Rn → R. In addition, we may also assume that the
graph of u has no edges parallel to the coordinate axis. We will use induction on the
number k of vertices of the graph of u.
If .k = 1, then u must be of the form
u(x) = gP (x − x0 ) + t
.
for some polytope .P ∈ Pno , .t ∈ R and .x0 ∈ R. Thus, it follows from translation
invariance and (2.38) that .Z1 (u) = Z2 (u).
Assume now that the statement is true for .k − 1 and let the graph of u have k
vertices. Denote by .(x̄, t¯) ∈ R2 one of the highest vertices. It is easy to see that we
can find a polytope .P̄ ∈ Pno such that
where .ū ∈ Convcoe (Rn ) is given by .ū(x) := gP̄ (x − x̄) + t¯ for .x ∈ Rn . See Fig. 2.4.
Since the graph of .u ∨ ū has only one vertex, it follows from our induction
assumption and the valuation property that
By the first part of the proof, this is a continuous and translation invariant valuation
such that .Z̄(gP + t) = Z(gP + t) for every .P ∈ Pno and .t ∈ R. The result now
follows from Lemma 2.40. u
n
2.7.4 Homogeneity
{λ o u ≤ t} = λ{u ≤ t}.
.
. Z(λ o u) = λp Z(u)
Obviously, .Convsc (Rn ) is a subspace of .Convcoe (Rn ). Note that for differentiable
.u ∈ Convsc (R ), the property
n
u(x)
. lim = +x
|x|→+x |x|
The space of super-coercive convex functions is, in the following way, closely
connected to the space of finite-valued convex functions,
Conv(Rn ; R) := {v : Rn → R : v is convex}.
.
This relation allows us to translate results for valuations on .Convsc (Rn ) easily to
results on .Conv(Rn ; R) and vice versa.
A valuation .Z : Convsc (Rn ) → R is epi-translation invariant if it is vertically
translation invariant in addition to having the usual translation invariance, that is, if
. Z(u ◦ τ −1 + α) = Z(u)
for every .u ∈ Convsc (Rn ), translation .τ on .Rn and .α ∈ R. Note that this means that
Z is invariant under translations of the epi-graph of u in .Rn+1 , where the epi-graph
.
of u is given by
denote their infimal convolution at .x ∈ Rn . Note that also .u O v ∈ Convsc (Rn ) and
that
where the addition on the right side is Minkowski addition of closed, convex sets in
Rn+1 . The infimal convolution is also called epi-addition. It naturally induces the
.
following operation. For .λ > 0 and .u ∈ Convsc (Rn ), define .λ u ∈ Convsc (Rn ) as
(x )
(λ u)(x) := λ u
.
λ
. epi(λ u) = λ epi u
. k u = 'u O ·''
· · O u'
k times
. Z(λ u) = λp Z(u)
for every .u ∈ Convsc (Rn ) and .λ > 0. We will present two examples of continuous,
epi-homogeneous, and epi-translation invariant valuations on .Convsc (Rn ).
First, it is easy to see that the constant map .u |→ c for .c ∈ R defines a continuous,
epi-translation invariant valuation that is epi-homogeneous of degree 0. In fact, it is
the only valuation with these properties (see [28, Theorem 25]).
Next, let .ζ ∈ Cc (Rn ), that is, .ζ is continuous with compact support. Consider
the map
f
. Z(u) = ζ (∇u(x)) dx (2.42)
dom u
for .u ∈ Convsc (Rn ). Because of (2.40), it is at least plausible that .Z(u) is well-
defined and finite for every .u ∈ Convsc (Rn ). On the other hand, the example .u(x) :=
|x| shows that .Z defined by (2.42) is not a well-defined (finite) map on the larger
space . Convcoe (Rn ).
We will state the following result from [28, Proposition 20] without proof.
66 M. Ludwig and F. Mussnig
for .v ∈ Conv(Rn ; R). Here, .MA(v; ·), the Monge–Ampère measure of v, is a Radon
measure on .Rn , which can be defined as a continuous extension of the measure
.det(D v(x)) dx from .C (R ) to . Conv(R ; R).
2 2 n n
. Z = Z0 + · · · + Zn
Similar to the proof of Theorem 2.37, we first show that the result is true on a
restricted set of functions and then use a reduction argument. For .y ∈ Rn , define
.ly : R → R by .ly (x) := <y, x>.
n
for every .K ∈ Kn and .x, x0 , y ∈ Rn . In other words, the epi-graph of .ly + IK+x0 is
a translate of the epi-graph of .ly + IK . Thus, by the epi-translation invariance of .Z,
we now obtain that
. Zi (λ u) = λi Z(u)
E
n
.Z̃y (K) = Z̃y,i (K)
i=0
E
n
. Z(λ (ly + IK )) = Z(ly + IλK ) = Z̃y (λK) = λi Z̃y,i (K)
i=0
⎛ ⎞ ⎛ 0 ⎞⎛ ⎞
Z(0 (ly + IK )) 0 ··· 0n Z̃y,0 (K)
⎜ . ⎟ ⎜ . .. ⎟ ⎜ .. ⎟
.⎝ .. ⎠ = ⎝ .. . . . . ⎠⎝ . ⎠
Z(n (ly + IK )) n0 · · · nn Z̃y,n (K)
E
n
Z̃y,i (K) =
. αij Z(j (ly + IK ))
j =0
E
n
. Zi (u) := αij Z(j u)
j =0
for every .0 ≤ i ≤ n. It easily follows from the properties of .Z that also the
functionals .Zi for .0 ≤ i ≤ n are continuous and epi-translation invariant valuations.
We now have
E
n
. Zi (ly + IK ) = αij Z(j (ly + IK )) = Z̃y,i (K)
j =0
and thus
E
n E
n
. Z(ly + IK ) = Z̃y (K) = Z̃y,i (K) = Zi (ly + IK )
i=0 i=0
. Z(ly + IP ) = 0 (2.44)
A
m
u=
. (wi + IPi ), (2.45)
i=1
are convex (for a more detailed discussion of such a partition, we refer to [21,
Section 7.1]). Let .u1 , u2 ∈ Convsc (Rn ) be defined as
A A
u1 :=
. (wi + IPi ) and u2 := (wi + IPi ).
i∈I1 i∈I2
70 M. Ludwig and F. Mussnig
A
k
ū =
. (w̄i + IP̄i ).
i=1
Using the induction assumption again, we see that also .Z(ū) = 0. Thus, by the
valuation property of .Z,
we have
E
.Z̄(u1 ∧ · · · ∧ um ) = (−1)|I |−1 Z̄(uI ),
∅/=I ⊂{1,...,m}
V
where .uI := i∈I ui . Thus, in order to show that .Z vanishes on functions of the
form (2.45), it suffices to show that
(v )
. Z (wi + IPi ) = 0
i∈I
V
for every .∅ /= I ⊂ {1, . . . , m}. Since every such function . i∈I (wi + IPi ) is again
an affine function restricted to a polytope, the statement follows from (2.44) and the
vertical translation invariance of .Z. u
n
We now have all ingredients to prove the homogeneous decomposition theorem.
Proof of Theorem 2.42 Let the valuations .Z0 , . . . , Zn : Convsc (Rn ) → R be given
by Lemma 2.44 and define .Z̄ : Convsc (Rn ) → R as
E
n
Z̄(u) := Z(u) −
. Zi (u).
i=0
Z̄(ly + IP ) = 0
.
2 Valuations on Convex Bodies and Functions 71
E
n
. Z(u) = Zi (u)
i=0
for .u ∈ Convsc (Rn ). Note that .Z̄λ,i is a valuation on .Convsc (Rn ). Using the same
arguments as above, we obtain that .Z̄λ,i ≡ 0, which shows that for each .0 ≤ i ≤ n,
the valuation .Zi is epi-homogeneous of degree i. u
n
With an approach similar to that for Theorem 2.18, we obtain the following result
by considering .u |→ Zi (u O ū) for fixed .ū ∈ Convsc (Rn ).
Theorem 2.46 Let .1 ≤ m ≤ n. If .Z : Convsc (Rn ) → R is a continuous, epi-
translation invariant valuation that is epi-homogeneous of degree m, then there is a
symmetric function .Z̄ : (Convsc (Rn ))m → R such that
E ( )
m
. Z(λ1 u1 O · · · O λk uk ) = λi1 · · · λikk Z̄(u1 [i1 ], . . . , uk [ik ])
i 1 · · · ik 1
i1 ,...,ik ∈{0,...,m}
i1 +···+ik =m
u |→ Z̄(u[j ], u1 , . . . , um−j )
.
for every .u, v ∈ Convsc (Rn ). The special case .m = 1 in the previous theorem leads
to the following result, which is a functional version of Corollary 2.19.
Corollary 2.47 If . Z : Convsc (Rn ) → R is a continuous, epi-translation invariant
valuation that is epi-homogeneous of degree 1, then .Z is epi-additive.
72 M. Ludwig and F. Mussnig
In this section, we will show that the valuations described in Lemma 2.41 are indeed
the only continuous and epi-translation invariant valuations on .Convsc (Rn ) which
are epi-homogeneous of degree n. The following result was established in [28,
Theorem 2].
Theorem 2.48 A map .Z : Convsc (Rn ) → R is a continuous and epi-translation
invariant valuation that is epi-homogeneous of degree n, if and only if there exists
.ζ ∈ Cc (R ) such that
n
f
. Z(u) = ζ (∇u(x)) dx (2.46)
dom u
By Lemma 2.41 and our assumptions on .Z, the operator .Z̄ is a continuous and epi-
translation invariant valuation. Furthermore, it follows from (2.47) that
f
.Z̄(ly + IK ) = Zy (K) − ζ (y) dx = 0
K
Z̄(u) = 0
.
can be seen as a functional analog of the n-dimensional volume on .Convsc (Rn ). This
interpretation is further supported by Theorem 2.48, which (up to the assumption of
continuity) is a functional version of Theorem 2.20. In the following, we restrict to
the rotation invariant case, where we say that a valuation .Z : Convsc (Rn ) → R is
rotation invariant if
. Z(u ◦ ϑ −1 ) = Z(u)
E
n
Vn,α (u O r IB n ) =
. r n−j κn−j Vj,α (u) (2.51)
j =0
for every .u ∈ Convsc (Rn ) and .r ≥ 0. Observe that (2.51) corresponds to the
classical Steiner formula (2.5) where we have replaced the n-dimensional volume
with .Vn,α and where now .IB n plays the role of the unit ball.
In many ways, the functionals .Vj,α behave like the classical intrinsic volumes.
First, it follows from the rotation invariance of .Vn,α and the radial symmetry of .IB n
that also .Vj,α is rotation invariant for every .0 ≤ j ≤ n − 1. Next, since
IK O r IB n = IK+rB n ,
.
for every .0 ≤ j ≤ n and .K ∈ Kn . Last but not least, the functionals .Vj,α are
characterized by a Hadwiger-type theorem. The version that is stated here follows
from [26, Theorem 1.3] and [30, Theorem 1.4].
Theorem 2.49 Let .n ≥ 2. A functional .Z : Convsc (Rn ) → R is a continuous,
epi-translation and rotation invariant valuation if and only if there are functions
.α0 , . . . , αn ∈ Cc ([0, x)) such that
Theorem 2.49 is a functional analog of the Hadwiger theorem, Theorem 2.24, and
shows that the valuations .Vj,α clearly play the role of the intrinsic volumes on
n
.Convsc (R ).
In [26], a different approach and notation are used. The functionals there take the
form
f
.u |→ ζ (|∇u(x)|)[D2 u(x)]n−j dx (2.52)
Rn
Acknowledgments M. Ludwig was supported, in part, by the Austrian Science Fund (FWF):
P 34446, and F. Mussnig was supported by the Austrian Science Fund (FWF): J 4490.
References
11. A. Bernig, Algebraic integral geometry, in Global Differential Geometry. Springer Proceedings
in Mathematics, vol. 17, pp. 107–145 (Springer, Heidelberg, 2012)
12. A. Bernig, J.H.G. Fu, Hermitian integral geometry. Ann. Math. (2) 173, 907–945 (2011)
13. W. Blaschke, Differentialgeometrie II (Springer, Berlin, 1923)
14. W. Blaschke, Vorlesungen über Integralgeometrie. H. 2 (Teubner, Berlin, 1937)
15. S.G. Bobkov, A. Colesanti, I. Fragalà, Quermassintegrals of quasi-concave functions and
generalized Prékopa-Leindler inequalities. Manuscripta Math. 143, 131–169 (2014)
16. V. Boltianskii, Hilbert’s Third Problem (Wiley, New York, 1978)
17. K.J. Böröczky, M. Ludwig, Valuations on lattice polytopes, in Tensor Valuations and Their
Applications in Stochastic Geometry and Imaging. Lecture Notes in Mathematics, vol. 2177
(Springer, Cham, 2017), pp. 213–234
18. A. Braides, r-Convergence for Beginners. Oxford Lecture Series in Mathematics and its
Applications, vol. 22 (Oxford University Press, Oxford, 2002)
19. J. Brandts, S. Korotov, M. Křížek, J. Šolc, On nonobtuse simplicial partitions. SIAM Rev. 51,
317–335 (2009)
20. E. Calabi, P. Olver, A. Tannenbaum, Affine geometry, curve flows, and invariant numerical
approximation. Adv. Math. 124, 154–196 (1996)
21. L. Cavallina, A. Colesanti, Monotone valuations on the space of convex functions. Anal. Geom.
Metr. Spaces 3, 167–211 (2015)
22. A. Colesanti, N. Lombardi, Valuations on the space of quasi-concave functions, in Geometric
Aspects of Functional Analysis. Lecture Notes in Math., vol. 2169 (Springer, Cham, 2017), pp.
71–105
23. A. Colesanti, N. Lombardi, L. Parapatits, Translation invariant valuations on quasi-concave
functions. Studia Math. 243, 79–99 (2018)
24. A. Colesanti, M. Ludwig, F. Mussnig, Minkowski valuations on convex functions. Calc. Var.
Partial Differ. Equ. 56, Paper No. 162 (2017)
25. A. Colesanti, M. Ludwig, F. Mussnig, Valuations on convex functions. Int. Math. Res. Not.
IMRN, 2384–2410 (2019)
26. A. Colesanti, M. Ludwig, F. Mussnig, The Hadwiger theorem on convex functions, I (2020).
arXiv:2009.03702
27. A. Colesanti, M. Ludwig, F. Mussnig, Hessian valuations. Indiana Univ. Math. J. 69, 1275–
1315 (2020)
28. A. Colesanti, M. Ludwig, F. Mussnig, A homogeneous decomposition theorem for valuations
on convex functions. J. Funct. Anal. 279, 108573 (2020)
29. A. Colesanti, M. Ludwig, F. Mussnig, The Hadwiger theorem on convex functions, II: Cauchy–
Kubota formulas (2021). Amer. J. Math. (in press)
30. A. Colesanti, M. Ludwig, F. Mussnig, The Hadwiger theorem on convex functions, III: Steiner
formulas and mixed Monge–Ampère measures. Calc. Var. Partial Differ. Equ. 61, Paper No.
181 (2022)
31. A. Colesanti, M. Ludwig, F. Mussnig, The Hadwiger theorem on convex functions, IV: the
Klain approach (2022). Adv. Math. 413, Paper No. 108832 (2023)
32. A. Colesanti, D. Pagnini, P. Tradacete, I. Villanueva, A class of invariant valuations on
Lip(S n−1 ). Adv. Math. 366, Paper No. 107069 (2020)
33. A. Colesanti, D. Pagnini, P. Tradacete, I. Villanueva, Continuous valuations on the space of
Lipschitz functions on the sphere. J. Funct. Anal. 280, Paper No. 108873 (2021)
34. G. Dal Maso, An Introduction to r-Convergence. Progress in Nonlinear Differential Equations
and Their Applications, vol. 8 (Birkhäuser Boston, Inc., Boston, 1993)
35. M. Dehn, Ueber den Rauminhalt. Math. Ann. 55, 465–478 (1901)
36. R.J. Gardner, The Brunn-Minkowski inequality. Bull. Amer. Math. Soc. (N.S.) 39, 355–405
(2002)
37. P. Goodey, W. Weil, Distributions and valuations. Proc. Lond. Math. Soc. (3) 49, 504–516
(1984)
38. C. Haberl, L. Parapatits, The centro-affine Hadwiger theorem. J. Amer. Math. Soc. 27, 685–705
(2014)
2 Valuations on Convex Bodies and Functions 77
39. H. Hadwiger, Vorlesungen über Inhalt, Oberfläche und Isoperimetrie (German) (Springer,
Berlin, 1957)
40. D.A. Klain, A short proof of Hadwiger’s characterization theorem. Mathematika 42, 329–339
(1995)
41. D.A. Klain, G.C. Rota, Introduction to Geometric Probability (Cambridge University Press,
Cambridge, 1997)
42. J. Knoerr, The support of dually epi-translation invariant valuations on convex functions.
J. Funct. Anal. 281, Paper No. 109059 (2021)
43. J. Knoerr, Smooth valuations on convex functions. J. Differential Geom. (in press)
44. K. Kusejko, L. Parapatits, A valuation-theoretic approach to translative-equidecomposability.
Adv. Math. 297, 174–195 (2016)
45. K. Leichtweiß, Zur Affinoberfläche konvexer Körper. Manuscripta Math. 56, 429–464 (1986)
46. K. Leichtweiß, On the affine rectification of convex curves. Beiträge Algebra Geom. 40, 185–
193 (1999)
47. J. Li, D. Ma, Laplace transforms and valuations. J. Funct. Anal. 272, 738–758 (2017)
48. M. Ludwig, A characterization of affine length and asymptotic approximation of convex discs.
Abh. Math. Semin. Univ. Hamb. 69, 75–88 (1999)
49. M. Ludwig, Valuations on polytopes containing the origin in their interiors. Adv. Math. 170,
239–256 (2002)
50. M. Ludwig, Fisher information and matrix-valued valuations. Adv. Math. 226, 2700–2711
(2011)
51. M. Ludwig, Valuations on function spaces. Adv. Geom. 11, 745–756 (2011)
52. M. Ludwig, Valuations on Sobolev spaces. Amer. J. Math. 134, 827–842 (2012)
53. M. Ludwig, Covariance matrices and valuations. Adv. Appl. Math. 51, 359–366 (2013)
54. M. Ludwig, Geometric valuation theory, in European Congress of Mathematics, ed. by A.
Hujdurović, K. Kutnar, D. Marušič, Š. Miklavič, T. Pisanski, P. Šparl (ESM Press, 2023), pp.
93–123
55. M. Ludwig, M. Reitzner, A characterization of affine surface area. Adv. Math. 147, 138–172
(1999)
56. M. Ludwig, M. Reitzner, A classification of SL(n) invariant valuations. Ann. Math. (2) 172,
1219–1267 (2010)
57. M. Ludwig, M. Reitzner, SL(n) invariant valuations on polytopes. Discrete Comput. Geom.
57, 571–581 (2017)
58. E. Lutwak, Extended affine surface area. Adv. Math. 85, 39–68 (1991)
59. E. Lutwak, The Brunn-Minkowski-Firey theory. II. Affine and geominimal surface areas. Adv.
Math. 118, 244–294 (1996)
60. D. Ma, Real-valued valuations on Sobolev spaces. Sci. China Math. 59, 921–934 (2016)
61. P. McMullen, Valuations and Euler-type relations on certain classes of convex polytopes. Proc.
Lond. Math. Soc. (3) 35, 113–135 (1977)
62. P. McMullen, Continuous translation-invariant valuations on the space of compact convex sets.
Arch. Math. 55, 377–384 (1980)
63. P. McMullen, Weakly continuous valuations on convex polytopes. Arch. Math. (Basel) 41,
555–564 (1983)
64. P. McMullen, Monotone translation invariant valuations on convex bodies. Arch. Math. (Basel)
55, 595–598 (1990)
65. V. Milman, L. Rotem, Mixed integrals and related inequalities. J. Funct. Anal. 264, 570–604
(2013)
66. F. Mussnig, Volume, polar volume and Euler characteristic for convex functions. Adv. Math.
344, 340–373 (2019)
67. F. Mussnig, SL(n) invariant valuations on super-coercive convex functions. Canad. J. Math.
73, 108–130 (2021)
68. F. Mussnig, Valuations on log-concave functions. J. Geom. Anal. 31, 6427–6451 (2021)
69. M. Ober, Lp -Minkowski valuations on Lq -spaces. J. Math. Anal. Appl. 414, 68–87 (2014)
78 M. Ludwig and F. Mussnig
70. R.T. Rockafellar, R.J.B. Wets, Variational Analysis. Grundlehren der Mathematischen Wissen-
schaften, vol. 317 (Springer, Berlin, 1998)
71. R. Schneider, Simple valuations on convex bodies. Mathematika 43, 32–39 (1996)
72. R. Schneider, Convex Bodies: The Brunn-Minkowski Theory. Encyclopedia of Mathematics
and Its Applications, vol. 151, second expanded edn. (Cambridge University Press, Cambridge,
2014)
73. R. Schneider, W. Weil, Stochastic and Integral Geometry. Probability and Its Applications
(New York) (Springer, Berlin, 2008)
74. C. Schütt, E. Werner, The convex floating body. Math. Scand. 66, 275–290 (1990)
75. P. Tradacete, I. Villanueva, Radial continuous valuations on star bodies. J. Math. Anal. Appl.
454, 995–1018 (2017)
76. P. Tradacete, I. Villanueva, Continuity and representation of valuations on star bodies. Adv.
Math. 329, 361–391 (2018)
77. A. Tsang, Valuations on Lp spaces. Int. Math. Res. Not. 20, 3993–4023 (2010)
78. A. Tsang, Minkowski valuations on Lp -spaces. Trans. Amer. Math. Soc. 364, 6159–6186
(2012)
79. I. Villanueva, Radial continuous rotation invariant valuations on star bodies. Adv. Math. 291,
961–981 (2016)
80. T. Wang, Semi-valuations on BV(Rn ). Indiana Univ. Math. J. 63, 1447–1465 (2014)
Chapter 3
Geometric and Functional Inequalities
Abstract In this chapter, some of the fundamental inequalities for the basic
geometric functionals on convex bodies are described. For some of these, functional
analogs and extensions have been obtained. Some of these are motivated by the
calculus of variations
3.1 Introduction
A. Colesanti
Dipartimento di Matematica e Informatica “U. Dini”, Università di Firenze, Firenze, Italy
e-mail: andrea.colesanti@unifi.it
D. Hug (O)
Karlsruhe Institute of Technology (KIT), Karlsruhe, Germany
e-mail: [email protected]
.Lp Brunn–Minkowski theory provided the ground for new formulations of existing
inequalities. On the other hand, as it has been found out, some functionals coming
from different areas satisfy inequalities very similar to those valid e.g. for the
volume, in convex geometry. Examples are provided by the Brunn–Minkowski type
inequalities that have been established for some classic functionals in the calculus
of variations.
Moreover, in recent times, for several of the main inequalities in convex geom-
etry, an analytic counterpart has been found. For instance, the Prékopa–Leindler
inequality can be seen as a functional form of the Brunn–Minkowski inequality.
As a second example, we mention the functional forms of the Blaschke–Santaló
inequality and its converse (we refer to Chap. 4 for this topic).
The goal of this chapter is to provide a general overview of this area, selecting
some specific and representative examples from the wide landscape of geometric
and functional inequalities in convex geometry.
Our starting point will be an elementary construction of volume, surface area and
mixed volumes of convex bodies, as well as of the class of mixed area measures,
contained in the first section.
Then we present the two most emblematic inequalities in convex geometry: the
Brunn–Minkowski and the Aleksandrov–Fenchel inequalities. This part includes
various proofs of these inequalities, applications, and the general Brunn–Minkowski
and Minkowski inequalities.
In the subsequent section we illustrate the basics of the so-called .Lp Brunn–
Minkowski theory, and we present a selection of results coming from this area.
In particular, we focus on the .Lp version of the Brunn–Minkowski inequality, for
.p ≥ 1, and on the conjecture concerning the log-Brunn–Minkowski inequality,
In this section, we give a more detailed introduction to some of the basic functionals
and measures which have proved to be key objects in convex geometry. We start by
constructing volume and surface area of convex bodies by using induction over the
dimension of the space. This can be done in an elementary way for polytopes, the
extension to general convex bodies is achieved by approximation of general convex
bodies by polytopes (from inside and outside). The approach to mixed volumes and
mixed area measures follows a similar path (in principle) but requires more refined
arguments.
We recall from the introductory chapter that by a convex body in .Rn we mean a
non-empty compact convex subset of .Rn . The class of all convex bodies in .Rn is
denoted by .Kn . The topology on .Kn , induced by the Hausdorff metric .δ(·, ·), allows
us to introduce and study geometric functionals on convex bodies by first defining
them for a special subclass, for example the class .Pn of polytopes or the class of
convex bodies with sufficiently smooth boundaries. Such an approach requires that
the geometric functionals under consideration have a continuity or monotonicity
property and also that the class .Pn of polytopes is dense in .Kn . The following result
ensures that the latter property is indeed available.
Theorem 3.1 Let .K ∈ Kn and .ε > 0.
(a) There exists a polytope .P ∈ Pn with .P ⊂ K and .δ(K, P ) ≤ ε.
(b) There exists a polytope .P ∈ Pn with .K ⊂ P and .δ(K, P ) ≤ ε.
(c) If . o ∈ relint K, then there exists a convex polytope .P ∈ Pn which satisfies
- ∈ Pn with .P
.P ⊂ K ⊂ (1 + ε)P . There is even a convex polytope .P - ⊂ relint K
-
and which satisfies .K ⊂ relint((1 + ε)P ).
For the proof of (a), one can use a simple compactness argument to see that .∂K
(the boundary of K) can be covered by finitely many Euclidean balls of radius .ε
with center in .∂K. The convex hull of these centers is a suitable choice of a convex
polytope P , as required in (a). Proofs of the remaining assertions can be found in
[43, Theorem 3.5].
Sometimes better or more specific approximation results are required. Examples
of such cases are:
. Approximation by combinatorially equivalent (simple and strongly isomorphic)
polytopes (see [59, Theorem 2.4.15]).
. Approximation by smooth bodies (see [59, Section 3.4]).
82 A. Colesanti and D. Hug
. Economic approximation (by polytopes having few vertices or few facets; see,
e.g., [13, 26, 56, 57, 63]).
. Simultaneous approximation with additional features (for instance, preserving
the convexity of the union; see [43, Exercise 3.1.15]).
Volume and surface area of convex bodies can be defined via Lebesgue measure and
Hausdorff measure of the appropriate dimension, respectively. Here we indicate a
more elementary approach, using polytopal approximation and induction over the
dimension. This also serves as a preparation for the introduction of mixed volumes.
We start with some preparatory remarks:
. A support set of a convex body K is a non-empty intersection of K with a
supporting hyperplane, i.e. a hyperplane H which has non-empty intersection
with K and is such that K is contained in one of the two half-spaces determined
by H .
. The support set .K(u) := {x ∈ K : <x, u> = hK (u)}, .u ∈ Sn−1 , of a non-empty
convex body K with (exterior) normal u lies in a hyperplane parallel to .u⊥ , the
orthogonal complement of u.
. The orthogonal projection .K(u)|u⊥ of the support set .K(u) to .u⊥ is a uniquely
determined translate of .K(u), and we can consider .K(u)|u⊥ as a convex body in
.R
n−1 if we identify .u⊥ with .Rn−1 .
U
m
P =
. conv({o} ∪ P (ui )).
i=1
3 Geometric and Functional Inequalities 83
The volume of P equals the sum of the volumes of the simplices and the volume
of the simplex .conv({o} ∪ P (ui )) is .1/n times the .(n − 1)-dimensional volume of
the base .P (ui ) times the height, which is the distance of o from the affine subspace
spanned by .P (ui ). The latter is equal to the value of the support function of P
evaluated at .ui . Some argument is required to see that the volume thus defined is
indeed translation invariant and that it still works properly if o is not an interior
point of P .
After these explanations, the following definitions are natural.
Definition 3.2 Let .P ∈ Pn be a polytope.
If .n = 1, then .P = [a, b] with .a ≤ b and
V (1) (P ) := b − a
. and S (1) (P ) := 2.
For .n ≥ 2, let
⎧ E
⎪ 1
⎨ hP (u)V (n−1) (P (u)), if dim P ≥ n − 1,
.V
(n)
(P ) : = n
(∗)
⎪
⎩
0, if dim P ≤ n − 2,
and
⎧E
⎪
⎨ V (n−1) (P (u)), if dim P ≥ n − 1,
S (n) (P ) : =
. (∗)
⎪
⎩0, if dim P ≤ n − 2,
where the summation .(∗) is over all .u ∈ Sn−1 for which .P (u) is a facet of P ; here,
in .Rn , by a facet of a polytope we mean a support set of dimension .n − 1. In .Rn , we
shortly write .V (P ) for .V (n) (P ) and call this the volume of P . Similarly, we write
.S(P ) instead of .S
(n) (P ) and call this the surface area of P .
V+ (K) := inf V (P ),
. V− (K) := sup V (P ), where P ∈ Pn ,
P ⊃K P ⊂K
and
and
Most of the properties are again easy to derive from the corresponding ones stated
in Theorem 3.3. However, part (a) and the continuity property require additional
arguments (see [43, Theorem 3.6] for the details). The surface area functional is also
3 Geometric and Functional Inequalities 85
continuous with respect to the Hausdorff metric. In the special case of a sequence
of convex bodies .Ki , .i ∈ N, converging to a convex body .K ∈ Kn with non-empty
interior, we can provide a direct argument: Since S is translation invariant, we can
assume that .K ∈ Kn(o) . Then we also have .Ki ∈ Kn(o) , if i is large enough. Let .ε > 0
be given. By Theorem 3.1 (c) there is a polytope .P ∈ Pn(o) such that
P ⊂ int(K) and
. K ⊂ int((1 + ε)P ).
Ki ⊂ K + rB n ⊂ (1 + ε)P
. and P + rB n ⊂ K ⊂ Ki + rB n
hence
( )
|S(Ki ) − S(K)| ≤ (1 + ε)n−1 − 1 S(K).
.
This proves the assertion under the assumption that K has non-empty interior.
The general case will follow later via the connection to mixed volumes (see
Theorem 3.14 (a) and (f)).
The classical Steiner formula has already been discussed in Sect. 1.9 of Chap. 1. An
elementary proof for the polynomial expansion of the volume of the parallel set of
a general convex body can be given by first showing this expansion for the parallel
sets of a convex polytope. In this case, the parallel set decomposes into (essentially
disjoint) wedges over the faces of the polytope whose volumes can be calculated by
means of Fubini’s theorem. The situation is easily illustrated for a convex polygon
in dimension two (see Fig. 3.1).
Now we consider the more general problem of calculating the volume of
Minkowski combinations of finitely many (general) convex bodies in .Rn . More
specifically, we are considering the following question. Let .Ki ∈ Kn and .αi ≥ 0 for
.i = 1, . . . , m. How does the volume
.V (α1 K1 + · · · + αm Km )
Fig. 3.1 Polynomial volume growth of .V (P +qB 2 ) = V (P )+S(P )·q +V (B 2 )·q2 as a function
of .q ≥ 0, where .S(P ) is the perimeter and .V (P ) is the area of P
Proof The argument in [43, Proposition 3.2] works without changes for general
convex bodies, but will be needed only for polytopes. u
n
Also the next lemma is taken from [43], see Lemma 3.2 there. It identifies the
intersection of two given adjacent support sets as a support set of one of the given
support sets.
Lemma 3.7 Let .K ∈ Kn , let .u, v ∈ Sn−1 be linearly independent unit vectors, and
let .w = λu + μv with some .λ ∈ R and .μ > 0. Then .K(u) ∩ K(v) /= ∅ implies that
.K(u) ∩ K(v) = K(u)(w).
While Lemma 3.7 holds for general convex bodies, the restriction to polytopes is
crucial in the following lemma.
Lemma 3.8 Let .P ∈ Pn with .dim(P ) = n, let F be a facet of P , and let Q be a
facet of F (with respect to the .(n − 1)-dimensional affine subspace spanned by F ,
as the ambient space). Then there is a (unique) facet G of P such that .Q = F ∩ G.
3 Geometric and Functional Inequalities 87
The assertion follows from Exercise 1.4.6 or from Exercise 1.5.5 (d) in [43].
Here we provide another argument. A minor modification of this argument shows
that support sets of support sets of a polytope are support sets (and consequently the
proper faces of a polytope are support sets). In the argument, we use the important
notion of an extreme point. A point x of a polytope P (or of a general compact
convex set) is said to be an extreme point of P if .P \ {x} is convex. It is a well-
known fact that a polytope P has finitely many extreme points and P equals the
convex hull of its extreme points (which form the minimal subset of P having this
property).
Proof We may denote the extreme points (vertices) .x0 , . . . , xt of P in such a way
that .x0 , . . . , xr are the vertices of Q and .x0 , . . . , xs are the vertices (extreme points)
of F , where .r < s < t. Let .F = P (u) for some .u ∈ Sn−1 and .Q = P (u)(w) for
some vector .w ∈ Sn−1 orthogonal to u.
Then
{
= 0 for i = 1, . . . , s,
. <xi − x0 , u>
< 0 for i = s + 1, . . . , t,
{
= 0 for i = 1, . . . , r,
<xi − x0 , w>
< 0 for i = r + 1, . . . , s.
Define
{ }
<xi − x0 , w>
.λ0 := max : i = s + 1, . . . , t .
−<xi − x0 , u>
Then
⎧
⎪
⎨= 0
⎪ for i = 1, . . . , r,
.<xi − x0 , λ0 u + w> = λ0 <xi − x0 , u> + <xi − x0 , w> <0 for i = r + 1, . . . , s,
⎪
⎪
⎩≤ 0 for i = s + 1, . . . , t,
is a facet of P and .G ∩ F = Q. u
n
For polytopes .P1 , . . . , Pk ∈ Pn , let .N(P1 , . . . , Pk ) denote the set of all .u ∈ Sn−1
such that .(P1 + · · · + Pk )(u) is a facet (an .(n − 1)-dimensional face) of .P1 + · · · + Pk
in .Rn .
88 A. Colesanti and D. Hug
and, for .n ≥ 2,
1E
V (P1 , . . . , Pn ) :=
. hPn (u)V (P1 (u), . . . , Pn−1 (u)),
n
(∗)
E
m E
m
V (α1 P1 + · · · + αm Pm ) =
. ··· αi1 · · · αin V (Pi1 , . . . , Pin ). (3.1)
i1 =1 in =1
(e) The mixed volume is linear with respect to positive Minkowski combinations in
each argument.
Proof The proof proceeds by induction on the dimension n. Thus the assertions
(a)–(e) can be established simultaneously. We focus on two points.
3 Geometric and Functional Inequalities 89
First We Prove (d) It is sufficient to consider the case where α1 > 0, . . . , αm > 0.
By the definition of volume and by Proposition 3.6,
(( m ) )
1 E E
.V (α1 P1 + · · · + αm Pm ) = hEm i=1 αi Pi
(u) v αi Pi (u)
n
u∈N (P1 ,...,Pm ) i=1
( m )
E
m
1 E E
= αin hPin (u) v αi Pi (u) ,
n
in =1 u∈N (P1 ,...,Pm ) i=1
1 E
= hPn (u) V (P1 (u), . . . , Pn−1 (u))
n
u∈N (P1 ,...,Pn )
1 E
=
n(n − 1)
u,v∈N (P1 ,...,Pn ),v/=±u
[ 1 1 ]
× hPn (u)hPn−1 (v) − hPn (u)hPn−1 (u)
sin γ (u, v) tan γ (u, v)
× V (P1 (u) ∩ P1 (v), . . . , Pn−2 (u) ∩ Pn−2 (v))
= V (P1 , . . . , Pn−2 , Pn , Pn−1 ),
1 E E
n
V (P1 , . . . , Pn ) =
. (−1)n+k V (Pr1 + · · · + Prk ). (3.2)
n!
k=1 1≤r1 <···<rk ≤n
Finally, we define mixed volumes of general convex bodies and summarize their
main basic properties.
(j )
Theorem 3.14 Let .K1 , . . . , Kn ∈ Kn be convex bodies. Let .(Pi )j ∈N for .i ∈
(j )
{1, . . . , n} be arbitrary approximating sequences of polytopes such that .Pi → Ki
as .j → ∞, for .i = 1, . . . , n. Then the limit
(j ) (j )
V (K1 , . . . , Kn ) = lim V (P1 , . . . , Pn )
.
j →∞
(j )
exists and is independent of the choice of the approximating sequences .(Pi )j ∈N .
The number .V (K1 , . . . , Kn ) is called the mixed volume of the convex bodies
.K1 , . . . , Kn . The mapping .V : (K ) → R, which is defined by .(K1 , . . . , Kn ) |→
n n
1 E E
n
V (K1 , . . . , Kn ) =
. (−1)n+k V (Kr1 + · · · + Krk ), (3.3)
n!
k=1 1≤r1 <···<rk ≤n
E
m E
m
V (α1 K1 + · · · + αm Km ) =
. ··· αi1 · · · αin V (Ki1 , . . . , Kin ). (3.4)
i1 =1 in =1
(j ) (j )
V (K1 , . . . , Kn ) → V (K1 , . . . , Kn ),
.
(j )
whenever .Ki → Ki as .j → ∞, for .i = 1, . . . , n.
(g) .V ≥ 0 and V is increasing in each argument.
The proofs of these assertions are partly based on Theorem 3.13 and properties
of mixed volumes (or the volume functional) of polytopes. The representation of
the surface area as a mixed volume is done in two steps, by first choosing K as
a polytope and approximating the unit ball by a sequence of polytopes; see [43,
Theorem 3.9] for a detailed argument.
Using the symmetry of the mixed volumes, we obtain for .K1 , . . . , Km ∈ Kn and
.α1 , . . . , αm ≥ 0 that
( m ) ( )
E E
n
n
V
. αi Ki = α r1 · · · αm
rm
V (K1 [r1 ], . . . , Km [rm ]),
r1 , . . . , rm 1
i=1 r1 ,...,rm =0
(3.5)
where the number in brackets is the multiplicity .ri of the convex body .Ki for .i =
1, . . . , m. Here the multinomial coefficient is defined as usual, in particular it is zero
if .r1 + · · · + rm /= n.
In the special case of the parallel body .K + αB n , .α ≥ 0, of a (non-empty) body
.K ∈ K , we retrieve the Steiner formula. In fact, with the choices .m = 2, .α1 = 1,
n
V (K + αB n ) = V (α1 K1 + α2 K2 )
.
E
2 E
2
= ··· αi1 · · · αin V (Ki1 , . . . , Kin )
i1 =1 in =1
E
n ( )
n
= αi V (K[n − i], B n [i]).
i
i=0
For the following definition, let .κk denote the volume of the k-dimensional unit
ball.
Definition 3.15 For .K ∈ Kn ,
E
n ( )
n
V (K + αB n ) =
. αi Wi (K)
i
i=0
and
E
n
V (K + αB n ) =
. α n−j κn−j Vj (K).
j =0
The intrinsic volumes were introduced by Peter McMullen (1975). Their obvious
advantages are that the index j of .Vj equals the degree of homogeneity,
Vj (αK) = α j Vj (K),
. K ∈ Kn , α ≥ 0
and, more importantly, intrinsic volumes are independent of the dimension of the
ambient Euclidean space. By this we mean that for a body .K ∈ Kn with .dim K =
k < n, we have
(n) (k)
Vj (K) = Vj (K),
. j = 0, . . . , k,
where the upper index indicates the dimension of the ambient space in which
(k)
the intrinsic volume is considered. In other words, by .Vj (K) we mean that the
j th intrinsic volume of the k-dimensional convex body K is calculated in the k-
(n)
dimensional affine subspace containing K, whereas .Vj (K) requires that the j th
intrinsic volume of K is determined with respect to the ambient space .Rn (even if
K is lower-dimensional).
