(Jaroslav Lukeš, Jan Malý, Ivan Netuka, Jiří S
(Jaroslav Lukeš, Jan Malý, Ivan Netuka, Jiří S
≥ Walter de Gruyter
Berlin · New York
Authors
Jaroslav Lukeš Jan Malý
Department of Mathematical Analysis Department of Mathematical Analysis
Faculty of Mathematics and Physics Faculty of Mathematics and Physics
Charles University Charles University
Sokolovská 83 Sokolovská 83
18675 Prague 8, Czech Republic 18675 Prague 8, Czech Republic
E-Mail: [email protected] and
Department of Mathematics
Faculty of Science
J. E. Purkyně University
České mládeže 8
40096 Ústı́ nad Labem, Czech Republic
E-Mail: [email protected]
Ivan Netuka Jiřı́ Spurný
Mathematical Institute Department of Mathematical Analysis
Faculty of Mathematics and Physics Faculty of Mathematics and Physics
Charles University Charles University
Sokolovská 83 Sokolovská 83
18675 Prague 8, Czech Republic 18675 Prague 8, Czech Republic
E-Mail: [email protected] E-Mail: [email protected]
Series Editors
Carsten Carstensen Niels Jacob
Department of Mathematics Department of Mathematics
Humboldt University of Berlin Swansea University
Unter den Linden 6 Singleton Park
10099 Berlin, Germany Swansea SA2 8PP, Wales, United Kingdom
E-Mail: [email protected] E-Mail: [email protected]
Nicola Fusco Karl-Hermann Neeb
Dipartimento di Matematica Department of Mathematics
Università di Napoli Frederico II Technische Universität Darmstadt
Via Cintia Schloßgartenstraße 7
80126 Napoli, Italy 64289 Darmstadt, Germany
E-Mail: [email protected] E-Mail: [email protected]
Mathematics Subject Classification 2000: 46-02, 31-02, 52-02, 46A55, 52A07, 46B99, 31B05, 31A05, 31A25, 31B20,
31C05, 31D05, 35K05, 35K20, 28A05, 54H05
Keywords: Convex sets, Choquet theory, Banach spaces, descriptive set theory, measure theory, potential theory,
Dirichlet problem
앪
앝 Printed on acid-free paper which falls within the guidelines of the ANSI
to ensure permanence and durability.
ISBN 978-3-11-020320-2
쑔 Copyright 2010 by Walter de Gruyter GmbH & Co. KG, 10785 Berlin, Germany.
All rights reserved, including those of translation into foreign languages. No part of this book may be reproduced in
any form or by any means, electronic or mechanical, including photocopy, recording, or any information storage and
retrieval system, without permission in writing from the publisher.
Printed in Germany.
Cover design: Martin Zech, Bremen.
Printing and binding: Hubert & Co. GmbH & Co. KG, Göttingen.
Introduction
basic properties. The most important fact is that many descriptive properties are sta-
ble with respect to perfect mappings, which allows us to transfer abstract Borel affine
functions to the setting of compact convex sets.
Simplicial function spaces are studied in Chapter 6. We discuss several classes of
simplicial function spaces, namely the Bauer and Markov simplicial function spaces
and spaces with boundary of type Fσ . Among other results, the abstract Dirichlet
problem for continuous and non-continuous functions is considered. Choquet sim-
plices are presented at the end of the chapter.
Next we generalize the basic concepts for function cones since they are indispens-
able in potential theory. We focus in particular on ordered compact convex sets.
Analogues of faces in a non-convex setting, so-called Choquet sets, are investigated
in Chapter 8. The main result is a characterization of simplicial spaces by means of
Choquet sets.
Suitably chosen families of closed extremal sets generate interesting boundary
topologies on the set of extreme points. Chapter 9 studies these topologies and func-
tions continuous with respect to them. It turns out that maximal measures induce
measures on sets of extreme points that are regular with respect to boundary topolo-
gies. The last section is devoted to a study of a facial topology and facially continuous
functions.
Chapter 10 collects several deeper results on function spaces and compact convex
sets. Among others, study of Shilov and James boundaries, Lazar’s improvement of
the Banach–Stone theorem, results on automatic boundedness of affine and convex
functions, embedding of `1 in Banach spaces, metrizability of compact convex sets
and their open images and some topological properties of the set of extreme points.
The Lazar selection theorem and its consequences occupy the first part of Chap-
ter 11. The second part is devoted to a presentation of Debs’ proof of Talagrand’s
theorem on measurable selectors.
Chapter 12 is concerned with two methods of constructing new function spaces:
products and inverse limits. We show that both operations preserve simpliciality and
describe resulting boundaries. The inverse limits lead to an interesting description
of metrizable simplices as inverse limits of finite-dimensional simplices. The general
results are illustrated by a construction of the Poulsen simplex and a couple of compact
convex sets due to Talagrand.
In Chapter 13, general results from the Choquet theory are applied to potential
theory and several of its basic notions are investigated from this perspective. Impor-
tant function cones and spaces appearing in potential theory are studied in detail, in
particular, in connection to various solution methods for the Dirichlet problem. The
functional analysis approach makes it possible to provide an interesting interpretation,
for instance, of balayage and regular points in terms of representing measures and the
Choquet boundary of suitable spaces and cones. The exposition covers potential the-
Introduction vii
ory for the Laplace equation and the heat equation as well as a more general setting
(harmonic spaces, fine potential theory etc.).
The final Chapter 14 presents several applications of the integral representation
theorems, such as for doubly stochastic matrices, the Riesz–Herglotz theorem, the
Lyapunov theorem on the range of a vector measure, the Stone–Weierstrass theorem,
positive-definite functions and invariant and ergodic measures.
Each chapter concludes with a series of exercises with sketches of proofs and with
concluding notes and comments where we try to give precise references and due cred-
its for the results presented in the main body of the text, and discuss additional mate-
rial which is related to the topics of the chapter in question, but was not included with
complete proofs. Open problems are also mentioned.