3 Geometric and Functional Inequalities 93
The intrinsic volumes .Vj : Kn → [0, ∞), .j ∈ {0, . . . , n}, are important
geometric functionals of a convex body.
Examples
. The volume of K:
If K has interior points, then .2Vn−1 (K) = S(K) = Hn−1 (∂K). For a body K
with .dim(K) = n − 1, .Vn−1 (K) = Hn−1 (K) is the .(n − 1)-dimensional content
of K.
. The mean width of K is proportional to .V1 (K). We introduce it next.
We now define the mean width of .K ∈ Kn and relate it to the first intrinsic
volume.
By the special case .j = 1 of the preceding definition,
κn−1
. V1 (K) = V (K, B n , . . . , B n ). (3.6)
n
Approximation of the unit ball by polytopes shows that
f
1
.V (K, B n , . . . , B n ) = hK (u) dHn−1 (u), (3.7)
n Sn−1
where the integration is with respect to the spherical Lebesgue measure, that is, the
(n − 1)-dimensional Hausdorff measure .Hn−1 , restricted to the unit sphere .Sn−1 . A
.
rigorous derivation of (3.7) follows from Theorem 3.21 (or by means of the method
of rotation averaging, as described for Theorem 2.26, and a characterization result
for Minkowski additive functionals, see Theorem 2.27).
Combining (3.6) and (3.7), we get
f
κn−1 V1 (K) =
. hK (u) dHn−1 (u).
S
n−1
Since .wK (u) := hK (u) + hK (−u) is the width of K in direction u (the distance
between the two parallel supporting hyperplanes with common unit normal u),
f f
1 n−1 1 κn
. hK (u) dH (u) = wK (u) dHn−1 (u) = w(K),
n S n−1 2n S n−1 2
94 A. Colesanti and D. Hug
where
f
1
.w(K) := wK (u) dHn−1 (u)
nκn Sn−1
denotes the mean width of K (with .nκn = Hn−1 (Sn−1 )). It follows that
nκn
V1 (K) =
. w(K),
2κn−1
that is, the first intrinsic volume .V1 and the mean width .w are proportional
functionals on convex bodies.
1 E
V (P1 , . . . , Pn−1 , Pn ) =
. hPn (u)V (P1 (u), . . . , Pn−1 (u)).
n
u∈S
n−1
On the right side, the summation extends over all unit vectors u for which
.V (P1 (u), . . . , Pn−1 (u)) > 0 or, alternatively, over all facet normals of the polytope
.P1 + · · · + Pn−1 . The (effectively finite) summation is independent of .Pn .
1 E
V (P1 , . . . , Pn−1 , Kn ) =
. hKn (u)V (P1 (u), . . . , Pn−1 (u)). (3.8)
n
u∈S
n−1
We define
E
S(P1 , . . . , Pn−1 , ·) :=
. V (P1 (u), . . . , Pn−1 (u))δu , (3.9)
u∈S
n−1
where the Dirac measure .δu with unit point mass in .u ∈ Sn−1 is considered to be a
Borel measure on .Sn−1 .
In the following, we only need a special version of Prokhorov’s theorem (see
[14, Theorem 8.6.2]), which can be deduced from the Helly–Bray theorem (see [45,
Theorem 6.20]). Since the unit sphere is compact, the statement of the theorem
does not involve the condition of “tightness” explicitly, since it is trivially satisfied.
On the other hand, since the present application is not restricted to probability
3 Geometric and Functional Inequalities 95
measures, a uniform bound on the total mass of the measures is required. For another
application of Prokhorov’s theorem (for sets of probability measures), see Sect. 4.2.4
of this volume.
Theorem 3.17 Let .μi , .i ∈ N, be a sequence of Borel measures on .Sn−1 . If there is
a constant C such that .μi (Sn−1 ) ≤ C for .i ∈ N, then there exists an infinite subset
.I ⊂ N and a Borel measure .μ on .S such that .μi → μ as .I e i → ∞, where
n−1
for .K ∈ Kn .
Proof We already know that
f
1
.V (P1 , . . . , Pn−1 , Kn ) = hK (u) dS(P1 , . . . , Pn−1 , u)
n Sn−1 n
96 A. Colesanti and D. Hug
holds for .P1 , . . . , Pn−1 ∈ Pn and .Kn ∈ Kn . Let .K1 , . . . , Kn−1 ∈ Kn be arbitrary.
(j ) (j )
Choose .Pi ∈ Pn , for .j ∈ N and .i ∈ {1, . . . , n − 1}, with .Pi → Ki as .j → ∞.
An application of the preceding relation with .Kn = B shows that
n
(j ) (j ) (j ) (j )
S(P1 , . . . , Pn−1 , Sn−1 ) = nV (P1 , . . . , Pn−1 , B n ),
. j ∈ N.
The mixed volumes on the right-hand side are bounded by the continuity (or
monotonicity) of the mixed volumes.
Hence an application of Prokhorov’s theorem (stated as Theorem 3.17) yields the
existence of a Borel measure .μ on .Sn−1 and of an infinite subset .I ⊂ N such that
with respect to the weak convergence of measures we have
(j ) (j )
S(P1 , . . . , Pn−1 , ·) → μ
.
The limit measure .μ depends on .K1 , . . . , Kn−1 , but could possibly also depend on
the choice of the approximating sequence. However, the (uniqueness) Lemma 3.18
shows that .μ is independent of the polytopes used for the approximation and of the
chosen convergent subsequence. Therefore we set .S(K1 , . . . , Kn−1 , ·) := μ. u
n
Lemma 3.18 and the symmetry property of mixed volumes imply that also
S(K1 , . . . , Kn−1 , ·) is independent of the order of the bodies .K1 , . . . , Kn−1 .
.
For polytopes .P1 , . . . , Pn−1 , the measures .S(P1 , . . . , Pn−1 , ·) are just again the
measures from which we started in (3.9).
Definition 3.20 The measure .S(K1 , . . . , Kn−1 , ·) is called the mixed surface area
measure or simply mixed area measure of the bodies .K1 , . . . , Kn−1 ∈ Kn . In
particular,
is called the j th order surface area measure or simply j th area measure of K for
j ∈ {0, . . . , n − 1}.
.
Remarks
we have
f
. <x, u> dS(K1 , . . . , Kn−1 , u) = 0
Sn−1
in particular,
nκn−j
Sj (K, Sn−1 ) = nV (K[j ], B n [n − j ]) =
. (n) Vj (K)
j
and
S0 (K, ·) = S(B n , . . . , B n , ·) = Sj (B n , ·)
.
Further properties of mixed area measures follow from the relation between
mixed volumes and mixed area measures (via the uniqueness property stated in
Lemma 3.18).
Theorem 3.21 The mapping S : (K1 , . . . , Kn−1 ) |→ S(K1 , . . . , Kn−1 , ·), defined
on (Kn )n−1 and with values in the space of finite Borel measures on the unit sphere,
has the following properties:
(a) S is symmetric, that is,
+ βS(L, K2 , . . . , Kn−1 , ·)
= nV (K1 , . . . , Kn−1 , Kn )
f
= hKn (u) dS(K1 , . . . , Kn−1 , u),
Sn−1
where Kn ∈ Kn is arbitrary.
3 Geometric and Functional Inequalities 99
(e) For ε > 0 and f ∈ C(Sn−1 ), choose K, L ∈ Kn with ||f − (hK − hL )|| ≤ ε.
Here we use the maximum norm ||g|| := max{|g(u)| : u ∈ Sn−1 } for g ∈
C(Sn−1 ).
(m)
Further, choose m0 ∈ N such that for m ≥ m0 we have Ki ⊂ Ki + B n ,
for i = 1, . . . , n − 1, and
(m) (m)
|V (K1 , . . . , Kn−1 , M) − V (K1 , . . . , Kn−1 , M)| ≤ ε
.
Sn−1 (α1 K1 + · · · + αm Km , ·)
.
E
m E
m
= ... αi1 · · · αin−1 S(Ki1 , . . . , Kin−1 , ·)
i1 =1 in−1 =1
E
n−1 n−1 (
E )
n−1
= ... α r1 · · · αm
rm
S(K1 [r1 ], . . . , Km [rm ], ·). (3.11)
r1 , . . . , rm 1
r1 =0 rm =0
100 A. Colesanti and D. Hug
S(K1 , . . . , Kn−1 , ·)
.
1 E n−1 E
= (−1)n−1+k Sn−1 (Kr1 + · · · + Krk , ·). (3.12)
(n − 1)!
k=1 1≤r1 <···<rk ≤n−1
As a special case of (3.11) and the preceding results, we summarize some of the
properties of area measures.
Corollary 3.22 For j = 0, . . . , n − 1, the mapping K |→ Sj (K, ·) on Kn is
translation invariant, rotation-covariant and continuous.
Moreover,
E
n−1 ( )
n−1
Sn−1 (K + αB , ·) =
.
n
α n−1−j
Sj (K, ·)
j
j =0
is concave.
. Consequences of the Brunn–Minkowski inequality are: several inequalities for
mixed volumes, in particular the Minkowski inequality and the celebrated
isoperimetric inequality.
3 Geometric and Functional Inequalities 101
. The inequality holds for more general sets, for instance for compact sets. This
is in contrast to other inequalities which involve mixed volumes and require
convexity.
. Several proofs are available in the literature. They are based on cubical approx-
imation, mass transportation, symmetrization, the use of pairs of coordinated
parallel planes (see the classical proof below), or on the Prékopa–Leindler
inequality (a functional version of the Brunn–Minkowski inequality), to mention
just a few. We provide various proofs in this section (see also the contribution by
Shiri Artstein-Avidan in Chap. 4).The proofs given here all involve an induction
argument as a crucial step.
Theorem 3.23 (Brunn–Minkowski Inequality, BMI) Let .K, L ∈ Kn be convex
bodies and .α ∈ (0, 1). Then
1 1 1
.V (αK + (1 − α)L) n ≥ αV (K) n + (1 − α)V (L) n
with equality if and only if K and L lie in parallel hyperplanes or K and L are
homothetic.
Before we provide a (in fact several) proof(s), we start with a sequence of
preparatory comments.
. We say that .K, L ∈ Kn are homothetic if and only if .K = αL+x or .L = αK +x,
for some .x ∈ Rn and some .α ≥ 0. Hence, if K or L is a point, then K and L are
homothetic.
. The Brunn–Minkowski inequality also holds for .α ∈ {0, 1}. But then the
inequality is an equality for all convex bodies .K, L.
. Another (seemingly weaker but in fact equivalent) way to state the inequality is
V (αK + (1 − α)L) ≥ 1,
. (3.13)
for .α ∈ [0, 1] and all convex bodies .K, L ⊂ Rn with .V (K) = V (L) = 1.
To see that (3.13) implies the BMI, let .K, L have positive volume and .α ∈ (0, 1).
Then we can apply (3.13) with
1
− n1 − n1 αV (K) n
K = V (K)
. K, L = V (L) L, α= 1 1
,
αV (K) n + (1 − α)V (L) n
to deduce the general BMI for .K, L, α. Note that this extension step does not
require convexity of the sets.
102 A. Colesanti and D. Hug
. The BMI is true for arbitrary compact sets. To state the result in this generality,
we write .Cn for the class of non-empty compact sets in .Rn . The volume of a
compact set is its Lebesgue measure (that is, n-dimensional Hausdorff measure).
Theorem 3.24 (Brunn–Minkowski Inequality, General Version) Let .X, Y ∈ Cn
with .V (X), V (Y ) > 0. Then
1 1 1
V (X + Y ) n ≥ V (X) n + V (Y ) n
.
H1 (A + B) ≥ H1 (A) + H1 (B).
. (3.14)
H1 (A + B) ≥ H1 (A ∪ B) = H1 (A) + H1 (B),
.
. The inequality (3.14) can be strict. To see this, choose .A = [−3, −2] ∪ [−1, 0],
.B = [0, 1], hence .A + B = [−3, 1].
A = [0, a1 ] × · · · × [0, an ],
. B = [0, b1 ] × · · · × [0, bn ],
A + B = [0, a1 + b1 ] × · · · × [0, an + bn ].
.
3 Geometric and Functional Inequalities 103
Lemma 3.26 Let .X, Y ∈ Cn , .V (X), V (Y ) > 0, .e ∈ Sn−1 , and .α, β ∈ R. Define
.X
± := X ∩ H ± (e, α), .Y ± := Y ∩ H ± (e, β). Suppose that .V (X + )/V (X) =
V (Y + )/V (Y ). Then
If X is the union of at least two boxes, then the dissecting hyperplane .H (e, α)
can be chosen as a coordinate hyperplane which separates two of the boxes. Thus
the total number of boxes in .X+ , Y + and the total number of boxes in .X− , Y − is
reduced by at least one, so that the induction hypothesis can be applied.
The discussion of the equality case requires additional efforts.
We now turn to a functional version of the BMI.
Theorem 3.27 (Prékopa–Leindler Inequality, PLI) Let .λ ∈ (0, 1), and let
f, g, h : Rn → [0, ∞) be measurable functions such that
.
Then
f (f )1−λ (f )λ
. h(x) dx ≥ f (x) dx g(x) dx .
Rn Rn R n
for .λ ∈ [0, 1] and .X, Y ∈ Cn . The general case then is obtained as described in the
third remark after Theorem 3.23.
Proof The assertion is proved by induction over .n ∈ N. We start with .n = 1.
We can assume that .f /≡ 0 and .g /≡ 0 are bounded. Otherwise, we first consider
.min{f, m} instead of f and .min{g, m} instead of g for .m ∈ N. The general case
[γ ≥ t] := {x ∈ R : γ (x) ≥ t}.
.
For .t ∈ (0, 1), the sets .[f ≥ t], [g ≥ t] are non-empty. If they are also compact,
the BMI for compact sets on the real line yields
Otherwise we approximate the sets .[f ≥ t] and .[g ≥ t] from inside by compact sets
using the fact that Lebesgue measure is a regular measure.
3 Geometric and Functional Inequalities 105
where in the last step we used that .a 1−λ bλ ≤ (1 − λ)a + λb for .a, b ≥ 0.
We continue the induction argument. Let .n > 1 and suppose that the assertion has
already been proved in lower dimensions. We use the identification .Rn = Rn−1 × R
and define, for .s ∈ R and .z ∈ Rn−1 ,
that is,
for .a, b ∈ R.
106 A. Colesanti and D. Hug
The case .n = 1 has already been proved, hence by Fubini’s theorem we get
f f f f
. h(x) dx = h s (z) dz ds = H (s) ds
Rn R Rn−1 R
(f )1−λ (f )λ
≥ F (a) da G(b) db
R R
(f )1−λ (f )λ
= f (x) dx g(x) dx .
Rn Rn
This completes the induction argument and thus the proof. u
n
The classical proof of the BMI also has its benefits. For instance, it has been
extended to obtain stability improvements of the BMI. In particular, it clarifies the
equality cases of the BMI for convex sets.
Proof (Classical Proof of the BMI) The main case to be considered deals with the
situation where .V (K) = V (L) = 1.
Volume is translation invariant. Hence we can assume that K and L have their
center of mass at the origin o. Recall that the center of mass of an n-dimensional
convex body M is the point .c = c(M) ∈ Rn for which
f
1
.<c, u> = <x, u> dx holds for u ∈ Sn−1 .
V (M) M
Since .V (K) = V (L) = 1, assuming that .c(K) = c(L) = o means that
f f
. <x, u> dx = <x, u> dx = 0 for u ∈ Sn−1 .
K L
Under these assumptions, the equality case then reduces to the claim that .K = L.
We now prove the Brunn–Minkowski theorem by induction on n.
For .n = 1, the assertion is already clear.
The argument employs a method that can be described as using sections of K
and L by coordinated parallel hyperplanes.
For the induction step we assume that .n ≥ 2 and that the assertion of the Brunn–
Minkowski theorem is true in dimension .n − 1. We fix an arbitrary unit vector
.u ∈ S and denote by .Eη := H (u, η) = {x ∈ Rn : <u, x> = η}, .η ∈ R,
n−1
the hyperplane in direction u with (signed) distance .η from the origin. Recall that
−
.H (u, η) = {x ∈ R : <u, x> ≤ η} is the halfspace bounded by .H (u, η) having u
n
and .f ' (β) = v(K ∩ Eβ ). Since f is invertible, the inverse function .β : [0, 1] →
[−hK (−u), hK (u)], which is also strictly increasing and continuous, satisfies
.β(0) = −hK (−u), .β(1) = hK (u), and
1 1
.β ' (τ ) = = , τ ∈ (0, 1).
f ' (β(τ )) v(K ∩ Eβ(τ ) )
By the same argument, for the convex body L we obtain a function .γ : [0, 1] →
[−hL (−u), hL (u)] with
1
γ ' (τ ) =
. , τ ∈ (0, 1).
v(L ∩ Eγ (τ ) )
Because of
τ |→ αβ(τ ) + (1 − α)γ (τ ),
f 1
= v((αK + (1 − α)L) ∩ Eαβ(τ )+(1−α)γ (τ ) )(αβ ' (τ ) + (1 − α)γ ' (τ )) dτ
0
f 1 ( )
≥ v α(K ∩ Eβ(τ ) ) + (1 − α)(L ∩ Eγ (τ ) )
0
( )
α 1−α
× + dτ
v(K ∩ Eβ(τ ) ) v(L ∩ Eγ (τ ) )
f 1[ / / ]n−1
≥ α n−1
v(K ∩ Eβ(τ ) ) + (1 − α) n−1 v(L ∩ Eγ (τ ) )
0
( )
α 1−α
× + dτ . (3.15)
v(K ∩ Eβ(τ ) ) v(L ∩ Eγ (τ ) )
≥ 1, (3.16)
108 A. Colesanti and D. Hug
with equality if and only if .r = s. (Note that the function .x |→ log x is strictly
concave.)
Now we discuss the equality case and assume that
V (αK + (1 − α)L) = 1.
.
Then we must have equality in (3.16), which implies that the integrand in (3.15)
equals 1, for all .τ (since the integrand is a continuous function of .τ ). In turn this
yields that
Therefore .β ' = γ ' on .(0, 1), hence the function .β−γ is a constant on .[0, 1]. Because
the center of gravity of K is at the origin, we obtain
f f β(1) f β(1) f 1
.0= <x, u> dx = ηv(K ∩ Eη ) dη = ηf ' (η) dη = β(τ ) dτ,
K β(0) β(0) 0
Consequently,
f 1
. (β(τ ) − γ (τ )) dτ = 0
0
Since .u ∈ Sn−1 was arbitrary, .V (αK + (1 − α)L) = 1 implies that .hK = hL , and
hence .K = L.
Conversely, it is clear that .K = L implies that .V (αK + (1 − α)L) = 1. u
n
3 Geometric and Functional Inequalities 109
1
= V (α[xK + (1 − x)L] + (1 − α)[yK + (1 − y)L]) n
1 1
≥ αV (xK + (1 − x)L) n + (1 − α)V (yK + (1 − y)L) n
1
The same is true for the function .t |→ V (K + tL) n with .t ∈ [0, 1].
As an important consequence, we obtain an inequality for mixed volumes which
was first proved by Hermann Minkowski.
Theorem 3.28 (Minkowski’s Inequality, MI) Let .K, L ∈ Kn . Then
1 1
f + (0) =
. V (K) n −1 n V (K[n − 1], L).
n
110 A. Colesanti and D. Hug
Hence we arrive at
1 1 1 1
.V (K) n −1 V (K[n − 1], L) ≥ V (K + L) n − V (K) n ≥ V (L) n ,
where we used the BMI for the second inequality (with .t = 12 ). This yields the
required inequality.
If equality holds, then equality holds in the BMI, which implies that K and L are
homothetic. Conversely, if K and L are homothetic, then equality holds. u
n
The isoperimetric inequality is one of the fundamental classical results in
mathematics. It states that among all (convex) bodies of given volume, precisely
the Euclidean balls minimize the surface area F .
Corollary 3.29 (Isoperimetric Inequality) If .K ∈ Kn is n-dimensional, then
( ) ( )
S(K) n V (K) n−1
. ≥ .
S(B n ) V (B n )
Equality holds if and only if K is a ball.
Proof We put .L := B n in the Minkowski inequality and get
or, equivalently,
For .n = 2 and using the common terminology .A(K) for the area (the “volume” in
R2 ) and .L(K) for the boundary length (the “surface area” in .R2 ), we obtain
.
1
A(K) ≤
. L(K)2 ,
4π
3 Geometric and Functional Inequalities 111
and, for .n = 3,
1
V (K)2 ≤
. S(K)3 .
36π
An exchange of K and .B n in the proof above leads to a similar inequality for the
mixed volume .V (B n [n − 1], K), whence we obtain the following corollary for the
mean width .w(K).
Corollary 3.30 Let .K ∈ Kn be a convex body. Then,
( )n
w(K) V (K)
. ≥ .
w(B n ) V (B n )
Equality holds if .dim(K) ≤ n − 2 or if K and L are homothetic, but these are not
the only cases.
[The characterization of the case of equality involves the .(n − 2)-tangential
bodies of L.]
1
Proof We assume that .V (K) > 0 (why?). The function .f (t) := V (K + tL) n is
concave on .[0, 1] by the BMI. Since
n ( )
E n i
.t →
| h(t) := V (K + tL) = t V (K[n − i], L[i]), t ∈ [0, 1],
i
i=0
is of class .C 2 , f is of class .C 2 as well (note that .V (K) > 0). Hence, .f '' ≤ 0 on
1
.[0, 1]. Since .f (t) = h(t) n , we get
[ ]
' 1 1 1 1
−2 1 − n '
.f (t) = h(t) n −1 h' (t), ''
f (t) = h(t) n ''
h (t) + h(t)h (t) .
2
n n n
n−1 ' 2
. h(0)h'' (0) ≤ h (0) . (3.18)
n
112 A. Colesanti and D. Hug
Furthermore,
n ( )
E
' n i−1
.h (t) = it V (K[n − i], L[i]),
i
i=1
n ( )
E n
h'' (t) = i(i − 1)t i−2 V (K[n − i], L[i]).
i
i=2
Plugging .h'' (0) = n(n − 1)V (K[n − 2], L, L), .h' (0) = nV (K[n − 1], L) and
.h(0) = V (K) into (3.18), it follows that
n−1 2
V (K)n(n − 1)V (K[n − 2], L, L) ≤
. n V (K[n − 1], L)2 ,
n
and hence
1 1
. (H (u, t) ∩ K) + (H (u, −t) ∩ K) ⊂ H (u, 0) ∩ K.
2 2
Hence (writing again v for the volume in a hyperplane)
( ) 1
1 1 1 n−1
v(H (u, 0) ∩ K)
. n−1 ≥v (H (u, t) ∩ K) + (H (u, −t) ∩ K)
2 2
1 1 1 1
≥ v (H (u, t) ∩ K) n−1 + v (H (u, −t) ∩ K) n−1
2 2
1 1 1 1
= v (H (u, t) ∩ K) n−1 + v (H (u, t) ∩ (−K)) n−1
2 2
1
= v(H (u, t) ∩ K) n−1 ,
The second application concerns a uniqueness property of the top order area
measures.
Theorem 3.32 (Uniqueness of Top Order Area Measures) Let .K, L ∈ Kn with
.dim K = dim L = n. Then .Sn−1 (K, ·) = Sn−1 (L, ·) if and only if K and L are
translates.
Proof If .K, L are translates of each other, the equality of the area measures is clear.
Assume now .Sn−1 (K, ·) = Sn−1 (L, ·). Then
f
1
V (K[n − 1], L) =
. hL (u) dSn−1 (K, u)
n Sn−1
f
1
= hL (u) dSn−1 (L, u) = V (L).
n Sn−1
which implies that .V (K) = V (L). But then we have equality in both inequalities,
and hence K and L are homothetic. Since K and L have the same volume, they
must be translates of each other. u
n
The Brunn–Minkowski inequality has been discussed also for general compact
sets. A similar extension can be obtained for the Minkowski inequality (see [42,
Sections 5.2.1-5]).
Definition 3.33 (Outer Relative Surface Area for Compact Sets) For .X, Y ∈ Cn
the lower (outer) relative surface area of X with respect to Y is defined by
V (X + εY ) − V (X)
S+ (X, Y ) := lim inf
. .
ε↓0 ε
V (X + εY ) − V (X)
S + (X, Y ) := lim sup
. .
ε↓0 ε
If .S + (X, Y ) = S+ (X, Y ) =: S(X, Y ), then we say that the outer relative surface
area of X with respect to Y exists.
In Chapter essentially the same concept is considered, where the .lim inf and the
lim sup are denoted as lower and upper anisotropic outer Minkowski content of X
.
with respect to Y .
114 A. Colesanti and D. Hug
Let .X, Y ∈ Cn with .V (X), V (Y ) > 0. Suppose that .S(X, Y ) exists. Then .S(X, Y ) =
n−1 1
nV (X) n V (Y ) n holds if and only if .X, Y are homothetic convex bodies.
n−1 1
If only .S+ (X, Y ) = nV (X) n V (Y ) n is available, then still the volume kernels
of .X, Y are homothetic convex bodies.
Here the volume kernel of a set .X ∈ Cn is the set of all .x ∈ X such that .X ∩ U
has positive volume for each neighborhood U of x. The volume kernel of X is again
a compact set which has the same volume as X, which justifies its name.
for convex bodies .K, L, M ∈ Kn . The case considered in Theorem 3.31 is recovered
by choosing .M = K. The resolution of the long-standing open problem of charac-
terizing the equality case for the general quadratic Minkowski inequality (3.19) has
been achieved by Shenfeld and van Handel [60].
The even more general Aleksandrov–Fenchel inequality (AFI) is obtained if the
sequence of the .n − 2 bodies .M, . . . , M in (3.19) is replaced by arbitrary convex
bodies .M1 , . . . , Mn−2 , so that
(3.20)
in the following.
For .m ∈ {2, . . . , n} and convex bodies .K1 , K2 , Km+1 , . . . , Kn ∈ Kn , we
consider the function defined by
1
fm (t) := V ((K1 + tK2 ) [m], Km+1..n ) m ,
. t ≥ 0.
We will see now that the validity of (AFI) is closely related to the fact that .fm is a
concave function on .[0, ∞).
The following argument is adjusted from Cordero-Erausquin, Klartag, Merigo,
Santambrogio [33] (see also [43, Section 3.5] for further details).
We start by determining the second derivative of the function .fm .
Lemma 3.36 Let .m ∈ {2, . . . , n}, .K1 , K2 , Km+1 , . . . , Kn ∈ Knn and .t ≥ 0. Then
Hence we get
1−m '
h''m (t)hm (t) +
. hm (t)2
m
(
= −m(m − 1) V (K t , K2 , K t [m − 2], Km+1..n )2
)
−V (K t [2], K t [m − 2], Km+1..n )V (K2 [2], K t [m − 2], Km+1..n )
( )
= −m(m − 1) V (K1 , K2 , K3..n )2 − V (K1 [2], K3..n )V (K2 [2], K3..n ) ,
We write .Pnn for the set of n-dimensional polytopes in .Rn . Polytopes in .Pnn are
called n-polytopes. For vectors .u1 , . . . , uN ∈ Sn−1 and .h = (h1 , . . . , hN )T ∈ RN ,
we now consider polyhedral sets of the form
n
N
P[h] :=
. H − (ui , hi ).
i=1
Clearly, if .h ∈ RN + , then .o ∈ P[h] and .P[h] is a polytope if and only if the vectors
.u1 , . . . , uN ∈ S are not contained in any hemisphere. Further, .o ∈ int(P[h] ) if
n−1
(c) If .P = P[h] ∈ Pnn is simple and has exterior facet normals .u1 , . . . , uN ∈ Sn−1 ,
then there is some .β > 0 such that any two of the polytopes .P[h+α] with .α =
(α1 , . . . , αN )T and .|αi | ≤ β are strongly isomorphic.
(d) For any .(K1 , . . . , Kn ) ∈ (Kn )n there is a sequence .(P1(m) , . . . , Pn(m) ) ∈ (Pnn )n ,
(m)
.m ∈ N, such that .P → Kj as .m → ∞ (in the Hausdorff metric), for
j
.j = 1, . . . , n, and .P1(m) , . . . , Pn(m) are simple and strongly isomorphic for each
.m ∈ N.
Let .P ∈ Pnn be a simple polytope with facet normals .u1 , . . . , uN ∈ Sn−1 . Then
{ }
C(P ) := h ∈ (0, ∞)N : P[h] ∈ a(P )
.
is an open convex cone. In fact, if .h, h' ∈ C(P ), then .P[h] + P[h' ] ∈ a(P ) by
Lemma 3.40 (b), and thus
implies that .P[h] + P[h' ] = P[h+h' ] . The fact that .C(P ) is open follows from
Lemma 3.40 (c).
Lemma 3.41 Let .P ∈ Pnn be a simple polytope with exterior facet normals
.u1 , . . . , uN ∈ Sn−1 . For .i = 1, . . . , n, let .Pi = P[h(i) ] ∈ a(P ) with
(i)
.hj = h(P[h(i) ] , uj ) for .j = 1, . . . , N . Then there are real numbers .aj1 ···jn , for
.j1 , . . . , jn ∈ {1, . . . , N }, depending only on .a(P ) (and independent of the support
numbers of the polytopes) and symmetric in the lower indices, such that
E
N
(1) (n)
V (P1 , . . . , Pn ) =
. aj1 ···jn hj1 · · · hjn .
j1 ,...,jn =1
Proof (of the AFI) We proceed by induction. The theorem has already been proved
in the cases where .n ∈ {2, 3}. Hence let .n ≥ 3 (or even .n ≥ 4) and assume the
theorem holds in smaller dimensions.
3 Geometric and Functional Inequalities 119
1
the map .C(P ) e h |→ Fi (h) 2 is concave.
(iv) Let .J := {(i, j ) ∈ {1, . . . , N}2 : dim(P (ui ) ∩ P (uj )) = n − 2}. For .(i, j ) ∈ J
and .h ∈ C(P ), it follows from an explicit formula for the mixed volume of
strongly isomorphic polytopes that
3 2 1
Fij (h) =
. V (n−2)
n n − 1 sin (ui , uj )
/
( )
P[h] (ui ) ∩ P[h] (uj ), Kn−3 (ui , uj ) > 0,
The following Lemma 3.42 shows that these properties imply that .C(P ) e h |→
1
F(h) 3 is concave. Now from the approximation Lemma 3.40 (d) and the continuity
of mixed volumes it follows that
1
L |→ V (L[3], K3 , . . . , Kn ) 3 ,
. L ∈ Kn ,
120 A. Colesanti and D. Hug
is a concave map, and therefore .f3 is concave for all convex bodies. Hence, AFI
holds in n-dimensional Euclidean space, which completes the induction step.
Lemma 3.42 Let .C ⊂ (0, ∞)N be an open, convex cone. Let .F : C → R be a
function of class .C 3 such that
(a) .F is positively 3-homogeneous,
1
(b) .C e h |→ Fi (h) 2 is concave and .Fi (h) > 0 for .h ∈ C,
(c) .T F(h) is irreducible and .Fij (h) ≥ 0 for .i /= j and .h ∈ C.
2
1
Then .C e h |→ F(h) 3 is a concave function.
In the proof of the analytic Lemma 3.42, the following fact from linear algebra,
which is based on a special case of the Perron–Frobenius theorem, is used. For
details, we refer to [43, Section 3.5].
Lemma 3.43 Let .M = (mij ) ∈ RN,N be symmetric, irreducible and such that
.mij ≥ 0 for .i /= j . Suppose that .v1 ∈ S
n−1
is a positive eigenvector of M with
corresponding eigenvalue .λ1 > 0. If M does not have any eigenvalues in .(0, λ1 ),
then
M ≤ λ1 v1 ⊗ v1 .
.
is concave.
From the inequalities derived up to this point, a variety of strong geomet-
ric inequalities can be deduced. For example, we obtain a general version of
Minkowski’s inequality.
3 Geometric and Functional Inequalities 121
1 1
fm' (t) |t=0+ =
. V (K1 [m], K) m −1 · m · V (K1 [m − 1], K2 , K) ≥ fm (1) − fm (0).
m
This and the GBMI for mixed volumes imply that
1
V (K1 [m], K) m −1 V (K1 [m − 1], K2 , K)
.
1 1
≥ V ((K1 + K2 ) [m], K) m − V (K1 [m], K) m
1 1 1
≥ V (K1 [m], K) m + V (K2 [m], K) m − V (K1 [m], K) m
1
= V (K2 [m], K) m ,
and hence
m−1 1
V (K1 [m − 1], K2 , K) ≥ V (K1 [m], K)
. m V (K2 [m], K) m ,
In fact, using the GMI and the Minkowski linearity of mixed volumes, we get
In the beginning of this section we briefly describe the basics of the .Lp Brunn–
Minkowski theory. Thereafter we present the .Lp Brunn–Minkowski inequality, in
the case .p ≥ 1 and in the more delicate case .0 ≤ p < 1, and we focus in particular
on the so-called log-Brunn–Minkowski inequality, corresponding to the case .p = 0.
The .Lp Brunn–Minkowski theory is based on the .Lp addition of convex bodies,
introduced by Firey in [35], that we are now going to recall.
Throughout this part we will be working with convex bodies containing the
origin. Recall that
Kno = {K ∈ Kn : o ∈ K}.
.
Moreover, .Kn(o) denotes the family of convex bodies containing the origin in their
interior. Note that .K ∈ Kno if and only if .hK ≥ 0, and .K ∈ K(o) if and only if
.hK (x) > 0 for every .x /= o.
M0 (a, b; t) := a 1−t bt .
.
Mp (a, b; t) := 0.
.
Remark 3.47 Let .p, q ∈ [−∞, ∞] with .p ≤ q. For every .t ∈ [0, 1] and for every
a, b ≥ 0:
.
Proposition 3.48 Let .p ≥ 1. For every .K, L ∈ Kno , and for every .α, β ≥ 0, the
function .h : Sn−1 → R+ defined by
( )1/p
h(x) = αhK (x)p + βhL (x)p
.
α · K +p β · L,
.
whence
α · K +p β · L ⊂ α · K +q β · L
.
124 A. Colesanti and D. Hug
(clearly, in the previous relations, the symbol “.·” is the product corresponding to the
Lp addition on the left-hand side, and the product corresponding to the .Lq addition
.
In analogy with the formula which expresses the first variation of the volume with
respect to the Minkowski addition, (see formula (1.2) in Chap. 1), we present the
corresponding formula for the .Lp addition.
Theorem 3.52 Let .K, L ∈ Kno , and let .p ≥ 1. Then
f
V (K +p ε · L) − V (K) 1
. lim = hL (u)p hK (u)1−p dSn−1 (K, u).
ε→0+ ε p Sn−1
(3.21)
tells us that the .Lp addition behaves linearly with respect to the pth power of support
functions. Hence formula (3.21) shows that the measure
1−p
hK (·)Sn−1 (K, ·)
.
is the first variation of the volume with respect to the .Lp addition. This measure
is called the .Lp surface area measure of K. A corresponding Minkowski-type
problem can be posed; for more details we refer the reader to [59, Chapter 9].
hλ·K = λ1/p hK .
.
Hence the multiplication by a non-negative real number .λ in the .Lp sense, coincides
with the standard dilation of a factor .λ1/p . In particular this changes the degree of
homogeneity of the volume: for .K ∈ Kno and .λ ≥ 0,
. Vn (λ · K) = λn/p Vn (K).
3 Geometric and Functional Inequalities 125
Hence the inequality is proved under the assumption that .K0 and .K1 have volume
1. The general case follows from a standard argument based on homogeneity (see
also (3.13) and the related remark). u
n
In this part we extend the definition of .Lp addition (or better, of .Lp -convex linear
combination) to the case .0 ≤ p < 1. The construction requires the notion of a Wulff
shape, also known as the Aleksandrov body.
Clearly, .K[f ] is closed and convex, being the intersection of closed half-spaces.
Moreover, as f is continuous on .Sn−1 , it is bounded; this implies that .K[f ] is
bounded. Note also that .o ∈ K[f ], since .f ≥ 0; if we assume that f is strictly
positive, then the origin is an interior point of K. We conclude that .∅ /= K[f ] ∈ Kn .
The convex body .K[f ] is called the Wulff shape or the Aleksandrov body of f .
It is easy to see that
hK[f ] ≤ f
.
126 A. Colesanti and D. Hug
and we have equality (for every point of .Sn−1 ) if and only if the 1-homogeneous
extension of f is a convex functions (i.e. f is a support function). .K[f ] can also be
characterized as follows: its support function is the largest 1-homogeneous convex
function which is bounded from above by f .
that is, .ft is the p-mean of the support functions of .K0 and .K1 (with weights .1 − t
and t). If .p < 1, in general this is not a support function. The p convex linear
combination of .K0 and .K1 is defined as
(1 − t) · K0 +p t · K1 := K[ft ].
.
(1−t)·K0 +p t ·K1 = {x ∈ Rn : <x, u> ≤ Mp (hK0 (u), hK1 (u); t) for every u ∈ Sn−1 }.
.
This is the so-called log Minkowski convex linear combination of .K0 and .K1 .
Exercise 3.55 Let .K0 and .K1 be two orthogonal segments in the plane, centered at
the origin. What is
1 1
. · K0 +0 · K1 ?
2 2
In [18], Böröczky, Lutwak, Yang and Zhang proved the following version of the
Brunn–Minkowski inequality for the 0 addition, in the two-dimensional case.
Theorem 3.56 Let .K0 , K1 ∈ K2o be centrally symmetric with respect to the origin.
If .t ∈ [0, 1], then
For .t ∈ (0, 1) equality holds if and only if .K0 and .K1 are dilates of each other, or
they are parallelograms with parallel sides.
Remark 3.57 Similarly to the case .p ≥ 1, by the monotonicity of p-means
with respect to p (see Remark 3.10), and by the definition of .Lp convex linear
combination for .p ≤ 1, we have the inclusion
. (1 − t) · K0 +0 t · K1 ⊂ (1 − t) · K0 +p t · K1
This conjecture has been intensively studied in recent years, and much progress
has been made. An updated survey on the state of the art can be found in [46].
In this part we present a result due to Saroglou [58], which proves the validity of the
log-Brunn–Minkowski conjecture for unconditional convex bodies.
A convex body .K ∈ Kn is called unconditional if it is symmetric with respect to
all coordinate hyperplanes of .Rn :
x = (x1 , . . . , xn ) ∈ K
. ⇒ (±x1 , . . . , ±xn ) ∈ K,
Theorem 3.60 If .K0 , K1 ∈ Kno are unconditional and .t ∈ [0, 1], then
Again, all choices of the signs are considered; in particular, .K01−t K1t is symmetric
with respect to all coordinate hyperplanes. We will prove the following two facts:
(i)
(1 − t) · K0 +0 t · K1 ⊃ K01−t K1t ;
.
(ii)
Then, we need to prove that .z = (z1 , . . . , zn ) ∈ K01−t K1t implies, for .u ∈ Sn−1 ,
As .K01−t K1t , .K0 and .K1 are symmetric with respect to all coordinate hyperplanes, it
is sufficient to consider the case in which all coordinates of z and u are non-negative.
There exist .x = (x1 , . . . , xn ) ∈ K0 and .y = (y1 , . . . , yn ) ∈ K1 (which we may also
assume to have non-negative coordinates) such that, for .i ∈ {1, . . . , n},
zi = xi1−t yit .
.
3 Geometric and Functional Inequalities 129
Then
E
n
. <z, u> = xi1−t yit u1−t
i ui
t
i=1
E
n
= (xi ui )1−t (yi ui )t
i=1
( n )1−t ( )t
E E
n
≤ xi ui yi ui
i=1 i=1
where we used Hölder’s inequality and the definition of support function. Hence we
have proved claim (i).