Since the presented material originates in an amalgamation of functional analy-
sis, measure theory, topology, descriptive set theory and potential theory, we collect
the needed notions and facts in the Appendix, sometimes even with proofs. We se-
lected the following books for each subject as the key references: W. Rudin [403]
and M. Fabian, P. Habala, P. Hájek, V. Montesinos Santalucı́a, J. Pelant and V. Zizler
[173] for functional analysis, D. H. Fremlin [182], [181] and [183] for measure the-
ory, R. Engelking [169] and K. Kuratowski [285] for topology, A. S. Kechris [262]
and C. A. Rogers and J. E. Jayne [394] for descriptive set theory, D. H. Armitage and
S. J. Gardiner [21] for classical potential theory and J. Bliedtner and W. Hansen [66]
for abstract potential theory.
Next we point out what is omitted from the book. First, we focus on integral repre-
sentation theorems for compact sets, and thus the readers interested in theory of sets
with the Radon–Nikodym property are referred to R. D. Bourgin [82], and those inter-
ested in Choquet theory in sets of measures are referred to G. Winkler [473]. Second,
although we consider several geometric aspects of simplicial spaces, they are not at
the center of our attention. They are thoroughly investigated in H. E. Lacey [290] and
P. Harmand, D. Werner and W. Werner [216]. Further, we do not pursue applications
of integral representation theory in C ∗ -algebras and thus we refer the interested reader
to E. M. Alfsen and F. W. Schultz [10] and [9], M. Rørdam [395] and M. Rørdam and
E. Størmer [396], B. Blackadar [59] and H. Lin [303] and the references therein. And
last but not least, our applications to potential theory do not require the full strength
of abstract potential theory and thus we restrict ourselves to a less general framework
than the one presented in J. Bliedtner and W. Hansen [66].
Except on a few explicitly stated occasions, we consider only real vector spaces and
apart from Chapter 9 we deal only with Hausdorff topologies and Radon measures.
We use the standard notation and terminology:
• N, Q, Z, R, C denote the usual sets of numbers,
• Re z and Im z denote the real and imaginary part of a complex number z, respec-
tively,
viii Introduction
• ∇f is the gradient of f .
A function f is positive if f ≥ 0, it is strictly positive if f > 0. Similarly we use
increasing, strictly
R increasing and so on. If µ is a measure, we often write µ(f ) for
the integral f dµ.
In preparation of the present book, we have received many valuable suggestions
from many colleagues. In particular, we would like to express our thanks to P. Hájek,
P. Holický, M. Johanis, O. Kalenda, P. Kaplický, M. Kraus, E. Murtinová, P. Simon,
J. Tišer, L. Zajı́ček and M. Zelený for stimulating and fruitful discussions, and to
E. Crooks for linguistic assistance. We are also indebted to the publishers for their
care and cooperation.
The preparation of the manuscript was supported by the grant 201/07/0388 of the
Grant Agency of the Czech Republic and partly by the grant MSM21620839 of the
Czech Ministry of Education.
Finally, our thanks go to Jana, Jarka, Hana and Hanka for encouragement and pa-
tience during the preparation of this book.
Introduction v
1 Prologue 1
1.1 The Korovkin theorem . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Notes and comments . . . . . . . . . . . . . . . . . . . . . . . . . . 3
14 Applications 563
14.1 Representation of convex functions . . . . . . . . . . . . . . . . . . 564
14.2 Representation of concave functions . . . . . . . . . . . . . . . . . . 567
14.3 Doubly stochastic matrices . . . . . . . . . . . . . . . . . . . . . . . 572
14.4 The Riesz–Herglotz theorem . . . . . . . . . . . . . . . . . . . . . . 573
14.5 Typically real holomorphic functions . . . . . . . . . . . . . . . . . . 575
14.6 Holomorphic functions with positive real part . . . . . . . . . . . . . 580
14.7 Completely monotonic functions . . . . . . . . . . . . . . . . . . . . 586
14.8 Positive definite functions on discrete groups . . . . . . . . . . . . . 589
14.9 Range of vector measures . . . . . . . . . . . . . . . . . . . . . . . . 593
14.10 The Stone–Weierstrass approximation theorem . . . . . . . . . . . . 595
14.11 Invariant and ergodic measures . . . . . . . . . . . . . . . . . . . . . 597
14.12 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 603
14.13 Notes and comments . . . . . . . . . . . . . . . . . . . . . . . . . . 605
A Appendix 608
A.1 Functional analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 608
A.1.A Locally convex spaces . . . . . . . . . . . . . . . . . . . . . 608
A.1.B Banach spaces . . . . . . . . . . . . . . . . . . . . . . . . . 609
A.1.C Ordered Banach spaces and lattices . . . . . . . . . . . . . . 610
A.2 Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 615
A.2.A Compact spaces and Čech–Stone compactification . . . . . . 616
A.2.B Baire and Borel sets . . . . . . . . . . . . . . . . . . . . . . 619
A.2.C Semicontinuous functions . . . . . . . . . . . . . . . . . . . 621
A.2.D Baire spaces and sets with the Baire property . . . . . . . . . 623
xvi Contents
Bibliography 669
Index 703
Chapter 1
Prologue
The task is to show that the sequence {Bn f } converges uniformly to f on [0, 1]. This
can be easily verified in the case when
f (x) = 1, x or x2 ,
since
n−1 2 1
Bn 1 = 1, Bn x = x and Bn x2 = x + x.
n n
Surprisingly, this is all that we need to compute, since these three tests are enough
to guarantee the uniform convergence of Bn f to f for all f in C([0, 1]). Indeed, one
of the current proofs of the classical Weierstrass approximation theorem is based on
the Korovkin theorem about linear operators. The Weierstrass theorem is an easy
consequence since the mappings
Bn : f 7→ Bn f, f ∈ C([0, 1]),
Proof. Pick f ∈ C([0, 1]) and ε > 0. By the uniform continuity of f , there exists
δ ∈ (0, 1) such that |f (s) − f (t)| ≤ ε for any s, t ∈ [0, 1], |s − t| ≤ δ. Now fix
t ∈ [0, 1] and set
2kf k
p∗ (x) := f (t) − ε − (x − t)2 , x ∈ [0, 1],
δ2
and
2kf k
p∗ (x) := f (t) + ε + (x − t)2 , x ∈ [0, 1].