In order to prove claim (ii), let
and let
(we recall that .1A is the standard indicator function of a set A). We also set
and we give similar definitions for .ḡ = ḡ(ȳ) and .h̄ = h̄(z̄). By the definition of
K01−t K1t , it follows that
.
for .x̄, ȳ ∈ Rn and .z̄ = (1 − t)x̄ + t ȳ. Then, by the Prékopa–Leindler inequality (see
Theorem 3.27),
f (f )1−t (f )t
. h̄(z̄) dz̄ ≥ f¯(x̄) dx̄ ḡ(ȳ) dȳ .
Rn R n
Rn
and, similarly,
f f
. f¯(x̄) dx̄ = f (x) dx = V (K0 ∩ Rn+ ),
Rn Rn+
f f
ḡ( ȳ) dȳ = g(y) dy = V (K1 ∩ Rn+ ).
Rn Rn+
In this section we present some functionals, coming from the world of calculus of
variations and elliptic PDE’s, which, when restricted to convex bodies, surprisingly
share some significant properties with the ordinary volume, including inequalities
of Brunn–Minkowski type, representation formulas, and variational formulas.
Our first and main example will be the electrostatic capacity; the second part of
the section is devoted to other well-known functionals.
The standard definition (see, for instance, [47]) of the electrostatic capacity of a
convex body .K ∈ Kn , .n ≥ 3, is
{f }
. Cap(K) := inf |Tu(x)| dx : u ∈
2
Cc1 (Rn ), u ≥ 1 on K
Rn
(in the 2-dimensional case this notion is naturally replaced by logarithmic capacity,
see, for instance, [30]). Here .Cc1 (Rn ) denotes the set of functions in .C 1 (Rn ) with
compact support. For simplicity, let us assume that K has non-empty interior. Then
.Cap(K) can also be written as
f
. Cap(K) = |Tu(x)|2 dx
Rn \K
The function u is uniquely determined (by problem (3.23) and by the maximum
principle), and it is called the equilibrium potential of K. Another consequence of
the maximum principle is that .0 ≤ u(x) ≤ 1 for every .x ∈ Rn \ K.
As a functional from .Kn to .R, the capacity has the following properties:
. it is continuous with respect to the Hausdorff metric;
. it is rigid motion invariant;
. it is homogeneous of degree .n − 2: for .K ∈ Kn and .λ ≥ 0,
If .K0 , K1 ∈ Knn and equality holds in (3.24), then .K0 and .K1 are homothetic.
The previous theorem is due to Borell (see [15]), who proved (3.24), and to
Caffarelli, Jerison and Lieb (see [27]), who characterized equality conditions.
The similarity between (3.24) and the Brunn–Minkowski inequality for the
volume is clear. Both assert the concavity in .Kn of a certain functional, raised to
the reciprocal of its homogeneity order, with respect to the Minkowski addition.
Here we sketch a proof of the Brunn–Minkowski inequality for the capacity. The
complete argument can be found in [32].
Step 1. Let K be a convex body with non-empty interior in .Rn , .n ≥ 3, and let u
be the equilibrium potential of K, i.e. the solution to problem (3.23). Extend u to
be identically 1 on K. A simple consequence of the maximum principle is that
0<u<1
.
in .Rn \ K. Moreover, u is quasi-concave, i.e. its super-level sets are convex: for
every .s ∈ (0, 1)
Step 2. We have already seen some formulas which permit to compute .Cap(K)
in terms of u. Here is another one, which turns out to be useful in this argument:
.ũt : R → R as follows:
n
It can be easily seen that .ũt is the (uniquely determined) function such that, for
every .s ∈ [0, 1],
In other words, the super-level sets of .ũt are the Minkowski convex linear
combinations of the corresponding super-level sets of .u0 and .u1 . In particular
we have
ũt = 1 on ∂K,
. and lim ũ(x) = 0. (3.27)
|x|→∞
Step 4. The most delicate part of the proof is to compare .ũt with the equilibrium
potential .ut of .Kt . This is done showing that .ũt is a viscosity sub-solution of the
Laplace equation in .Rn \ Kt :
Aũt ≥ 0
. in Rn \ Kt . (3.28)
In order to prove (3.28) one needs to evaluate and estimate the Hessian matrix of
.ũ at a point z, in terms of the Hessian matrices of .u0 at x and of .u1 at y, where
.x, y, z are such that .z = (1 − t)x + ty and
ũt ≤ ut
. (3.29)
in .Rn .
3 Geometric and Functional Inequalities 133
We start this part with a formula for capacity which parallels the representation
formula (1.1) for the volume.
Let .K ∈ Kn , and assume that it has non-empty interior. Let u be the equilibrium
potential of K. Then .Tu can be appropriately extended to almost every point of .∂K,
with respect to the .(n − 1)-dimensional Hausdorff measure restricted to .∂K, and the
extension is such that
Tu ∈ L2 (∂K).
.
We refer the reader to the article [47] for a detailed presentation of these facts. Hence
Cap
we can define a measure .Sn−1 (K, ·) on .Sn−1 , as follows: for every Borel subset .ω
of .Sn−1
f
Cap
.S
n−1 (K, ω) := |Tu(x)|2 dHn−1 (x).
−1
νK (ω)
−1
Recall that .νK is the Gauss map of K, and .νK (ω) is the set of points of .∂K where
the outer unit normal is defined, and it belongs to .ω.
The following property can be proved (if K has non-empty interior):
Cap
Sn−1 (K, ·) is not concentrated on any great subsphere of Sn−1 .
. (3.30)
In this section we present two more functionals defined on .Kn , namely the torsion
and the first Dirichlet eigenvalue of the Laplacian, which, like the capacity, have
several features in common with the volume. In particular,
(a) they satisfy an inequality of Brunn–Minkowski type;
3 Geometric and Functional Inequalities 135
(b) they admit a representation formula and a formula for the first variation, similar
to (3.32) and (3.33);
(c) a Minkowski-type problem has been posed and solved for these functionals.
As for the capacity, these functionals are classical examples in the calculus of
variations.
The Torsion
Let .K ∈ Kn , and assume that K has non-empty interior. The variational definition
of the torsion of K , denoted by .τ (K), is the following:
{f f }
K |Tu(x)| dx
2
1 1,2
. := inf (f )2 : u ∈ W0 (int(K)), |u(x)| dx > 0 .
τ (K) |u(x)| dx K
K
Here, following the standard notation, we denote by .W 1,2 (int(K)) the Sobolev space
of functions with (weak) gradient in .L2 (int(K)), and by .W01,2 (int(K)) the closure
of .Cc∞ (int(K)) with respect to the norm of .W 1,2 (int(K)). The previous minimum
problem admits in fact a minimizer u, which is the unique solution of the following
boundary value problem:
{
Au = −2 in int(K),
. (3.34)
u=0 on ∂K.
The torsion is rigid motion invariant, continuous with respect to the Hausdorff
metric, and homogeneous of order .n + 2: for .K ∈ Kn and .λ > 0,
The Brunn–Minkowski inequality for .τ was established by Borell (see [16]), and
equality conditions were characterized in [29].
Theorem 3.65 Let .K0 , K1 ∈ Kn and let .t ∈ [0, 1]. Then
If .K0 , K1 ∈ Knn and equality holds in (3.35), then .K0 and .K1 are homothetic.
136 A. Colesanti and D. Hug
As in the case of the capacity, the gradient of u admits an .L2 (∂K) extension to
.∂K, and we have the following formulas (see [31, Theorem 3.1]):
f
1
τ (K) =
.
τ
hK (y) dSn−1 (K, y);
n + 2 Sn−1
then there exists a convex body K with non-empty interior such that
τ
Sn−1
. (K, ·) = μ(·).
Let .K ∈ Kn , and assume that K has non-empty interior. Consider the following
minimum problem for the so-called Rayleigh quotient:
{f f }
|Tu(x)|2 dx
λ(K) := inf
. f
K
:u∈W 1,2
(int(K)), 2
u (x) dx > 0 .
2
K u (x) dx K
There exists a minimizer u for this problem, which can be chosen to be the solution
of
⎧
⎨ Au = −λ(K)u in int(K),
. u>0 in K, (3.36)
⎩
u=0 on ∂K.
The (positive) number .λ(K) is called the first Dirichlet eigenvalue of the Laplacian.
It can be also characterized as the best constant .λ such that the following Poincaré
inequality,
f f
1
. u2 dx ≤ |Tu|2 dx
K λ K
If .K0 , K1 ∈ Knn and equality holds in (3.37), then .K0 and .K1 are homothetic.
Following the lines that we have seen for the capacity and for the torsion, a
Minkowski-type problem can be posed for .λ. This problem was firstly considered
by Jerison in [48], who proved the existence of a solution. Uniqueness was then
established in [29], based on the equality conditions in (3.37).
138 A. Colesanti and D. Hug
3.5.6 Remarks
. We have seen that the validity of the classical Brunn–Minkowski inequality goes
beyond convex sets. By this point of view, the picture for the functionals that we
have seen in this part is not complete. The Brunn–Minkowski inequality for the
torsion and for the first Dirichlet eigenvalue of the Laplacian can be extended to
non-convex sets (with sufficiently smooth boundary). A corresponding extension
for the capacity is an interesting open problem.
. The solutions to the boundary value problems (3.23), (3.34) and (3.36) inherit
specific qualitative properties from the convexity of the domain K (we have
already seen it in the case of the capacity). The capacitary equilibrium is quasi-
concave; the solution of (3.34) is power concave, the solution of (3.36) is
log-concave (see [29] for references about these results). It is interesting to
notice that the argument of many of the proofs of geometric properties of the
solutions, can be adapted to prove the Brunn–Minkowski inequality for the
relevant functionals.
. The results that we have seen in this section have been object of numerous
extensions. References to a first group of results, where Cap, τ or λ are replaced
by other variational functionals (like p-capacity, p-torsion, eigenvalues of other
elliptic operators), can be found in [29]. More recently, extensions where the
usual Minkowski addition is replaced by the p addition have been studied. The
literature in this area is rapidly growing; possible references are [50, 62, 65].
In this section, we discuss a dual pair of analytic inequalities which had a major
impact on convex geometry. Conversely, research in convexity has motivated to a
large extent the further development of these and related analytic inequalities.
f || k (f )ci
1 ||
k
. fi (B i x) ci
dx ≤ √ f (x
i i ) dxi
Rn i=1 D i=1 Rni
and
f { k }
∗ || E
k
. sup fi (zi ) : x =
ci
ci Bi∗ zi , zi ∈R
ni
dx
Rn i=1 i=1
√ k (f
|| )ci
≥ D fi (xi ) dxi ,
i=1 Rni
where
⎧ (E ) ⎫
⎨ det k ∗ ⎬
i=1 ci Bi Ai Bi
.D := inf ||k : Ai ∈ Rpd (ni , ni )
⎩ ci ⎭
i=1 (det Ai )
and .Rpd (ni , ni ) are the positive definite (real) .ni × ni -matrices.
Remarks
n
. If ki=1 Ker(Bi ) /= {o}, then D = 0 and the stated inequalities hold trivially.
E
. If ni = n, Bi = In , ci = 1/pi and ki=1 1/pi = 1 and if fi is replaced by
1/c
fi i then BLI turns into Hölder’s inequality. In this situation we have D = 1.
To see this, one can use an inequality for determinants: Let Ai ∈ Rpd (n, n),
E
i = 1, . . . , k, and ci > 0 with ki=1 ci = 1. Then
( k )
E ||
k
. det ci Ai ≥ (det Ai )ci
i=1 i=1
. Meanwhile continuous versions [6, 8, 25] and extensions to other spaces [9, 21]
have been dealt with by various authors, and a variety of arguments has been
developed in order to study the structure and to obtain generalizations of
Brascamp–Lieb and Barthe type inequalities and to resolve the problem of
obtaining a classification of the cases where equality occurs in these inequalities
(see [8, 11, 12, 28, 34]).
. From a geometric viewpoint, the following special rank one case is useful:
Choose ni = 1, let Bi : Rn → R be given by Bi (x) := <x, ui > for some ui ∈
Rn \ {o}, where span{u1 , . . . , uk } = Rn (otherwise D = 0). Let ci > 0 with
E k ∗ ∗
i=1 ci = n and Bi : R → R , Bi (t) = tui , which is the Euclidean adjoint
n
f || k (f )ci
1 ||
k
. fi (<x, ui >) ci
dx ≤ √ fi (t) dt
Rn i=1 D i=1 R
and
f { } (f )ci
∗ ||
k E
k √ ||
k
. sup fi (zi ) : x =
ci
ci zi ui , zi ∈ R dx ≥ D fi (t) dt ,
Rn i=1 i=1 i=1 R
where
⎧ (E ) ⎫
⎨ det k
i=1 ci ai u i ⊗ ui ⎬
.D = inf ||k : ai > 0 .
⎩ ci ⎭
i=1 ai
E
k
. ci ui ⊗ ui = id. (3.38)
i=1
Then
f ||
k k (f
|| )ci
. fi (<x, ui >) dx ≤
ci
fi (t) dt
Rn i=1 i=1 R
3 Geometric and Functional Inequalities 141
and
f { } k (f )ci
∗ ||
k E
k ||
. sup fi (zi ) : x =
ci
ci zi ui , zi ∈ R dx ≥ fi (t) dt .
Rn i=1 i=1 i=1 R
Remarks
and span{u1 , . . . , uk } = Rn .
We show that D = 1. Choosing a1 = · · · = ak = a > 0, we get
( k ) ( )
E E
k
. det ci ai ui ⊗ ui = det a · ci ui ⊗ ui = an
i=1 i=1
142 A. Colesanti and D. Hug
and
||
k Ek
. aici = a i=1 ci = an,
i=1
√ E
hence D ≤ 1. Next we show that D ≥ 1. Set wi := ci ui . Then ki=1 wi ⊗wi = id
and thus
( k )
E ( )
.1 = det(id) = det wi ⊗ wi = det W W T , W := (w1 . . . wk ) ∈ Rn,k .
i=1
where (3.39) and the inequality of arithmetic and geometric means were used. The
exponent of ai in the product on the right side of (3.40) is given by
⎛ ⎞
E E E Ek
. WI = WI − WI = 1 − det ⎝ wj ⊗ wj − wi ⊗ wi ⎠
i∈I,|I |=n |I |=n i ∈I,|I
/ |=n j =1
) (
= 1 − det (In − wi ⊗ wi ) = 1 − 1 − |wi |2 = |wi |2 = ci .
E
k
. ci ui ⊗ ui = id.
i=1
3 Geometric and Functional Inequalities 143
Ek
(ii) If z = i=1 ci θi ui for θ1 , . . . , θk ∈ R, then
E
k
|z|2 ≤
. ci θi2 .
i=1
For the proof of Theorem 3.69 we first assume that each fi is a positive
continuous probability density. Let g(t) = e−π t for t ∈ R be the Gaussian density.
2
E
k
o(x) =
. ci Ti (<ui , x>) ui , x ∈ Rn ,
i=1
which satisfies
E
k
do(x) =
. ci Ti' (<ui , x>) ui ⊗ ui .
i=1
Therefore, using first (3.41), then Lemma 3.70 (i) with ti = Ti' (<ui , x>), the
definition of o and Lemma 3.70 (ii), and finally the transformation formula, the
following argument leads to the BLI
f ||
k
. fi (<ui , x>)ci dx
R
n
i=1
f ( k )( )
|| ||
k
= g(Ti (<ui , x>)) ci
Ti' (<ui , x>)ci dx
Rn i=1 i=1
f ( k ) ( k )
|| E
−π ci Ti (<ui ,x>)2
≤ e det ci Ti' (<ui , x>) ui ⊗ ui dx
Rn i=1 i=1
f
e−π |o(x)| det (do(x)) dx
2
≤
R n
f
e−π |y| dy = 1.
2
≤
R n
The BLI for arbitrary non-negative integrable functions fi follows by scaling and
approximation.
For the reverse BLI (BI), we consider the inverse Si of Ti , and hence
f t f Si (t)
. g(s) ds = fi (s) ds,
−∞ −∞
whence
In addition,
E
k
dw(x) =
. ci Si' (<ui , x>) ui ⊗ ui
i=1
E
k
w(x) =
. ci Si (<ui , x>) ui , x ∈ Rn .
i=1
Therefore, the transformation formula, Lemma 3.70 (i), and (3.42) imply that
f ∗ ||
k
. sup fi (θi )ci dx
R
n E
x= ki=1 ci θi ui i=1
⎛ ⎞
f ∗ ||
k
≥ ⎝ sup fi (θi )ci ⎠ det (dw(y)) dy
Rn E
w(y)= ki=1 ci θi ui i=1
f ( k ) ( k )
|| E
≥ fi (Si (<ui , y>)) ci
det ci Si' (<ui , y>) ui ⊗ ui dy
Rn i=1 i=1
f ( k )( )
|| ||
k
≥ fi (Si (<ui , y>)) ci
Si' (<ui , y>)ci dy
Rn i=1 i=1
f ( k ) f
||
e−π |y| dy = 1.
2
= g(<ui , y>) ci
dy =
Rn
i=1 R n
For the proof, we consider for i ∈ {1, . . . , k} the indicator function fi (t) :=
1[−βi /ci ,βi /ci ] (t), t ∈ R, which is 1 if t ∈ [−βi /ci , βi /ci ] and zero otherwise. If
x ∈ Rn is such that
{ k }
|| E
k
.1 = sup fi (zi ) : x = ci zi ui , zi ∈ R
ci
i=1 i=1
{ }
||
k E
k
= sup 1[−βi ,βi ] (zi ci ) : x = ci zi ui , zi ∈ R ,
i=1 i=1
146 A. Colesanti and D. Hug
Ek
then x ∈ i=1 βi [−ui , ui ]. Hence we deduce from BI that
( k ) f { }
E ∗ ||
k E
k
V
. βi [−ui , ui ] ≥ sup fi (zi ) : x =
ci
ci zi ui , zi ∈ R dx
i=1 Rn i=1 i=1
k (f
|| )ci m (
|| ) k ( )ci
||
2βi ci βi
≥ fi (t) dt = =2 n
.
i=1 R i=1
ci
i=1
ci
{ } p1
E
k
||x||α,p :=
. αi |<ui , x>| p
, x ∈ Rn ,
i=1
K := {x ∈ Rn : ||x||α,p ≤ 1},
.
2n
V (K) =
. (α1 · · · αn )−1 .
n!
For general parameters, we now provide the sharp upper bound
( )n
2n r 1 + p1 || k ( ) ci
ci p
.V (K) ≤ ( ) ,
r 1 + pn αi
i=1
for which equality is achieved in the above special case of a cross polytope.
3 Geometric and Functional Inequalities 147
To verify this, we first derive a general representation for the volume of a convex
body K which contains the origin in its interior. We claim that
f
1 ( p)
V (K) =
. ( ) exp −||x||K dx (3.43)
r 1+ n R
n
p
f || k
1
= ( ) fi (<ui , x>)ci dx
r 1+ p
n R n
i=1
m (f
|| )ci
1
≤ ( ) fi (t) dt
r 1+ n
i=1 R
p
[ ( )1 ( )]ci
1 ||
k
ci p 1
= ( ) 2 r 1+ ,
r 1 + pn i=1 αi p
E
which yields the claim since ki=1 ci = n.
Again equality holds if and only if K is a cross polytope.
In this section, we consider n-dimensional compact convex sets in .Rn . The class
of these bodies was denoted by .Knn . It has already been emphasized that the
isoperimetric ratio .S(K)n /V (K)n−1 is minimized among all .K ∈ Knn by Euclidean
balls, but the functional .S n /V n−1 on .Knn is clearly unbounded from above. This
can be seen, for .n = 2, by considering a sequence of rectangles with increasingly
larger aspect ratios. To prevent such a degeneration, we consider for each convex
body .K ∈ Knn a transformation of K by an invertible linear map for which the
isoperimetric ratio is as small as possible. More formally, we define
{ }
S(oK)n
. ir(K) := inf : o ∈ GL(n) .
V (oK)n−1
Here the minimization is over all bijective linear transformations (translations are
irrelevant, since volume and surface area are translation invariant).
The new functional .ir(·) has the following properties:
. .ir is affine invariant, upper semi-continuous and hence attains its maximum.
. Petty [55] (see also Giannopoulos, Papadimitrakis [37]) has shown the following
facts: The infimum is attained, in fact, there is a unique .K0 ∈ {oK : o ∈ GL(n)},
the .GL(n) class of K, such that
S(K0 )n
. ir(K) = .
V (K0 )n−1
3 Geometric and Functional Inequalities 149
The determination of the maximum for the new functional .ir(·) and of its
extremizers turned out to be an inspiring problem. In the Euclidean plane, the
solution was achieved by elementary geometric arguments (see Gustin [41] and
Behrend [10]). In the following, we write .T2 for a regular triangle and .W2 for a
square in .R2 . Since the functionals under consideration are affine invariant, we can
assume that .T2 and .W2 are circumscribed to the unit ball .B 2 .
Theorem 3.71 (Gustin, Behrend (Symmetric Case))
(a) If .K ∈ K22 , then
. ir(K) ≤ ir(T2 ).
. ir(K) ≤ ir(W2 ).
δBM (K, L):= log min{λ≥1 : K − x⊂o(L − y) ⊂ λ(K − x), o∈ GL(n), x, y∈Rn }.
.
If .K, L ∈ Knn are both symmetric with respect to the origin, that is, if .K = −K and
.L = −L, this definition simplifies, and we arrive at
If K and L are origin symmetric, the translations in the definition of .δvol (K, L) are
omitted.
These definitions induce metrics on the affine equivalence classes of convex
bodies, where two convex bodies .K, L ∈ Knn are said to be affinely equivalent,
if .K = αL for an invertible affine transformation .α of .Rn (in the case of origin
symmetric convex bodies, we consider linear equivalence classes). We refer to [17,
Lemma 8.2] and the literature cited there for relations between .δvol and .δBM .
Theorem 3.72 (Böröczky, Hug and Böröczky, Fodor, Hug (Symmetric Case))
(a) Let .K ∈ K22 , and let .ε ∈ [0, 1/72). If .δBM (K, T2 ) ≥ 72 ε, then
. ir(K) ≤ (1 − ε) ir(T2 ).
(b) Let .K ∈ K22 be origin symmetric, and let .ε ∈ [0, 1). If .δvol (K, W2 ) ≥ ε or
.δBM (K, W2 ) ≥ ε, then
( ε )
. ir(K) ≤ 1 − ir(W2 ).
54
Theorem 3.73
(a) For each .K ∈ Knn there is a unique ellipsoid .EJ (K) which is contained in K
and has maximal volume among all ellipsoids contained in K.
(b) For each .K ∈ Knn there is a unique ellipsoid .EL (K) which contains K and has
minimal volume among all ellipsoids containing K.
(c) Both ellipsoids are affinely associated with K, that is, if .α is an invertible affine
transformation of .Rn and .K ∈ Knn , then .EJ (αK) = αEJ (K) and .EL (αK) =
αEL (K).
Next we state John’s classical result [44] (see also [3, Proposition 1]) which
provides necessary and sufficient conditions which ensure that the unit ball is the
John ellipsoid (the Loewner ellipsoid) of .K ∈ Knn . These conditions involve the
common contact points of the unit sphere and the boundary of K.
Theorem 3.74 (F. John, K. Ball)
(a) Let .K ∈ Knn and .EJ (K) = B n . Then there are .s ∈ N with .n + 1 ≤ s ≤
2 n(n + 3), .c1 , . . . , cs > 0 and .u1 , . . . , us ∈ S ∩ ∂K such that
1 n−1
E
s E
s
. ci ui = o and ci ui ⊗ ui = id. (3.44)
i=1 i=1
The converse is also true. If ⊂ K and there are .s, ci , ui as above, then
.B
n
EJ (K) = B n .
.
(b) If .K = −K, then the centeredness condition is omitted and .n ≤ s ≤ 12 n(n + 1).
(c) An analogous result holds for the Loewner ellipsoid. Moreover, .EJ (K) = B n if
and only if .EL (K ◦ ) = B n , where .K ◦ is the polar body of K.
In honor of Fritz John, the decomposition of the identity in (3.44) has been
called John decomposition (of the identity). Various proofs for Theorem 3.74 and
characterization results of John-type for general pairs of convex bodies have been
suggested (see, e.g., Gordon, Litvak, Meyer, Pajor [38] and Gruber, Schuster [40]).
For a given convex body .K ∈ Knn , we consider the inner and the outer volume ratio
( )1 ( )1
V (K) n V (EL (K)) n
. vr(K) := , VR(K) := .
V (EJ (K)) V (K)
Clearly, the functionals .vr and .VR are
. bounded from below by 1 (sharp) and bounded from above,
. affine invariant,
. continuous.
152 A. Colesanti and D. Hug
The following theorem solves the problem of maximizing the volume ratio
vr. The result is due to Ball [1, 2], the characterization of the equality cases is
.
due to Barthe [5]. We write .[K] for the class of all convex bodies which are
affinely equivalent to K (that is, which are obtained from K by an invertible
affine transformation). We write .Tn for a regular simplex circumscribed to the
unit ball .B n and .Qn for a cube circumscribed to .B n . Hence the class .[Tn ]
contains all n-dimensional simplices and the class .[Qn ] contains all n-dimensional
parallelepipeds.
Theorem 3.75 (Volume Ratios, Ball, Barthe)
(a) If .K ∈ Knn , then .vr(K) ≤ vr(Tn ) with equality if and only if .[K] = [Tn ].
(b) If .K ∈ Knn and .K = −K, then .vr(K) ≤ vr(Qn ) with equality if and only if
.[K] = [Qn ].
Proof We start with the easier part (b). We show: If .K = −K, .EJ (K) = B n and
.Qn is a cube centred at o with edge length 2, then
V (K) ≤ V (Qn ) = 2n .
.
In fact, by the characterization theorem for .EJ (K) we get .k ≥ n, .c1 , . . . , ck > 0
and vectors .u1 , . . . , uk ∈ Sn−1 ∩ ∂K with
E
k
. ci ui ⊗ ui = id and ui ∈
/ {±uj : j /= i} for i = 1, . . . , k.
i=1
nk − (u , 1),
By elementary geometry .K ⊂ i=1 H i and by symmetry even
n
k n
k
K⊂
. H − (ui , 1) ∩ H − (−ui , 1) =: L. (3.45)
i=1 i=1
||
k ||
k
1L (y) =
. 1[−1,1] (<y, ui >) = fi (<y, ui >)ci .
i=1 i=1
3 Geometric and Functional Inequalities 153
Hence, first using (3.45) and then Theorem (3.69), we deduce that
f ||
k
V (K) ≤ V (L) =
. fi (<x, ui >)ci dx
R n
i=1
k (f
|| )ci
≤ fi (t) dt
i=1 R
||
k
= 2ci = 2n . (3.46)
i=1
If .V (K) = 2n , then .(3.46) must hold with equality. Since .fi /≡ 0 and .fi are not
Gaussian, the equality condition in the geometric BLI (due to Barthe) implies that
.k = n and .u1 , . . . , un are an orthonormal basis of .R , hence .L = Qn . On the other
n
E
k
. di vi ⊗ vi = idRn+1 .
i=1
and define
||
k
F (x) :=
. fi (<vi , x>)di , x ∈ Rn+1 .
i=1
154 A. Colesanti and D. Hug
f k (f
|| )di
. F (x) dx ≤ fi (t) dt = 1. (3.47)
Rn+1 i=1 R
We next determine the value of the left side of (3.47) by means of Fubini’s theorem.
For this we write
x = (y, r) ∈ Rn × R ∼
. = Rn+1
such that
/
r n
<vi , x> = √
. − <ui , y>.
n+1 n+1
For each .y ∈ Rn there is a .j ∈ {1, . . . , k} with .<uj , y> ≥ 0. In fact, if not then
.<ui , y> < 0 for .i = 1, . . . , k, hence
/ k \
E
k E
0>
. ci <ui , y> = ci ui , y = 0, a contradiction.
i=1 i=1
Consider .x = (y, r) ∈ Rn+1 . If .r < 0, then .<vj , x> < 0 for some .j ∈ {1, . . . , k},
and hence .F (x) = 0. If .r ≥ 0, then .F (x) /= 0 if and only if .<vj , x> ≥ 0 for
.j = 1, . . . , k, hence if and only if
r r
<uj , y> ≤ √
. for j = 1, . . . , k or y ∈ √ L.
n n
{ }
E
k { √ }
F (x) = exp −
. di <vi , x> = exp −r n + 1 .
i=1
( )n f ∞ ( )n
1 2 t dt
= V (L) √ e−t √
n 0 n + 1 n +1
n!
= V (L) · n n+1
n (n + 1)
2 2
V (L) V (K)
= n
≥ .
V (T ) V (T n )
If .V (K) = V (Tn ), then also .V (L) = V (Tn ), and equality must hold in (3.47).
Hence .v1 , . . . , vk are among the vertices of an origin symmetric regular cross
polytope in .Rn+1 . This yields .k = n + 1 and .v1 , . . . , vn+1 is an orthonormal basis in
.R
n+1 . From .<v , v > = δ it follows that .<u , u > + 1 = 0 for .i /= j , hence .L = T .
i j ij i j n n
Since also .K ⊂ L and .V (K) = V (L), we conclude that .K = L = Tn . u
n
Similar results hold for outer volume ratios. Then the analysis is based on the BI.
Having resolved the extremal problem for the volume ratio, it is now surprisingly
easy to resolve also the reverse isoperimetric problem.
Theorem 3.76 (Reverse Isoperimetric Inequality, Ball, Barthe)
(a) If .K ∈ Knn , then .ir(K) ≤ ir(Tn ) with equality if and only if .[K] = [Tn ].
(b) If .K ∈ Knn and .K = −K, then .ir(K) ≤ ir(Qn ) with equality if and only if
.[K] = [Qn ].
Proof
(a) Let .K ∈ Knn and choose an affine transformation .α of .Rn such that .αEJ (K) =
B n = EJ (αK) ⊂ αK. Then
V (αK + εB n ) − V (αK)
S(αK) = lim
.
ε↓0 ε
V (αK + εαK) − V (αK)
≤ lim
ε↓0 ε
(1 + ε)n − 1
= V (αK) lim =
ε↓0 ε
n−1 1
= V (αK)n = V (αK) n nV (αK) n
n−1 1
≤ V (αK) n nV (Tn ) n ,
where we used the preceding theorem on volume ratios for the last inequality.
Hence
S(αK)n S(Tn )n
. ir(K) ≤ ≤ nn
V (T n ) = = ir(Tn ).
V (αK)n−1 V (Tn )n−1
156 A. Colesanti and D. Hug
S(K)n 4 S(T )
n n
. ≤ (1 − γ ε ) ,
V (K)n−1 V (T n )n−1
References
1. K. Ball, Volumes of sections of cubes and related problems, in Geometric Aspects of Functional
Analysis. Lectures Notes in Mathematics, vol. 1376 (Springer, Berlin, 1989), pp. 251–260
2. K. Ball, Volume ratios and a reverse isoperimetric inequality. J. Lond. Math. Soc. 44, 351–359
(1991)
3. K. Ball, Ellipsoids of maximal volume in convex bodies. Geom. Dedicata 41, 241–250 (1992)
4. K. Ball, Convex geometry and functional analysis, in Handbook of the Geometry of Banach
Spaces, vol. I (North-Holland, Amsterdam, 2001), pp. 161–194
5. F. Barthe, On a reverse form of the Brascamp-Lieb inequality. Invent. Math. 134, 335–361
(1998)
6. F. Barthe, A continuous version of the Brascamp–Lieb inequalities, in Geometric Aspects of
Functional Analysis. Lecture Notes in Mathematics, vol. 1850 (Springer, Berlin, 2004), pp.
53–63
7. F. Barthe, P. Wolff, Positive Gaussian kernels also have Gaussian minimizers. Mem. Am. Math.
Soc. 276(1359), v+90 pp. (2022)
8. F. Barthe, D. Cordero-Erausquin, Inverse Brascamp–Lieb inequalities along the heat equation,
in Geometric Aspects of Functional Analysis. Lecture Notes in Mathematics, vol. 1850
(Springer, Berlin, 2004), pp. 65–71
9. F. Barthe, D. Cordero-Erausquin, M. Ledoux, B. Maurey, Correlation and Brascamp–Lieb
inequalities for Markov semigroups. Int. Math. Res. Not. 10, 2177–2216 (2011)
10. F. Behrend, Über einige Affininvarianten konvexer Bereiche (German). Math. Ann. 113, 713–
747 (1937)
11. J. Bennett, A. Carbery, M. Christ, T. Tao, The Brascamp–Lieb inequalities: finiteness, structure
and extremals. Geom. Funct. Anal. 17, 1343–1415 (2017)
12. J. Bennett, N. Bez, S. Buschenhenke, M.G. Cowling, T.C. Flock, On the nonlinear Brascamp–
Lieb inequality. Duke Math. J. 169, 3291–3338 (2020)
3 Geometric and Functional Inequalities 157
13. F. Besau, S. Hoehner, An intrinsic volume metric for the class of convex bodies in Rn .
Commun. Contemp. Math. (in press)
14. V.I. Bogachev, Measure Theory, vol. II (Springer, Berlin, 2007)
15. C. Borell, Capacitary inequalities of the Brunn–Minkowski type. Math. Ann. 263, 179–184
(1983)
16. C. Borell, Greenian potentials and concavity. Math. Ann. 272, 155–160 (1985)
17. K.J. Böröczky, D. Hug, Isotropic measures and stronger forms of the reverse isoperimetric
inequality. Trans. Am. Math. Soc. 369, 6987–7019 (2017)
18. K.J. Böröczky, E. Lutwak, D. Yang, G. Zhang, The log-Brunn–Minkowski inequality. Adv.
Math. 231, 1974–1997 (2012)
19. K.J. Böröczky, F. Fodor, D. Hug, Strengthened volume inequalities for Lp zonoids of even
isotropic measures. Trans. Am. Math. Soc. 371, 505–548 (2019)
20. K.J. Böröczky, F. Fodor, D. Hug, Strengthened inequalities for the mean width and the l-norm.
J. Lond. Math. Soc. (2) 104, 233–268 (2021)
21. R. Bramati, Brascamp–Lieb inequalities on compact homogeneous spaces. Anal. Geom. Metr.
Spaces 7, 130–157 (2019)
22. H.J. Brascamp, E.H. Lieb, Best constants in Young’s inequality, its converse and its general-
ization to more than three functions. Adv. Math. 20, 151–173 (1976)
23. H.J. Brascamp, E.H. Lieb, On extensions of the Brunn–Minkowski and Prékopa–Leindler
theorems, including inequalities for log concave functions, and with an application to the
diffusion equation. J. Funct. Anal. 22, 366–389 (1976)
24. H.J. Brascamp, E.H. Lieb, J.M. Luttinger, A general rearrangement inequality for multiple
integrals. J. Funct. Anal. 17, 227–237 (1974)
25. S. Brazitikos, A. Giannopoulos, Continuous version of the approximate geometric Brascamp–
Lieb inequalities. J. Geom. Anal. 32, Paper No. 174, 23 pp. (2022)
26. E.M. Bronshtein, L.D. Ivanov, The approximation of convex sets by polytopes. Siberian Math.
J. 16, 852–853 (1975)
27. L. Caffarelli, D. Jerison, E. Lieb, On the case of equality in the Brunn–Minkowski inequality
for capacity. Adv. Math. 117, 193–207 (1996)
28. E.A. Carlen, D. Cordero-Erausquin, Subadditivity of the entropy and its relation to Brascamp–
Lieb type inequalities. Geom. Funct. Anal. 19, 373–405 (2009)
29. A. Colesanti, Brunn–Minkowski inequalities for variational functionals and related problems.
Adv. Math. 194, 105–140 (2005)
30. A. Colesanti, P. Cuoghi, The Brunn–Minkowski inequality for the n-dimensional logarithmic
capacity of convex bodies. Potential Anal. 85, 45–66 (2006)
31. A. Colesanti, M. Fimiani, The Minkowski problem for torsional rigidity. Indiana Univ. Math.
J. 59, 1013–1039 (2010)
32. A. Colesanti, P. Salani, The Brunn–Minkowski inequality for p-capacity of convex bodies.
Math. Ann. 327, 459–479 (2003)
33. D. Cordero-Erausquin, B. Klartag, Q. Merigot, F. Santambrogio, One more proof of the
Alexandrov–Fenchel inequality. C. R. Math. Acad. Sci. Paris 357, 676–680 (2019)
34. T.A. Courtade, J. Liu, Euclidean forward-reverse Brascamp–Lieb inequalities: finiteness,
structure, and extremals. J. Geom. Anal. 31, 3300–3350 (2021)
35. W.J. Firey, p-Means of convex bodies. Math. Scand. 10, 17–24 (1962)
36. R.J. Gardner, The Brunn-Minkowski inequality. Bull. Am. Math. Soc. (N.S.) 39, 355–405
(2002)
37. A. Giannopoulos, M. Papadimitrakis, Isotropic surface area measures. Mathematika 46, 1–3
(1999)
38. Y. Gordon, A.E. Litvak, M. Meyer, A. Pajor, John’s decomposition in the general case and
applications. J. Differ. Geom. 68, 99–119 (2004)
39. P.M. Gruber, Convex and Discrete Geometry. Grundlehren der mathematischen Wissen-
schaften, vol. 336 (Springer, Berlin, 2007)
40. P.M. Gruber, F.E. Schuster, An arithmetic proof of John’s ellipsoid theorem. Arch. Math.
(Basel) 85, 82–88 (2005)
158 A. Colesanti and D. Hug
Shiri Artstein-Avidan
Abstract In these three lectures we shall review and discuss in detail the fascinating
connection between duality transforms (mainly Legendre and polarity, but also those
associated with other cost functions), measure transportation, cost sub-gradient
mappings and concentration inequalities. The topics of the three letures are:
1. Introduction
2. Transportation of measure
3. Concentration of measure
In the introductory lecture we review some background topics that put the main
subjects of the course in context, and these are measure transportation—the topic of
the second lecture, and measure concentration—the topic of the third one.
The two main objects of study are convex bodies and convex functions. One
could offer a whole semester working only on the basic facts about these objects, as
there is a vast literature, in particular the textbooks of Schneider [48], Gruber [27],
Gardner [26], Hug and Weil [30] and my books with Giannopoulos and Milman [6]
and [8]. The topic of duality, present in the title of our lecture series, will be present
throughout the talks as a connecting theme, to which we point every now and again.
We start with a few facts, and even fewer proofs, which will put the main topic of
the course in context, and are useful and beautiful on their own right. Let us mention
that the intersection of this lecture with the other courses of this program is quite
large, but one should see this, hopefully, not as a bug but rather as a feature, helping
readers absorb the material better.
S. Artstein-Avidan (O)
Tel Aviv University, Tel Aviv, Israel
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 159
A. Colesanti, M. Ludwig (eds.), Convex Geometry, C.I.M.E. Foundation Subseries
2332, https://doi.org/10.1007/978-3-031-37883-6_4
160 S. Artstein-Avidan
(usually one assumes that .A + B measurable as well, and writes .Vol(A + B)1/n ≥
Vol(A)1/n + Vol(B)1/n ).
Many proofs exist for this theorem, the simplest one, which was given by Hadwiger
and Ohmann [28] uses induction, and approximation by finite unions of disjoint
boxes.