δ2
Dealing separately with the cases |x − t| ≤ δ and |x − t| > δ, we get
kf k
|f (x) − f (t)| ≤ ε + 2(x − t)2
δ2
for each x ∈ [0, 1]. Hence,
Tn p∗ ≤ Tn f ≤ Tn p∗ .
Tn pj − pj < εδ 2 , j = 0, 1, 2.
Since
2kf k 4kf kt 2kf k
p∗ (x) = f (t) + ε + t2 2 − 2
x + 2 x2 , x ∈ [0, 1],
δ δ δ
for n ≥ N we have the estimate
In particular,
and similarly
Tn f (t) ≥ Tn p∗ (t) ≥ p∗ (t) − Cε = f (t) − Cε − ε, n ≥ N.
Hence
kTn f − f k ≤ ε(C + 1), n ≥ N,
and the proof is complete.
1.2 Notes and comments 3
In stating the Korovkin theorem, which sometimes bears the name the first Korovkin
theorem, it is possible to go further, replacing the interval [0, 1] by a suitable space,
and the set of three functions {p0 , p1 , p2 } by a more general family of functions. In
the sequel, we will take a deeper look at this issue.
H ⊂ Kor(H) = Kor(P).
λ0 e0 + · · · + λn en = 0 and λ0 + · · · + λn = 0,
Definition 2.3 (Affine hulls and subspaces, hyperplanes). For a set A ⊂ W , the affine
hull of A, denoted by aff A, is the set of all affine combinations of points of A. (A
linear combination α1 x1 + · · · + αn xn , where α1 + · · · + αn = 1, is called an affine
combination of points x1 , . . . , xn .)
A set A ⊂ W is said to be an affine subspace of W if aff A = A. Affine subspaces
are just the translations (of type) x + F , where F is a linear subspace of W and
x ∈ W . By definition, the dimension (or, the codimension) of x + F is the dimension
(or, the codimension, respectively) of F .
A set H is a hyperplane if there exists a nonzero linear functional f on W and
α ∈ R such that
H = {w ∈ W : f (w) = α}.
Since a subspace F of W is a maximal proper subspace of W if and only if there
exists a nonzero linear functional f on W such that F = ker f , we see that H is a
hyperplane if and only if there exist a maximal proper subspace F of W and w ∈ W
such that H = w + F . In other words, hyperplanes are exactly affine subspaces of
codimension 1.
6 2 Compact convex sets
Corollary 2.7. The convex hull co A of any set A ⊂ Rd is the union of all n-simplices
(n ≤ d) with vertices in A.
Proof. Obviously, any n-simplex with vertices in A, where n ≤ d, is a subset of co A.
The reverse inclusion immediately follows from Carathéodory’s theorem 2.6.
co K = F (D × K × · · · × K).
ext S = {e0 , . . . , en } .
Proof. Let x ∈ ext S. Since S = co {e0 , . . . , en } and the set S \ {x} is convex,
x = ek for some k ∈ {0, 1, . . . , n}. Conversely, select ek and assume that
1 1
ek = s + t
2 2
where s, t ∈ S = co {e0 , . . . , en }. Write
n
X n
X
s= αj ej , t= βj ej ,
j=0 j=0
8 2 Compact convex sets
with
n
X n
X
αj , βj ≥ 0, αj = βj = 1.
j=0 j=0
Then
n
X 1
ek = (αj + βj )ej
2
j=0
or
X1
(αj + βj )(ej − ek ) = 0.
2
j6=k
Consequently,
αj + β j = 0 for j ∈ {0, . . . , n} \ {k} ,
thus s = t = ek .
The following assertion shows the prominent role of extreme points in finite-dim-
ensional compact convex sets. Infinite-dimensional situation is more complicated, see
Example 2.15 and the Krein–Milman theorem 2.22.
Theorem 2.11 (Minkowski). Each point of a compact convex set C ⊂ Rd is a convex
combination of extreme points of C.
Proof. We proceed by induction on the dimension d. For the dimension d = 0, the
set C reduces to a one-point set and the assertion holds. So assume that d > 0 and
that the assertion is valid for compact convex sets in spaces of dimension smaller than
d. We may also assume that the interior of C is nonempty, for otherwise C is a subset
of an affine subspace of a smaller dimension (cf. Exercise 2.107(c)) and the assertion
follows by the induction assumption.
We distinguish two cases. If x is a boundary point of C, then by Proposition 2.4
there exists a support hyperplane L of C at x. Then the compact convex set F :=
C ∩ L lies in the affine subspace L of dimension smaller than d. By the induction
assumption, x ∈ co ext F . Since obviously ext F ⊂ ext C, the induction step is
finished.
Now suppose that x ∈ Int C. There exists a segment [a, b] ⊂ C such that x ∈ (a, b)
and a, b ∈ ∂C. Since a, b ∈ co ext C by the previous argument, we see that x can be
expressed as a convex combination of extreme points of C.
Corollary 2.13. Let x be a point of a compact convex set C ⊂ Rd . Then there exists
an n-simplex S, n ≤ d, such that x ∈ S ⊂ C and ext S ⊂ ext C.
Proof. The assertion is a rewording of the previous Minkowski–Carathéodory theo-
rem.
Remark 2.14. In [1] E. M. Alfsen constructed a non-simplicial compact polyhedron in
`1 , showing that the conclusion of the previous Corollary 2.13 in infinite-dimensional
spaces fails. At the same time, he posed a question that “it would be of some interest
to find sufficient conditions for a compact convex set X to admit a decomposition” as
in Corollary 2.13: Given x ∈ X, there would exists a set S ⊂ X such that x ∈ S,
ext S ⊂ ext C, and a unique representing measure for x carried by ext S. We present
in Exercise 6.93 an example illustrating this phenomenon.
Definition 2.16 (Extremal sets and faces). A generalization of the notion of extreme
points leads to an important concept: A nonempty subset F of a set C ⊂ W is an
extremal subset of C if x, y ∈ F provided that x, y ∈ C and λx + (1 − λ)y ∈ F for
some λ ∈ (0, 1). It is needless to say that one-point extremal sets are exactly extreme
points of C, and that C itself is an extremal set.
Convex extremal sets are called faces.
Definition 2.17 (Affine, concave and convex functions). Let C be a convex subset of
a vector space W . A real-valued function s on C is said to be concave if
s λx + (1 − λ)y ≥ λs(x) + (1 − λ)s(y)
Proof. Obvious.
Proof. Assume that there exists a point x ∈ K \co ext K. Using the geometric version
of the Hahn–Banach theorem, there exists f ∈ E ∗ such that
If
H := {z ∈ K : f (z) = min f (K)},
then H is a (nonempty) extremal subset of K by Lemma 2.19. Hence, by Proposi-
tion 2.20, H ∩ ext K 6= ∅. Since H ∩ co ext K = ∅, this is impossible. Therefore,
K ⊂ co ext K.
Proof. It is pretty clear that co ext X ⊂ X. The reverse inclusion follows from Theo-
rem 2.21.
Remarks 2.23. (a) The Krein–Milman theorem also holds in locally convex spaces
over complex numbers. For a proof see, for example, W. Rudin [403], Theorem 3.21.
(b) If K is a compact subset of a locally convex space, then co K = co ext K. This is
an immediate consequence of Theorem 2.21.
Proof. Denote
D := {x ∈ X : s(x) = min s(X)} .
Then D is a nonempty compact subset of X and, by Lemma 2.19, an extremal subset
of X. By Proposition 2.20, D ∩ ext X 6= ∅, and thus the proof is complete.
Integral representation. Now we would like to show how to use the Krein–Milman
theorem for establishing integral-type representation theorems.
Let x be a point of a compact convex set X in a locally convex space. Our aim is
to find a Radon measure µ on X so that
Z
f (x) = f dµ
X
holds for any f ∈ Ac (X). Since the functionals from E ∗ separate the points of E, and
since the restrictions to X of such functionals are elements of Ac (X), we see that the
2.1 Geometry of convex sets 13
Obviously, as noted above, the Dirac measure εx always represents the point x. In
what follows, we answer the following questions:
(a) Does any Radon measure have a barycenter ?
(b) Is any point of X a barycenter of a Radon measure carried by ext X ?
Proposition 2.27. The space M1 (K) consisting of all probability measures on a com-
pact space K is a compact convex subset of M(K) and
then
1 = αµ({x}) + (1 − α)ν({x}).
This implies that µ({x}) = ν({x}) = 1, and therefore µ = ν = εx .
If µ ∈ M1 (K) is not a Dirac measure, then there exists a compact set F ⊂ K such
that the measures ν := µ|F and λ := µ|K\F are nontrivial and distinct from µ. Since
ν λ
µ = ν(F ) + λ(K \ F ) ,
ν(F ) λ(K \ F )
Corollary 2.28. The set of all convex combinations of Dirac measures is dense in
M1 (K).
14 2 Compact convex sets
Proof. As an easy consequence of the Krein–Milman theorem 2.22 and the charac-
terization of extreme points given by Proposition 2.27, we have
Proof. With the uniqueness part already out of the way, we nowPconcentrate on an
existence proof. If a measure µPin question is molecular, µ = nj=1 P λj εxj , where
xj ∈ X, λj ≥ 0, j = 1, . . . , n, nj=1 λj = 1, then obviously r(µ) := nj=1 λj xj ∈
X is the barycenter of µ. Given a measure µ ∈ M1 (X), there exists a net {µγ } of
molecular measures on X such that µγ → µ. Since X is a compact set, there exists a
subnet {r(µα )} of {r(µγ )} converging to an element z ∈ X. Pick f ∈ Ac (X). Then
Z Z
f (z) = lim f (r(µα )) = lim f dµα = f dµ,
α α X X
In Proposition 2.38 we show that the barycenter mapping is a continuous and affine
mapping from M1 (X) into X. This mapping is surjective since r(εx ) = x for any
x ∈ X.
Proof. From the Krein–Milman theorem 2.22, we can see the following fact: if f ∈
Ac (X) and f = 0 on ext X, then f = 0 on X. We denote by B the subspace of
C(ext X) consisting of all restrictions of functions in Ac (X) to ext X. Then, for every
h ∈ B, there exists, by the above mentioned fact, a unique function b h ∈ Ac (X) which
coincides with h on ext X. We fix x ∈ X and set
ϕ : h 7→ b
h(x), h ∈ B.
for every g ∈ Ac (X), we see that the barycenter of the measure µ is exactly the
point x.
Remarks 2.32. (a) Theorem 2.31 can be proved by an alternative manner. In fact, we
can follow the proof of Proposition 2.39 step by step.
(b) In concrete applications, we are often able to characterize the set ext X. However,
the character of the elements of the set ext X \ ext X is generally rather obscure. Con-
sequently, the information concerning the support of the measure from the theorem on
integral representation is problematic, unless the set ext X is closed. Moreover, there
is another problem. Let us imagine that the set of extreme points of a compact convex
set X is dense in this set, that is, ext X = X. Then, naturally, the Krein–Milman
theorem says nothing, and equally useless is the theorem on integral representation.
Indeed, it suffices to take the Dirac measure εx at the point x for the measure repre-
senting the point x . This situation can actually occur. As an example, we can take
the closed unit ball B in an arbitrary infinite-dimensional Hilbert space, which we of
course consider to be equipped with the weak topology. The extreme points of B are
then the points of the unit sphere, and its (weak) closure is equal to the whole ball B.
A more sophisticated example with ext X = X is the Poulsen simplex in the Hilbert
space `2 (see Subsection 12.3.A).
However, much more is known. Namely, if we consider the so-called Hausdorff
metric on the set F of all nonempty compact convex subsets ofa given Banach space
X of infinite dimension, the space F is complete and the set C ∈ F : ext C 6= C
is merely meager in F. This assertion was proved by V. L. Klee in [271]. Thus, in a
certain sense, for the majority of compact convex sets we have ext C = C.