Proof We start with the case in which A and B are coordinate ||boxes, and as the
inequality is invariant to translation we may assume .A = [0, ai ] and .B =
|| ||
[0, bi ] in which case .A + B = [0, ai + bi ]. The inequality then amounts to
( n )1/n ( n )1/n ( n )1/n
|| || ||
. (ai + bi ) ≥ ai + bi , (4.1)
i=1 i=1 i=1
To prove the inequality in the case where A and B are unions of disjoint coordinate
boxes we use an induction argument on the total number of boxes in A and B
together. If there are m boxes altogether, we just solved for .m = 2, so .m ≥ 3 and
we assume to know the theorem when A and B are unions of disjoint coordinate
boxes with altogether at most .(m − 1) boxes. So, at least one of the two, say A, has
at least two boxes in it. Pick a coordinate hyperplane .H = {xi = c} which separates
two of these boxes (it exists as the boxes are disjoint). Let .H + = {x : xi ≥ c} and
.H
− = {x : x ≤ c}, and let .A+ = A ∩ H + , .A− = A ∩ H − . We next make the same
i
splitting for B, using a hyperplane .E = {x : xi = d} parallel to H , with d chosen
in such a way that
Vol(A ∩ H − ) Vol(B ∩ E − )
. = .
Vol(A) Vol(B)
observation is that .A+ + B + and .A− + B − are disjoint, and are both Minkowski
sums of a pair of sets which satisfy the induction hypothesis. This gives
completing the proof. For the general case we approximate A and B by disjoint
unions of coordinate boxes and use the continuity of volume. u
n
Let us stop for a moment to appreciate what we have seen. The main ingenious
part of the proof is in the induction step where, after choosing the hyperplane H ,
162 S. Artstein-Avidan
so as to allow induction, we choose a parallel E in such a way that the measure (of
B) to its left and to its right correspond to the measures (of A) to the left and to the
right of H . This is a very primitive “forefather” of transportation of measure, as we
shall see shortly.
Before moving on to the next topic, let us mention that by homogeneity the
Brunn–Minkowski inequality can also be written as
for any .λ ∈ (0, 1) which by the geometric arithmetic means inequality implies
It is an easy exercise to show that knowing the latter for all .λ ∈ (0, 1) implies
the original formulation of the Brunn–Minkowski inequality for the same A and B
(simply choose the right .λ and use homogeneity).
Another very famous and basic inequality for convex bodies is concerned with the
operation of duality for convex bodies. Given a convex body which includes the
origin, it defines a gauge function
x
||x||K = inf{r > 0 :
. ∈ K},
r
which is positively 1-homogeneous and convex. If .0 ∈ int(K) then this function is
finite, if K is bounded then the function is bounded away from 0, and if in addition
.K = −K then this function is a norm (with unit ball K, of course). A convex body
It is easy to check that the only self-dual convex body is the Euclidean ball .B2n =
E 2
{x : xi = 1} (we show this below, after Theorem 4.11). The Blaschke–Santaló
inequality (see e.g. [6, Section 1.5.4]) states that among convex bodies, the volume
product, namely the product of the volume of a body and its polar, which is easily
seen to be an affine invariant, is maximal for ellipsoids.
4 Dualities, Measure Concentration and Transportation 163
Concentration of measure is a very strong tool, which since its discovery has been in
use in a multitude of areas in mathematics, and proving concentration results serves
as a good motivation for many developments in Asymptotic Geometric Analysis,
see e.g. [6, Chapter 3] and the Notes and Remarks section of the same Chapter with
many references. Concentration is the topic of the third talk, to which this section
serves as a short introduction.
The general notion is as follows: Consider a metric probability space .(X, d, μ),
with the measurable sets being the Borel .σ -algebra. Concentration of measure is the
phenomenon (which may or may not occur) that sets .A ⊆ X of measure .1/2 (say),
when extended slightly, already have very large measure. By an extension we mean
to consider
and “large measure” means that the measure of this extension is close to 1. Note that
we always have that .μ(At ) → 1 as .t → ∞, so the issue is usually regarding “how
fast” this convergence occurs in terms of t.
To illustrate the importance and usefulness of this phenomenon, which was
developed by V. Milman and has been a topic of wide study since then (with
applications throughout vast areas of mathematics) let us address two issues:
164 S. Artstein-Avidan
1. When this happens, Lipschitz functions are “virtually constant” on the space.
2. In spaces like the sphere, Gauss space, and the discrete cube, this is indeed the
case.
M} and .A2 = {x : f (x) ≥ M}, see Fig. 4.3. Note that for y in the intersection
of their t-extensions, .|f (y) − M| < t, since there is some x with .f (x) ≤ M and
.f (y) − f (x) ≤ d(x, y) < t, so .f (y) < M + t and there is some x with .f (x) ≥ M
and .f (x) − f (y) ≤ d(x, y) < t so .f (y) > M − t. Under the assumption of “high
concentration” this intersection has a big measure since
Examples of Concentration
The sphere, embedded in .Rn , has a natural metric (geodesic; which is equivalent to
the one induced by distance in .Rn ). The normalized Lebesgue measure on it (called
Haar measure and denoted .σ ) is invariant under rotations, and can be realized by
volume of the associated hull-with-0: for .A ⊆ S n−1
1
.σ (A) = Vol(C(A)) where C(A) = {sx : s ∈ [0, 1], x ∈ A}.
Vol(B2n )
The following theorem [35, 47] is well known and can be proved in various ways
(see e.g. [23]):
Theorem 4.3 (Lévy, Schmidt) Let .A ⊆ S n−1 and .t > 0. Then
.σ (At ) ≥ σ (Bt )
where .B = {x ∈ S n−1 : d(x, x0 ) ≤ d} for some .x0 ∈ S n−1 and d is chosen so that
.σ (B) = σ (A).
In other words, among subsets of the sphere with fixed measure, spherical caps have
the smallest t-extension volume.
Theorem 4.3 gives a straightforward way to deduce concentration, provided
we compute .σ (x : x1 > t) and show that it is very small (here we denoted
.x = (x1 , . . . , xn ) so that we have chosen a specific cap, which by rotation invariance
of the measure is enough). This allows one to give very precise bounds,
/
π/8e−t n/2 ,
2
σ (At ) ≥ 1 −
.
166 S. Artstein-Avidan
but the calculation is tedious, and it is more fun to give a geometric bound as follows
(this argument was presented by Ball in [11, Chapter 3])
1
σ ({x ∈ S n−1 : x1 > t}) =
. Vol({sx : s ∈ [0, 1], x ∈ S n−1 , x1 > t})
Vol(B2n )
√ √
and so long as .t < 1 − t 2 (that is, .t < 1/ 2), we know
/
{sx : s ∈ [0, 1], x ∈ S n−1 , x1 > t} ⊆ {y ∈ Rn : d(y, (t, 0)) <
. 1 − t 2 },
1
Vol((1 − t 2 )1/2 B2n ) = (1 − t 2 )n/2 ≤ e−t n/2 .
2
σ ({x ∈ S n−1 : x1 > t}) ≤
.
n
Vol(B2 )
We conclude, using Theorem 4.3, that for any .A ⊆ S n−1 with .σ (A) ≥ 1/2, we have
.σ (At ) ≥ 1 − e
−t 2 n/2 .
Gauss Space
The term Gauss space here is used to mean .Rn with the usual metric and the measure
e−|x| /2 dx. This is a “model space” with super nice properties, and
2
.dγ (x) =
1
(2π )n/2
will be a main object of study for us in this short course. The measure .γ (when
we want to emphasize the dimension we write .γn ) is rotation invariant and also
a product measure (in some sense it is quite surprising that a rotation invariant
product-measure exists). Its marginals are the lower dimensional .γk . Here too the
exact solutions to the isoperimetric inequality are known, and are half-spaces. This
Theorem goes back to Borell, Sudakov–Tsirelson [16, 49]
Theorem 4.4 (Borell, Sudakov-Tsirelson) Let .A ⊂ Rn and .t > 0. Then
γ (At ) ≥ γ (Ht )
.
One may prove the theorem using symmetrizations. Once we have Theorem 4.4,
computing the measure of .Ht for a measure .1/2 half-space, say, is simply computing
one dimensional Gaussians, .γ1 ((−∞, t]) (which in particular does not depend on
the dimension). As a corollary we thus get that for .A ⊂ Rn with .γ (A) = 1/2 we
have
f ∞
1 1 2
e−s /2 ds ≥ 1 − e−t /2 .
2
.γ (At ) ≥ 1 − √
2π t 2
4 Dualities, Measure Concentration and Transportation 167
For combinatorics fans, there are many discrete versions of the theorems considered
in this course. The set .{−1, 1}n , consisting of vertices of the cube .Q = [−1, 1]n ,
endowed with the uniform probability measure (counting, normalized) and the
Hamming metric .d(x, y) = n1 #{i : xi /= yi } is such an object (called “the discrete
cube”). It satisfies concentration in the form
1
μ(At ) ≥ 1 − e−2t n .
2
.
2
Here, again, the exact isoperimetric family of extremal sets is known, and consists
of metric caps, by a theorem of Harper [29]. For more on these discrete cases see [6,
Section 3.1.5] and the references in the corresponding Notes and Remarks section.
The class of functions we shall consider is the class of proper lower semi-continuous
convex functions from .Rn to .(−∞, ∞], which we denote .Cvx(Rn ). A main reason
for considering this class comes from the fact that replacing convex bodies with
log-concave measures leads to a useful generalization of the theory. To this end we
recall briefly what these measures are.
168 S. Artstein-Avidan
Definition 4.5 A Borel measure .μ on .Rn is called log-concave if for any .0 < λ < 1
and any .A, B ⊆ Rn such that .(A, B, λA + (1 − λ)B) are measurable,
( )
μ λA + (1 − λ)B ≥ μ(A)λ μ(B)1−λ .
.
uniformly.
Another important example for a log-concave measure is the Gaussian measure
on .Rn which has the log-concave density .e−|x| /2 /(2π )n/2 and we denoted above by
2
.γ or .γn . As we shall see shortly, its density function is in some sense the functional
Definition 4.8 (Inf-Convolution) Given two lower semi continuous convex func-
tions, .ϕ, ψ : Rn → (−∞, ∞], let their inf-convolution be defined by
the inf-convolution is standard Minkowski addition, and we have, as one can prove
directly from the definitions (see [8, Chapter 9] once more) that
Lemma 4.9 Let .ϕ, ψ : Rn → (−∞, ∞] be lower semi continuous convex
functions. Then .ϕOψ is a lower semi continuous convex function which satisfies
Given two upper semi continuous log-concave functions, .f1 , f2 : Rn → [0, ∞),
and some .λ ∈ (0, 1) let their sup-.λ-average be defined by
( )
(f1 *λ f2 )(x) =
. sup f1(1−λ) (y)f2λ (z) .
(1−λ)y+λz=x
For the usual polarity on convex bodies, the only self-polar set is the Euclidean
ball .B2n . Indeed, clearly for any norm one has (letting .|| · ||∗ denote the dual norm,
that is, the norm corresponding to the polar body of the unit ball)
so that if .||x|| = ||x||∗ then .||x|| > |x|, and as this is true for all x we get .B2n ⊇ K for
the unit ball K of .||x||, in which case .K ◦ ⊆ B2n , so that .K = K ◦ implies .K = B2n .
Similarly, for the Legendre transform, the only self-dual function is .|x|2 /2.
Indeed, by the definition of .L we have for any .x, y ∈ Rn that
so that in particular
Assume .ϕ = Lϕ, then we get .ϕ(x) > |x|2 /2, but since the right-hand side is
self-dual, and .L reverses order, we get that .Lϕ(x) ≤ |x|2 /2 as well, and we have
equality.
4 Dualities, Measure Concentration and Transportation 171
For this reason, one often uses .|x|2 /2 as the analogue for the Euclidean ball in the
world of convex functions, and .exp(−|x|2 /2) (normalized, or not) as its analogue in
the world of log-concave functions.
Since we will be using the Legendre transform frequently, we discuss in more
depth some of its properties.
Fact For a convex function .ϕ : Rn → (−∞, ∞] we have: .Lϕ(0) = − inf ϕ, and
.L(ϕ + c) = Lϕ − c. In addition for a closed convex K we have .L1
∞ = h and
K K
.LhK = 1 .
∞
K
Here we denote .1∞ K the function attaining 0 on the set K and .+∞ on its complement.
Indeed, for any y there is some .x ∈ ∂K which has y as a “direction of normal” at x,
that is, .hK (x) = supz∈K <z, x> is attained at .z = αy for some .α ≥ 0. If .y /∈ K then
.α < 1 and so .<x, y> − hK (x) > 0 and by 1-homogeneity the supremum satisfies
If on the other hand .y ∈ K then .<x, y> ≤ hK (x) for any x and by picking .x = 0 the
supremum is 0.
Lemma 4.12 Let .ϕ : Rn → (−∞, ∞] and not constant .+∞. Then .LLϕ is the
largest lower semi continuous convex function which is below .ϕ (also called the
“convex enelope” of .ϕ). In particular, for proper lower semi continuous convex .ϕ
we have that .LLϕ = ϕ.
Proof We start by discussing what we have to prove. First of all, .LLϕ is below .ϕ
since
LLϕ(x) = sup <y, x> − Lϕ(y) = sup infn <y, x> − <z, y> + ϕ(z) ≤ ϕ(x).
.
y∈Rn y∈Rn z∈R
Secondly, it is in the right class, obviously (we stated this in the previous fact). If
we show that on the class of proper lower semi continuous convex functions .L is
an involution, then for any function we know that .LLϕ is in this class. Any .ψ ≤ ϕ
which is in this class satisfies .Lψ ≥ Lϕ so .LLψ ≤ LLϕ and as .LLψ = ψ
we know .LLϕ is bigger. So, all we need to prove is that on the class .Cvx(Rn ) of
proper lower semi continuous convex functions (with values in .(−∞, ∞]) it holds
that .LLϕ = ϕ. To this end we use that .ϕ is the supremum of affine functions below
it (this is “separation” which one always needs for claims of this sort, it is a claim
about the epi-graph of .ϕ, which is a convex set in .Rn+1 , together with the fact that
we can “forget” about hyperplanes which are vertical). So, we use that
The latter condition, after moving terms from one side to another, can be written as
c ≥ Lϕ(y). Concluding,
.
Proof Indeed, by the previous theorem we can take .L of both sides and show
equality. Compute:
u
n
Based on the functional analogues introduced in the previous section for volume and
for Minkowski addition, we can now present the inequality of Prékopa and Leindler
which is considered the “functional analogue” for the Brunn–Minkowski inequality.
It is the following statement [34, 42] (see also [6, Chapter 1]).
Theorem 4.14 (Prékopa–Leindler) Let .f, g, h : Rn → R+ be measurable
functions, and let .λ ∈ (0, 1). We assume that f and g are integrable, and that
for every .x, y ∈ Rn
Then,
f (f )1−λ (f )λ
. h≥ f g .
Rn Rn Rn
Let us make a few remarks. First, it is indeed a generalization for the Brunn–
Minkowski inequality. Indeed,
Proof of the Brunn–Minkowski Inequality, Using Prékopa–Leindler Let K and
T be non-empty compact subsets of .Rn , and .λ ∈ (0, 1). We define .f = 1K , .g = 1T ,
and .h = 1(1−λ)K+λT . It is easily checked that the assumptions of Theorem 4.14 are
satisfied, therefore
f (f )1−λ (f )λ
Voln ((1 − λ)K + λT ) =
. h≥ f g = Voln (K)1−λ Voln (T )λ .
Rn Rn Rn
which is an inequality in the opposite direction (but for a function smaller than h).
174 S. Artstein-Avidan
Once we know that the measure on .Rn+1 given by .e−z dzdx is log-concave then
Theorem 4.14 follows directly, and we do know this if we accept the fact mentioned
above that a log-concave density makes for a log-concave measure. However, since
to prove this fact one usually uses the Prékopa–Leindler inequality, this argument is
somewhat circular (not completely: it means if we know this for a specific log-linear
density, we deduce it for all log-concave densities). The full proof of Theorem 4.14
appears in the course of Colesanti and Hug (it is given as Theorem 3.27 in this
volume). However, as the proof involves a part directly connected with one of the
main topics in this course, in the following section we discuss some (main) elements
of this proof in dimension one.
The classical Blaschke–Santaló inequality (Theorem 4.2) states that the volume
product .s(K) is maximized for Euclidean balls and their linear images. Recall once
more that Euclidean balls (of radius 1) are the only self-dual bodies in .Rn . For log-
concave functions, with duality given by the Legendre duality, one thus anticipates
an inequality of similar form. Define for an even function .ϕ the quantity
f f
s(ϕ) =
. e−ϕ e−Lϕ .
Note that when considering an upper bound we may as well restrict to convex
functions, since .LLϕ ≤ ϕ and thus .s(ϕ) ≤ s(Lϕ). The only self-dual function
.ϕ(x) = |x| /2, gives .s(ϕ) = (2π ) . It turns out that also in the functional setting,
2 n
the product .s(ϕ) is maximized when the function .ϕ is self-dual. The not-necessarily
even case is the following theorem from [5]. The even version was given first by
Ball [10]. See also [24] and [32, 33].
4 Dualities, Measure Concentration and Transportation 175
Theorem
f 4.16 Let .ϕ : Rn → (−∞, +∞] be a proper function, and assume that
. x exp(−ϕ(x))dx = 0 then
f f
. exp(−ϕ) exp(−Lϕ) ≤ (2π )n .
We will not prove this theorem, which is more in the vein of theorems from the
course of Colesanti and Hug. See [8, Chapter 9] for a detailed discussion, proofs
and many more references.
In this last part of the first lecture we detail some specific elements from the proof
of Theorem 4.14. The full proof appears in the course “Geometric and Analytic
Inequalities” and we will here only address the case of dimension .n = 1, with the
additional assumption that f and g are continuous and strictly positive. This will
lead us smoothly to the second lecture. f f
Without loss of generality one may assume . f = g = 1 by dividing both by
constants (and h by the corresponding average constant). We may thus think of f
and of g as densities of probability measures. We can take the uniform measure m
on .[0, 1] as a model space. A monotone increasing transport f map from .([0,
f 1], dm)
to .(R, f (x)dx) is a mapping .x : (0, 1) → R such that . A f (x)dx = x −1 (A) dm,
f x(t)
which is equivalent to . −∞ f = t. See Fig. 4.4.
In other words, we define .x, y : (0, 1) → R by the equations
f x(t) f f y(t) f
. f =t f = t and g=t g = t.
−∞ R −∞ R
In view of our assumptions, x and y are differentiable, and for every .t ∈ (0, 1) we
have
f f
'
.x (t)f (x(t)) = f = 1 and y ' (t)g(y(t)) = g = 1.
R R
Since x and y are strictly increasing, z too is strictly increasing, and the arithmetic-
geometric means inequality implies that
z' (t) = (1 − λ)x ' (t) + λy ' (t) ≥ (x ' (t))1−λ (y ' (t))λ .
.
Hence, we can estimate the integral of h making the change of variables .s = z(t),
as follows:
f f 1
. h= h(z(t))z' (t)dt
R 0
f 1
≥ h((1 − λ)x(t) + λy(t))(x ' (t))1−λ (y ' (t))λ dt
0
f 1 ( )1−λ ( )λ
1 1
≥ f 1−λ λ
(x(t))g (y(t)) dt = 1.
0 f (x(t)) g(y(t))
and so
where the last inequality is the original assumption on h joined with the fact that
z(t) = (1 − λ)x(t) + λy(t).
.
Note that the choice of .([0, 1], dm) is quite arbitrary, and any other reference
measure (with strictly positive continuous density function) would work just as well.
4 Dualities, Measure Concentration and Transportation 177
Monotonicity of the transport map was crucial here, since we used the arithmetic-
geometric means inequality which can only be applied to positive numbers. To prove
Theorem 4.14 in higher dimensions one usually uses induction, but in fact one can
provide a direct proof with a similar transportation argument, after developing these
tools in .Rn . For us this serves as motivation to understand what general measure
transportation is, and more importantly, what monotone transportation should mean
in higher dimensions. This is the topic of the second lecture.
We are given sets X and Y , and on them probability measures .μ ∈ P (X) and
.ν ∈ P (Y ). Here we usually assume that X and Y are subsets of .Rn with the Borel
.σ -algebra, but much of the theory works in the setting of general measure spaces. A
Remark 4.17 Whether it is enough to take only continuous bounded .ϕ, ψ, and
further restrict to functions tending to 0 at .∞, can depend on how general a setting
one wants to work with. We here assume X and Y to be Polish (complete separable
metric spaces), .ν and .μ to be Borel probability measures (so, the .σ -algebras
generated by open sets) and further that .X, Y are locally compact (each point has
a compact neighborhood). For these spaces the Riesz representation theorem states
that .M(X) = C0 (X)∗ .
Remark 4.18 We would like to emphasize a detail which may give some extra
motivation. In the case where .μ and .ν are absolutely continuous on .Rn , say
4 Dualities, Measure Concentration and Transportation 179
(.DT (z) denoting the differential of T at the point .z ∈ Rn ). This holding for any test
function .ϕ means that
Cost Functions
An optimal plan is one that minimizes this total cost, and a priori it is not clear
under which conditions such an optimal plan exists, or is unique (up to measure
zero). When the plan is induced by a mapping .T : X → Y , the total cost can be
written as
f
. c(x, T x)dμ,
X
Definition 4.19 The optimal transportation cost between .μ and .ν is the value
f
C(μ, ν) = inf{
. c(x, y)dπ : π ∈ ||(μ, ν)}.
X×Y
f
A plan .π ∈ ||(μ, ν) for which .C(μ, ν) = X×Y c(x, y)dπ , when exists, is called
an optimal transport plan.
Not much needs to be assumed in order to show that an optimal plan exists, since
||(μ, ν) is convex, and since the functional being infimized is linear in .π . Therefore
.
When the sets X and Y are finite and of the same size (say, .{xi }ni=1 , {yi }ni=1 ), and
the measures are normalized counting measures, a transport map is simply a pairing
(or matching, in combinatorial language) of the two sets of points. A transport plan
can then be represented by a matrix .π = (πi,j ) with all elements non-negative (with
this being a permutation matrix in the case of a plan induced by a map). The fact
that the plan transports one measure to the other is captured in the E fact that the
sum
E of every row, and every column, of the matrix, is 1, namely . i πi,j = 1 =
j π i,j . Such non-negative matrices are called bi-stochastic. Clearly the class of
2
bi-stochastic matrices is convex and bounded (as a subset of .Rn , as .πi,j ∈ [0, 1]).
In particular it is the convex hull of its extreme points. Clearly permutation matrices
are bi-stochastic and are extreme points of this set. The fact that they are the only
extreme points is a nice exercise, see e.g. [15]. This polytope is called the Birkhoff
polytope. E
The transportation problem in this case amounts to infimizing . πi,j c(xi , yj )
on the Birkhoff polytope. Since such a linear function is minimized on an extreme
4 Dualities, Measure Concentration and Transportation 181
We also mention that denseness of functions of the form .f (x) + g(y) cannot be
replaced by equality, namely there can be elements of .L1 (π ) which do not “split”
in this way, as the following example, again from [37], shows.
Example (J. Lindenstrauss) Let .λi be a sequence of positive numbers E2n−1 which
sum to 1. Take .x0 = y0 = 0, .x1 = λ1 , .y1 = λ1 + λ2 , .xn = i=1 λi ,
E2n
.yn = i=1 λi . Define the measure .π to be supported on segments joining .(xn , yn )
with .(xn+1 , yn+1 ), and joining .(xn+1 , yn ) with .(xn+2 , yn+1 ). On each segment
.[(xn , yn ), (xn+1 , yn+1 )] we take the uniform measure with total mass .λ2n+1 , and
on the segment .[(xn+1 , yn ), (xn+2 , yn+1 )] uniform with total mass .λ2n+2 . The total
mass is 1, so that .π is a probability measure, and moreover the marginals are m, that
is, uniform measures on .[0, 1]. The fact that .π is extreme follows from the theorem
we have just shown (this is an easy exercise). However, not every .h ∈ L1 (π ) is
a sum of .f (x) ∈ L1 (m) and .g(y) ∈ L1 (m), .μ almost everywhere. Indeed, we
can take h to be constant on the corresponding segments, and the condition on the
constants making it in .L1 (π ) is easy to write. Such a function cannot be split as a
sum.
Note that in the discussions above no special attention was given yet to the cost
function c. It turns out that when transport problems are considered for special
spaces .X, Y and a cost function which is well-behaved, not only can one show
that optimal plans are induced by transport maps, these transport maps are of very
special and “monotone” form. In particular, the most important and useful case for
convexity is when .X = Y = Rn and .c(x, y) = |x − y|2 . This is the subject matter
of the next section.
An extremely useful theorem in our field states, essentially, that under mild
assumptions on two probability measures on .Rn , when the quadratic cost function
is considered .c(x, y) = |x − y|2 , one may find a transport plan between the two
measures which is not only optimal, it is concentrated on a set (in the product
space) which is the graph of a function (in other words, the transport plan is in
fact a transport map). Moreover, the transport map .T : Rn → Rn is given by the
gradient of a convex function. This means that the differential of T is positive semi
definite at every point, which is a very useful property, as we shall see below.
follow from our analysis below that it is optimal with respect to the quadratic cost.
We shall see that it is equivalent to the fact that the support of the transport plan .π
is a cyclically monotone set with respect to c, which in turn is equivalent to the fact
that it lies on the graph of a sub-gradient of a convex function. All of these notions
are explained below.
In the next sections we will show some applications of the above theorem. In
the sections following them we develop some of the general theory, which will then
allow us to give a short proof of Brenier’s theorem (along with many other Brenier-
type theorems).
Before embarking on this journey, however, we remark on a connection with the
Legendre transform. Assume some transport map is given by .∇ϕ. Note that for a
convex function .ϕ, its gradient .∇ϕ is defined almost everywhere in its domain. In
fact, at every interior point of the domain the sub-gradient .∂ϕ(x) is defined, and it
is not single-valued only in a set of Lebesgue measure zero. Here
we see that it is attained at x for a certain y, if and only if .∂ϕ(x) e y. It is not a hard
exercise to check that the domain of .Lϕ is precisely the image of .∇ϕ. Using that
.LL = Id (see Lemma 4.12), we have that .∇Lϕ(∇ϕ(x)) = x, that is, the gradients
of .ϕ and of .Lϕ are inverse maps. So, if .∇ϕ maps .μ to .ν, then .∇Lϕ maps .ν back to .μ.
One may use Brenier’s theorem to give yet another proof for the Brunn–Minkowski
inequality for two convex bodies. Given two convex bodies .K0 and .K1 in .Rn , define
.μ =
Vol(K0 ) Vol|K0 and .ν = Vol(K1 ) Vol|K1 . Use Theorem 4.25 to find a measure
1 1
figured out (in the induction basis for our box-proof of the Brunn–Minkowski
inequality), using the Arithmetic-Geometric inequality, see (4.1), that
( n )1/n ( n )1/n ( )1/n
|| || Vol(K1 )
. (1 + λi ) ≥1+ λi =1+ .
Vol(K0 )
i=1 i=1
Note that we could have instead used the more general fact that for positive
definite .A, B we have .det(A + B)1/n ≥ det(A)1/n + det(B)1/n , which is one of
Aleksandrov’s inequalities for positive definite matrices (see e.g. [6, Appendix B]).
Mixed volumes are be discussed in depth the course of Ludwig and Mussnig (in
this volume). Let us show briefly how Brenier’s theorem gives a short proof for
Minkowski’s theorem on the polynomiality of volume, namely thatE as a function of
.(λi )
m ∈ Rm , for fixed convex bodies .(K )m , the quantity .Vol( m λ K ) is a
i=1 + i i=1 i=1 i i
homogeneous polynomial of degree n. Usually one calls its coefficients the “mixed
volumes” and writes
E
m E
m E
m ||
n
Vol(
. λ i Ki ) = ··· V (Ki1 , . . . , Kin ) λij ,
i=1 i1 =1 in =1 j =1
where in the notation we use that the coefficient of a monomial only depends on the
.Ki ’s where i belongs to the monomial. In addition, we normalize V to be symmetric
with respect to its arguments. It is part of Minkowski’s theorem that .V ≥ 0 as a
function on n-tuples.
We can show all of this using the Brenier map of Theorem 4.25. Choose
(somewhat arbitrarily) one measure to be the Gaussian measure .γ on .Rn , and for
.i = 1, . . . , m let .νi =
1
Vol(Ki ) Vol|Ki . Brenier’s theorem applied to the pair .γ and
.νi gives us for each i a convex function .ϕi such that .∇ϕi transports .γ to .νi . In
particular, and strangely this is all we will be using, this means that .∇ϕi (Rn ) = Ki ,
as sets, and as a consequence .∇(λi ϕi )(Rn ) E = λ i Ki . Em
The key observation here is that .∇( m i=1 λi ϕi )(R ) =
n
i=1 λi Ki . One
inclusion is immediate, just as we did in the proof above, where .(I + ∇ϕ)(K0 )
4 Dualities, Measure Concentration and Transportation 185
was clearly a subset of .K0 + K1 . However, we claim that here we in fact have an
equality,
E
m E
m
∇(
. λi ϕi )(Rn ) = ∇(λi ϕi )(Rn ).
i=1 i=1
transform
E maps summation to Einfimum convolution, the domain of this sum is
simply . m i=1 iλ dom(Lϕ i ) = m
λ K
i=1 i i .
Once this is settled, we get that
E
m E
m
Vol(
. λi Ki ) = Vol(∇ λi ϕi (Rn ))
i=1 i=1
f f E
m
= det(D∇ Em (x))dx = det( λi ∇ 2 ϕi (x))dx
i=1 λi ϕi
Rn Rn i=1
which is clearly a polynomial, and in fact, as the matrices are positive definite, one
can use the theory of mixed discriminants to show that the coefficients of the various
monomials are non-negative.
f ||
m f ∗
c
I (f1 , . . . , fm ) =
. fj j (Bj x)dx and K(h1 , . . . , hm ) = m(x)dx,
Rn j =1 Rm
186 S. Artstein-Avidan
f∗
where . denotes outer integral and where
{ ||
m E
m }
c
.m(x) = sup hjj (yj ) | yj ∈ R nj
and cj BjT yj =x .
j =1 j =1
A denote by .gA (x) = exp(−<Ax, x>) a “centered Gaussian”. Franck Barthe proved
the following theorem.
Theorem 4.26 (Barthe) The constants E and F can be computed using centered
Gaussian functions. That is,
{ }
K(g1 , . . . , gm )
E = inf ||
. (f )cj | gj is a centered Gaussian, j = 1, . . . , m
m
R j gj
n
j =1
and
{ }
I (g1 , . . . , gm )
. F = sup || (f )cj | gj is a centered Gaussian, j = 1, . . . , m .
m
R j gj
n
j =1
(E
m ) ||
m
. det cj BjT Aj Bj > D · (det Aj )cj , (4.4)
j =1 j =1
√ 1
E=
. D and F = √ .
D
The applications of this theorem are far reaching, especially in its geometric form
(a case where D can be computed and is equal to 1). This appears in the course of
Colesanti and Hug. For the proof, we note that letting
{ }
K(g1 , . . . , gm )
.Eg = inf )cj | gj is a centered Gaussian , j = 1, . . . , m
||m ( f
R j gj
n
j =1
4 Dualities, Measure Concentration and Transportation 187
and
{ }
I (g1 , . . . , gm )
.Fg = sup )cj | gj is a centered Gaussian , j = 1, . . . , m ,
||m ( f
R j gj
n
j =1
√ 1
E = Eg =
. D and F = Fg = √ .
D
The second inequalities from the left and from the right are trivial. The rightmost
and leftmost are sophisticated linear-algebra, and we do not discuss them here (see
the original papers of Barthe [12–14], or [8, Section 4.4])
The following proposition is in fact the main step for the proof of Theorem 4.26,
and implies the middle inequality .E ≥ D · F in (4.5). Recall that D is defined in
(4.4), and we may clearly assume that .D > 0. We will show (using Brenier’s map)
the following proposition.
Proposition 4.27 Assume that the functions .hj , fj ∈ L+
1 (R ), .1 ≤ j ≤ m, satisfy
nj
f f
. fj = hj = 1.
nj nj
R R
Then,
Proof We omit some technical details regarding the domain, range and differentia-
bility of the involved mappings, as we mainly want to illustrate the usefulness of
Brenier’s map. By Theorem 4.25, for every j there exists a mapping .Tj = ∇ϕj
mapping the density .fj (x)dx to .hj (y)dy, where .ϕj is convex. This means (at
differentiability points) that
( )
. det DTj (x) · hj (Tj x) = fj (x). (4.6)
E
m
O(y) :=
. cj BjT (Tj (Bj (y))).
j =1
188 S. Artstein-Avidan
E
By linearity its differential is given by .DO (y) = m T
j =1 cj Bj DTj (Bj y)Bj . Using
the definition of the constant D in (4.4), the determinant of .DO is bounded from
below
( ) ||
m
. det DO (y) > D · | det DTj (Bj y)|cj > 0, (4.7)
j =1
and in particular .DO is positive definite (and symmetric of course). It follows that
O is injective. Using (4.6) and (4.7) we may write
.
f { ||
m E
m }
c
K(h1 , . . . , hm ) =
. sup hjj (xj ) |x= cj BjT xj dx
Rn j =1 j =1
f { ||
m E
m }
c
> sup hjj (xj ) | x = cj BjT xj dx
O(Rn ) j =1 j =1
f { ||
m E
m }
c
= | det DO (y)| sup hjj (yj ) | O(y) = cj BjT yj dy
j =1 j =1
f ||
m
c
> | det DO (y)| hjj (Tj (Bj y)) dy
O(Rn ) j =1
f ||
m ||
m
c
>D· | det DTj (Bj y)| cj
hjj (Tj (Bj y)) dy
j =1 j =1
f ||
m
c
=D· fj j (Bj y) dy
j =1
= D · I (f1 , . . . , fm ).
u
n
The proof of Brenier’s theorem which we shall show is structuredf as follows. Pick,
among all measures .π ∈ ||(μ, ν), the one which minimizes . cdπ . The fact that
a minimizer exists follows from some compactness argument, and we show this
(using Prokhorov’s theorem) in Sect. 4.2.4. Here the specific choice of the quadratic
cost is of no special importance, only some regularity of the cost is needed.
We then show that a minimizing measure is concentrated on a subset of the
product space with special structure, a so-called “cyclically monotone set”. This
4 Dualities, Measure Concentration and Transportation 189
.σ . It is quite intuitive that this be the case, as otherwise one could make some
perturbation of the measure, not changing the marginals, and lowering the total
cost. Here, again, changing the condition of “cyclically monotone” to a form of
cyclic monotonicity involving a more general cost c, works quite well, under some
regularity assumptions on the cost.
Finally, the last step is to use Rockafellar’s theorem, which we will presently
describe, which states that any cyclically monotone set is in fact the graph of the
sub-gradient of a convex function. Together with the first steps, this means that the
support of the optimal plan .π lies on the graph of the gradient of some convex .ϕ, and
in particular, .π is not only a transport plan, but is a map. Here the generalization to
other costs is also possible, but one must change the notion of “sub-gradient” to the
less obvious notion of a “c-sub-gradient”. This is called the Rockafellar–Rochet–
Rüschendorf theorem, and applies to costs with values in .R. We will discuss this
part in depth as it is less well-known. In fact, quite recently the author jointly with
Sadovsky and Wyczesany [9] generalized Rockafellar’s theorem to capture the case
of costs which are allowed to obtain the value .+∞, and at the same time found a new
and transparent proof for the original Rockafellar–Rochet–Rüschendorf theorem.
We will describe this development here as well.
topology (that is, its closure is sequentially compact) if and only if K is tight, that
is, for any .ε > 0 there exists some .Aε ⊆ S with .μ(Aε ) ≥ 1 − ε for any .μ ∈ K.
Consider .K = ||(μ, ν), which is non-empty as it includes the product measure
.μ × ν. Let us explain why K is “tight". Fixing some .δ > 0 we can pick compact
.A ⊆ X and .B ⊆ Y with .μ(A), ν(B) ≥ 1 − δ (this is called inner regularity of the
This means that we can find a compact set such that for all .π ∈ ||(μ, ν) its measure
is big. The name for this property of .||(μ, ν) is, as explained above, “tight”, and
Prokhorov’s theorem implies we can extract a converging sub-sequence for any
sequence in .||(μ, ν).
Next, .||(μ, ν) is closed in weak* topology. This is trivial: take a sequence .πk ∈
||(μ, ν) which converges in weak* to .π . This means that
f f
. f (x, y)dπk (x, y) → f (x, y)dπ(x, y)
for every .f ∈ C0(b) (X × Y ) (continuous, compact support). Using this for functions
depending only on x and then for functions depending only of y we see that .π ∈
||(μ, ν).
Next consider a minimizing sequence .πk for the total cost. As .||(μ, ν) is
sequentially compact, we may extract a converging sub-sequence, and re-index to
have .πk → π in the weak* sense.
The second step relies on the lower semi continuity of the cost function c, in
this case the quadratic cost (but one may consider a lower semi continuous cost
function). It is a well-known fact that weak convergence .μk → μ is equivalent to
the inequality
f f
. lim inf f dμk ≥ f dμ
for all non-negative lower semi continuous f . Alternatively, we can just use that as
c is non-negative and lower semi continuous we can find a sequence of continuous
functions which is non-decreasing and converges point-wise to c, .cj → c. We may
now use monotone convergence to get
f f
. cdπ = lim cj dπ
j
4 Dualities, Measure Concentration and Transportation 191
f f
but by weak* convergence . cj dπ = limk cj dπk and by monotonicity of .cj we
get
f f f
. cdπ = lim lim cj dπk ≤ lim cdπk
j k k
as claimed. Note that in general we would have to move to a .lim inf with respect to
k in the last inequality but we know that the limit exists having chosen a minimizing
sequence. This completes the proof of the existence of a minimizing measure.
Remark 4.29 For the students less familiar with Prokhorov’s theorem, let us
illustrate the part most relevant to us, in .Rn . First note that tightness is necessary:
For example in .R, if a set of probability measures is not tight, we can choose a
sequence of probability measures such that for some .ε0 , .μk ((−k, k)) < 1 − ε0 . This
means that if a limit measure (weak*) existed it would have to satisfy (note that an
open indicator is lower semi continuous)
and this for any x, impossible. So tightness is indeed necessary for weak*
compactness.
To see how tightness is sufficient (which is the condition we use) let us give a
sketch for the proof in .Rn . A probability measure is determined by the function
.F (x) = μ(w : wi ≤ xi ∀i) which is a function on .R , with values in .[0, 1],
n
||
n E
μ(
. (ai , bi ]) = (−1)#(θ) F ((ai + θi di )ni=1 ) ≥ 0
i=1 θ∈{0,1}n
usual convergence of .Fk to F . Finally, putting all of these together we get that the
measures .μk converge weakly to .μ.
192 S. Artstein-Avidan
To see more clearly why Brenier’s theorem has that particular form, namely
why a gradient appears as the structural form of an optimal transport, and more
interestingly to anticipate which other forms of transport maps can one expect if we
optimize with respect to another cost function, we introduce the Kantorovich duality
theorem along with the cost-transform for functions. The duality theorem serves as
a motivation for the transform and, as we shall see, also for Brenier’s theorem.
(Below we sometimes denote this quantity also by .Wc (μ, ν), see Sect. 4.3.6) It turns
out that one may express this infimum in a dual way, as a supremum of integrals of
pairs of functions satisfying some constraint. (Note the visual resemblance with the
pairs of functions in the Lindenstrauss theorem).
Theorem 4.30 Let .X, Y be Polish spaces and .μ ∈ P (X), .ν ∈ P (Y ) probability
measures. Assume .c : X × Y → [0, ∞] is lower semi continuous. Then
f f
.C(μ, ν) = sup{ ϕdμ + ψdν : (ϕ, ψ) ∈ L1 (μ) × L1 (ν),
X Y
ϕ(x) + ψ(y) ≤ c(x, y) (μ, ν)−a.e.}.