Hence the problem of whether it is possible to find a measure µ which is carried just
on the set of extreme points in the theorem on integral representation is crucial. This
problem was solved successfully by G. Choquet in the fifties of the 20th century and
laid the foundations of the Choquet theory. We will devote the next chapters to it, in
the more general setting of function spaces. The Choquet theory has provided many
insights for abstract analysis, infinite-dimensional geometry, descriptive set theory,
16 2 Compact convex sets
potential theory and other fields of mathematics. It has remained fruitful ever since
and has found new applications again and again in deriving new results.
(c) In Sections 14.5 and 14.6, we exceptionally consider locally convex spaces over
the field of complex numbers. Note that the Integral representation theorem 2.31
extends trivially to the complex case. Indeed, consider a compact convex subset X
of a complex locally convex space E. Of course, E can be regarded as a locally
convex space over the field of real numbers. Then, given x ∈ X, there exists, by
Theorem 2.31, a measure µ ∈ M1 (X) carried by ext X such that
Z
f (x) = f dµ (2.1)
X
The Integral representation theorem will be applied several times in the sequel. For
this purpose, the following easy consequence of the previous theorem will be useful.
for each continuous linear functional F on E. Let us define µ = Φ−1 ] µ̃. By Proposi-
tion A.92,
Z Z Z
F dµ̃ = (F ◦ Φ) ◦ Φ−1 dµ̃ = F ◦ Φ dΦ−1 ] µ̃
ext K ext K Q
Z (2.3)
= F ◦ ϕy dµ(y).
Q
2.1 Geometry of convex sets 17
Lemma 2.34. Let X be a compact convex subset of a locally convex space E. Then
the space (E ∗ + R)|X is dense in Ac (X).
Proof. Let X be a nonempty compact convex set. Fix a function h ∈ Ac (X) and
ε > 0. Denote
J1 := {(x, t) ∈ X × R : t = h(x)}
and
J2 := {(x, t) ∈ X × R : t = h(x) + ε} .
Then J1 , J2 is a pair of disjoint nonempty compact convex subsets of E × R. By the
Hahn–Banach theorem, there exist a functional F ∈ (E × R)∗ and λ ∈ R such that
sup F (J1 ) < λ < inf F (J2 ).
There are ϕ ∈ E ∗ and β ∈ R such that F (t, r) = ϕ(t) + βr for any t ∈ E and
any r ∈ R. From the separation of J1 and J2 it easily follows that β 6= 0. Put
ψ(t) := β1 (λ − ϕ(t)) for t ∈ E. Since
ϕ(t) + βh(t) < λ < ϕ(t) + βh(t) + βε, t ∈ X,
we easily get kh − ψk < ε.
Remark 2.35. In general, there might exist a continuous affine function on a compact
convex set X that is not of the form (E ∗ + R)|X ; see Exercise 2.111.
Proposition 2.36. A Radon measure µ ∈ M1 (X) represents x ∈ X if and only if
Z
ϕ(x) = ϕ dµ for any ϕ ∈ E ∗ .
X
R
Proof. Recall that, by definition, µ represents x ∈ X if h(x) = X h dµ for any
h ∈ Ac (X). Hence the assertion is an easy consequence of Lemma 2.34 and the
Lebesgue dominated convergence theorem.
Definition 2.37 (The barycenter revisited). Proposition 2.36 enables us to extend sli-
ghtly the definition of a barycenter to the case of nonconvex sets and points not be-
longing to this set.
Let K be a compact subset of a locally convex space E. We say that x ∈ E is the
barycenter of a Radon measure µ ∈ M1 (K), or that µ represents x, in a symbol
x = r(µ), if Z
ϕ(x) = ϕ dµ for any ϕ ∈ E∗.
K
18 2 Compact convex sets
Then µα ∈ M1 (K) and r(µα ) = xα . Since the set M1 (K) is compact, we may
assume that {µα } converges to µ ∈ M1 (K). For any ϕ ∈ E ∗ ,
Theorem 2.40 (Bauer’s characterization of ext X). Let X be a compact convex subset
of a locally convex space and x ∈ X. Then x ∈ ext X if and only if the Dirac measure
εx is the only measure from M1 (X) with a barycenter x.
Proof. We follow the same lines as in the proof of Corollary 2.8. Since
co(K1 ∪ · · · ∪ Kn ) = F (D × K1 × · · · × Kn )
where
n n
X o
D := λ ∈ Rn : λ = (λ1 , . . . , λn ), λj = 1 and λj ≥ 0 for j = 1, . . . , n
j=1
20 2 Compact convex sets
and
n
X
F : (λ, x1 , . . . , xn ) 7→ λ j xj , (λ, x1 , . . . , xn ) ∈ D × K1 × · · · × Kn
j=1
Theorem 2.43 (Milman). Let B be a subset of a locally convex space E such that the
set X := co B is compact. Then ext X ⊂ B.
Proposition 2.44. Let X be a compact convex subset of a locally convex space E and
∅ 6= B ⊂ X. Then the following assertions are equivalent:
(i) co B = X,
(ii) inf f (B) = min f (X) for each f ∈ E ∗ ,
(iii) ext X ⊂ B.
Proof. Assertions (i) and (iii) are equivalent by the Krein–Milman and Milman theo-
rems 2.22 and 2.43. The implication (i) =⇒ (ii) is obvious. Conversely, the proof
of the implication (ii) =⇒ (i) follows the same lines as the end of the proof of the
Krein–Milman theorem: if co B 6= X, the Hahn–Banach separation theorem yields
the assertion.
Proof. Since X is a metrizable compact set, the space C(X) is separable, so it is the
space Ac (X) of all continuous affine functions on X. Let {fn : n ∈ N} be a dense
subset of the closed unit ball of Ac (X). If
n1 o
T : x 7→ fn (x) , x ∈ X,
n
Proposition 2.54. The measure Λ from Example 2.53 is carried by the (closed) set
ε(spt λ) and its barycenter equals λ.
Proof. Since
Λ(ε(spt λ)) = λ(ε−1 (ε(spt λ))) = λ(spt λ) = 1,
∗ by ε(spt λ).
we see that Λ is carried
Pick ϕ ∈ M(K) . By duality theory, there exists f ∈ C(K) such that
and r(Λ) = λ.
2.2 Interlude: On the space M(K) 23
Proposition 2.55. Let F be a closed subset of a compact space K and b > 0. Then
the function
is open and contains µ. It remains to show that ϕb (ν) < c for every ν ∈ W .
Given a measure ν ∈ W , let Lν := {x ∈ F : ν({x}) ≥ b}. It follows from the
choice of W that Lν ⊂ U . Thus
Proposition 2.56. Let F be a closed subset of a compact space K. Then the function
X
ψ : µ 7→ µd (F ) := µ({x}), µ ∈ M1 (K),
x∈F
Proof. For n ∈ N, let ψn be the function ϕb from Proposition 2.55 for b = 1/n. Then
it is easy to check that ψn % ψ on M1 (K). As ψn are upper semicontinuous and
positive functions, the proof is finished.
Definition 2.57 (Fσ and Gδ faces). A face which is simultaneously an Fσ set is called
an Fσ face. Analogously, a Gδ face is a face which is a Gδ set.
24 2 Compact convex sets
Proposition 2.58. Let F be a closed subset of a compact space K. Then the set
(cf. Proposition 2.27) and Dirac measures are discrete, we have G ∩ ext M1 (K) =
∅.
ψ : µ 7→ µs (K), µ ∈ M1 (K),
where µs is the singular part of µ with respect to the measure ω. Then ψ is a limit of
a decreasing sequence of lower semicontinuous functions on M1 (K).
Proof. For n ∈ N, set
1
ψn (µ) := sup µ(G) : G ⊂ K open and ω(G) < , µ ∈ M1 (K).
n
Obviously, {ψn } is a decreasing sequence of lower semicontinuous functions. Recall
that, by Theorem A.85(b), the function
µ 7→ µ(G), µ ∈ M1 (K),
L := µ ∈ M1 ([0, 1]) : µ ⊥ λ ,
Proof. It is easy to check that L is convex, extremal and dense in M1 ([0, 1]).
Therefore all that needs to be proved is that L is a Gδ set. Let {ψn } be a sequence
of functions as in Proposition 2.60 for K := [0, 1] and ω := λ. The assertion then
follows from the following equalities
∞
\
L = µ ∈ M1 ([0, 1]) : µs ([0, 1]) = 1 = µ ∈ M1 ([0, 1]) : ψn (µ) = 1
n=1
∞ \
∞
\
1 1
= µ ∈ M ([0, 1]) : ψn (µ) > 1 − ,
k
n=1 k=1
Remark 2.62. In Exercise 2.118 we indicate another reasoning of the fact that L is a
Gδ set.
26 2 Compact convex sets
and
ψ : µ 7→ µs ([0, 1]), µ ∈ M1 ([0, 1]).
Then ϕ and ψ are bounded affine functions on M1 ([0, 1]) of the second Baire class
and there exists a probability measure Λ on M1 ([0, 1]) such that
Recall that convex extremal sets are called faces. Closed extremal sets occasionally
bear the name absorbent sets.
Definition 2.65 (Generated faces). It is easy to see that any intersection of faces of
X is again a face. Hence, given a set A ⊂ X, there exists the smallest face of X
containing A. It equals the intersection of all faces containing A and is denoted by
face A. Given x ∈ X, we will write simply face x instead of face {x} where no
confusion can arise.
Proposition 2.66. If F is a convex subset of X, then
[
face F = {(λF − (λ − 1)X) ∩ X : λ ≥ 1} .
y + (λ − 1)−1 (x − y) ∈ X.
is an Fσ set.
Proposition 2.68. Let F be a subset of X. Then the following assertions are equiva-
lent:
(i) F is extremal,
(ii) the characteristic function cF of F is convex,
(iii) F is a union of faces.
Proof. The equivalence of (i) and (ii) is clear from the definition. Since it is easy to
check that any union of extremal sets is extremal, we have (iii) =⇒ (i).
If (i) holds, then [
F = face x,
x∈F
Proof. The equivalence of (i), (ii) and (v) follows immediately; see Proposition 2.68.
Assertions (ii) and (iii) are obviously equivalent.
To see that (i) =⇒ (iv), suppose that F is a closed extremal set and µ ∈ M1 (X)
such that r(µ) ∈ F . If spt µ is not a subset of F , then there exists a closed convex
set C ⊂ X \ F such that µ(C) > 0. We have µ(C) < 1, since otherwise r(µ) =
r(µ|C ) ∈
/ F . Set
1 1
µ1 := µ|C and µ2 := µ| .
µ(C) 1 − µ(C) X\C
Now,
µ = µ(C)µ1 + (1 − µ(C))µ2 ,
hence
r(µ) = µ(C)r(µ1 ) + (1 − µ(C))r(µ2 ) ∈ F.
Since F is extremal, r(µ1 ), r(µ2 ) ∈ F , which is a contradiction.
2.3 Structures in convex sets 29
F := {C ⊂ F : x ∈ C, C convex}
where c ∈ C and λ1 , λ2 ∈ [0, 1]. It is clearly sufficient to assume that λ1 ∈ (0, 1).
(If λ1 = 0, e
c = c. If λ1 = 1, e c belongs to the segment joining a and b, and thus
c ∈ F by the extremality of F .) Moreover, we may assume that λ2 ≥ λ (otherwise,
e
1 − λ2 ≥ λ). With λ1 and λ2 chosen in this manner, we set
λ1 λ2 λ
α := and β := .
λ1 λ2 + λ(1 − λ1 ) λ1 λ2 + λ(1 − λ1 )
Since β > 0,
c + (1 − β)b = αz + (1 − α)c ∈ C ⊂ F
βe
c ∈ F.
and since F is extremal, we get e
Corollary 2.71. Let F be a closed convex subset of X. Then F is a face if and only if
for every measure µ ∈ M1 (X) with barycenter r(µ) in F , we have spt µ ⊂ F .