Moreover, the infimum in the definition of .C(μ, ν) is attained, and the supremum in
the above formula does not change if one allows only bounded and continuous pairs
.(ϕ, ψ).
The fact that the infimum is attained was discussed in the previous section. The
fact that the supremum does not change if we restrict to bounded and continuous
functions is due to the fact that we can approximate any function in .L1 by bounded
and continuous ones. We shall prove the theorem using an infinite dimensional min-
max principle, but it is useful to notice that an inequality is immediate:
Remark 4.31 If .ϕ, ψ are as above then for .π ∈ ||(μ, ν),
f f f f
. ϕdμ + ψdν = (ϕ(x) + ψ(y)) dπ ≤ c(x, y)dπ,
X Y X×Y X×Y
since the inequality holds .μ, ν a.e. which implies it holds .π a.e. as well. Since this is
true for any .π ∈ ||(μ, ν), one may infimize over all these, and we get an inequality
in the Kantorovich Duality Theorem.
4 Dualities, Measure Concentration and Transportation 193
The proof of the theorem itself uses an infinite dimensional linear programming
duality theorem, in which, strangely, Legendre transform on infinite dimensional
spaces comes up. Theorem 4.30 is an easy consequence of the following theorem,
where E is a general linear space, which can be (and will be, in our application)
infinite dimensional. For the proof of Theorem 4.32 see [51]. We do comment on
the geometry behind it after the statement. Here for a function .η : E → (−∞, ∞]
where E is a linear space with a canonical pairing with its dual space .<·, ·> : E ×
E ∗ → R, the Legendre transform .η∗ is defined on the dual space .E ∗ by
Note that part of the statement in the theorem is that on the right hand side there is
an actual extremizer. To give some geometric intuition, let us see what the theorem
means in finite dimensions, where we have already established some facts regarding
the Legendre transform. Note first that
Therefore, the left hand side in the formula claimed in Theorem 4.32 is simply
−L(ϕ + ψ)(0). However, by Lemmas 4.12 and 4.13 (which we discussed in finite
.
which is very similar to the claim in the theorem. The missing ingredients are that
we did not show that the infimum is in fact attained as a minimum, and that we have
used lemmas that we discussed only in finite dimensions. This paragraph serves only
to provide some intuition, for the proof of the theorem, which is essentially a clever
use of the Hahn–Banach theorem, see [51]. We next use this theorem to prove the
Kantorovich Duality Theorem.
Proof of Theorem 4.30 Using Theorem 4.32 We shall define our two convex func-
(b)
tions, on the linear space .E = C0 (X × Y ) of continuous functions .f : X × Y → R
194 S. Artstein-Avidan
with compact support, equipped with the supremum norm. Its dual space is the space
f all Radon measures (signed) on .X × Y , with the canonical pairing .<π, f > =
of
f dπ. The convex functions we choose are
o(f ) = 1∞
. {f ≥−c}
and
f f
w(f ) =
. ϕdμ + ψdν if f (x, y) = ϕ(x) + ψ(y)
and .+∞ on functions f for which there is no such representation. Notice that such
a representation is unique up to adding a constant to .ϕ and subtracting it from .ψ,
so that .w is well defined. Convexity of .o is immediate, as is convexity of .w (in
fact, linearity on its domain). In order to use Theorem 4.32 we need to find a point
in their common domain at which one of the two convex functions is continuous.
This point will be the function .f ≡ 1 which is clearly a bounded and continuous
function on .X × Y . The function f belongs to the domain of .o since c is non-
negative, and .o is continuous at f since on a .1/2-neighborhood (in the .sup norm)
of f , the function .o is constant 0. Finally, f is in the domain of .w since 1 can be
written as the combination of two constant functions.
Next we compute the Legendre transforms of .o and .w, which are now functions
on the dual space, namely on the space of Radon measures. We expect (and this will
indeed be the case) that .o∗ be a linear function (as .o was an indicator) and that
the dual of .w be an indicator as .w was linear on its domain (we can write .w as
.<f, μ ⊗ ν> on its domain). Indeed, for a measure .π we get
(f ) f
∗
o (π ) = sup
. f dπ − o(f ) = sup f dπ
f f ≥−c
f f
o∗ (−π ) = sup (−f )dπ = sup f dπ.
f ≥−c f ≤c
If .π is not a non-negative measure then there exists some very negative f (in the
area were .π “is negative”) which forces the .sup to be f .+∞. On the other hand if .π
is non-negative then we can get arbitrarily close to . cdπ by choosing appropriate
f (we might not be able to take .c itself if it is not bounded, but we can get as close
as we want), so we see that
f
∗
.o (−π ) = cdπ
4 Dualities, Measure Concentration and Transportation 195
for .π a non-negative measure, and .+∞ for measures outside the cone of non-
negative measures. As for .w ∗ , let us compute
(f )
w ∗ (π ) = sup
. f dπ − w(f )
f
(f f f )
= sup f (x, y)dπ − ϕdμ − ψdν .
f (x,y)=ϕ(x)+ψ(y)
If .π ∈ ||(μ, ν) then the expression on the right hand side is 0, and if .π /∈ ||(μ, ν)
then for any M there exists a function f for which the difference
f f f
. (ϕ(x) + ψ(y)) dπ − ϕdμ − ψdν > M.
We conclude that .w ∗ is the indicator of the set of measures .π ∈ ||(μ, ν). Plugging
these in we get that
f
∗ ∗
. − min∗ {o (−π ) + w (π )} = − min{ cdπ : π ∈ ||(μ, ν)}.
π ∈E
(Note that the fact that the marginals are .μ and .ν is not enough for .π ∈ ||(μ, ν)
because we also require .π ≥ 0 which we obtained be the condition of belonging to
the domain of .o∗ ). By Theorem 4.32 we get that
f
. inf (o(f ) + w(f )) = − min{ cdπ : π ∈ ||(μ, ν)}
The Kantorovich Duality Theorem, Theorem 4.30, tells us that when searching for
f at pairs .ϕ(x), ψ(y) satisfying .ϕ(x) +
the total cost .C(μ, ν) we can finstead look
ψ(y) ≤ c(x, y) and supremize . ϕdμ+ ψdν. (Here we let .−∞+∞ = −∞.) We
can call a pair satisfying this inequality an “admissible pair”. Given some function
.ϕ : X → [−∞, +∞], we may associate to it the largest .ψ such that the two
When .X = Y and .c(x, y) = c(y, x), these amount to the same transform. We
abuse notation slightly by using the same notation for both transforms, as we usually
assume .X = Y = Rn and assume the cost to be symmetric. (Here, to match the
above definition of admissible pairs, one must let .∞ − ∞ = ∞.)
The following facts follow directly from the definition.
Fact If .ϕ and .ψ are admissible then .ϕ c ≥ ψ.
Fact If .ϕ1 ≤ ϕ2 then .ϕ1c ≥ ϕ2c .
A third simple fact is that on the image of the transform, namely for .ψ = ϕ c , the
transform is an involution. More precisely,
Fact For any .ϕ we have that .ϕ ccc = ϕ c .
Proof Indeed, the pair .ϕ, ϕ c is admissible and thus by the first Fact above we see
that .ϕ ≤ ϕ cc . We may now insert into this inequality the function .ϕ c , getting .ϕ c ≤
ϕ ccc . We may, instead, take the c transform of both sides of this inequality, and by
the second Fact above get that .ϕ c ≥ ϕ ccc . u
n
This last Fact implies that to each cost function there corresponds a natural class
of functions on which the cost-induced transform is an order reversing involution.
Definition 4.34 (The c-Class) Given a symmetric cost function .c(x, y) : X×X →
(−∞, ∞], the corresponding c-class is the image of the mapping .ϕ |→ ϕ c on the
class of all functions .f : X → [−∞, ∞].
It readily follows from the above facts and definition that
Fact A function .ϕ is in the c-class if and only if .ϕ = ϕ cc .
Among functions in the c-class, there is a subfamily which is of special
importance.
Definition 4.35 (Basic Functions) A function of the form .ϕ(x) = c(x, y0 )+β, for
a fixed .y0 ∈ Y and a constant .β ∈ R, is called a basic function for the cost c. When
c is not a symmetric cost and X is different than Y , one must distinguish between
X-basic functions of the form .ϕ(x) = c(x, y0 ) + β, and Y -basic functions of the
form .ψ(y) = c(x0 , y) + β.
4 Dualities, Measure Concentration and Transportation 197
( )
ψ c (x) = inf c(x, y) − ψy0 ,β (y) = c(x, y0 ) + β.
.
y
Therefore, the function .ψ(x) = c(x, y0 ) + β is in the c-class. (One may also check
directly that .ϕ = ϕ cc ). u
n
It is useful to notice that the c-class is always closed under the operation of point-
wise infimum.
Fact Assume .(ϕα )α∈I is some family of functions which are all in the c-class, where
I is some indexing set. Then the point-wise infimum
Regardless of whether .ψα is in the c-class or not, we may consider its c-transform
and we see that
( )
.ψ (x) = inf (c(x, y) − ψ(y)) = inf c(x, y) − sup ψα (y)
c
y y α
u
n
The basic functions generate the c-class since, using the definition of the c-
transform, we see that
Fact Any function .ϕ in the c-class is the infimum of basic functions and vice-versa.
Finally, whenever one encounters a class of functions which is closed under
ˆ
point-wise infimums, one may define within this class the operation of .sup(S)
(where S is some subset of functions) which denotes the minimal element in the
1 Note that we use the fact that .c(x, y) /= −∞, since for .y /= y0 the function .ψy0 ,β gave the
result .+∞ and under our convention, if added to .−∞ this would result in .−∞ and be taken as the
infimum, had we allowed .c(x, y) to attain the value .−∞.
198 S. Artstein-Avidan
class which is greater than all the elements in the set. For example, the class of
ˆ a
convex bodies is closed under intersection, and the “convex hull” is the .sup,
“replacement” for the union operation, which does not preserve convexity. The
general definition is as follows:
ˆ Let .S be a class of functions which is closed
Definition 4.36 (The Operation .sup)
under the operation of infimum. Given a family of functions .(ϕα )α∈I for some index
set I , one defines
ˆ α : α ∈ I ) = inf{ϕ ∈ S : ϕ ≥ ϕα ∀α}.
sup(ϕ
.
is, up to sign, a cost transform. Indeed, it can be recovered using one of the two
following possibilities for cost functions: .c(x, y) = −<x, y> or .c̃(x, y) = |x−y|
2
2 . To
see that .−ϕ c = L(−ϕ) write
One may easily check that the corresponding c-class is the class of upper semi
continuous concave functions (which are allowed to attain the value .−∞).
Regarding the cost function .c̃, we note that
so that
This means that .c̃-class are simply functions in c-class with an added fixed function
|x|2 /2, and the induced transform is, up to this “change of appearance”, the same.
.
Here one must specify that division by 0 of a positive number gives .+∞ whereas
the fraction . 00 gives 0 (this corresponds, after taking a logarithm, to the .±∞ calculus
presented above for the cost setting).
One may easily verify that the image of .A is the set of so-called geometric convex
functions, .Cvx0 (Rn ), namely lower semi continuous convex functions which vanish
4 Dualities, Measure Concentration and Transportation 199
at the origin and which are non-negative (and allowed to attain the value .+∞, but
not everywhere). As a transform on epi-graphs, .A corresponds to standard polarity
and reflection, see [3].
It turns out that .A is induced by the cost function
in the following sense .e−v 0 = A(e−v ), as can be verified with a direct computation.
p
In Brenier’s theorem, Theorem 4.25, the gradient of a convex function plays a key
role: it is the form of the optimal transport map. For more general cost functions
(Brenier’s theorem deals of course with the quadratic cost function) the analogue of
the gradient mapping is the so-called c-sub-gradient as we explain below.
Definition 4.37 (The c-Sub-Gradient) Given a cost .c : X × Y → (−∞, ∞] and
a function .ϕ which is in the c-class, we define
One motivation for the definition of the c-sub-gradient comes, as expected, from
the Kantorovich Duality Theorem. Indeed, in searching for a pair .(ϕ, ϕ c ) for which
the supremum in Kantorovich Duality Theorem is attained, namely
f f f
Wc (μ, ν) =
. c(x, y)dπ0 (x, y) = ϕdμ + ϕ c dν,
X×Y
we note that the inequality .ϕ(x)+ϕ c (y) ≤ c(x, y) always holds, so that for equality
to hold we expect the support of the measure .π0 to be concentrated on the equality
cases for this inequality, which is precisely the definition of the c-sub-gradient of .ϕ.
More precisely
Lemma 4.38 Assume .π ∈ ||(μ, ν) and there is some c-class .ϕ ∈ L1 (μ) such that
π(∂ c ϕ) = 1. Then .π is an optimal transport plan with respect to c.
.
Proof Indeed, as .π is some plan, an upper bound for .Wc (μ, ν) is given by
f f
. c(x, y)dπ(x, y) = c(x, y)dπ
X×Y ∂cϕ
f f f
= (ϕ(x) + ϕ c (y))dπ = ϕdμ + ϕ c dν,
∂cϕ X Y
which by Kantorovich Duality Theorem is also a lower bound. This means that not
only is .π optimal, but the supremum in Kantorovich’s theorem is attained, and in
the case where .∂ c ϕ is a graph (namely, is uni-valued almost everywhere) .π is in fact
a transport map, not just a plan. u
n
Note that by definition .∂ c ϕ c = {(y, x) : (x, y) ∈ ∂ c ϕ} which is another way
of stating that as (set-valued) mappings, the two mappings .x |→ ∂ c ϕ(x) and .y |→
∂ c ϕ c (y) are inverse to one another. This is precisely what we discussed regarding
the Legendre transform in Sect. 4.2.2. We rewrite it here in fuller generality.
Lemma 4.39 Let .ϕ : Rn → (−∞, ∞] be a convex lower semi continuous function,
and set .ψ(x) = |x|2 /2 − ϕ(x). Let .c(x, y) = |x − y|2 /2, so that .|y|2 /2 − Lϕ(y) =
ψ c (y). Then it holds that
∂ c ψ(x) = ∂ϕ(x)
.
where
As we have seen above, understanding optimal maps has to do with the possibility
of finding a plan supported on the c-sub-gradient of some c-class function. We thus
turn to the question of understanding the structure of supports of c-sub-gradients.
To this end, we define a property of a subset of .X × Y called c-cyclic monotonicity.
Definition 4.40 Given a cost .c : X × Y → (−∞, ∞] , a set .G ⊂ X × Y is called
i=1 ⊆ G it holds that for any
c-cyclically monotone if for any N and any .(xi , yi )N
permutation .σ of .{1, . . . , N}
E
N E
N
. c(xi , yi ) ≤ c(xi , yσ (i) ).
i=1 i=1
Proof By definition
and
Summing over all i’s we get that the left hand sides have the same sum, and we end
up with
E
N E
N
. c(xi , yi ) ≤ c(xi , yσ (i) ).
i=1 i=1
u
n
Our main goal in this section is to reverse this observation, and show that in
many cases any c-cyclically monotone set has a so-called “potential”, that is, some
.ϕ such that .G ⊆ ∂ ϕ. When this is possible, and further, when .∂ ϕ is actually a
c c
mapping, we will not only get a maximizer in Kantorovich Duality Theorem, but
this maximizer will correspond to a transport map (rather than a mere plan). In the
case of a quadratic cost, this map will be the gradient of a convex function (which is
well-defined almost everywhere) which in addition means its differential is positive
definite.
It turns out that the above lemma can be reversed in the case of so called
“traditional” costs, namely when .c(x, y) ∈ R for all .x ∈ X, y ∈ Y . For costs
that are allowed to attain the value .+∞, a slightly stronger condition than c-
cyclic monotonicity is needed for the existence of a potential, as we will discuss
in Sect. 4.2.8.
However, before embarking on this important goal of reversing the implication
in Lemma 4.41, we show that in the classical case of the quadratic cost on .Rn (or,
more generally, any continuous cost), there exists a plan with c-cyclically monotone
support. Thus, a reverse Lemma to Lemma 4.41 will in fact produce an optimal
transport plan supported on a c-sub-gradient (recall Lemma 4.38).
Lemma 4.42 Let .μ and .ν be Borel probability measures on .Rn , and let .c : Rn ×
Rn → R be a continuous cost function. Then there exists a probability measure
.π ∈ ||(μ, ν) with c-cyclically monotone support.
E
Proof This is clearly the case for finite discrete measures .μ = m1 m i=1 δxi and
E m
.ν =
1
m δ
i=1 yi since we simply optimize on a finite set of possibilities and select
the best permutation (see Sect. 4.2.1). Given some general .μ and .ν we construct
discrete measures .μk and .νk which converge to them in the weak* topology. For
each k we find .πk ∈ ||(μk , νk ) with c-cyclically monotone support. By a standard
compactness argument (see Sect. 4.2.4) there exists a weak* sequential limit .π of
.πk , which by definition of weak* convergence is in .||(μ, ν) and by continuity of
c has c-cyclically monotone support. Indeed, for .(zi , wi ) ∈ support(π ) there are
(k) (k)
.(z
i , wi ) ∈ support(πk ) converging to them, so by continuity of the cost
E
N E
N
(k) (k) (k) (k)
. c(zi , wi ) − c(zi , wσ (i) ) = lim c(zi , wi ) − c(zi , wσ (i) ) ≤ 0.
i=1 i=1
Note that we have not claimed that the plan .π we constructed is optimal with
respect to c, though we do expect this as a limit of optimal plans for very close-by
measures.
The first main constructive theorem we present states that when the cost considered
attains only finite values, given a c-cyclically monotone set G, one may find a
“potential” .ϕ which is a c-class function such that .G ⊆ ∂ c ϕ.
Theorem 4.43 (Rockafellar–Rochet–Rüschendorf) Let .c : X × X → R be a
symmetric cost function, .c(x, y) = c(y, x). Let .G ⊂ X × X be non-empty and
c-cyclically monotone. Then there exists some c-class function .ϕ : X → [−∞, ∞]
such that .G ⊆ ∂ c ϕ (and in particular, .ϕ(x) /= ±∞ for x such that .(x, y) ∈ G).
We provide two proofs for this theorem. Our main goal is to show a new proof
from [9] which we find both insightful and simple in nature, and which, more
importantly, can be generalized to non-traditional costs in a clear fashion. This is
given in the next section. However, for completeness of the exposition, we first
present the classical proof, see [44–46].
Proof Fix some element .(x0 , y0 ) ∈ G which we shall call the “pivot”. We will make
sure .ϕ(x0 ) = 0. Define
( )
E
m
.ϕ(x) = inf c(x, ym ) − c(x0 , y0 ) + (c(xi , yi−1 ) − c(xi , yi )) ,
i=1
where the infimum runs over all .m ∈ N and all m-tuples .(xi , yi ) ∈ G, i = 1, . . . , m.
The first observation is that this function is indeed in the c-class, namely it is
c
.ψ for some .ψ. Indeed, this holds because the c-class is closed under infimum, and
each of the functions in this infimum is simply of the form .c(x, ym ) − β which (we
showed before) is in the c-class as well.
The second observation is that .ϕ(x0 ) = 0. Indeed, by picking .m = 0 (or .m = 1
with .(x1 , y1 ) = (x0 , y0 )) it is clear that .ϕ(x0 ) ≤ 0. On the other hand the c-cyclic
monotonicity implies that
E
m E
m
. c(xi , yi ) ≤ c(x0 , ym ) + c(xi , yi−1 )
i=0 i=1
Our goal is thus to show that if .(x, y) ∈ G then .ϕ(x) + ϕ c (y) = c(x, y). We
shall do this without computing .ϕ c , but instead showing that for any .z ∈ X we have
that
Before doing this, let us have a short discussion on whether these two claims, (4.8)
and .(x, y) ∈ ∂ c ϕ, are indeed the same. The answer is an obvious yes, since this
means that the infimum is attained at x. A problem may occur only if we start
worrying that perhaps .ϕ(x) = ±∞ (which, the reader will recall, is not allowed for
x with .(x, y) ∈ ∂ c ϕ). We clearly have .ϕ(x) ≤ c(x, y0 )−c(x0 , y0 ) which is finite. So
one might nevertheless worry that maybe .ϕ(x) = −∞. This definitely could happen
if x is some arbitrary point, however here we are assuming .(x, y) ∈ G. In particular,
once we show the above inequality, .c(x, y) − ϕ(x) ≤ c(x0 , y) − ϕ(x0 ) = c(x0 , y)
so .ϕ(x) cannot be .−∞.
To show (4.8) we take some .t > ϕ(x) and use the definition of .ϕ(x) as an
infimum to find some .m ∈ N and some .(xi , yi )m i=1 ⊆ G such that
E
m
t > c(x, ym ) − c(x0 , y0 ) +
. (c(xi , yi−1 ) − c(xi , yi )).
i=1
Since .(x, y) ∈ G we may rename them .(xm+1 , ym+1 ) and use the new .(m + 1)-tuple
for an upper bound in the definition of .ϕ(z) as an infimum. Namely
E
m+1
ϕ(z) ≤ c(z, ym+1 ) − c(x0 , y0 ) +
. (c(xi , yi−1 ) − c(xi , yi )).
i=1
. ϕ(z) ≤
E
m
c(z, y) − c(x0 , y0 ) + c(x, ym ) − c(xm , ym ) + (c(xi , yi−1 ) − c(xi , yi )),
i=1
. ϕ(z) − c(z, y) ≤
E
m
−c(x0 , y0 ) + c(x, ym ) − c(x, y) + (c(xi , yi−1 ) − c(xi , yi )) < t − c(x, y).
i=1
as claimed. This means that .ϕ c (y) + ϕ(x) = c(x, y), namely .(x, y) ∈ ∂ c ϕ, which
completes the proof. u
n
Remark 4.44 The c-cyclic monotonicity of G was only used in the part where we
showed that .ϕ(x0 ) = 0.
While the above classical proof is quite short and beautiful, it is not immediate how
one may use it to tackle costs which are non-traditional, namely costs for which
there exist pairs .(x, y) with .c(x, y) = +∞. To do this, we first reformulate the
problem of finding a potential for a given set .G ⊂ X × Y as a question regarding
the existence of a solution to a linear system of inequalities.
Theorem 4.45 Let .c : X × Y → R ∪ {+∞} be a cost function and let .G ⊆ X × Y .
Then there exists a potential for G, namely a c-class function .ϕ : X → [−∞, ∞]
such that .G ⊆ ∂ c ϕ, if and only if the following system of inequalities,
Letting .a(x) = ϕ(x) for .x ∈ PX G, we have found a solution for the system of
inequalities.
For the other direction, assume we have a solution to the system of inequalities.
We would like to extend it to some c-class function defined on X. To this end let
ϕ(z) =
. inf {c(z, y) − c(x, y) + a(x)}.
(x,y)∈G
206 S. Artstein-Avidan
We will show that the function .ϕ, which is clearly in the c-class, is an extension of
the function .a : PX G → R, and that it is a potential for G, namely .G ⊆ ∂ c ϕ.
The assumption (4.9) implies that for .z ∈ PX G we have
.αi,j ≤ vi − vj , i, j ∈ I (4.10)
has a solution if and only if forE any m and any .i1 , · · · , im ∈ I , and a permutation .σ
of .[m] = {1, . . . , m}, we have . m
k=1 αik ,iσ (k) ≤ 0.
To make the translation, let us consider the set of indices I to be our original set
EmIf .i = (x, y) and .j = (z, w) we let .αi,j = c(x, y) − c(z, y). The condition
G.
. k=1 αik ,iσ (k) ≤ 0 amounts to
E
m
. c(xi , yi ) − c(xσ (i) , yi ) ≤ 0,
k=1
Let us repeat this delicate point: The index set for the inequalities are pairs
.((x, y), z) ∈ G × PX G (or, equivalently, pairs .((x, y), (z, w)) ∈ G × G, where
we ignore w as it does not appear in the inequalities). The solution vector we are
looking for is indexed by .PX G, and we denoted it .(a(x))x∈PX G . In fact, formally, we
4 Dualities, Measure Concentration and Transportation 207
and
which means
αi,j ≤ vi − vj
.
and E
(b) For any permutation .σ of .[m] we have that . mi=1 αi,σ (i) ≤ 0.
Proof Clearly (a) implies (b), by summing over the pairs .(i, σ (i)), as explained
above:
E E
. αi,σ (i) ≤ vi − vσ (i) = 0.
i∈J i∈J
For the other direction, we shall use induction. Note that we are given a set of at
most .m(m − 1) inequalities and we would like E to show they have a joint solution.
Without loss of generality we may assume . vi = 0, so in fact these .m(m − 1)
inequalities are on a vector which is essentially in .Rm−1 . By Helly’s theorem, it
is enough to make sure that any m of these inequalities have a joint solution (the
existence of a solution amounts to the intersection of the corresponding half-spaces
being non-empty).
208 S. Artstein-Avidan
Using induction on m, we may assume that the theorem is true for .(m − 1). In
particular, given a subset P of m pairs .{(i, j )}, if there is an index within .{1, . . . , m}
which does appear in any of these, then by induction we know that the intersection
of the corresponding half-spaces in not empty. We may thus assume that the family
of m inequalities which we are trying to satisfy simultaneously include each of the
integers in .{1, . . . , m} at least once.
Further, we argue that if one of these integers, say k, appears only once, say then
the constraint on .vk is one sided. In particular, we may consider the vector v without
its kth coordinate, solve the system on inequalities using the induction assumption,
and then solve the single remaining inequality for .vk (since a single inequality in a
single variable always has a solution). Moreover, the same argument applies if one
of the integers, call it k again, appears (as many times as it likes) only as the first
(resp. only as the second) in any pair .(i, j ) ∈ P .
We may thus assume, without loss of generality, that each of the indices in
.{1, . . . , m} appears at least once as a first index and at least once as a second index,
We fix .v1 arbitrarily and let .vσ (1) = v1 −α1,σ (1) , .vσ (σ (1)) = vσ (1) −ασ (1),σ (σ (1)) and
inductively, having defined .vj , let .vσ (j ) = vj − αj,σ (j ) . After m steps we will have
defined .vσ m−1 (1) , say .σ m−1 (1) = k. Clearly .σ (k) = 1. All inequalities are satisfied
(as equalities, in fact) except possibly the last one: .αk,1 ≤ vk − v1 . Since we have
defined everything explicitly, we can write this inequality explicitly as well:
E
i−1
= · · · = vσ −i (k) − ασ −(j +1) (k),σ −j (k) − v1
j =0
E
m−1
= · · · = v1 − ασ −(j +1) (k),σ −j (k) − v1 .
j =0
E
m
. αi,σ (i) ≤ 0
i=1
4 Dualities, Measure Concentration and Transportation 209
which holds by our assumption (b). We thus showed that any m inequalities can
be satisfied simultaneously (the condition that the sum is 0 can be satisfied simply
be adding a constant to all coordinates of the vector), and conclude, using Helly’s
theorem and the fact that the problem is essentially .(m − 1) dimensional, that if (b)
holds then the system of inequalities in (a) has a solution. u
n
To move from finite sets of inequalities to an infinite one, we use compactness in
a standard fashion.
Proposition 4.48 Let .αi,j ∈ R for .i, j ∈ I with .αi,i = 0 and I is some index set.
Consider the family of inequalities
αi,j ≤ vi − vj .
.
Assume that for any finite subset .J ⊆ I , one may find a solution .(vj )j ∈J satisfying
jointly all the inequalities for .i, j ∈ J . Then there exists a solution .(vi )i∈I solving
jointly all the inequalities.
Proof Recall Tychonoff’s theorem which states that the product of any collection of
compact topological spaces is compact with respect to the product topology. We fix
some .i0 ∈ I and let .vi0 = 0. The space
|| where we are searching || for solutions to the
family of inequalities is now .X = i∈I \{i0 } [αi,i0 , −αi0 ,i ] = i∈I Xi , since among
the inequalities we will have .αi,i0 ≤ vi − vi0 ≤ −αi0 ,i (which, by the way, we know
is a non-empty interval by our assumption of solvability for any finite subset, in the
case the subset .{i0 , i}). Note that X is compact with respect ||to the product topology,
in which a basis for open sets is the Cartesian product . i∈I Ai where .Ai is the
whole interval .Xi except finitely many indices in which .Ai is an open set. The set
of elements .v ∈ X for which a certain inequality .α(i, j ) ≤ vi − vj is not satisfied is
clearly an open set, as it is an open set in the two participating coordinates (product
with the full intervals in all the other components).
Assume there does not exist a solution to the (possibly infinite) system of
inequalities. This means that every point in the product space lies in at least one
of these open sets, namely for every choice of v at least one of the inequalities is not
satisfied. We thus have a cover of the space X, and by compactness there is some
finite sub-cover. As it is finite, a finite number of indices, say only the finite subset
J , participate in the corresponding inequality. The fact that it is a cover means that
not all these inequalities can be satisfied simultaneously within X. But, after adding
the inequalities .αi,i0 ≤ vi − vi0 ≤ −αi0 ,i for the indices .i ∈ J , we know by our
assumption that there is a solution vector .(vi )i∈J , and we may add to it any .vi ∈ Xi
for .i /∈ J , getting a contradiction to the covering property of the finite sub-cover.
We conclude that the original collection was not a cover, namely, there is a solution
to the infinite family of inequalities. u
n
We have thus completed the new and simple proof for the Rockafellar–Rochet–
Rüschendorf theorem for real valued costs.
210 S. Artstein-Avidan
Our motivation for providing the new proof above was that for non-traditional costs
(costs that are allowed to assume .+∞, such as the polar cost) it may happen that a
set is c-cyclically monotone but fails to have a potential .ϕ, even when c is continuous
and the spaces considered are simply .Rn . The proof above, however, crystallized
what the “right” condition for the existence of a potential is, and we call it c-path-
boundedness.
Definition 4.50 Fix sets .X, Y and .c : X × Y → (−∞, ∞]. A subset .G ⊆ X × Y
will be called c-path-bounded if .c(x, y) < ∞ for any .(x, y) ∈ G, and for any
.(x, y) ∈ G and .(z, w) ∈ G, there exists a constant .M = M((x, y), (z, w)) ∈ R such
E
m−1
( )
. c(xi , yi ) − c(xi+1 , yi ) ≤ M.
i=1
It is not hard to see that a c-path-bounded set must be c-cyclically monotone (indeed,
if .(x, y) = (z, w) then if there is some path for which the sum is positive, one can
duplicate it many times to get paths with arbitrarily large sums). It is also not hard
to check that c-path-boundedness is a necessary condition for the existence of a
potential. Our main theorem is that the condition of c-path-boundedness is in fact
equivalent to the existence of a potential.
Theorem 4.51 Let .X, Y be two arbitrary sets and .c : X × Y → (−∞, ∞] be an
arbitrary cost function. For a given subset .G ⊆ X ×Y there exists a c-class function
.ϕ : X → [−∞, ∞] such that .G ⊆ ∂ ϕ if and only if G is c-path-bounded.
c
Note that in Theorem 4.45 we did allow the cost to assume infinite values, so that
we may use it to once again reformulate the problem of finding a potential for a
given set .G ⊆ X × Y as a question regarding the existence of a solution to a linear
system of inequalities:
indexed by .(x, y), (z, w) ∈ G. The main difference between this system and the
one given in (4.9) is that here some of the inequalities are in fact of the form .−∞ ≤
a(x) − a(z), which are satisfied automatically. While this seems at first sight to be
a helpful feature, it ruins the compactness argument we used in Proposition 4.48.
Indeed, one may easily construct a set of inequalities such that any finite subset
of inequalities has a solution but the infinite system does not (here is an example:
.v1 ≤ vi and .vi ≤ v2 − i for .i = 3, 4, . . .). This is precisely the reason we need a
αi,j ≤ xi − xj ,
. i, j ∈ I (4.12)
has a solution if and only if for any .i, j ∈ I there exists some constant .M(i, j )
Em−1that for any m and any .i2 , · · · , im−1 , letting .i = i1 and .j = im one has that
such
.
k=1 αik ,ik+1 ≤ M(i, j ).
Instead of proving Theorem 4.52 directly, we shall prove the following theorem,
which at first glance might seem weaker.
Theorem 4.53 Let .{ai,j }i,j ∈I ∈ [−∞, ∞), where I is some arbitrary index
set. Assume that for any .m ≥ 1 and any .i1 , i2 , · · · , im it holds that .ai1 ,im ≥
Em−1
k=1 aik ,ik+1 . Then the system of inequalities
. ai,j ≤ xi − xj , i, j ∈ I
has a solution.
Clearly, Theorem 4.52 implies Theorem 4.53. In fact, the reverse implication
holds as well. We will show this implication first, namely that Theorem 4.52 follows
from Theorem 4.53.
Proof that Theorem 4.53 implies Theorem 4.52 The “only if” part of Theo-
rem 4.52 is easy and does not require Theorem 4.53. Indeed, let .{αi,j }i,j ∈I ∈
[−∞, ∞), where I is some arbitrary index set, and with .αi,i = 0. Assume that the
system of inequalities
αi,j ≤ xi − xj ,
. i, j ∈ I
has a solution, .(xi )i∈I . Summing the relevant inequalities we see that .M(i, j ) =
xi − xj provides the required bound.
212 S. Artstein-Avidan
For the opposite direction, we will use Theorem 4.53. Assume that for any .i, j ∈
m−1
I there exists some .M(i, j ) such that for any m and any .{ik }k=2 , letting .i1 = i and
.im = j it holds that
E
m−1
. αik ,ik+1 ≤ M(i, j ).
k=1
E
m−1
ai,j = sup{
. αik ,ik+1 : m ∈ N, m ≥ 2, i2 , . . . , im−1 ∈ I }.
k=1
By the above condition, the right hand side is bounded from above and so the
supremum is not .+∞.
We first claim that the system of inequalities .ai,j ≤ xi − xj , satisfies the
conditions of Theorem 4.53.EAssume we are given .i1 , i2 , · · · , im−1 , im , and we
m−1
want to prove that .ai1 ,im ≥ k=1 aik ,ik+1 . Fix .ε > 0. For each .k ∈ [m] use the
(k) (k) (k)
definition of .aik ,ik+1 to pick some .mk and .i2 , . . . , imk −1 such that, letting .i1 = ik
(k)
and .imk = ik+1 , we have
k −1
mE
aik ,ik+1 ≤
. αi (k) ,i (k) + ε/m.
l l+1
l=1
which is naturally arranged as a path from .i1 to .im . Using again the definition of
ai,j , the path thus defined participates in the supremum, and we have that
.
E
m
(E
m
)
ai1 ,im ≥
. (aik ,ik+1 − ε/m) = aik ,ik+1 − ε.
k=1 k=1
As this holds for any .ε, we get the inequality in the condition of Theorem 4.53.
Applying Theorem 4.53, we see that the system of inequalities
. ai,j ≤ xi − xj , (4.13)
admits a solution. Moreover, since .ai,j ≥ αi,j by definition, the resulting vector x
is also a solution of the original system of inequalities. u
n
4 Dualities, Measure Concentration and Transportation 213
Having made the reduction from Theorem 4.52 to Theorem 4.53, we proceed by
proving the latter.
Proof of Theorem 4.53 We use Zorn’s Lemma. Consider the partially ordered set
of pairs .(J, fJ ) where .J ⊆ I and .fJ : J → R are such that for any .i, j ∈ J
we have .fJ (i) − fJ (j ) ≥ ai,j . We know the set is non-empty because it contains
pairs .({i0 }, 0). The partial order we consider is .(J, fJ ) ≤ (K, fK ) if .J ⊆ K and
.fK |J = fJ .
First let us notice that every chain has an upper bound. Assume .(Jα , fJα )α∈A
is a chain (namely any two elements are comparable). Consider .J = ∪α Jα and
.fJ = ∪α fJα . This function is well defined because of the chain property (at a point
.i ∈ J it is defined as .fJα (i) for any .α with .i ∈ Jα ). The pair .(J, fJ ) is in our
set because if .i, j ∈ J then for some .α we have .i, j ∈ Jα , so .f |Jα satisfies the
inequality on .fJ (i) − fJ (j ) ≥ ai,j and so does .fJ . Finally, .(J, fJ ) is clearly an
upper bound for the chain. So, we have shown that every chain has an upper bound,
and we may use Zorn’s lemma to find a maximal element. Denote the maximal
element by .(J0 , fJ0 ).
Assume towards a contradiction that .J0 /= I , that is, there exists some element
.i0 ∈ I such that .i0 /∈ J0 . If we are able to extend .fJ0 to be defined on .{i0 } in
such a way that all inequalities with indices of the form .(i0 , j ) and .(j, i0 ) with
.j ∈ J0 still hold, we will contradict maximality and complete the proof. Note that
the inequalities that need to be satisfied in order to extend the function are
We can rewrite the condition as .ai0 ,j + ak,i0 ≤ fJ0 (k) − fJ0 (j ). Recall that under
our assumptions .ak,j ≥ ak,i0 + ai0 ,j . Since .fJ0 already satisfies the inequality
.ak,j ≤ fJ0 (k) − fJ0 (j ), we know the above inequality holds for any .j, k, and so the
inequality (4.14) holds and we may extend the function .fJ0 . This is a contradiction
to the maximality, and we conclude .J0 = I , so that we have found a solution to the
full system of inequalities. u
n
214 S. Artstein-Avidan
We are now ready to prove our new Rockafellar-type theorem for non-traditional
costs.
Proof of Theorem 4.51 One direction is immediate: assume .G ⊆ X × Y satisfies
that for some c-class .ϕ : X → [−∞, ∞], .G ⊆ ∂ c ϕ. Then
holds for all .(x, y), (z, w) ∈ G. So, given a pair .(x, y), (z, w) ∈ G we set
M = M((x, y), (z, w)) = ϕ(x) − ϕ(z). Any sum, as in the definition of c-path-
.
We discussed in Lecture I that usually one cannot find exact extremizers for the
volume of a t-extension of a set of fixed measure, and the fact that finding such
estimates is desirable since when these are strong, in the sense that the t-extension
has large measure (namely, when we have “concentration of measure”), this is a
great tool for proving many beautiful results. In this lecture we shall see other ways
for obtaining concentration, without knowing exact extremizers, some of which are
intimately connected with the transportation of measure results from Lecture II.
We will not be discussing the extremely rich array of examples for applications
of concentration of measure, which started with the proof of Dvoretzky’s theorem
by V. Milman and these days spans a vast amount of literature. Many examples
and insights into this topic can be found in the books [6, 8, 21] and in the
references therein. In this lecture, we will mainly concentrate (!) on methods to
obtain concentration.
As mentioned, concentration of measure is a phenomenon of high dimensions
which is responsible to many counter-intuitive (until intuition changes and these
become intuitive) results. Let us approach it, as an introduction to this last lecture,
from a somewhat non-standard angle. We recall the beautiful formula for computing
the surface area of a convex body using its projections.
4 Dualities, Measure Concentration and Transportation 215
Sometimes one uses integration on the sphere with respect to usual Lebesgue
measure, the relation being .dσ (u) = du/Voln−1 (S n−1 ) = du/(nκn ) rewriting the
formula as
f
1
.Voln−1 (∂K) = Voln−1 (Pu⊥ K)du.
κn−1 S n−1
Proof We work with a polytope P first. For a generic .u ∈ S n−1 , each facet .Fi of P
has some angle .θi between its normal and u which is not .π/2. When such a facet is
projected onto .u⊥ , its area when projected is .| cos(θ )| times its original area. Clearly
the projection .Pu⊥ (P ) is covered twice by the projections of the facets of P . We get
that
1E
m
Voln−1 (Pu⊥ (P )) =
. | cos(θi )|Voln−1 (Pi ).
2
i=1
By rotation invariance, the latter integral does not depend on .ni and in simply a
constant depending on the dimension (for example, let .n = (1, 0, . . . , 0), so that
.cos(θ ) = u1 ). We have thus shown that there exists some constant .cn such that for
Since both sides are well-defined for convex bodies and are clearly monotone, and
since we have equality for all polytopes, we have equality for bodies as well. Thus,
for the same .cn (which is half the integral of .|u1 | over .u ∈ S n−1 ) we get for all
convex K that
f
. Voln−1 (Pu⊥ (K))dσ (u) = cn Voln−1 (∂K).
S n−1
Looking at the proof above, we see that (along with proving a nice formula for
surface area) we computed
f f
2 2κn−1
. |x1 |dσ (x) = Voln−1 (Pu K)dσ (u) = .
S n−1 Voln−1 (∂K) S n−1 nκn
Since we know what is the asymptotic behavior of .κn we may use it to compute:
n−1
2κn−1 2 π 2 r( n2 + 1) 2 n + 1 1/2 / 1
. = n = √ ( ) = 2/π √ .
nκn n π 2 r( n−1 + 1) n π 2 n
2
So, the average of .|u1 | on the sphere is of the order .n−1/2 , which is quite small. If
one considers the median of .|u1 | instead of average, and up to factor 2 the median of
a positive quantity is less than the average, this means that about .1/2 of the measure
the whole sphere is concentrated near the hyperplane .x1 = 0, in a strip of width
of √
.1/ n. Since the sphere is rotation invariant, this applies to any hyperplane through
the origin.
Concentration is not just about half the volume, but about the majority of volume.
Indeed, one may compute .σ {x ∈ S n−1 : |x1 | < r}, to see how it behaves with
respect to r. The above computation shows that this integral will be .1/2 when r
is of the order .n−1/2 . One may write out this integral and estimate it. However, to
avoid these computations, we can estimate it rather well (say for .r < 1/2) by the
method explained in Sect. 4.1.3 to get
Our first example for a concentration type inequality will be a direct application
of the Brunn–Minkowski inequality, which implies a useful form of concentration.
Here the neighborhood is not of euclidean distance, but captured in the form of
the volume belonging to a dilate of the set considered. Borell’s lemma from [17]
describes some form concentration of volume in convex bodies in .Rn : if .A ∩ K
captures more than half of the volume of K, then the percentage of K that stays
outside tA, when .t > 1, decreases exponentially with respect to t as .t → ∞, with a
bound that does not depend at all on the body K or the dimension n.
4 Dualities, Measure Concentration and Transportation 217
Theorem 4.55 (Borell’s Lemma) Let K be a convex body in .Rn with volume
.Voln (K) = 1, and let A be a closed, convex and centrally symmetric set such that
( ) t+1
1−δ 2
Voln (K ∩ (R \ tA)) ≤ δ
.
n
.
δ
Remark 4.56 The same is true when .Vol(A ∩ K) is replaced by any log-concave
probability measure .μ (see Definition 4.5). Such measures serve as a natural habitat
for geometric inequalities, as we discussed in the section on functional forms of
geometric inequalities, Sect. 4.1.4.
Proof We first show that
2 t −1
(Rn \ A) ⊇
. (Rn \ tA) + A.
t +1 t +1
2 t −1
a=
. y+ a1 ,
t +1 t +1
1 t +1 t −1
. y= a+ (−a1 ) ∈ A,
t 2t 2t
because of the convexity and symmetry of A. This means that .y ∈ tA, which is a
contradiction.
Since K is convex, we have
2 ( n ) t − 1( )
(Rn \ A) ∩ K ⊇
. (R \ tA) ∩ K + A∩K .
t +1 t +1
2 t−1
= Voln ((Rn \ tA) ∩ K) t+1 δ t+1 .
which by Markov’s inequality has measure at least .2/3. Splitting the integral for
q
||f ||q to two parts, and making some simple estimates, will help the interested reader
.
As mentioned in Sect. 4.1.3, the fact that for a set .A ⊂ Rn with .γn (A) = 1/2, we
have that .γn (At ) ≥ 1−e−t /2 follows from comparison to half-spaces which are the
2
isoperimetric extremizers. However, let us demonstrate another proof for the same
result (with a slightly worse estimate) which avoids using exact extremizers, as such
a method generalizes to other spaces as well.
Theorem 4.58 Let .A ⊂ Rn with .γn (A) = 1/2, then for any .t > 0, .γn (At ) ≥
1 − 2e−t /4 .
2
Proof Consider the following three functions: .f (x) = exp(d(A, x)2 /4)e−|x| /2 ,
2
.g(y) = 1A (y)e
−|y|2 /2 and .h(z) = e−|z|2 /2 , and fix .λ = 1/2. These satisfy the
f f f
Therefore, . h ≥ ( f g)1/2 andf the same is true whenf both sides are normalized
by .(2π )−n/2 . Note that .(2π )−n/2 h = 1 and .(2π )−n/2 g = γn (A) = 1/2, so the
inequality implies
f
2 /4
. ed(x,A) dγn ≤ 2.
This type of inequality (if it is given for some probability measure, possibly with
other constants) implies a concentration inequality by Markov’s inequality
f 2 /4
ed(x,A) dx
≤ 2e−r
2 /4
1 − γ (Ar ) = γ ({x : d(x, a) ≥ r}) ≤
.
2
.
er /4
This completes the proof. u
n
We mention that if .c(x, x) = 0 then .ϕ c ≤ −ϕ, and we note that for every bounded
measurable function .ϕ : X → R we have
f f
. eϕ dμ · e−ϕ dμ ≥ 1
by Hölder’s inequality.
The Cost-Santaló inequality is to do with the possibility of reversing this Hölder
inequality when .−ϕ is replaced by .ϕ c .
Definition 4.59 (Cost-Santaló Inequality) Let .c : X × X → R. We say that
(X, μ) satisfies a Cost-Santaló inequality with respect to the cost function c if for
.
given probability measure and cost, we need only consider pairs .ϕ, ϕ c where .ϕ is in
the c-class.
A good example to keep in mind is that of the quadratic cost, .c(x, y) = |x−y|2 /2
on .Rn × Rn . In this case the c-class consists of functions .|x|2 /2 − φ(x) where .φ
is convex lower semi continuous, and the c-transform of this function is simply
.|y| /2 − Lφ(y). Using this together with the Gaussian measure, a Cost-Santaló
2
(where the normalizing factor for the Gaussian measure was moved to the right).
This is nothing other than the functional Blaschke–Santaló inequality, which
we have encountered in Sect. 4.1.4. We discussed its geometric counterpart
◦
.Vol(K)Vol(K ) ≤ Vol(B ) in Sect. 4.1.4. It is important to note here that this
n 2
2
inequality is not true. As we shall see shortly, the Gaussian measure does satisfy
a Cost-Santaló inequality but with respect to the cost .|x − y|2 /4. Still, the above
inequality is almost true, in particular it is true when .ϕ is an even function, and it is
true if some centroid is the origin. Our current aim is to show two claims. The first is
that a certain class of measures, which include Gaussian but are more general, satisfy
the Cost-Santaló inequality with the quadratic cost (properly normalized). Second,
that satisfying a Cost-Santaló inequality with respect to some cost function, implies
a form of concentration where the t-extension of a set is measured with respect to
the cost function. When the cost is quadratic, the extension is the standard Euclidean
one. Let us start with the latter. We start with the following easy statement.
Proposition 4.60 Let .(X, μ) be a probability space. Assume that .μ satisfies the
Cost-Santaló inequality (4.15) with respect to some cost function c. Then, for any
measurable .A ⊆ X and .t > 0,
({ }) 1 −t
1 − μ x : inf c(x, y) < t ≤
. e .
y∈A μ(A)
Proof Consider the function .ϕ which is a .{0, −∞} indicator of the set A, namely
is equal to 0 on the set A and .−∞ outside. (If one insists on real valued .ϕ, let it be
.{0, −N } valued and eventually take .N → ∞). In this case
as claimed. u
n
In particular we get that if .μ satisfies the Cost-Santaló inequality with respect to
c(x, y) = κ4 d(x, y)2 then for a set A of measure .1/2
.
x+y κ
U (x) + U (y) − 2U (
. ) ≥ |x − y|2 .
2 4
Then .μ satisfies the Cost-Santaló inequality with respect to the cost function
c(x, y) = κ4 |x − y|2 .
.
x+y
U (x) + U (y) − 2U (
. ) ≥ c(x, y).
2
Then .μ satisfies the Cost-Santaló inequality with respect to the cost function c.
Proof We show that the conditions in the Prékopa–Leindler inequality hold for .λ =
1/2 and the triplet of functions
c −U
f = eϕ
. , g = eϕ−U and h = e−U .
which is immediate from the definition of the cost transform. The inequality
which the Prékopa–Leindler theorem implies is precisely the desired inequality,
completing the proof. u
n
It turns out that one can weaken the inequality and still obtain a similar, but not
identical, form of concentration.
Definition 4.63 (Weak Cost-Santaló Inequality) We say that .(X, μ) satisfies a
Weak Cost-Santaló inequality with respect to the cost function c if for any bounded
measurable function .ϕ : X → R,
f f
c
. eϕ dμ · e ϕdμ ≤ 1. (4.16)
f f
Note that by Jensen’s inequality . eϕ dμ ≥ e ϕdμ , so that the new condition
is indeed weaker (namely, any measure satisfying the Cost-Santaló inequality will
automatically satisfy the Weak Cost-Santaló inequality). Nevertheless, satisfying a
Weak Cost-Santaló inequality also implies a form of concentration, as the following
claim indicates.
Proposition 4.64 Let .(X, μ) be a probability space. Assume that .μ satisfies the
Weak Cost-Santaló inequality (4.16) with respect to some cost function .c ≥ 0. Then,
for any measurable .A ⊆ X and .t > 0,
({ })
1 − μ x : inf c(x, y) < t ≤ e−μ(A)t .
.
y∈A
Proof We shall plug into the inequality the function .ϕ which is 0 on A and .−t on
X \ A. Note that if .infy∈A c(x, y) < t then .ϕ c (x) = infy∈X (c(x, y) − ϕ(y)) < t.
.
On the other hand, if .infy∈A c(x, y) ≥ t then .ϕ c (x) = infy∈X (c(x, y) − ϕ(y)) ≥ t
as this inequality is true both for .y ∈ A and .y /∈ A (one of them due to the cost,
and the other due to .ϕ being t outside A). Therefore .μ(x : infy∈A c(x, y) ≥ t) =
μ (x : ϕ c (x)f ≥ t).
Clearly . ϕdμ = −tμ(X \ A) = −(1 − μ(A))t and by applying (4.16) to the
function .ϕ we get that
f
c
. eϕ dμ ≤ et (1−μ(A)) .
4 Dualities, Measure Concentration and Transportation 223
as claimed. u
n
It turns out that the Weak Cost-Santaló inequality is completely equivalent to an
inequality between cost and entropy. This type of inequality concerns, again, a fixed
cost and some base measure .μ, and relates the Wasserstein distance from .μ of any
.ν which is absolutely continuous with respect to .μ with the relative entropy of .ν
with respect to .μ (we define these notions below). To approach this topic we start
by defining entropy and relative entropy, and showing this connection. We will get
more acquainted with inequalities about entropy and concentration in the section
after the next one.
4.3.5 Entropy
Definition 4.65 (Entropy) Let (X, μ) be a probability space. For every non-
negative integrable function f : X → R, the entropy of f with respect to μ is
the quantity
f f (f )
. Entμ (f ) = f ln f dμ − f dμ · ln f dμ ∈ [0, +∞].
X X X
Note that Entμ (f ) ≥ 0 by Jensen’s inequality for the convex function x ln x and that
entropy is homogeneous of degree 1, that is, Entμ (λf ) = λ Entμ (f ) for all λ > 0.
We mention that in the literature sometimes the sign of entropy is reversed. Also,
occasionally the reference measure is Lebesgue and not a probability measure.
Definition 4.66 (Relative Entropy) In the case where f = dμ dν
for a probability
measure ν, the entropy of f is called the relative entropy of μ with respect to ν. We
define
f
.H (ν|μ) := Entμ (f ) = ln f dν.
X
224 S. Artstein-Avidan
f
Remark 4.67 We mention that, when X f dμ = 1 (as is in the case where f =
dν/dμ),
f
p
. lim ln ||f ||p = f ln f dμ,
p→1+ p−1 X
and the supremum remains the same if we consider only g which is bounded from
above and from below.
f
Proof Since the relation is homogeneous, we may assume that Eμ (f ) = f dμ =
1. Young’s inequality (which is due to the fact that the function u ln u − u on R+ is
the Legendre dual of the function ev ) states that
uv ≤ u ln u − u + ev , u ≥ 0, v ∈ R
.
f
and hence, if g is a function on X such that eg dμ ≤ 1, we have
f f f f f
. fg dμ ≤ f ln f dμ − f dμ + eg dμ ≤ f ln f dμ
f
due to the assumption that f dμ = 1. Taking the supremum we get
{f f } f
. sup f g dμ : eg dμ ≤ 1 ≤ f ln f dμ = Entμ (f ).
Finally, plugging in g =
f ln f (for which
f Young’s inequality becomes an equality),
which is allowed since eln f dμ = f dμ = 1, we have that
{f f } f
. sup f g dμ : eg dμ ≤ 1 ≥ f ln f dμ,
and the equality follows. For the last assertion of the lemma note that if we set fN =
min{max{f, N1 }, N } and gN = ln(fN /Eμ (fN )) then gN is bounded from above and
f f
below, eln gN dμ = 1 for all N, and as N → ∞ we get fgN dμ → Entμ (f ).
u
n
4 Dualities, Measure Concentration and Transportation 225
With the notion of entropy and relative entropy, we may now discuss yet another
inequality which a probability space might satisfy with respect to a given cost,
and which in turn is connected with concentration. As we shall shortly see, this
inequality is equivalent to one of the two previous ones, although at this point it is
not necessarily very easy to see this.
Definition 4.69 (Cost-Entropy Inequality) We say that .(X, μ) satisfies a Cost-
Entropy inequality with respect to the cost function c if for any bounded measurable
function .ϕ : X → R,
Wc (μ, ν) ≤ H (ν|μ).
. (4.17)
Here we are using the notation .Wc (μ, ν) to represent the total cost .C(μ, ν) (so
that the role of the cost function c is more apparent.) The letter W corresponds to
Wasserstein.
Remark 4.70 This is a good time to take a short break and see what the total cost is
for certain special cost functions, and whether we can better understand it. When the
cost function is defined using a distance function .c : X×X → R, .c(x, y) = d(x, y),
the c-class is the class of all 1-Lip. functions, and the transform is simply .ϕ |→ −ϕ.
In particular, using Kantorovich Duality Theorem,
f f
Wc (μ, ν) = sup{
. ϕdμ − ϕdν : ϕ is 1 Lip.}
A special case is where .d(x, y) = d0 (x, y) is the trivial metric, satisfying .d0 (x, x) =
0 and .d0 (x, y) = 1 if .x /= y. In this case 1-Lip. functions are simply functions whose
image is in a unit interval, and as
f f
Wc (μ, ν) = sup{
. ϕdμ − ϕdν : ϕ(X) ⊆ [0, 1]}
where we have used the fact that adding a constant to .ϕ does not change the integral
difference. Since this is a linear function it is extremized on extremal point of this
cone which are precisely function attaining values in .{0, 1}, so we get that
is the total variation distance between the measures. In particular if .ν << μ and
dν(x) = f (x)dμ(x) then
.
f
1
Wc (μ, ν) = ||μ − ν||T V
. = |1 − f |dμ.
2
226 S. Artstein-Avidan
Going back to our main theme, namely the newly introduced notion of a Cost-
Entropy inequality, there are two main claims we aim to show. The first is that
satisfying such an inequality implies concentration (with respect to a “distance”
determined by the cost, as was the case in the previous sections). The second is
a connection between this notion and the previous one. We will in fact show that
they are equivalent. This makes the “first” part redundant (as we have already
shown Weak Cost-Santaló implies concentration, and we will get exactly the same
estimate) but we do this nevertheless as we believe that seeing different proofs
makes the theory more transparent. Moreover, the proof involves measure transport
(naturally) which is especially relevant to this note. Finally, we can show, using
Brenier map, that Gaussian-type measures satisfy this condition.
We start by showing how a Cost-Entropy inequality implies concentration.
Proposition 4.71 Let .(X, μ) be a probability space. Assume that .μ satisfies the
Cost-Entropy inequality (4.17) with respect to some cost function .c ≥ 0. Then, for
any measurable .A ⊆ X and .t > 0,
({ })
1 − μ x : inf c(x, y) < t ≤ e−μ(A)t .
.
y∈A
{ }
Proof Fix some A of a given measure, and let .B = x : infy∈A c(x, y) ≥ t .
We would like to bound .μ(B) from above. To this end, consider the measure .ν
defined as .μ restricted to B, and normalized. So, letting .m = μ(B), we have that
.ν = m
−1 μ| . Given some transport plan .π ∈ ||(μ, ν), it clearly has to transport all
B
the measure inside A to the support of .ν, namely to B. The cost .c(x, y) for .x ∈ A
and .y ∈ B is at least t, and so
Wc (μ, ν) ≥ tμ(A).
.
On the other hand, the entropy of .ν with respect to .μ is particularly easy to compute.
Indeed,
f
.H (ν|μ) = m−1 ln(m−1 )dμ = μ(B)m−1 ln(m−1 ) = − ln μ(B).
B
tμ(A) ≤ − ln μ(B),
. so μ(B) ≤ e−tμ(A) .
Proof of Proposition 4.72 We will make heavy use of the Kantorovich Duality
Theorem, which we recall allows us to represent
f f
Wc (μ, ν) =
. sup ϕdμ + ψdν.
ϕ,ψ∈Cb (X), admissible
In fact, as we have discussed, the supremum can, instead, run over all .ϕ, ψ = ϕ c
which are in the c-class.
Assume first that .μ satisfies the Cost-Entropy inequality. To show that the Weak
Cost-Santaló inequality holds, consider a bounded continuous function .ϕ : X → R.
Together with .ϕ c they form an admissible pair. Letting .g = dμdν
, the Cost-Entropy
inequality implies
f f f
. ϕdμ + ϕ c gdμ ≤ Wc (μ, ν) ≤ Ent(g) = g ln gdμ.
c f c
As we are free to chose .ν (namely g), we pick .g = eϕ / eϕ dμ.
Plugging it into the inequality and canceling some terms we obtain
f (f )
c
. ϕ dμ + ln eϕ dμ ≤ 0.
X X
{f f }
H (ν | μ) = Entμ (g) = sup
. hg dμ : eh dμ ≤ 1 .
Given an admissible
f pair .ϕ, ψ for the cost c we recall that .ψ ≤ ϕ c and we let
.h = ϕ +
c ϕdμ. Note that by the Weak Cost-Santaló inequality
f f f
c
. eh dμ = e ϕdμ
eϕ dμ ≤ 1
so we may use h in the equivalent formulation for entropy above, and get
f f f f f f
H (ν|μ) ≥
. gh = ϕ c gdμ + ϕdμ gdμ = ϕ c dν + ϕdμ
and as this was true for any .ϕ, using Kantorovich Duality Theorem we complete the
proof. u
n
228 S. Artstein-Avidan
As promised in the previous section, we will next show how Brenier’s theorem
allows us to prove that the Gaussian measure satisfies a Weak Cost-Santaló
inequality for the quadratic cost, in the usual normalization. These theorems are due
to Marton and to Talagrand [36, 50]. In particular this show that these inequalities
are not equivalent, since the Cost-Santaló inequality does hold for the Gaussian
measure and quadratic costs .|x − y|2 /2 but only for .|x − y|2 /4.
Proposition 4.73 (Marton-Talagrand) Consider .X = Rn with the standard
Gaussian measure .γn , let .c(x, y) = |x − y|2 /2, and let .ν = f (x)dγn (x). Then
Wc (γn , ν) ≤ H (ν|γn ).
.
e−|x| = e−|T x|
2 /2 2 /2
dγn (x) = dν(T x) det(DT (x)),
. i.e. f (T x) det(DT (x)).
Computing the relative entropy, using that .DT (x) = ∇ 2 ϕ > 0 and the differential
relation we get that
f
H (ν|γn ) =
. f (y) ln(f (y))dγn (y)
f
1
f (T x) ln(f (T x))e−|T x|
2 /2
= det(DT (x))dx
(2π )n/2
f ( )
e|T x| /2−|x| /2 −|x|2 /2
2 2
1
= ln e dx
(2π )n/2 det(DT (x))
f ( )
|T x|2 |x|2
= − − ln(det(DT (x))) dγn (x).
2 2
f
We claim that the term in the last integral is greater than . |T (x) − x|2 /2dγn (x). If
we show this, we are done, as this precisely was the expression for .Wc we got using
the Brenier map. We are thus left, after rearrangement, with showing that
f f
. ln(det(DT (x)))dγn (x) ≤ <T (x) − x, x>dγn (x).
4 Dualities, Measure Concentration and Transportation 229
To show this we use that .DT = ∇ 2 ϕ is diagonalizable, and its diagonal entries,
in the relevant basis, are positive, namely some .(ti (x))ni=1 . Note that here .ti (x) =
∂2
ϕ(x)or, letting .T = ∇ϕ = (T1 , . . . , Tn ), we have that .ti (x) = ∂x∂ i Ti (x). So,
∂xi2
En
.ln(det(DT (x))) = i=1 ln ti (x). We use the inequality .ln(ti (x)) ≤ ti (x) − 1 and
get
f f E
n n f
E
. ln(det(DT (x)))dγn (x) ≤ (ti (x) − 1)dγn (x) = (ti (x) − 1)dγn (x)
i=1 i=1
n f
E
= (2π )−n/2 e−|x|
2 /2
(ti (x) − 1)dx
i=1
n f
E ∂(Ti − Ii )
= (2π )−n/2 e−|x|
2 /2
(x)dx.
∂xi
i=1
Here we used the notation .Ii (x) = xi , namely the ith coordinates of the identity
operator I . We next use integration by parts, justified as the Gaussian measure
decreases rapidly, to get that
n f
E E n f
∂(Ti − Ii ) ∂ −|x|2 /2
e−|x|
2 /2
. (x)dx = − (Ti − Ii )(x) e dx
∂xi ∂xi
i=1 i=1
n f
E
((T x)i − xi )xi e−|x|
2 /2
= dx
i=1
References
1. S. Artstein-Avidan, V.D. Milman, A new duality transform. C. R. Math. Acad. Sci. Paris 346,
1143–1148 (2008)
2. S. Artstein-Avidan, V.D. Milman, The concept of duality in convex analysis, and the charac-
terization of the Legendre transform. Ann. of Math. (2) 169, 661–674 (2009)
230 S. Artstein-Avidan
3. S. Artstein-Avidan, V.D. Milman, Hidden structures in the class of convex functions and a new
duality transform. J. Eur. Math. Soc. 13, 975–1004 (2011)
4. S. Artstein-Avidan, Y. Rubinstein, Differential analysis of polarity: polar Monge Ampère,
Hamilton–Jacobi and conservation laws. J. Anal. Math. 132, 133–156 (2017)
5. S. Artstein-Avidan, B. Klartag, V.D. Milman, The Santaló point of a function, and a functional
form of the Santaló inequality. Mathematika 51, 33–48 (2004)
6. S. Artstein-Avidan, A. Giannopoulos, V.D. Milman, Asymptotic Geometric Analysis. Part I.
Mathematical Surveys and Monographs, vol. 202 (American Mathematical Society, Provi-
dence, 2015)
7. S. Artstein-Avidan, D.I. Florentin, A. Segal, Functional Brunn-Minkowski inequalities induced
by polarity. Adv. Math. 364, 107006 (2020)
8. S. Artstein-Avidan, A. Giannopoulos, V.D. Milman, Asymptotic Geometric Analysis. Part II.
Mathematical Surveys and Monographs, vol. 261 (American Mathematical Society, Provi-
dence, 2021)
9. S. Artstein-Avidan, S. Sadovsky, K. Wyczesany, A Rockafellar-type theorem for non-
traditional costs. Adv. Math. 395, 108157 (2022)
10. K.M. Ball, Ph.D. Dissertation. Cambridge, 1986
11. K.M. Ball, An elementary introduction to modern convex geometry, in Flavors of Geometry.
Mathematical Sciences Research Institute Publications, vol. 31 (Cambridge University Press,
Cambridge, 1997)
12. F. Barthe, Inégalités fonctionelles et géomėtriques obtenues par transport des mesures. Thèse
de Doctorat de Mathématiques, Université de Marne-la-Vallée, 1997
13. F. Barthe, Inégalités de Brascamp-Lieb et convexité. C. R. Acad. Sci. Paris Ser. I Math. 324,
885–888 (1997)
14. F. Barthe, On a reverse form of the Brascamp-Lieb inequality. Invent. Math. 134, 335–361
(1998)
15. A. Barvinok, A Course in Convexity. Graduate Studies in Mathematics, vol. 54 (American
Mathematical Society, Providence, 2002)
16. C. Borell, The Brunn-Minkowski inequality in Gauss space. Invent. Math. 30, 207–216 (1975)
17. C. Borell, Convex set functions in d-space. Period. Math. Hungar. 6, 111–136 (1975)
18. H.J. Brascamp, E.H. Lieb, On extensions of the Brunn-Minkowski and Prékopa-Leindler
theorems, including inequalities for log-concave functions, and with an application to the
diffusion equation. J. Funct. Anal. 22, 366–389 (1976)
19. Y. Brenier, Décomposition polaire et réarrangement monotone des champs de vecteurs. C. R.
Acad. Sci. Paris Ser. I Math. 305, 805–808 (1987)
20. Y. Brenier, Polar factorization and monotone rearrangement of vector-valued functions.
Commun. Pure Appl. Math. 44, 375–417 (1991)
21. S. Brazitikos, A. Giannopoulos, P. Valettas, B.H.Vritsiou, Geometry of Isotropic Convex
Bodies. Mathematical Surveys and Monographs, vol. 202 (American Mathematical Society,
Providence, 2014)
22. R.G. Douglas, On extremal measures and subspae density. Mich. Math. J. 11, 243–246 (1964)
23. T. Figiel, J. Lindenstrauss, V.D. Milman, The dimension of almost spherical sections of convex
bodies. Acta Math. 139, 53–94 (1977)
24. M. Fradelizi, M. Meyer, Some functional forms of Blaschke-Santaló inequality. Math. Z. 256,
379–395 (2007)
25. M. Fradelizi, A. Hubard, M. Meyer, R. Roldan-Pensado, A. Zvavitch, Equipartitions and
Mahler volumes of symmetric convex bodies. Am. J. Math. 144, 1201–1219 (2022)
26. R.J. Gardner, The Brunn-Minkowski inequality. Bull. Am. Math. Soc. (N.S.) 39, 355–405
(2002)
27. P.M. Gruber, Convex and Discrete Geometry. Grundlehren der mathematischen Wissen-
schaften, vol. 336 (Springer, Berlin, 2007)
28. H. Hadwiger, D. Ohmann, Brunn-Minkowskischer Satz und Isoperimetrie. Math. Z. 66, 1–8
(1956)
4 Dualities, Measure Concentration and Transportation 231
29. L.H. Harper, Optimal numbering and isoperimetric problems on graphs. J. Comb. Theory 1,
385–393 (1966)
30. D. Hug, W. Weil, Lectures on Convex Geometry. Graduate Texts in Mathematics, vol. 286
(Springer, Cham, 2020)
31. H. Iriyeh, M. Shibata, Symmetric Mahler’s conjecture for the volume product in the 3-
dimensional case. Duke Math. J. 6, 1077–1134 (2020)
32. J. Lehec, A direct proof of the functional Santaló inequality. C. R. Math. Acad. Sci. Paris 347,
55–58 (2009)
33. J. Lehec, Partitions and functional Santaló inequalities. Arch. Math. 92, 89–94 (2009)
34. L. Leindler, On a certain converse of Hölder’s inequality. II. Acta. Sci. Math. Szeged 33, 217–
223 (1972)
35. P. Lévy, Problémesconcrets d’analyse fonctionnelle, in Avec un complément sur les fonction-
nelles analytiques par F. Pellegrino (French) 2nd edn. (Gauthier-Villars, Paris, 1951)
36. K. Marton, Bounding d-distance by informational divergence: a method to prove measure
concentration. Ann. Probab. 24(2), 857–866 (1996)
37. J. Lindenstrauss, A remark on extreme doubly stochastic measures. Am. Math. Mon. 72, 379–
382 (1965)
38. R.J. McCann, A convexity theory for interacting gases and equilibrium crystals. Ph.D. Thesis,
Princeton University, 1994
39. R.J. McCann, Existence and uniqueness of monotone measure preserving maps. Duke Math.
J. 80, 309–323 (1995)
40. V.F. Milman, New proof of the theorem of Dvoretzky on sections of convex bodies. Funct.
Anal. Appl. 5, 23–37 (1971)
41. G. Monge, Mémoire sur la théorie des déblais et des remblais. Histoire de l’Académie Royale
des Sciences de Paris, avec les Memoires de Mathematique et de Physique pour la meme annee
(1781), pp. 666–704
42. A. Prékopa, On logarithmic concave measures and functions. Acta Sci. Math. Szeged 34,
335–343 (1973)
43. Y.V. Prokhorov, Convergence of random processes and limit theorems in probability theory.
Theory Probab. Appl. 1(2), 157–214 (1956)
44. R.T. Rockafellar, Convex Analysis. Princeton Mathematical Series, vol. 28 (Princeton
University Press, Princeton, 1970)
45. J.C. Rochet, A necessary and sufficient condition for rationalizability in a quasi-linear context.
J. Math. Econ. 16, 191–200 (1987)
46. L. Rüschendorf, On c-optimal random variables. Stat. Probab. Lett. 27, 267–270 (1996)
47. E. Schmidt, Die Brunn-Minkowskische Ungleichung und ihr Spiegelbild sowie die
isoperimetrische Eigenschaft der Kugel in der euklidischen und nichteuklidischen Geometrie. I
(German) Math. Nachr. 1, 81–157 (1948)
48. R. Schneider, Convex Bodies: the Brunn-Minkowski Theory. Encyclopedia of Mathematics and
its Applications, vol. 151, 2nd expanded edn. (Cambridge University Press, Cambridge, 2014)
49. V.N. Sudakov, B.S. Cirel’son, Extremal properties of half-spaces for spherically invariant
measures (Russian). Problems in the theory of probability distributions, II. Zap. Naucn. Sem.
Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 41, 14–24, 165 (1974)
50. M. Talagrand, Transportation cost for Gaussian and other product measures. Geom. Funct.
Anal. 6(3), 587–600 (1996)
51. C. Villani, Optimal Transport: Old and New. Grundlehren der mathematischen Wissen-
schaften, vol. 338 (Springer-Verlag, Berlin, 2009)
Chapter 5
Symmetrizations
5.1 Introduction
The idea of replacing an object by one that retains some of its features but is
in some sense more symmetrical has been extremely fruitful over the years. The
object may be a set or a function, for example, and the process is then often called
symmetrization or rearrangement, respectively. Steiner symmetrization, introduced
by Jakob Steiner around 1836 in his attempt to prove the isoperimetric inequality,
is still today a potent tool for establishing crucial inequalities in geometry. The
influence of such inequalities, which often have analytical versions, extends far
beyond geometry to other areas such as analysis and PDEs, and even outside
mathematics, to economics and finance.
The topic received a huge boost in 1951 from the classic text of Pólya and Szegő
[72]. By this time, many other types of symmetrization had been introduced, with
similar applications. The general idea is to find a symmetrization that preserves one
physical quantity, while not increasing (or sometimes not reducing) another. As well
as volume, surface area, and mean width, the book by Pólya and Szegő considers
electrostatic capacity, principal frequency (the first eigenvalue of the Laplacian), and
torsional rigidity, thereby extending the scope to mathematical physics.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 233
A. Colesanti, M. Ludwig (eds.), Convex Geometry, C.I.M.E. Foundation Subseries
2332, https://doi.org/10.1007/978-3-031-37883-6_5
234 G. Bianchi and P. Gronchi
5.2 Preliminaries
Throughout this chapter we use the notation and many of the notions which have
been introduced in Chap. 1, but we need to introduce some more notation.
Let .D n be the open unit ball in .Rn . If .x, y ∈ Rn we write .[x, y] for the line
segment with endpoints x and y. If X is a set, we denote by .conv X, .clo X, and
.dim X the convex hull, closure, and dimension (that is, the dimension of the affine
hull) of X, respectively.
Throughout the paper, the term subspace means a linear subspace. The Grass-
mannian of k-dimensional subspaces in .Rn is denoted by .G(n, k). If H is a subspace
of .Rn , then .X|H is the (orthogonal) projection of X on H and .x|H is the projection
of a vector .x ∈ Rn on H . Moreover, .X† denotes the reflection of X in H , i.e., the
image of X under the map that takes .x ∈ Rn to .2(x|H ) − x. If .x ∈ Rn \ {o}, then
.x
⊥ is the .(n − 1)-dimensional subspace orthogonal to x.
We write .Hk for k-dimensional Hausdorff measure in .Rn , where .k ∈ {1, . . . , n}.
We denote by .Cn , .Mn , and .Ln the class of non-empty compact sets, .Hn -measurable
sets, and .Hn -measurable sets of finite .Hn -measure, respectively, in .Rn . Let .Kn be
the class of convex bodies, i.e. non-empty compact convex subsets of .Rn , and let .Knn
be the class of members of .Kn with interior points. For .K ∈ Kn , .S(K) denotes its
surface area, defined in Sect. 1.9. If .K ∈ Knn then .S(K) = Hn−1 (∂K). By .κn we
denote the volume .Hn (B n ) of the unit ball in .Rn .
The Blaschke addition .K # L of .K, L ∈ Knn is a convex body whose surface area
measure is
A∗ = {x ∈ Rn : 0(A, x) = 1}.
.
236 G. Bianchi and P. Gronchi
Elements of .A∗ are called Lebesgue density points, or simply density points, of
A. Note that .A∗ = A, essentially, by the Lebesgue density theorem. Given .A ∈
∗
M(Rn ) and .C ∈ Knn containing o in its interior, let .MC (A) and .M∗C (A) denote,
respectively, its upper and lower anisotropic outer Minkowski content with respect
to C, i.e.,
∗ Hn (A + εC) − Hn (A)
MC (A) = lim sup ,
ε→0+ ε
. (5.4)
Hn (A + εC) − Hn (A)
M∗C (A) = lim inf .
ε→0+ ε
essentially the same concept is considered, where the .lim sup and the .lim inf are
denoted as the upper and lower (outer) relative surface area of A with respect to C.
Let .M(Rn ) (or .M+ (Rn )) denote the set of real-valued (or non-negative, respec-
tively) measurable functions on .Rn and let .S(Rn ) denote the set of functions f in
n
.M(R ) such that .H ({x : f (x) > t}) < ∞ for .t > ess inf f . By .V(R ), we denote
n n
n
the set of functions f in .M+ (R ) such that .H ({x : f (x) > t}) < ∞ for .t > 0.
n
Members of .S(Rn ) have been called symmetrizable and those of .V(Rn ) are often
said to vanish at infinity. If .f ∈ M(Rn ), we denote its graph by .Gf and define its
subgraph .Kf ⊂ Rn+1 by
(♦∗ K) + dB n ⊂ ♦∗ (K + dB n ) ⊂ ♦(K + dB n ),
. (5.6)
♦∗ A = (♦A)∗ .
.
(♦∗ K) + dD n ⊂ ♦∗ (K + dD n ).
. (5.7)
and that .♦ reduces the upper and lower outer Minkowski content (and the perimeter,
when K and .♦∗ K are set whose outer and lower Minkowski content coincide with
the perimeter).
Invariance under translations orthogonal to H of H -symmetric sets, as well as
being idempotent, are satisfied by most natural symmetrizations, so in discussing
examples we shall only mention these properties when they do not hold.
Two special cases are of particular importance: .i = 0 and .i = n−1. If .i = 0, then
.H = {o} and 0-symmetrization is the same as the o-symmetrization. One example
1 1
.♦K = AK = K + (−K). (5.9)
2 2
The central symmetral .AK differs from the ubiquitous difference body .DK =
K + (−K) only by a dilatation factor of .1/2. It is a particular instance of
Minkowski symmetrizations that we will define in a moment. Other examples of
0-symmetrizations are the pth central symmetrization, given for .K ∈ Kn and .p ≥ 1
by
( ) ( )
♦K = Ap K = 2−1/p K +p 2−1/p (−K)
.
(here .+p denotes the general .Lp addition introduced in Sect. 3.4.1) and the M-
symmetrization. For its definition and for more on 0-symmetrizations we refer the
reader to Gardner et al. [38] and to Bianchi, Gardner, and Gronchi [8, Section 4].
The other case of particular importance is .i = n − 1, to which we will devote
more attention.
is a (possibly degenerate) closed line segment with midpoint in H and .H1 -measure
equal to that of .G ∩ K. An extension of this definition to Lebesgue measurable
subsets of .Rn is possible and is presented below in the more general setting of
Schwarz symmetrization. We list some of its properties.
(a) If .K ∈ Cn , then .SH K ∈ Cn , and if .K ∈ Kn , then .SH K ∈ Kn . The first claim
is elementary and we prove the second one. Let .u ∈ Sn−1 be orthogonal to H .
There are two functions .f, g : K|H → R such that we can describe .K ∈ Kn as
K = {x + tu ∈ Rn : g(x) ≤ t ≤ f (x)},
. x ∈ K|H.
5 Symmetrizations 239
SH K + SH L ⊂ SH (K + L).
. (5.10)
(SH K + SH L) ∩ G ⊂ SH (K + L) ∩ G .
.
Since all the segments .(SH K ∩Gy )+(SH L∩G−y ) are contained in G and centered
in H , their union equals the largest of its elements, .(SH K ∩ Gȳ ) + (SH L ∩ G−ȳ ).
Its length equals the sum of the lengths of the two segments
Section 5.4 presents many other properties of the Steiner symmetrization. It also
contains a proof of (5.10) valid for convex bodies and which uses shadow systems.
.SH K, i.e.
(g) The existence of the approximating sequence presented in item (f) allows to
prove that properties valid for Steiner symmetrization and which are maintained
in the passage to the limit, are also valid for Schwarz symmetrization. For
instance, it can be used to prove that if .K ∈ Kn then .SH K ∈ Kn . Indeed, if
.(Hm ) is as in item (f), then, for .m ∈ N, .SHm . . . SH2 SH1 K ∈ K , since Steiner
n
/
where .ry = d 2 − |y − x|2 . Indeed, .p ∈ (L + dD n ) ∩ (x + H ⊥ ) if and only
if .p|H = x and there is a .z ∈ L such that .p ∈ z + dD n . If .z|H = y, then this
holds if and only if .|y − x| < d and
|p − ||x z| < ry ,
.
We use (5.12) with .L = SH A, (5.11), the fact that the action of .SH is the same
for each y, the pointwise monotonicity of .SH , and (5.12) with .L = A, to obtain
( ) ( )
. SH A + dD n ∩ x + H ⊥
|| ([ ] )
= ||x SH A ∩ (y + H ⊥ ) + ry D n−i
y∈H,|y−x|<d
|| ( [ ] )
= ||x SH A ∩ (y + H ⊥ ) + ry D n−i
y∈H,|y−x|<d
|| ( [ ])
⊂ ||x SH (A ∩ (y + H ⊥ )) + ry D n−i
y∈H,|y−x|<d
|| ( [ ])
= SH ||x (A ∩ (y + H ⊥ )) + ry D n−i
y∈H,|y−x|<d
⎛ ⎞
|| [ ]
⊂ SH ⎝ ||x (A ∩ (y + H ⊥ )) + ry D n−i ⎠
y∈H,|y−x|<d
( )
= SH (A + dD n ) ∩ (x + H ⊥ )
= SH (A + dD n ) ∩ (x + H ⊥ ).