Proof. A closed convex set is a face if and only if it is extremal. Hence the assertion
follows immediately from Proposition 2.69, (i) ⇐⇒ (iv).
Proof. Let H ⊂ Y be an extremal set and αx1 + (1 − α)x2 ∈ ϕ−1 (H), where x1 ,
x2 ∈ X, α ∈ [0, 1]. Then αϕ(x1 ) + (1 − α)ϕ(x2 ) ∈ H, which yields ϕ(x1 ), ϕ(x2 ) ∈
H. Hence x1 , x2 ∈ ϕ−1 (H), concluding the proof of (a).
Since (b) is a straightforward consequence of (a), we proceed to the proof of (c).
Given a point y ∈ ext Y , the set ϕ−1 (y) is a closed extremal set by (a). By Proposi-
tion 2.20, it intersects ext X. Hence y ∈ ϕ(ext X), and we are done.
We will show that, for a universally measurable set F ⊂ X, the relations between
“F measure convex” (labelled as MC) and “F convex” (labelled as C) are as follows:
MC ⇒ C (2.74)
MC : C (2.81)
MC ⇐ C F is closed or open (2.74, 2.76)
MC ⇐ C F is resolvable (2.80)
MC ⇐ C dim E < ∞ (2.77)
MC ⇐ C F is a resolvable face (2.91)
MC : C F is Fσ face (2.82, 2.83)
MC : C F is Gδ face (2.84)
2.3 Structures in convex sets 31
Since ∞ ∞
P
n=1 αn = 1 − α0 < 1, there is a sequence {βn }n=1 of real numbers in (0, 1]
such that
∞
X αn
βn → 0 and = 1.
βn
n=1
Put
∞
[
L := K1 ∪ (βn Kn+1 + (1 − βn )K1 ) .
n=1
Then
s := sup f (L) = max s0 , sup βn sn + (1 − βn )s0
n∈N
= s0 + s − s0 = s.
Proof. Let G ⊂ X be an open convex set. By Theorem 2.75, it is enough to show that
co K ⊂ G whenever K ⊂ G is a compact set. So fix such K. For every x ∈ K there
exists a compact convex neighborhood Vx such that x ∈ Vx ⊂ G. By compactness,
the set K can be covered by finitely many compact convex sets Vx1 , . . . , Vxn . Then,
by Proposition 2.42,
T := {µ ∈ M+ (X) : µ ≤ λ, µ 6= 0}
and
µ
S := {r( ) : µ ∈ T },
kµk
then the closure of S equals co spt λ.
2.3 Structures in convex sets 33
We recall that resolvable sets are defined and their basic properties presented in
Section A.5. (In particular we note that any resolvable set in a compact space is
universally measurable by Proposition A.118.)
Lemma 2.79. Let F ⊂ X be a resolvable convex set and let λ ∈ M1 (X) be carried
by F . Then there exists a nonempty set G ⊂ F ∩ co spt λ which is open in co spt λ.
Proof. Let L := co spt λ. In order to find the required set G we note that L = F ∩ L
because the latter set is a closed convex set containing the support of λ. In particular,
F ∩ L is a dense resolvable set in L. Due to Proposition A.117(c), F ∩ L has a
nonempty interior (relative to L). Hence, the interior of F ∩ L is the sought set G.
and X
kλk = kµ0 k + kµn k.
n≥1
and P
c0 µ0 ck n≥k µn
ωk := · + · .
c0 + ck c0 c0 + ck ck
2.3 Structures in convex sets 35
k−1
µ0 X µn X µn
λ = c0 + kµn k + kµn k
c0 kµn k kµn k
n=1 n≥k
k−1
X µn
= kµn k + (c0 + ck ) · ωk ,
kµn k
n=1
and the last formula shows that λ is a finite convex combination of measures which
have their barycenters in F . Since F is convex, the barycenter of λ belongs to F as
well.
There are convex Borel subsets of a compact convex set which are not measure
convex. We present some of them.
Then B is a convex Borel set containing ext X which is not measure convex.
and
B := {xn } ∈ X : the set {n ∈ N : xn 6= 0} is finite .
Then X is a compact convex set and B is an Fσ face of X which is not measure
convex.
36 2 Compact convex sets
Proof. Since
∞
[
B= {xn } ∈ X : xn = 0 for all n ≥ j ,
j=1
Proposition 2.83. If
∞
[
µ ∈ M1 ([0, 1]) : spt µ ⊂ [ n1 , 1] ,
H :=
n=2
L := µ ∈ M1 ([0, 1]) : µ ⊥ λ .
We will show that, for a universally measurable set F ⊂ X, the relations between
“F measure extremal” (labelled as ME) and “F extremal” (labelled as E) are as fol-
lows:
ME ⇒ E (2.86)
ME ⇐ E F is closed or open (2.86, 2.88)
ME ⇐ E F is resolvable (2.92)
ME ⇐ E dim E < ∞ (2.89)
ME : E F is Fσ face (2.94, 2.96)
ME : E F is Gδ face (2.93, 2.95)
1 1
µ1 := µ|F and µ2 := µ| .
µ(F ) µ(X \ F ) X\F
Then
r(µ2 ) ∈ X \ F
and
r(µ) = µ(F )r(µ1 ) + µ(X \ F )r(µ2 ).
This is a contradiction since F is assumed to be extremal. Hence µ(X \ F ) = 0 and
F is measure extremal.
Since X \F is convex if F is extremal, Propositions 2.77, 2.74 and 2.76 yield using
Proposition 2.87 the following two corollaries.
Proof. Assume that z ∈ Int F and let x be any point of X \ F . By the continuity of
the vector operations in E, there is α ∈ (0, 1) so that y := αx + (1 − α)z ∈ Int F .
Since F is extremal, x ∈ F , which is a contradiction.
Proof. Let
G := µ ∈ M1 ([0, 1]) : µ = µc .
By Proposition 2.58, G is a Gδ face of M1 ([0, 1]) such that G ∩ ext M1 ([0, 1]) = ∅.