SH A + dD n ⊂ SH (A + dD n ).
.
and let H be the .xn+1 -axis. The proof lies in the following three observations:
i) for .t ∈ [0, 1],
( )
M ∩ {xn+1 = t} = (1 − t)K + tL × {t};
.
iii) Since .Hn (SH M ∩ ({xn+1 = t}) = )Hn (M ∩ {xn+1 = t}), .κn f (t)n = Hn (M ∩
{xn+1 = t}) = Hn (1 − t)K + tL .
We can conclude that the convexity of .SH M is equivalent to the concavity of the
( )1/n
function .Hn (1 − t)K + tL on .[0, 1].
defined by
1 1
MH K =
. K + K †, (5.13)
2 2
|| ( 1 1
)
.FH K = (K ∩ (H ⊥ + x)) + (K ∩ (H ⊥ + x))†
2 2
x∈H
|| ( 1 1 †
)
= (K ∩ (H ⊥ + x)) + (K ∩ (H ⊥ + x)) . (5.14)
2 2
x∈H
( ) ( )
1 1
FH K =
. ◦ K nH ◦K .
†
2 2
Finally, for .H ∈ G(n, i), .i ∈ {0, . . . , n − 1}, we can define the Blaschke symmetral
of .K ∈ Knn by
( ) ( )
BH K = 2−1/(n−1) K # 2−1/(n−1) K † .
.
1 1
. Sn−1 (BH K, ·) = Sn−1 (K, ·) + Sn−1 (K † , ·). (5.15)
2 2
246 G. Bianchi and P. Gronchi
These formulas define .BH K up to translation. We define the Blaschke sum so that
the centroids of .BH K and .K|H coincide. When .i = 0, we have .K † = −K and then
the body .BH K is often called the Blaschke body of K and denoted by .∇K. We list
some of the properties.
(a) Blaschke symmetrization is invariant on H -symmetric sets.
(b) When .n = 2, then up to translation and on .Knn , .BH coincides with .A (.i = 0) or
.MH (.i = 1), whose properties have already been discussed.
Proof Let .T n be an n-dimensional cone in .Rn with centroid at the origin, .xn -axis
as its axis, and radius and height (i.e., width in the direction .en ) both equal to 1. Let
and .H = {o} when .i = 0, and let H be the subspace of .Rn spanned by .e1 , . . . , ei ,
when .i ≥ 1. For each i, we have .(T n )† = −T n and, therefore, .BH T n = ∇T n , the
Blaschke body of .T n . We claim that when .n ≥ 3, the height of .∇T n is less than 1.
Suppose the claim is true. Let .0 < s < 1 and let .Ls ⊂ T n be the spherical cylinder
with base of radius s contained in the base of .T n , the .xn -axis as its axis, and with
maximal height .w = w(s). The set .Ls is centrally symmetric, so .BH Ls = ∇Ls is
a translate of .Ls and the height of .∇Ls is w; since .w → 1 as .s → 0, when s is
sufficiently small it is not possible that .∇Ls ⊂ ∇T n .
To prove the claim, let .n ≥ 3 and recall that the surface area of the curved
part √ of the boundary of an n-dimensional cone of radius r and height h is
n−2 h2 + r 2 κ
.r
√ n−1 . Therefore the surface area of the curved part of the boundary
of .T n is . 2κn−1 , while the area of the base of .T n is .κn−1 . The surface area measure
.Sn−1 (T , ·) consists of a point mass at .−en and a multiple of .(n − 2)-dimensional
n
that
√ √ n−1 √
. 2a n−1 κn−1 − 2h κn−1 = 2κn−1 /2
and hence .a = 1. Thus the height of .∇T n is .2(a − h) = 2(1 − 2−1/(n−1) ), which is
less than 1 when .n ≥ 3. This proves the claim. u
n
5 Symmetrizations 247
5.3.7 Polarization
respect to H .
The rearrangement .PH associated to this set operation is again called polariza-
tion. The superlevel set .{x : PH f ≥ t} is the (set) polarization of .{x : f (x) ≥ t},
for each .t > ess inf f . The rearrangement can also be defined as
{
max{f (x), f (x † )} if x ∈ H + ,
PH f (x) =
.
min{f (x), f (x † )} if x ∈ H − .
248 G. Bianchi and P. Gronchi
and
( ) ( )†
.PH A = (A ∩ A† ) ∪ (AAA† ) ∩ H + ∪ (AAA† ) ∩ H − ,
where the unions on the right-hand sides of these formulas are disjoint.
(d) On .Kn , polarization is perimeter preserving. This is essentially due to the
following simple structure of the polarization of a convex set. Let .K ∈ Kn ,
⊥
.x ∈ K|H and let .px be the midpoint of the segment .K ∩ (x + H ). Then
{
⊥ K ∩ (x + H ⊥ ) if px ∈ H + ,
.PH K ∩ (x + H ) =
(K ∩ (x + H ⊥ ))† if px ∈ / H +.
Thus the boundary of K can be decomposed in two disjoint parts so that the
boundary of .PH K is the disjoint union of one of these parts and of the reflection
with respect to H of the other part. More precisely, if .E1 = {x ∈ K|H : px ∈
H + } and .E2 = (K|H ) \ E1 then
and
Let .z ∈ PH A.
5 Symmetrizations 249
z + dB n = PH (z + dB n ) ⊂ PH (A + dB n ).
.
z + dB n = (z† + dB n )† = PH (z† + dB n ) ⊂ PH (A + dB n ).
.
Assume now .z ∈ H − . Then both z and .z† are in A. The set .{z, z† }+dB n ⊂ A+dB n
and, being symmetric, it coincides with its polarization. Thus
( )
.z + dB n ⊂ {z, z† } + dB n = PH {z, z† } + dB n ⊂ PH (A + dB n ).
(f) Both Steiner and Schwarz symmetrization, for any i, can be approximated in .Cn
by sequences of polarizations in the Hausdorff metric. See Van Schaftingen [88]
for the explicit construction of one sequence with this property. See also
Burchard and Fortier [25] for the importance and the history of these results.
This property has many consequences. In the literature it has been used to prove
that certain properties and inequalities which are valid for the rearrangement
associated to polarization are also valid for the rearrangements associated to
Steiner and Schwarz symmetrization.
In Sect. 5.5 we describe a duality between fiber (Steiner, if .i = n−1) and Minkowski
symmetrization. For .i = 0, . . . , n − 1 the fiber symmetral of .K ∈ Kn is the union
of all H -symmetric compact convex sets such that some translate orthogonal to
H is contained in K, while the Minkowski symmetral of K is the intersection
of all H -symmetric compact convex sets such that some translate orthogonal to
H contains K. Bianchi, Gardner, and Gronchi [8, Section 5] introduces two new
symmetrizations, the inner and outer rotational symmetrizations, which display
exactly the same duality in a rotationally symmetric setting.
Let .H ∈ G(n, i), .i ∈ {1, . . . , n−1} and .K ∈ Kn . The inner rotational symmetral
.IH K of K is such that for each .(n − i)-dimensional plane G orthogonal to H and
(g) .OH generally increases .Vj for .j ∈ {2, . . . , n} and also for .j = 1 when .i ∈
{1, . . . , n − 2}.
Some of the applications of the symmetrizations stem from containment rela-
tions. The following theorem summarizes all the known inclusions between the
various known symmetrals. Some are well known and some are observed and proved
in [8] or [11]. It should be added that .FH K = MH K when .i = 0.
Theorem 5.8 If .H ∈ G(n, i), .i ∈ {1, . . . , n − 1}, and .K ∈ Kn , then
IH K ⊂ FH K ⊂ MH K ⊂ OH K
. (5.18)
and
.IH K ⊂ SH K ⊂ M H K ⊂ OH K. (5.19)
and studied by Rogers and Shephard in 1958 [75]. Given a subset A of .Rn , a
direction .v ∈ Rn and a real valued function .α defined on A, we set
In 1964 Shephard [82] observed that such a family of convex sets can be seen as
a shadow system, a family of projections of the same .(n + 1)-dimensional convex
body onto a fixed hyperplane along a varying direction. If we define
we may recognize that .Kt is the projection (not an orthogonal one!) of .K̃ onto
⊥ along the direction of .e
Rn = en+1
. n+1 − tv.
This idea of visualizing shadow systems as projections of a larger body immedi-
ately clarifies that two or more shadow systems moving in the same direction v may
generate other shadow systems by taking their projections onto a fixed subspace or
their Minkowski addition or their convex hulls. In formulas, if .{Kt }t∈R and .{Lt }t∈R
are shadow systems in the same direction v and H is a subspace, then
are also shadow systems. While it is now clear that many examples of shadow
systems can be produced, we have still to explain the convenience and advantages
of this new concept. And so we come to the main feature of shadow systems, stated
in the following theorem by Rogers and Shephard.
Theorem 5.9 If .{Kt }t∈R is a shadow system, then .Vn (Kt ) is a convex function of t.
Proof A definitely concise and elegant proof of the statement is based on the
formula
the inclusion
t1 + t2 1 1
[0, en+1 −
. v] ⊂ [0, en+1 − t1 v] + [0, en+1 − t2 v]
2 2 2
and the monotonicity and continuity of mixed volumes (see Proposition 1.41).
For completeness, we present a second proof where we infer the convexity of
1
.Vn (Kt ) from that of .H (Kt ∩G), for any line G parallel to the direction of movement
f (x) + g(x)
Kt = {x + sv ∈ Rn : x ∈ K|H, g(x) ≤ s + t
. ≤ f (x)} (5.20)
2
shows that the average brightness of a convex body L is, up to a constant depending
on n, its surface area. Therefore, the function .S(Kt ) is convex in t. Since it assumes
the same value at 0 and 2, we deduce that its value at 1 is not larger than that at
the extreme points, which means that the surface area of .SH K is not larger than the
surface area of K. Formally, this statement reads as follows.
Theorem 5.10 For every hyperplane H and convex body K,
S(SH K) ≤ S(K) .
.
This result can be generalized in various ways. From the above argument,
stressed for the first time in [75], we can deduce the following principle.
Whenever we have a functional F which is
• convex with respect to the parameter t of every shadow system, and
• invariant under reflections,
we can conclude that .F (SH K) ≤ F (K).
This can be used, together with results on the convergence to a ball of sequences
of successive Steiner symmetrizations (see Sect. 5.7), to prove the classical isoperi-
metric inequality for convex sets. Indeed, Rogers and Shephard’s complete argu-
ment gave the following principle.
5 Symmetrizations 253
Therefore, all intrinsic volumes of a shadow system .{Kt }t∈I are convex functions
of t. Since they are also continuous and reflection invariant, we can conclude that
Steiner symmetrization does not increase any intrinsic volume and hence arrive to
the corresponding isoperimetric inequalities.
Theorem 5.12 (Isoperimetric Inequalities) For .K ∈ Knn and .i = 1, 2, . . . , n − 1
Vi (K) Vi (B n )
.
i
≥ i
.
Vn (K) n Vn (B n ) n
The discussion of the equality case is usually more involved (and sometimes it
is still an open problem). It can be proved that in Theorems 5.10 and 5.12 equality
holds if and only if K is a ball. In both cases the characterization of balls can be
proved showing that the intrinsic volume .Vi (Kt ) is, for .i < n, a strictly convex
function of t. A possible proof is postponed until after the next theorem.
The convexity of the intrinsic volumes of a shadow system with respect to its
parameter t is a special case of a more general result showed by Shephard in [82].
Theorem 5.13 If .{Kt1 }t∈I , .{Kt2 }t∈I , .. . . , .{Ktn }t∈I are shadow systems along the
same direction v, then the mixed volume .V (Kt1 , Kt2 , . . . , Ktn ) is a convex function
of t.
254 G. Bianchi and P. Gronchi
Proof A proof of this result could follow the path already made for the volume of a
shadow system, using the formula
⊥
where .K̃ j denotes the .(n + 1)-dimensional body whose shadows on .en+1 are in
j
{Kt }t∈I .
. u
n
Strict convexity of the .Vi (Kt ), .i < dim(K). In order to prove the strict convexity,
following the previous proof of the convexity of mixed volumes, it is enough to
show that
where we abbreviate .w1 = en+1 + t1 v and .w2 = en+1 + t2 v. By Theorem 3.19 and
Definition 3.20, this inequality can be written in terms of an integral with respect to
the ith area measure of .K̃:
f
1 ( )
. h[0,w1 ] (u) + h[0,w2 ] (u) − h[0,w1 +w2 ] (u) dS(K̃[i], B n [n − i], u) > 0 .
n+1 S n
SH K + SH L ⊂ SH (K + L) .
.
Here we provide another proof of this inclusion, valid if K and L are convex bodies,
which uses shadow systems.
Proof Let us consider the shadow systems .{Kt }t∈[0,2] and .{Lt }t∈[0,2] which connect
K and L to their reflections on H passing through their Steiner symmetral,
respectively. As we observed, .{Kt + Lt }t∈[0,2] is a shadow system and a proof of
Theorem 5.9 used the fact that .H1 ((Kt + Lt ) ∩ G) is a convex function of t for
every line G orthogonal to H . Since for .t = 0 and .t = 2 it attains the same value,
we deduce that
H1 ((SH K + SH L) ∩ G) ≤ H1 ((K + L) ∩ G) ,
.
SH (SH K + SH L) ⊂ SH (K + L) ,
.
Mm (K) Mm (B n )
. ≥ .
Vn (K) Vn (B n )
Proof Consider the shadow system .{Kt }t∈[0,2] connecting K to its reflection on H
and a polytope .P ⊂ K0 = K with at most m vertices. If we let the vertices of P
move solidly to the chord on which they are located, then each vertex has a constant
speed and .{Pt }t∈[0,2] (the convex hulls of these vertices at time t) form a shadow
system. Hence, the .Vn (Pt ) is a convex function of t and so is the maximum over
all polytopes with at most m vertices, that is .Mm (Kt ). Since .Vn (Kt ) is constant
and .Mm (K) is continuous with respect to the Hausdorff distance, the argument by
Rogers and Shephard concludes the proof. u
n
In Theorem 5.15 equality holds for ellipsoids but a complete characterization is
still missing. In 1939 Sas [78] proved that in the plane, for every m, equality holds
if and only if K is an ellipse. In 1917 Blaschke [16] characterized 3-dimensional
ellipsoids as the only minimizers for .m = 4. In 1986 Bianchi [5] proved the
corresponding result for .m = 5. For .n > 3 or .n = 3 and .m > 5 the problem is
still open.
If we denote by .Mm i (K) the maximal i-th intrinsic volume of a polytope
where the power on the volume of K is chosen so that the functional is scaling
invariant and convex along shadow systems. However, such a functional is not affine
invariant and has no upper bound.
256 G. Bianchi and P. Gronchi
i (K)
Mm i (B n )
Mm
.
i
≥ i
.
Vn (K) n Vn (B n ) n
which is the expected volume of a random polytope from K divided by .Vn (K)
(to ensure scaling invariance), where the vertices are selected uniformly and
independently in K.
If we consider the shadow system .{Kt }t∈[0,2] defined in (5.20), then the volume
of .Kt is constant and the functional can be rewritten as
f f ||
m
1
Sn (Kt ; m) =
. . . . Vn (conv[ xj + tα(xj |H )v]) dx1 . . . dxm .
Vn (K)m+1 K K j =1
.Sn (K; m) ≥ Sn (B n ; m) ,
is called Sylvester’s problem and it is solved only in the plane. For .p = 1, Dalla and
Larman [34] proved that triangles are maximizers and Giannopoulos [39] proved
they are the only ones. Campi et al. [27] showed that parallelograms are maximizers
among centrally symmetric figures and Saroglou [77] proved the uniqueness of such
maximizers.
In higher dimensions the main result is due to Bárány and Buchta [3], who proved
that for every .K ∈ Kn there exists .m̄, depending on K, such that .Sn (K; m) ≤
Sn (T ; m), for all .m ≥ m̄, where T is a simplex.
258 G. Bianchi and P. Gronchi
They further proved that the ball is the only minimizer if f is convex and strictly
increasing.
where, for .m ≥ n,
f f
1
.Bn (K; m, p) = . . . Vn (conv[o, x1 , . . . , xm ])p dx1 . . . dxm .
Vn (K)m+p K K
The functional .Bn (K; m, p) is called Busemann’s functional and differs from
Sylvester’s in having fixed a point in the origin.
It is no more translation invariant, but it clearly remains .GL(n) invariant,
continuous with respect to Hausdorff distance, and convex along shadow systems.
In [26] Busemann proved that .Bn (K; n, 1) attains its minimum if and only if K is
an origin symmetric ellipsoid. Such a result is known as Busemann random simplex
inequality.
Theorem 5.18 (Busemann Random Simplex Inequality) For every .m ≥ n, .p ≥
1 and .K ∈ Knn
Bn (K; m, p) ≥ Bn (B n ; m, p) ,
.
This fact suggests a different extension that we shall meet in next paragraph.
Bourgain et al. [21], in connection with some comparisons between norms in the
local theory of Banach spaces, considered the functional
f f ( m )p
1 E
.I (K; m, p) = . . . Vn [0, xi ] dx1 . . . dxm ,
Vn (K)m+p K K i=1
I (K1 , K2 , . . . , Km ; p)
.
f f (Em
)p
1
= m+p . . . Vn [0, xi ] dx1 . . . dxm ,
(Vn (K1 ) . . . Vn (Km )) m
i=1
K1 Km
I (K; m, p) ≥ I (B n ; m, p) ,
.
In fact, Bourgain et al. [21] proved the inequality for all .p ≥ 0 and also for the
functional involving more bodies:
I (K1 , . . . , Km ; p) ≥ I (B1 , . . . , Bm ; p) ,
.
where .Bi is the ball with the same volume as .Ki centered at the origin.
The Minkowski sum of a finite number of segments is called a zonotope. The
simplest zonotope is a parallelotope, the sum of n affinely independent segments,
that is an affine image of the n-dimensional cube. By increasing the number of
segments, zonotopes can approximate the unit ball. A set which is the limit, in the
Hausdorff metric, of a sequence of zonotopes is called a zonoid. Zonoids play a
basic role in the Brunn–Minkowski theory of convex bodies and appear in different
contexts of the mathematical literature. We refer to [80] for an exhaustive review on
this topic.
f
2π κn
hrK (x) =
. |<x, z>| dz ,
κn+1 Vn (K) K
where the constant in front of the integral is such that .r(λK) = λrK, for all .λ > 0,
and .rB n = B n .
This body (usually with a different normalization) is known in the literature as
the centroid body of K. Centroid bodies were first defined and investigated by Petty
[69], but the concept had previously appeared in work of Dupin, in connection with
problems for floating bodies (see Gardner [37, Chap. 9] and Schneider [79, Sect.
7.4] for references). When K is an origin symmetric body, the boundary of .rK is,
up to a dilatation, the locus of the centroids of all the halves of K obtained by cutting
K with hyperplanes through the origin.
One of the basic results obtained by Petty [69] is an integral representation of
the volume of .rK by means of Busemann’s functional .Bn (K; n, 1). Using the
Busemann random simplex inequality, Petty proved the well known Busemann–
Petty centroid inequality: For .K ∈ Knn ,
Vn (rK) ≥ Vn (K) ,
. (5.23)
Petty [70] proved that the Busemann–Petty centroid inequality implies the Petty
projection inequality:
where equality holds if and only if K is an ellipsoid. Here, .||◦ K is the polar
projection body of K, i.e., the polar body of the projection body .||K of K, that
can be defined by
A short way to show that (5.23) implies (5.24) was obtained by Lutwak in [54].
Zhang [91] proved a reverse form of (5.24), known as Zhang projection
inequality:
( )
1 2n
.Vn (K)
n−1
Vn (||◦ K) ≥ n ,
n n
with equality if and only if K is a simplex.
The .Lp extension of the classical Brunn–Minkowski theory for convex bodies
was initiated by Lutwak [55], in which the idea of Firey [35] of the p-Minkowski
addition for sets is widely developed.
As already seen in Sect. 3.4, if .p ≥ 1 and K, L are convex bodies containing the
origin in their interior, the p-sum of K and L is the convex body .K +p L defined
by
Bianchini and Colesanti [14] observed that the p-sum of shadow systems is again
a shadow system, since the projection of a p-sum is the p-sum of the projections.
Notice that the p-sum is the Minkowski sum for .p = 1 and tends to the convex hull
as p tends to infinity. Taking into account the p-sum of segments we can define
f
κ2 κp−1 κn
.hrp K (x) = |<x, z>|p dz ,
p
(5.25)
κn+p Vn (K) K
where the constant is so that .rp (λK) = λrp K, for all .λ > 0, and .rp B n = B n . The
body .rp K is known as the .Lp centroid body of K.
For .p = 1, .r1 K is the centroid body of K, while for .p = 2, the body defined
by (5.25) is also well known. Indeed, up to a constant, .r2 K is the ellipsoid of inertia
(or Legendre ellipsoid) of K, i.e., the ellipsoid having the same moments of inertia
as K about every axis. Many results concerning this body, which is fundamental in
classical mechanics, can be found in the literature (see, e.g., Milman and Pajor [68]
and Lindenstrauss and Milman [50] for references). In 1918 Blaschke [18] proved
that, for .n = 3,
Vn (r2 K) ≥ Vn (K) ,
. (5.26)
262 G. Bianchi and P. Gronchi
where equality holds if and only if K is an origin symmetric ellipsoid. In 1937 this
result was extended by John [46] in all dimensions; other proofs were given by Petty
[69] and by Lutwak, Yang, and Zhang [56].
Inequalities (5.23) and (5.26) are special instances of the more recent
.Lp Busemann–Petty centroid inequality.
Vn (rp K) ≥ Vn (K) ,
.
K ◦ = {x ∈ Rn |<x, y> ≤ 1, ∀y ∈ K} .
.
Notice that the polar body of K strongly depends on the location of the origin. If K
is an origin-symmetric convex body, then the product
Vn (K)Vn (K ◦ )
.
attained is called the Santaló point of K and is characterized by the fact that .(K −z)◦
has its centroid at the origin if and only if .z = s(K).
A sharp upper bound for the volume product of a convex body .K ∈ Kn with
centroid at the origin is given by the Blaschke–Santaló inequality.
Theorem 5.21 (Blaschke–Santaló Inequality) For .K ∈ Knn ,
Vn (K)Vn (K ◦ ) ≤ κn2 ,
.
This was proved by Campi and Gronchi [31] for centrally symmetric bodies
and by Meyer and Reisner [66] in full generality. We sketch here a proof in the
symmetric case.
One of the main ingredients in the proof of Theorem 5.22 is a consequence of the
Borell–Brascamp–Lieb inequality, which deals with p-means of functions and their
integrals. It can be interpreted as an inverse Hölder inequality, and its links with
other well-known inequalities are widely described in the survey article by Gardner
[36].
Theorem 5.23 (Borell–Brascamp–Lieb Inequality) If .0 < λ < 1, .−1/n ≤ p ≤
∞, and f , g, h are non-negative integrable functions on .Rn satisfying
f [ (f ) p (f ) p ] np+1
p
np+1 np+1
. h(x) dx ≥ (1 − λ) f (x) dx +λ g(x) dx .
Rn Rn Rn
Note that if .p < 0, then f is p-concave if and only if .f p is convex. The above
definition can be extended to the case .p = 0 by continuity.
Corollary 5.25 Let .F (x, y) be a non-negative p-concave function on .Rn × Rm ,
.p ≥ −1/n. If, for every y in .R , the integral
m
f
. F (x, y) dx
Rn
p
exists, then it is a . np+1 -concave function of y.
Proof Take .y0 , .y1 ∈ Rm and fix .λ ∈ (0, 1). Let .yλ = (1 − λ)y0 + λy1 , and
Kt = conv{x + α(x)t v : x ∈ K0 }
.
can be thought of as the projection along the direction .en+1 − tv of .K̃ onto .en+1⊥ .
The support functions .hKt , .t ∈ [0, 1], and that of .K̃ are clearly related. Precisely,
⊥ , we have
for .u ∈ en+1
We know that
f
1
Vn (Kt◦ ) =
. h−n (z) dz .
n Sn−1 Kt
Let .D n−1 = {x /∈ v ⊥ : |x| ≤ 1}; thus .Sn−1 + = {z ∈ Sn−1 : <z, v> ≥ 0} is the graph
of the function . 1 − |x|2 , .x ∈ D n−1 . Consequently,
f f /
h−n (x + 1 − |x|2 v)
h−n
Kt
.
Kt (z) dz = 2 / dx ,
Sn−1 D n−1 1 − |x|2
5 Symmetrizations 265
f
2
Vn (Kt◦ ) =
. h−n (y + v + ten+1 ) dy .
n Rn−1 K̃
Since the function .hK̃ is convex in .Rn+1 , by Corollary 5.25, we infer that .Vn (Kt◦ ) is
p-concave, with respect to t, with .p = (−1/n)/(1 − (n − 1)/n) = −1. u
n
In [58] Lutwak and Zhang dealt with the functional
where .rp◦ K is the polar of the .Lp centroid body of K. Using Steiner symmetrization
they proved the so-called .Lp Blaschke–Santaló inequality.
Theorem 5.26 (.Lp Blaschke–Santaló Inequality) For all .p ≥ 1 and .K ∈ Knn
(n + 1)n+1
Vn (K)Vn (K ◦ ) ≥
. . (5.29)
(n!)2
In 1939 Mahler [61] proved the conjecture in the plane and in 1991 Meyer [64]
showed that equality holds only for triangles.
266 G. Bianchi and P. Gronchi
4d
Vn (K)Vn (K ◦ ) ≥
. (5.30)
d!
is a conjecture as well, where the value on the right-hand side is the volume product
of a parallelotope. It was proved in the plane by Mahler [61] and Reisner [74]
characterized parallelograms as the only minimizers. Saint Raymond [76] showed
that in higher dimension there are convex bodies, other than parallelotopes and
their polars, giving equality in (5.30). He also proved that the conjecture holds
true in all dimensions for the affine images of convex sets symmetric with respect
to the coordinate hyperplanes (called unconditional bodies). Barthe and Fradelizi
[4] generalized to all bodies whose hyperplanes of symmetries have a one-point
intersection. Inequality (5.30) was proved by Reisner [73, 74] for all zonoids.
Different proofs were presented by Gordon et al. [40] and by Campi and Gronchi
[32] using shadow systems.
Bourgain and Milman [22] proved that there exists a constant c, not depending
on the dimension, such that
Vn (K)Vn (K ◦ ) ≥ cn κn2 .
.
In [31] Campi and Gronchi dealt with the lower bound of (5.28). It is easy
to check that .Gp is continuous, .(−1)-concave along shadow systems and .GL(n)
invariant. Besides, .Gp (K) tends to zero as K moves away from the origin. If .cK
denotes the centroid of K, Campi and Gronchi proved that in the two-dimensional
case the functionals
. min Gp (K − x) , maxn Gp (K − x) , Gp (K − cK )
x∈K x∈R
The intrinsic volumes (or their close relatives, the quermassintegrals) of a convex
body .K ∈ Knn are not invariant under volume preserving affine transformations. An
affine invariant version was defined by Lutwak in [52] by replacing the .L1 norm in
Kubota’s formula (5.22) by the .L−n norm:
(f )− 1
κn n
.0i (K) = Vi−n (K|E) dE .
κi G(n,i)
0i (BK ) ≤ 0i (K) ,
. (5.31)
0i (SH K) ≤ 0i (K) ,
.
but they need the definition of a new body, the Projection Rolodex of K, (a subset of
a vector bundle over a lower-dimensional Grassmannian) and appropriate measures
on Grassmannians. Last but not least, they solved the equality cases with a ten pages
proof. In short: The arguments they used are surely too involved to fit into these
notes.
The following theorem shows a duality between Steiner (or, more generally, fiber)
and Minkowski symmetrization. We recall that .FH K = AK = MH K if .i = 0 and
that, if .i = n − 1, Steiner and fiber coincide. It is proved in [8] and the proof is taken
from there.
Theorem 5.28 Let .H ∈ G(n, i), .i ∈ {0, . . . , n − 1}, and for .K ∈ Kn and .y ∈ H ⊥ ,
let
Ky = K + y
. and Ky† = (Ky )† = K † − y. (5.32)
and
n
MH K =
. conv(Ky ∪ Ky† ). (5.34)
y∈H ⊥
so .FH K is contained in the right-hand side of (5.33). For the reverse inclusion, note
first that .Ky ∩ Ky† is H -symmetric. From the invariance of .FH on H -symmetric sets
and the fact that .FH is monotonic and invariant on translations orthogonal to H of
H -symmetric sets, we obtain
1 1
M H K = M H Ky =
. Ky + Ky† ⊂ Qy ,
2 2
so .MH K ⊂ ∩y∈H ⊥ Qy . To prove the reverse containment in (5.34), observe that if
v ∈ Sn−1 , then by (5.32) and .hK±y (v) = hK (v) ± <y, v>, we obtain
.
{ }
= min max hK (v) + <y, v>, hK † (v) − <y, v>
y∈H ⊥
1 1
= hK (v) + hK † (v) = hMH K (v), (5.35)
2 2
as required, where the first equality in (5.35) results from observing that the
minimum occurs when the two expressions are equal, i.e., when .<y, v> = (hK † (v) −
hK (v))/2. u
n
Corollary 5.29 ([8]) If .H ∈ G(n, i), .i ∈ {0, . . . , n − 1}, and .K ∈ Kn , then the
fiber symmetral .FH K (and therefore the Steiner symmetral .SH K, if .i = n−1) is the
union of all H -symmetric compact convex sets such that some translate orthogonal
to H is contained in K, and the Minkowski symmetral .MH K is the intersection
of all H -symmetric compact convex sets such that some translate orthogonal to H
contains K.
Proof Let us prove the claim regarding .FH K. For each .y ∈ H ⊥ , the set .Ky ∩ Ky† is
H -symmetric and its translation by .−y is contained in K. On the other hand, if .M ∈
5 Symmetrizations 269
The containment results presented here have a crucial role in proving many of the
results for general i-symmetrizations described in the next sections. All of them
are proved in [8], which also contains a critical discussion of the necessity of each
hypothesis. The first one is a corollary of Theorem 5.28.
Corollary 5.30 Let .H ∈ G(n, i), .i ∈ {0, . . . , n − 1}, and let .B = Kn or .B = Knn .
Suppose that .♦ : B → BH is monotonic, invariant on H -symmetric sets, and
invariant under translations orthogonal to H of H -symmetric sets. Then
FH K ⊂ ♦K ⊂ MH K
. (5.36)
for all .K ∈ B.
Proof Let .K ∈ B and let .y ∈ H ⊥ . The set .Ky ∩ Ky† is H -symmetric. Hence, using
the monotonicity and invariance property of .♦, we have
This formula and (5.33) prove the inclusion on the left. The set .conv(Ky ∪ Ky† ) is
H -symmetric and .K ⊂ conv(Ky ∪ Ky† ) − y. Again, using the monotonicity and
invariance property of .♦, we have
( )
♦K ⊂ ♦ conv(Ky ∪ Ky† ) − y = ♦ conv(Ky ∪ Ky† ) = conv(Ky ∪ Ky† ).
. (5.38)
. SH K ⊂ ♦K ⊂ MH K (5.39)
for all .K ∈ B.
270 G. Bianchi and P. Gronchi
IH K ⊂ ♦K ⊂ OH K.
. (5.40)
The left-hand inclusion holds for all .K ∈ B if, in addition to the assumptions stated
before (5.40), .B = Knn and .♦ is monotonic. The right-hand inclusion holds for all
.K ∈ B if, in addition to the assumptions stated before (5.40), .♦ is strictly monotonic
and idempotent.
In this section we present some characterizations proved in [8] and in [9]. We refer
to these papers for a critical discussion of the necessity of each hypothesis.
Theorem 5.33 Let H ∈ G(n, i), i ∈ {0, . . . , n − 1}, and let B = Kn or B = Knn .
Suppose that ♦ : B → BH is monotonic. Assume in addition either that
(i) i = n−1 and ♦ is mean width preserving and either invariant on H -symmetric
spherical cylinders or projection invariant, or that
(ii) i ∈ {1, . . . , n − 1} and ♦ is mean width preserving, invariant on H -symmetric
sets, and invariant under translations orthogonal to H of H -symmetric sets,
or that
(iii) i = 0 and ♦ is invariant on o-symmetric sets and invariant under translations
of o-symmetric sets.
Then ♦ is Minkowski symmetrization with respect to H .
Proof Let K ∈ B. Let us prove part (i). Theorem 5.31 with F = V1 proves
♦K ⊂ MH K.
.
Theorem 5.35
(i) Let .H ∈ G(n, n − 1). Suppose that .♦ : Cn → CnH is an .(n − 1)-symmetrization
.
[−s, s] (we identify .H with .R so that .[−s, s] is a shorthand for .s(B ∩ H ⊥ )).
⊥ n
The set .Dr (x) + [−s, s] is an H -symmetric spherical cylinder which contains .K ∩
(Dr (x) + H ⊥ ). For .L ∈ Cn let
( )
mr,L (x) = Hn L ∩ (Dr (x) + [−s, s]) .
.
272 G. Bianchi and P. Gronchi
mr,♦K ≥ mr,K .
. (5.42)
From this inclusion, using the fact that .♦ is volume preserving, we obtain
( ) ( )
mr,♦K (x) = Hn (♦K) ∩ (Dr (x) + [−s, s]) ≥ Hn ♦(K ∩ (Dr (x) + [−s, s]))
.
( )
= Hn K ∩ (Dr (x) + [−s, s]) = mr,K (x).
This proves (5.42). Dividing both sides in (5.42) by .Hn−1 (Dr (x)) and passing to
the limit as .r → 0 we obtain (5.41) with the inequality (left-hand side .≥ right-
hand side), for .Hn−1 -almost all .x ∈ H . Integrating this inequality over H , using
Fubini’s theorem and the volume invariance of .♦, we conclude that (5.41) holds
with equality, for .Hn−1 -almost all .x ∈ H .
The proof of part Theorem 5.35 (ii) follows directly from (5.41) and the definition
of .SH K.
Theorem 5.36, under the assumption that .♦ is invariant on H -symmetric
spherical cylinders, is a corollary of Theorem 5.35, while under the assumption
that .♦ is projection invariant it requires other ideas, which we do not describe here.
We conclude this section with a characterization of Schwarz symmetrization
from [11].
Theorem 5.37 Let .i ∈ {1, . . . , n − 2}, let .H ∈ G(n, i), and let .♦H be an
i-symmetrization on .Knn . Suppose that .♦H is monotonic, volume preserving, rota-
tionally symmetric, and invariant on H -symmetric cylinders. Then .♦H is Schwarz
symmetrization with respect to H .
We consider the four maps .Id, .†, .PH , or .PH† = † ◦ PH , where .Id and .† denote the
identity map and reflection in H , respectively. In this section we present two results
from [9] which characterize these four maps among maps .♦ : E ⊂ Ln → Ln .
Theorem 5.38 Let .H = u⊥ , .u ∈ Sn−1 , be oriented with .u ∈ H + , let .E = Cn or
.L , and suppose that .♦ : E → L is monotonic, measure preserving, perimeter
n n
We say that .♦ is perimeter preserving on convex bodies if, for each .K ∈ Knn , .♦K
is a set of finite perimeter such that .S(♦K) = S(K) = 2Vn−1 (K), where S denotes
perimeter in the sense of De Giorgi. The condition of invariance on H -symmetric
union of two disjoint balls may seem peculiar, but it is much weaker than the natural
assumption that .♦ is invariant on all H -symmetric sets.
For maps .♦ : Kn → Ln , invariance on H -symmetric unions of two disjoint balls
is not available. The next result resorts to a different and rather strong condition; we
say that .♦ is convexity preserving away from H if .♦K is essentially convex (that is,
n
.♦K coincides with a convex set up to a set of .H -measure zero) for all .K ∈ K
n
with .K ∩ H = ∅.
Theorem 5.39 Let .H = u⊥ , .u ∈ Sn−1 , be oriented with .u ∈ H + and let
.♦ : Kn → L
n n
be monotonic, measure preserving, invariant on H -symmetric
sets, perimeter preserving on convex bodies, and convexity preserving away from
H . Then .♦ essentially equals .Id, .†, .♦PH , or .♦†PH .
Around 1836, Jacob Steiner introduced the symmetrization process that today bears
his name in an attempt to prove the isoperimetric inequality. As we have explained,
many other inequalities have since been proved by the same method. Two are the
main features that have allowed to prove that a certain functional is minimized (or
maximized) by a ball:
1. The functional at hand is always decreased (or increased) by a symmetrization;
2. There are sequences of hyperplanes such that the corresponding sequence of
successive symmetrals of a convex body converges to a ball of the same volume.
In this section we focus on item 2. Given an i-symmetrization .♦ we focus
on the convergence of successive applications of .♦ through a sequence of i-
dimensional subspaces. (In this section convergence always means convergence
in the Hausdorff metric.) We refer to this as a symmetrization process. We try to
summarize what is known up to date, to explain how the answer depends on the
specific symmetrization, whether it depends on the class, .Kn or .Cn , of the initial
seed, to describe which questions still remain unanswered.
It is well known that in the plane a sequence of successive Steiner symmetrals
may converge to a ball or to a regular polygon, and convergence may sometimes
seem simple or even obvious. In fact, in the literature there are few examples of
Steiner symmetrization processes which do not converge (see [13, Section 1], [25,
Lemma 6.3], [7, Examples 2.1], [86, Section 4]). These examples exhibit the same
behavior and here we describe one in dimension two. We remark that in this example
the directions orthogonal to the sequence of the lines of symmetrization form a
dense subset of .S1 .
274 G. Bianchi and P. Gronchi
E
Let .βk = km=1 αm and .vm = (cos βm , sin βm ). Besides, let .K ∈ K22 have area
smaller than .π r 2 and contain a horizontal segment L, with L of length 1 and o as
midpoint. The sequence .(Km ) defined as
Km = Svm⊥ . . . Sv ⊥ K , m ∈ N,
. (5.44)
1
does not converge. Indeed, let .Lm = Svm⊥ . . . Sv ⊥ L. Since .βm diverges, the segments
1
.Lm (which are contained in .Km ) spin in circles forever while their length decreases
the process as a cyclical repetition of the three symmetrizations .SU1 , .SU2 and .SU3 ,
i.e. as .SU1 , .SU2 , .SU3 , .SU1 , .SU2 , .SU3 , . . . . The universality of the sequence was the
right ingredient that Blaschke needed to present Schwarz symmetrization as a limit
of sequences of Steiner symmetrizations, as explained in Sect. 5.3.2, and to prove
the Brunn–Minkowski inequality.
The argument used by Blaschke was extended by Klain [47, Theorem 5.1] to
prove that every sequence of hyperplanes chosen from a finite set .F is Steiner-
stable in .Kn and the limit is symmetric under reflection in each hyperplane occurring
infinitely often in the sequence. Klain [47, Corollary 5.4] uses this result to construct
Steiner-universal sequences in .Kn , as described in the next theorem. We say that
.v1 , . . . , vn ∈ R form an irrational basis of .R if they span .R and the angle between
n n n
Vn (Km ) = Vn (K).
. (5.46)
276 G. Bianchi and P. Gronchi
The main idea is to construct a subsequence along which the subspaces .vj⊥ ∈ F
appear in a particular order. With each index m, we associate a permutation .πm
of .{1, . . . , n} that indicates the order in which the subspaces .v1⊥ , . . . , vn⊥ appear
for the first time among those .Hj with .j ≥ m. Since there are only finitely many
permutations, we can pick a subsequence .(Hmp ) such that the permutation .πmp
is the same for each p. By relabeling the subspaces, we may assume that this
permutation is the identity. Passing to a further subsequence, we may assume
that every subspace in .F appears in each segment .Hmp , Hmp +1 , . . . , Hmp+1 −1 .
By Blaschke selection theorem, there is a subsequence (again denoted by
.(Kmp )) that converges in the Hausdorff metric to some .L ∈ Kn .
n
We also have
Therefore, .Sv ⊥ L = L, i.e. L is .vj⊥ -symmetric and this concludes the inductive
j
step.
Once that the symmetry of L with respect to each .vj⊥ is proved, the fact that
Steiner symmetrization does not increase the symmetric difference distance can
be used again to prove that the entire sequence .(Km ) converges to L.
4) Blaschke and Klain used roughly the same hypothesis on the hyperplanes
mutual position. Burchard, Chambers, and Dranovski [24] tackled the problem
of characterizing the sets of reflections with respect to hyperplanes in .Rn that
generate a dense subgroup of .O(n). They proved that the set of reflections in .vi⊥ ,
.i = 1, 2, . . . , n, generate a dense subgroup of .O(n) if the .vi ’s
i) span .Rn ,
ii) cannot be partitioned into two mutually orthogonal non-empty subsets, and
iii) at least two of them form an angle that is an irrational multiple of .π .
5 Symmetrizations 277
Conditions i) and ii) are also necessary, but, when .n > 2, iii) is not, and [24]
explains that iii) implies the right necessary and sufficient condition. i.e. that the
group generated by the reflections is not a finite Coxeter group in .O(n).
u
n
The extension of Klain’s result to compact sets was first obtained in [7]. Now it can
also be seen as a consequence of Bianchi et al. [11, Theorem 7.3] which proves
that a sequence of hyperplanes is Steiner-universal in .Cn if and only if it is Steiner-
universal in .Knn , Volčič [90] extends Klain’s result to measurable sets.
To complete the picture regarding Steiner symmetrization, we recall some results
on the rate of convergence to a ball and on random sequences.
In 1986 Mani-Levitska [62] was the first to deal with a sequence of hyperplanes
uniformly, independently and randomly chosen. He proved that the sequence of
related Steiner symmetrizations almost surely rounds every convex body with pos-
itive volume and conjectured that the same holds for compact sets. The conjecture
was settled by van Schaftingen [87] and extended to measurable sets by Volčič [89].
The first result we know of on the rate of convergence (that is, on the deter-
mination of the least number of successive symmetrals required to transform a set
K of volume .κn within a certain distance from .B n ) goes back to Hadwiger [44],
even though the estimate was very rough. Such results often require very delicate
analysis, as evidenced by the deep work of Bourgain, Klartag, Lindenstrauss,
Milman, and others. (See [48, 49], and the references given there.) Klartag [49]
proves that there exist
Vn (MHk . . . MH1 K) → V ,
. (5.47)
for a suitable .V > 0. By the Blaschke selection Theorem, there exists a subsequence
(λk ) such that, as .k → ∞,
.
MHλk . . . MH1 K → E,
.
for a suitable .E ∈ Knn with .Vn (E) = V . For any positive integer .k, m, .m > k, the
inclusion between Steiner and Minkowski symmetrization (5.18) implies
= MHλm . . . MH1 K.
Passing to the limit with respect to m, and using the Steiner universality of the
sequence .(Hk ), we obtain that a ball with volume .Vn (MHλk . . . MH1 K) is contained
in E. The arbitrariness of k and (5.47) imply that E contains a ball of volume V .
Since .V = Vn (E), a comparison of the volume forces E to be a ball of volume V .
Therefore, any convergent subsequence has the same limit, that is .MHk . . . MH1 K
converges to a ball. The same argument can be repeated for the sequence
MHk . . . MHl K,
.
for any .l ∈ N. Observe that the limiting ball has the same mean width as K,
since Minkowski symmetrization is mean width preserving, and therefore it does
not depend on l.
The reverse is completely analogous. Assume that .(Hk ) is a Minkowski-universal
sequence of hyperplanes. Since Steiner symmetrization decreases the mean width,
the sequence .V1 (SHk . . . SH1 K) is nonincreasing and positive and, as .k → ∞,
V1 (SHk . . . SH1 K) → W
.
of mean widths forces E to be a ball with mean width W . Therefore, any convergent
subsequence has the same limit, that is, .SHk . . . SH1 K converges to a ball with the
same volume of K, since Steiner symmetrization is volume preserving. u
n
Bianchi et al. [11] studies some of the questions touched on in this section
for other known symmetrizations and for general i-symmetrizations. It proves
that Klain’s Theorem is valid in .Kn for fiber, Schwarz, and Minkowski–Blaschke
symmetrizations, as well as for any i-symmetrization satisfying certain hypotheses,
and it is valid in .Cn for Schwarz symmetrization (see [11, Section 5]). It also proves
results in the spirit of Coupier and Davydov; for .i = 1, . . . , n − 1, a sequence
of subspaces in .G(n, i) is Minkowski-universal in .Knn if and only if it is so in .Cn ,
and the same holds true for Steiner-universal and for Schwarz-universal sequences
(see [11, Section 7]). These results, together with a study of the problem of which
reflections with respect to i-dimensional subspaces generate a dense subgroup of
.O(n), carried out in [10], enable the authors to create universal sequences in .K and
n
in .C for many of the symmetrizations mentioned above (see [11, Section 6]).
n
Ulivelli [85] further extends Klain’s Theorem by proving its validity in .Cn
for Minkowski symmetrization (as a consequence of the proof that a Minkowski
symmetrizations process with seed .C ∈ Cn converges if and only if the one
with seed .conv(C) converges). The same author proves in [86] that the family
.F of all i-symmetrizations .♦ which are monotonic, invariant on H -symmetric
f f
. 0(|∇f #
(x)|) dx ≤ 0(|∇f (x)|) dx; (5.48)
Rn Rn
see, e.g., [2]. Here .f # denotes the symmetric decreasing rearrangement of f , the
function whose superlevel sets have the same .Hn -measure as those of f and such
that, for .t > 0, .{x : f # (x) > t} is a ball centered at the origin o of .Rn . The subgraph
of .f # is the Schwarz symmetrization of the subgraph of f with respect to the .xn+1 -
axis.
280 G. Bianchi and P. Gronchi
where .♦T denotes a map from .Ln to itself which is monotonic and measure
preserving. The superlevel set .{x : Tf (x) > t} depends only on .{x : f (x) > t}
and this relation, the map .♦T , is the same for each t. A rearrangement is a map from
function spaces and one may wonder which properties of this map make (5.49) valid.
An answer is given in the following theorem, proved in [9], but before stating it let
us define two properties.
Let .X ⊂ M(Rn ) and .T : X → X. We say that:
1. T is equimeasurable if .Hn ({x : Tf (x) > t}) = Hn ({x : f (x) > t}) for .t ∈ R;
2. T is monotonic if .f, g ∈ X, .f ≤ g, essentially, implies .Tf ≤ T g.
Theorem 5.42 Let .X = S(Rn ) or .V(Rn ), let .T : X → X be equimeasurable and
monotonic. Then there exists a map .♦T : Ln → Ln for which (5.49) is valid. This
map is defined for .A ∈ Ln by
♦T A = {x : T 1A (x) = 1},
.
We recall that the term essentially means up to a set of .Hn -measure zero. We can
thus define the notion of rearrangement as follows.
Let .X ⊂ M(Rn ). A map .T : X → X is called a rearrangement if it is
equimeasurable and monotonic.
Bianchi et al. [9] and [12] have studied rearrangements in an abstract setting,
based on the properties that they satisfy, independent from the specific rearrange-
ment. For the convenience of the reader, we now state five results proved in [9] as
Lemmas 4.1, 4.5, 4.7, Theorem 4.8 and the remarks that follow it, and Theorem 4.9,
respectively.
Proposition 5.43
(i) If .T : S(Rn ) → S(Rn ) is equimeasurable, then .ess inf Tf = ess inf f for
.f ∈ S(R ).
n
5 Symmetrizations 281
(ii) If .T : M(Rn ) → M(Rn ) is a rearrangement, then .ess inf Tf ≥ ess inf f for
.f ∈ M(R ). Hence, .T : S(R ) → S(R ).
n n n
constant functions.
Proposition 5.44 Let .X = M(Rn ), .M+ (Rn ), .S(Rn ), or .V(Rn ), and let .T : X →
X be equimeasurable.
(i) The induced map .♦T : Ln → Ln given by
♦T A = {x : T 1A (x) = 1}
.
T 1A = 1♦T A ,
.
essentially.
Proposition 5.45 Let .X = S(Rn ) or .V(Rn ) and let .T : X → X be a
rearrangement. For .X = S(Rn ), .A ∈ Ln , and .α, β ∈ R with .α ≥ 0, we have
T (α1A + β) = α T 1A + β,
.
.Tf (x) = max {sup{t ∈ Q, t > ess inf f : x ∈ ♦T {z : f (z) ≥ t}}, ess inf f } ,
essentially.
Proposition 5.47 Let .T : S(Rn ) → S(Rn ) be a rearrangement and let .f ∈ S(Rn ).
If .ϕ : R → R is right-continuous and increasing (i.e., non-decreasing), then .ϕ ◦ f ∈
S(Rn ) and
ϕ(Tf ) = T (ϕ ◦ f ),
.
essentially.
282 G. Bianchi and P. Gronchi
We recall that the symbol .A∗ , for a subset A of .Rn , denotes the set of points of .Rn
of density 1 for A.
We say that a rearrangement T is smoothing if the associated map .♦T is
smoothing, i.e. if
essentially, for each .d > 0 and bounded measurable set A, where .♦∗T A is defined
by
We recall (see Lemma 5.2) that in this definition one can equivalently require the
pointwise inclusion
(♦∗T A) + dD n ⊂ ♦∗T (A + dD n ).
. (5.51)
Bianchi et al. [12] prove that the notion of smoothing is equivalent to the rearrange-
ment reducing the modulus of continuity.
Theorem 5.48 Let .X = S(Rn ) or .V(Rn ). The rearrangement .T : X → X is
smoothing if and only if T reduces the modulus of continuity, that is, T is such that
.ωd (Tf ) ≤ ωd (f ) for .d > 0 and .f ∈ X, where
All special rearrangements mentioned in the lines following (5.48) are smooth-
ing, as we have proved in Sect. 5.3. Bianchi et al. [12] proves the Pólya–Szegő
inequality for all smoothing rearrangements.
Theorem 5.49 Let .X = S(Rn ) or .V(Rn ), let .T : X → X be a rearrangement, and
let .0 : [0, ∞) → [0, ∞) be convex with .0(0) = 0. If T is smoothing and .f ∈ X
is Lipschitz then Tf coincides with a Lipschitz function .Hn -almost everywhere on
.R , and
n
f f
. 0 (|∇Tf (x)|) dx ≤ 0 (|∇f (x)|) dx (5.52)
{x: Tf (x)≥a} {x: f (x)≥a}
The method of proof of Theorem 5.49 is new and we present it here. We divide the
proof in four steps and, for each step, we first describe the relevant ideas and then
write and prove the relative lemmas (with one exception).
Step 1. The function Tf coincides with a Lipschitz function .Hn -almost everywhere
on .Rn . Assume that L is the Lipschitz constant for f . Then Tf is Lipschitz with
a Lipschitz constant not larger than L, as T reduces the modulus of continuity,
by Theorem 5.48.
In the proof we use the following abbreviations: let .Kf,a = Kf ∩ {xn+1 ≥ a},
.KTf,a = KTf ∩ {xn+1 ≥ a} and .K
∗ ∗
Tf,a = (KTf,a ) .
Step 2. Let .C ∈ Kn+1 n+1 be an o-symmetric convex body of revolution about the
.xn+1 -axis, supported by the hyperplanes .{xn+1 = ±1}. This body in a later step
We prove this slice by slice, where by this we mean that we prove formula (5.58)
below, for .t > ess inf f . Taking the .Hn -measures of both sides of (5.58),
integrating with respect to t, and using Fubini’s theorem and the fact that
.♦T is measure preserving, we obtain (5.55). In order to prove (5.58) we need
.{x : f (x) ≥ a} bounded, and Lemma 5.51 proves that this is the case for any
In formula (5.58) we are slightly abusing notation by extending the action of .♦T
to horizontal hyperplanes in .R n+1 . To make this rigorous, if E is a subset of the
hyperplane .{xn+1 = t} = Rn + ten+1 in .Rn+1 such that .E|Rn ∈ Ln , we shall define
♦T E = (♦T (E | Rn )) + ten+1 .
. (5.56)
The action of .♦∗T can be extended in a similar fashion. Note that we have, for .t >
ess inf f ,
essentially, and
( )∗
. KTf ∩ {xn+1 = t} = ♦∗T (Kf ∩ {xn+1 = t}), (5.57)
where here and below, sets of Lebesgue density points are taken with respect to the
appropriate horizontal hyperplane identified with .Rn .
Lemma 5.50 Let .X = S(Rn ) or .V(Rn ), let .T : X → X be a rearrangement,
and let .d > 0. Let .C ∈ Kn+1
n+1 be an o-symmetric convex body of revolution about
the .xn+1 -axis, supported by the hyperplanes .{xn+1 = ±1}. If T is smoothing, .a >
d + ess inf f , and .f ∈ S(Rn ) is such that .{x : f (x) ≥ a} is bounded, then
( ∗ ) ∗
( )
. KTf,a + d int C ∩ {xn+1 = t} ⊂ ♦T (Kf,a + d int C) ∩ {xn+1 = t} (5.58)
where .rs = d g((t −s)/d) and .D n = int B n . Indeed, .p ∈ (L+d int C)∩{xn+1 = t}
if and only if .p | <en+1 > = ten+1 and there is a .z ∈ L such that .p ∈ z + d int C. If
.z | <en+1 > = sen+1 , then this holds if and only if .t − d < s < t + d and
( )
t −s
.|p − ||t z| < d g ,
d
that is, .p ∈ ||t (z + rs D n ).
Applying (5.59) with L replaced by .L ∩ {xn+1 ≥ a}, we obtain
Let .f ∈ X satisfy the hypotheses of the lemma. It is not difficult to prove that
( )∗
. Kf∗ ∩ {xn+1 = s} ⊂ Kf ∩ {xn+1 = s} , (5.61)
where the set of Lebesgue density points on the right is formed with respect to the
hyperplane .{xn+1 = s} = Rn + sen+1 , identified with .Rn . (See [12, Lemma 5.1] for
a detailed proof.) We have
∗ ∗
KTf,a
. ⊂ KTf ∩ {xn+1 ≥ a}∗ ⊂ KTf
∗
∩ {xn+1 ≥ a}.
whenever .s ≥ a, while the set on the left is clearly empty if .s < a. We use (5.59)
with .L = (KTf ∩ {xn+1 ≥ a})∗ , (5.62), (5.57), the fact that T is smoothing as
expressed by (5.51) with .A = Kf ∩ {xn+1 = s}, the fact that the action of .♦∗T
as extended by (5.56) is the same for each t, the pointwise monotonicity of .♦∗T ,
and (5.60) with .L = Kf , to obtain
( )
∗
. KTf,a + d int C ∩ {xn+1 = t}
|| ( )
∗
= ||t [KTf,a ∩ {xn+1 = s}] + rs D n
t−d<s<t+d
|| ( )
⊂ ||t [KTf ∩ {xn+1 = s}]∗ + rs D n
t−d<s<t+d, s≥a
|| ([ ∗ ( )] )
= ||t ♦T Kf ∩ {xn+1 = s} + rs D n
t−d<s<t+d, s≥a
|| ( [ ])
⊂ ||t ♦∗T (Kf ∩ {xn+1 = s}) + rs D n
t−d<s<t+d, s≥a
|| ( [ ])
= ♦∗T ||t (Kf ∩ {xn+1 = s}) + rs D n
t−d<s<t+d, s≥a
( || [ ])
⊂ ♦∗T ||t (Kf ∩ {xn+1 = s}) + rs D n
t−d<s<t+d, s≥a
( )
= ♦∗T (Kf,a + d int C) ∩ {xn+1 = t} .
u
n
Lemma 5.51 Let .f ∈ S(Rn ) be Lipschitz. If .a > ess inf f , then .{x : f (x) ≥ a} is
bounded.
286 G. Bianchi and P. Gronchi
Proof Let .ε > 0 be such that .a − ε > ess inf f and let L be the Lipschitz constant
of f .
Suppose that .{x : f (x) ≥ a} is unbounded. Then there are points .xk in this set
with .|xk+1 | > |xk | + 2ε/(1 + L) for .k ∈ N. The Lipschitz property implies that
.f (x) ≥ a − ε whenever .x ∈ B (xk , ε/(1 + L)), .k ∈ N. As these balls are disjoint,
n
.H ({x : f (x) ≥ a − ε}) = ∞, contradicting .f ∈ S(R ). u
n
n
Recall that .MC (A) is the anisotropic outer Minkowski content of .A ∈ M(Rn )
with respect to C defined in (5.4). We will apply this notion in .Rn+1 . Also recall
that .hC is the support function of C and .Gf denotes the graph of .f ∈ M(Rn ).
Step 3. Since
∗
Hn+1 (KTf,a
. ) = Hn+1 (KTf,a ) = Hn+1 (Kf,a ),
∗
Hn+1 (KTf,a ∗ )
+ d int C) − Hn+1 (KTf,a
.
d
( )
Hn+1 Kf,a + d int C − Hn+1 (Kf,a )
≤ .
d
Passing to the limit as .d → 0, if they exist, we obtain an inequality between the
respective anisotropic outer Minkowski contents, i.e.
∗
MC (KTf,a
. ) ≤ MC (Kf,a ). (5.63)
Lemma 5.52 below, applied to Tf and to f, proves that the limits exist and that
this inequality can be expressed in terms of integrals over the graphs of T f and
f. The inequality (5.63) becomes
f f
n
. hC (ν(x)) dH (x) ≤ hC (ν(x)) dHn (x), (5.64)
GTf ∩{xn+1 >a} Gf ∩{xn+1 >a}
where .ν(x) denotes the outer unit normal to the graph. The integral on the right
can be written as
f ( )/
(−∇f (y), 1)
. hC / 1 + |∇f (y)|2 dy
{y:f (y)>a} 1 + |∇f (y)|2
f
= hC (−∇f (y), 1) dy,
{y:f (y)>a}
5 Symmetrizations 287
where we used the 1-homogeneity of .hC . Similarly, the integral on the left can be
rewritten in the same form, with f replaced by Tf. Consequently, (5.64) yields
f f
. hC (−∇Tf (y), 1) dy ≤ hC (−∇f (y), 1) dy. (5.65)
{y:f (y)>a}
{y:Tf (y)>a}
In order to apply Lemma 5.52 we have to assume .Hn ({x : f (x) = a}) = Hn ({x :
Tf (x) = a}) = 0. In the next step we show that, for the purpose of proving
Theorem 5.49, this is not restrictive.
Lemma 5.52 Let .f ∈ S(Rn ) be Lipschitz and let C be as in Lemma 5.50. Let
n
.a > ess inf f be such that .H ({x : f (x) = a}) = 0. Then
f
( ∗ ) ( )
.MC Kf,a = MC Kf,a = hC (ν(x)) dHn (x)
Gf ∩{xn+1 >a}
μ(x + rD n ) ≥ γ r n ,
. (5.67)
then E has finite perimeter, the anisotropic outer Minkowski content of E with
respect to C is defined, and
f
MC (E) =
. hC (ν(x)) dHn (x), (5.68)
∂ eE
Step 4: conclusion. Lemma 5.53 below proves that, given any .M > 0, it is
possible to choose C so that
hC (y, 1) = 1 + b 0(|y|),
. ∀y ∈ Rn : |y| < M, (5.69)
288 G. Bianchi and P. Gronchi
for some .b > 0. If we choose M larger than the Lipschitz constant of f and of Tf,
we have
for .Hn -almost all .x ∈ Rn . Assume .Hn ({x : f (x) = a}) = 0. Note that, by the
equimeasurability of T, this is equivalent to .Hn ({x : Tf (x) = a}) = 0. Under
this assumption inequality (5.65) is valid and, using (5.69), we can rewrite it in
terms of .0 as
f f
. 1 + b0(|∇Tf (x)|) dx ≤ 1 + b0(|∇f (x)|) dx.
{y:Tf (y)>a} {y:f (y)>a}
(5.70)
Since .Hn ({y : Tf (y) > a}) = Hn ({y : f (y) > a}), the terms 1 in the integrands
give the same contribution, they cancel each other, and (5.70) implies the Pólya–
Szegő inequality (5.52).
If .Hn ({x : f (x) = a}) > 0 we argue by approximation. The set of values t
such that .Hn ({x : f (x) = t}) = 0 is dense in .(ess inf f, ∞), so there is an
increasing sequence .{am } contained in .(ess inf f, a) and converging to a such
that .Hn ({x : f (x) = am }) = Hn ({x : Tf (x) = am }) = 0 for each m. The
validity of (5.52) with .a = am , for each m, implies, in the limit, its validity for a.
Finally, by Proposition 5.43, we have .ess inf Tf = ess inf f . Letting .a →
ess inf f in (5.52), we arrive at (5.53).
Lemma 5.53 Let .0 : [0, ∞) → [0, ∞) be convex with .0(0) = 0 and let .M > 0.
Then there exist .b > 0 and an o-symmetric convex body .C ⊂ Rn+1 of revolution
about the .xn+1 -axis, such that
hC (y, 1) = 1 + b 0(|y|),
. (5.71)
where .m > 0 and .q ≤ 0 are such that .w : [0, ∞) → [0, ∞) is convex. Then, for
y ∈ Rn and .t ∈ R, define
.
{
|t| (1 + b w(|y|/|t|) , if t /= 0,
.g(y, t) = . (5.72)
b m|y|, if t = 0,
5 Symmetrizations 289
{
|t| (1 + b 0(|y|/|t|) , if |t| ≥ |y|/M,
= (5.73)
b m|y| + (1 + b q)|t|, if |t| ≤ |y|/M,
where .b > 0. It is enough to show that b can be chosen so that .g = hC is the support
function of a convex body C, since the origin symmetry and symmetry about the
.xn+1 -axis of C, and (5.71), then follow directly from the definition of g. To this
end, note that from (5.72), the positive homogeneity of g follows immediately, and
the subadditivity of g for .t > 0 or for .t < 0 is a routine exercise using the triangle
inequality and the convexity of .w. It is then enough to observe that if b is small
enough to ensure that .1 + b q > 0, then the function .b m|y| + (1 + b q)|t| in (5.73)
coincides with the support function of the o-symmetric spherical cylinder with the
.xn+1 -axis as its axis, with height .2(1 + b q) and radius .b m. u
n
References
1. A.D. Aleksandrov, On mean values of support functions. Soviet Math. Dokl. 8, 149–153 (1967)
2. A. Baernstein II, Symmetrization in Analysis, ed. by D. Drasin, R.S. Laugesen. New Mathe-
matical Monographs, vol. 36 (Cambridge University Press, Cambridge, 2019). https://doi.org/
10.1017/9781139020244
3. I. Bárány, C. Buchta, Random polytopes in a convex polytope, independence of shape, and
concentration of vertices. Math. Ann. 297, 467–497 (1993)
4. F. Barthe, M. Fradelizi, The volume product of convex bodies with many hyperplane
symmetries. Am. J. Math. 135, 311–347 (2013)
5. G. Bianchi, Poliedri di 5 vertici inscritti in convessi: una proprietá estremale. Boll. Un. Mat.
Ital. B (6) 5, 767–780 (1986)
6. G. Bianchi, P. Gronchi, Steiner symmetrals and their distance from a ball. Israel J. Math 135,
181–192 (2003)
7. G. Bianchi, A. Burchard, P. Gronchi, A. Volčič, Convergence in shape of Steiner symmetriza-
tions. Indiana Univ. Math. J. 61, 1695–1710 (2012)
8. G. Bianchi, R.J. Gardner, P. Gronchi, Symmetrization in geometry. Adv. Math. 306, 51–88
(2017)
9. G. Bianchi, R.J. Gardner, P. Gronchi, M. Kiderlen, Rearrangement and polarization. Adv.
Math. 374, 107380, 51 (2020)
10. G. Bianchi, R.J. Gardner, P. Gronchi, Full rotational symmetry from reflections or rotational
symmetries in finitely many subspaces. Indiana Univ. Math. J. 71, 767–784 (2022)
11. G. Bianchi, R.J. Gardner, P. Gronchi, Convergence of symmetrization processes. Indiana Univ.
Math. J. 71, 785–817 (2022)
12. G. Bianchi, R.J. Gardner, P. Gronchi, M. Kiderlen, The Pólya-Szegő inequality for smoothing
rearrangements. arXiv:2206.09833 (2022)
13. G. Bianchi, D.A. Klain, E. Lutwak, D. Yang, G. Zhang, A countable set of directions is
sufficient for Steiner symmetrization. Adv. Appl. Math. 47, 869–873 (2011)
14. C. Bianchini, A. Colesanti, A sharp Rogers and Shephard inequality for the p-difference body
of planar convex bodies. Proc. Amer. Math. Soc. 136, 2575–2582 (2008)
15. W. Blaschke, Über affine Geometrie IX: Verschiedene Bemerkungen und Aufgaben. Ber. Verh.
Sächs. Akad. Wiss. Leipzig Math.-Phys. Kl. 69, 412–420 (1917)
16. W. Blaschke, Über affine Geometrie X: Eine Minimumeigenschaft des Ellipsoids. Leipziger
Ber. 69, 421–435 (1917)
290 G. Bianchi and P. Gronchi
17. W. Blaschke, Über affine Geometrie XI: Lösung des “Vierpunktproblems” von Sylvester aus
der Theorie der geometrischen Wahrscheinlichkeiten. Leipziger Ber. 69, 436–453 (1917)
18. W. Blaschke, Affine Geometrie XIV: Eine Minimumaufgabe für Legendres Trägheitsellipsoid.
Ber. Verh. Sächs. Akad. Leipzig, Math.-Phys. Kl. 70, 72–75 (1918)
19. W. Blaschke, Kreis und Kugel (Chelsea Publishing, New York, 1949)
20. T. Bonnesen, W. Fenchel, Theory of Convex Bodies (BCS Associates, Moscow, 1987)
21. J. Bourgain, M. Meyer, V.D. Milman, A. Pajor, On a geometric inequality, in Geometric Aspects
of Functional Analysis. Lecture Notes in Mathematics, vol. 1317 (Springer, Berlin, 1988), pp.
271–282
22. J. Bourgain, V.D. Milman, New volume ratio properties for convex symmetric bodies in Rn .
Invent. Math. 88, 319–340 (1987)
23. H. Brunn, Über Kurven ohne Wendepunkte (Habilitationschrift, München, 1889)
24. A. Burchard, G.R. Chambers, A. Dranovski, Ergodic properties of folding maps on spheres.
Discrete Contin. Dyn. Syst. 37, 1183–1200 (2017)
25. A. Burchard, M. Fortier, Random polarizations. Adv. Math. 234, 550–573 (2013)
26. H. Busemann, Volumes in terms of concurrent cross-sections. Pac. J. Math. 3, 1–12 (1953)
27. S. Campi, A. Colesanti, P. Gronchi, A note on Sylvester’s problem for random polytopes in a
convex body. Rend. Istit. Mat. Univ. Trieste 31, 79–94 (1999)
28. S. Campi, P. Gronchi, The Lp -Busemann-Petty centroid inequality. Adv. Math. 167, 128–141
(2002)
29. S. Campi, P. Gronchi, On the reverse Lp -Busemann-Petty centroid inequality. Mathematika
49, 1–11 (2002)
30. S. Campi, P. Gronchi, Extremal convex sets for Sylvester-Busemann type functionals. Appl.
Anal. 85, 129–141 (2006)
31. S. Campi, P. Gronchi, On volume product inequalities for convex sets. Proc. Am. Math. Soc.
134, 2393–2402 (2006)
32. S. Campi, P. Gronchi, Volume inequalities for Lp -zonotopes. Mathematika 53, 71–80 (2006)
33. D. Coupier, Y. Davydov, Random symmetrizations of convex bodies. Adv. in Appl. Probab. 46,
603–621 (2014)
34. L. Dalla, D.G. Larman, Volumes of a random polytope in a convex set, in Applied Geometry
and Discrete Mathematics, ed. by P. Gritzmann, B. Sturmfels. DIMACS Series Discrete
Mathematics and Theoretical Computer Science, vol. 4 (American Mathematical Society,
Providence, 1991), pp. 175–180
35. W.J. Firey, p-means of convex bodies. Math. Scand. 10, 17–24 (1962)
36. R.J. Gardner, The Brunn-Minkowski inequality. Bull. Amer. Math. Soc. 39, 355–405 (2002)
37. R.J. Gardner, Geometric tomography, in Encyclopedia of Mathematics and its Applications,
vol. 58, 2nd edn. (Cambridge University Press, New York, 2006)
38. R.J. Gardner, D. Hug, W. Weil, Operations between sets in geometry. J. Eur. Math. Soc.
(JEMS) 15, 2297–2352 (2013)
39. A.A. Giannopoulos, On the mean value of the area of a random polygon in a plane convex
body. Mathematika 39, 279–290 (1992)
40. Y. Gordon, M. Meyer, S. Reisner, Zonoids with minimal volume-product—a new proof. Proc.
Amer. Math. Soc. 104, 273–276 (1988)
41. E.L. Grinberg, Isoperimetric inequalities and identities for k-dimensional cross-sections of
convex bodies. Math. Ann. 291, 75–86 (1991)
42. H. Groemer, On some mean values associated with a randomly selected simplex in a convex
set. Pac. J. Math. 45, 525–533 (1973)
43. P.M. Gruber, Convex and discrete geometry, in Grundlehren der mathematischen Wissen-
schaften, vol. 336 (Springer, Berlin, 2007)
44. H. Hadwiger, Vorlesungen über Inhalt, Oberfläche und Isoperimetrie (German) (Springer-
Verlag, Berlin/Göttingen/Heidelberg, 1957)
45. M. Hartzoulaki, G. Paouris, Quermassintegrals of a random polytope in a convex body. Arch.
Math. 80, 430–438 (2003)
46. F. John, Polar correspondence with respect to convex regions. Duke Math. J. 3, 355–369 (1937)
5 Symmetrizations 291
47. D.A. Klain, Steiner symmetrization using a finite set of directions. Adv. Appl. Math. 48, 340–
353 (2012)
48. B. Klartag, 5n Minkowski symmetrizations suffice to arrive at an approximate Euclidean ball.
Ann. Math. (2) 156, 947–960 (2002)
49. B. Klartag, Rate of convergence of geometric symmetrizations. Geom. Funct. Anal. 14, 1322–
1338 (2004)
50. J. Lindenstrauss, V.D. Milman, Local theory of normed spaces and convexity, in Handbook
of Convex Geometry, ed. by P.M. Gruber, J.M. Wills (North-Holland, Amsterdam, 1993), pp.
1149–1220
51. L. Lussardi, E. Villa, A general formula for the anisotropic outer Minkowski content of a Set.
Proc. Roy. Soc. Edinburgh Sect. A 146, 393–413 (2016)
52. E. Lutwak, A general isepiphanic inequality. Proc. Amer. Math. Soc. 90, 415–421 (1984)
53. E. Lutwak, Inequalities for Hadwiger’s harmonic Quermassintegrals. Math. Ann. 280, 165–
175 (1988)
54. E. Lutwak, Selected affine isoperimetric inequalities. in Handbook of Convex Geometry, ed. by
P.M. Gruber, J.M. Wills (North-Holland, Amsterdam, 1993), pp. 151–176
55. E. Lutwak, The Brunn-Minkowski-Firey theory I: Mixed volumes and the Minkowski problem.
J. Differ. Geom. 38, 131–150 (1993)
56. E. Lutwak, D. Yang, G. Zhang, A new ellipsoid associated with convex bodies. Duke Math. J.
104, 375–390 (2000)
57. E. Lutwak, D. Yang, G. Zhang, Lp -affine isoperimetric inequalities. J. Differ. Geom. 56, 111–
132 (2000)
58. E. Lutwak, G. Zhang, Blaschke-Santaló inequalities. J. Differ. Geom. 47, 1–16 (1997)
59. A.M. Macbeath, An extremal property of the hypersphere. Math. Proc. Cambridge Philos. Soc.
47, 245–247 (1951)
60. K. Mahler, Ein Übertragungsprinzip für konvexe Körper. Casopis Pêst. Mat. Fys. 68, 93–102
(1939)
61. K. Mahler, Ein Minimalproblem für konvexe Polygone. Mathematica (Zutphen) B 7, 118–127
(1939)
62. P. Mani-Levitska, Random Steiner symmetrizations. Studia Sci. Math. Hungar. 21, 373–378
(1986)
63. P. McMullen, New combinations of convex sets. Geom. Dedicata 78, 1–19 (1999)
64. M. Meyer, Convex bodies with minimal volume product in R2 . Monatsh. Math. 112, 297–301
(1991)
65. M. Meyer, A. Pajor, On the Blaschke-Santaló inequality. Arch. Math. 55, 82–93 (1990)
66. M. Meyer, S. Reisner, Shadow systems and volumes of polar convex bodies. Mathematika 53,
129–148 (2006)
67. E. Milman, A. Yehudayoff, Sharp isoperimetric inequalities for affine quermassintegrals. J.
Amer. Math. Soc. 36, 1061–1101 (2023)
68. V.D. Milman, A. Pajor, Isotropic position and inertia ellipsoids and zonoids of the unit ball of
a normed n-dimensional space, in Geometric Aspects of Functional Analysis. Lecture Notes in
Mathematics, vol. 1376 (Springer, Berlin, 1989), pp. 64–104
69. C.M. Petty, Centroid surfaces. Pacific J. Math. 11, 1535–1547 (1961)
70. C.M. Petty, Isoperimetric problems, in: Proceedings of the Conference on Convexity and
Combinatorial Geometry (Department of Mathematics, The University of Oklahoma, Norman,
1971), pp. 26–41
71. R.E. Pfiefer, The historical development of J. J. Sylvester’s four point problem. Math. Mag.
62, 309–317 (1989)
72. G. Pólya, G. Szegö, Isoperimetric Inequalities in Mathematical Physics. Annals of Mathemat-
ics Studies, vol. 27 (Princeton University Press, Princeton, 1951)
73. S. Reisner, Random polytopes and the volume product of symmetric convex bodies. Math.
Scand. 57, 386–392 (1985)
74. S. Reisner, Zonoids with minimal volume product. Math. Z. 192, 339–346 (1986)
292 G. Bianchi and P. Gronchi
75. C.A. Rogers, G.C. Shephard, Some extremal problems for convex bodies. Mathematika 5,
93–102 (1958)
76. J. Saint Raymond, Sur le volume des corps convexes symétriques, in Initiation Seminar on
Analysis: G. Choquet-M. Rogalski-J. Saint-Raymond, 20th Year: 1980/1981. Publications
mathématiques de l’Université Pierre et Marie Curie, vol. 46 (University of Paris VI, Paris,
1981)
77. C. Saroglou, Characterizations of extremals for some functionals on convex bodies. Canad. J.
Math. 62, 1404–1418 (2010)
78. E. Sas, Über eine Extremumeigenschaft der Ellipsen. Compos. Math. 6, 468–470 (1939)
79. R. Schneider, Convex bodies: the Brunn-Minkowski theory, in Encyclopedia of Mathematics
and Its Applications, vol. 151, 2nd expanded edn. (Cambridge University Press, Cambridge,
2014)
80. R. Schneider, W. Weil, Zonoids and related topics, in Convexity and Its Applications, ed. by
P.M. Gruber, J.M. Wills (Birkhäuser, Basel, 1983), pp. 296–317
81. P. Schöpf, Gewichtete Volumsmittelwerte von Simplices, welche zufällig in einem konvexen
Körper des Rn gewählt werden. Monatsh. Math. 83, 331–337 (1977)
82. G.C. Shephard, Shadow systems of convex bodies. Israel J. Math. 2, 229–236 (1964)
83. G. Talenti, The standard isoperimetric theorem, in Handbook of Convex Geometry, ed. by P.M.
Gruber, J.M. Wills, J.M. (North-Holland, Amsterdam, 1993), pp. 73–123
84. G. Talenti, The art of rearranging. Milan J. Math. 84, 105–157 (2016)
85. J. Ulivelli, Generalization of Klain’s theorem to Minkowski symmetrization of compact sets
and related topics. Canad. Math. Bull. 66, 124–141 (2023)
86. J. Ulivelli, Convergence properties of symmetrization processes. Adv. Appl. Math. 46, 102484
(2023)
87. J. Van Schaftingen, Approximation of symmetrizations and symmetry of critical points. Topol.
Methods Nonlinear Anal. 28, 61–85 (2006)
88. J. Van Schaftingen, Explicit approximation of the symmetric rearrangement by polarizations.
Arch. Math. 93, 181–190 (2009)
89. A. Volčič, Random Steiner symmetrizations of sets and functions. Calc. Var. Partial Differential
Equations 46, 555–569 (2013)
90. A. Volčič, On iterations of Steiner symmetrizations. Ann. Mat. Pura Appl. (4) 195, 1685–1692
(2016)
91. G. Zhang, Restricted chord projection and affine inequalities. Geom. Dedicata 39, 213–222
(1991)
Index
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 293
A. Colesanti, M. Ludwig (eds.), Convex Geometry, C.I.M.E. Foundation Subseries
2332, https://doi.org/10.1007/978-3-031-37883-6
294 Index
Editorial Policy
1. Lecture Notes aim to report new developments in all areas of mathematics and their
applications – quickly, informally and at a high level. Mathematical texts analysing new
developments in modelling and numerical simulation are welcome.
Manuscripts should be reasonably self-contained and rounded off. Thus they may, and
often will, present not only results of the author but also related work by other people.
They may be based on specialised lecture courses. Furthermore, the manuscripts should
provide sufficient motivation, examples and applications. This clearly distinguishes
Lecture Notes from journal articles or technical reports which normally are very concise.
Articles intended for a journal but too long to be accepted by most journals, usually do not
have this “lecture notes” character. For similar reasons it is unusual for doctoral theses to
be accepted for the Lecture Notes series, though habilitation theses may be appropriate.
4. In general, monographs will be sent out to at least 2 external referees for evaluation.
A final decision to publish can be made only on the basis of the complete manuscript,
however a refereeing process leading to a preliminary decision can be based on a pre-final
or incomplete manuscript.
Volume Editors of multi-author works are expected to arrange for the refereeing, to the
usual scientific standards, of the individual contributions. If the resulting reports can be
forwarded to the LNM Editorial Board, this is very helpful. If no reports are forwarded
or if other questions remain unclear in respect of homogeneity etc, the series editors may
wish to consult external referees for an overall evaluation of the volume.
6. Careful preparation of the manuscripts will help keep production time short besides
ensuring satisfactory appearance of the finished book in print and online. After
acceptance of the manuscript authors will be asked to prepare the final LaTeX source
files (see LaTeX templates online: https://www.springer.com/gb/authors-editors/book-
authors-editors/manuscriptpreparation/5636) plus the corresponding pdf- or zipped ps-
file. The LaTeX source files are essential for producing the full-text online version of
the book, see http://link.springer.com/bookseries/304 for the existing online volumes
of LNM). The technical production of a Lecture Notes volume takes approximately 12
weeks. Additional instructions, if necessary, are available on request from lnm@springer.
com.
7. Authors receive a total of 30 free copies of their volume and free access to their book on
SpringerLink, but no royalties. They are entitled to a discount of 33.3 % on the price of
Springer books purchased for their personal use, if ordering directly from Springer.
Addresses:
Professor Jean-Michel Morel, CMLA, École Normale Supérieure de Cachan, France
E-mail: [email protected]