Let λ denote Lebesgue measure on [0, 1]. If Λ := ε] λ (cf. Example 2.53), then by
Proposition 2.54, r(Λ) = λ ∈ G whereas Λ(G) = 0 since Λ is carried by ε([0, 1]).
Whence, G is not measure extremal.
Proof. Let again λ denote Lebesgue measure on [0, 1] and Λ = ε] λ (see Example
2.53). If F := face λ is the face generated by λ, then by Corollary 2.67, F is an Fσ
face.
Assume that µ ∈ F ∩ ε([0, 1]). Hence, µ = εx for some x ∈ [0, 1] and by
Corollary 2.67, there exist ν ∈ M1 ([0, 1]) and α ∈ [0, 1) so that
λ = αν + (1 − α)εx .
Then
0 = λ({x}) = αν({x}) + (1 − α),
which implies that α = 1, a contradiction. Therefore, F ∩ ε([0, 1]) = ∅. We see that
Λ(F ) = 0 while r(Λ) = λ ∈ F . Therefore F is not measure extremal.
2.3 Structures in convex sets 39
Proposition 2.95. There exists a Gδ face which is neither measure convex nor measure
extremal.
Proof. We combine Examples 2.84 and 2.93. Let λ be Lebesgue measure on [0, 1]
and C the Cantor ternary set. Set G := G1 ∩ G2 where
G1 := µ ∈ M1 ([0, 1]) : µ ⊥ λ
and
G2 := µ ∈ M1 ([0, 1]) : µ|C is continuous .
It follows from Propositions 2.58 and 2.61 that G is a Gδ set in M1 ([0, 1]). Further,
as an intersection of faces, G is a face.
Let Λ denote again the image ε] λ of Lebesgue measure λ on [0, 1] (see Exam-
ple 2.53). Then r(Λ) = λ by Proposition 2.54 and the barycenter r(Λ) does not
belong to G, although
Proposition 2.96. There exists an Fσ face which is neither measure convex nor mea-
sure extremal.
Proof. Here we combine examples constructed in Propositions 2.83 and 2.94. Set
F := F1 ∩ F2 , where
∞
[
µ ∈ M1 ([0, 1]) : spt µ ⊂ [ n1 , 1]
F1 := and F2 := face λ
n=2
(here, λ is again Lebesgue measure on [0, 1] and face λ denotes the face generated by
λ). According to the aforementioned examples, F is an Fσ face in M1 ([0, 1]). Let
is not contained in F . Thus F is not measure convex and the proof is finished.
2.4 Exercises
Exercise 2.97. Let {Cα }α∈A be a collection of nonempty convex subsets of a vector
space W . Prove that
[ [∞ n n
X n
X
co Cα = x∈W: x= λ j x αj , λj = 1, λj ≥ 0, xαj ∈ Cαj
α∈A n=1 j=1 j=1
o
for j = 1, . . . , n and αj 6= αk for j 6= k .
Exercise 2.98. Prove that x is an extreme point of a convex set X if and only if
x
P= x1 = x2 = · · · = xm whenever x = λ1 x1 + λ2 x2 + · · · + λm xm with m ∈ N,
m
j=1 λj = 1, xj ∈ X and λj > 0 for each j = 1, 2, . . . , m.
Exercise 2.99. Let X 6= ∅ be a compact convex subset of Rd . Prove directly (do not
use the Minkowski theorem 2.11) that ext X 6= ∅.
1 1
diam X = |x − y| = (x1 − y) + (x2 − y)
2 2
1
< |x1 − y| + |x2 − y| ≤ diam X.
2
Another hint. Prove that any point z ∈ X having the property that |z| ≥ |x| for any
x ∈ X is an extreme point of X.
Set
I + := {j : αj ≥ 0} and I − := {j : αj < 0}
and
M1 := xj : j ∈ I + M2 := xj : j ∈ I − .
and
Then M1 ∩ M2 = ∅. Since X
λ := αj > 0,
j∈I +
we get
X αj X αj
co M1 3 xj = − xj ∈ co M2 .
+
λ −
λ
j∈I j∈I
satisfies that any subfamily with d + 1 elements has nonempty intersection. Hence the
inductive assumption yields that the latter, and, consequently, the former family has
nonempty intersection.
Assume now that the family K consisting of closed convex sets is infinite and a
set Z ∈ K is compact. By the first part, every finite subfamily of K has nonempty
intersection. Hence K, and consequently {Z ∩K : K ∈ K}, has the finite intersection
property. By compactness,
\ \
{Z ∩ K : K ∈ K} = {K : K ∈ K}
is nonempty.
Exercise 2.102. (a) Let G be an open subset of Rd . Prove that co G is an open set.
(b) If F is a closed subset of Rd , the convex hull co F need not be closed.
Exercise 2.103. (a) Let C be a compact convex subset of R2 . Prove that the set ext C
is closed.
(b) The set ext C need not be closed if C is a compact convex subset of Rd for d ≥ 3.
Exercise 2.104. Let C be a closed convex subset of Rd containing a line. Prove that
ext C = ∅. If C contains no line, then ext C 6= ∅.
Hint. If C contains a line L and x ∈ C, then C contains also a line passing through
x parallel to L.
If C contains no line then use induction on the dimension d. Find a boundary point
of C and follow the reasoning of the proof of the Minkowski theorem 2.11.
2.4 Exercises 43
Exercise 2.105 (Closed convex hulls). Let C be a subset of a locally convex space E.
Prove that the closed convex hull co C defined as
\
co C := {F : F is a convex closed subset of E, F ⊃ C}
Hint. For the proof of (a) consider K to be the convex hull of the union of two circles
Exercise 2.109. Let F be a closed subset of a compact convex set X such that ext X ⊂
F . Assume that for any x ∈ X there exists a unique measure µ ∈ M1 (F ) such that
r(µ) = x. Prove that ext X = F .
44 2 Compact convex sets
Exercise 2.110 (Proof of the Milman theorem 2.43). Verify the following indication
of an alternative proof of Theorem 2.43.
Exercise 2.111. Find an example of a compact convex set X in a locally convex space
E such that (E ∗ + R)|X 6= Ac (X).
Exercise 2.112. Prove that in any infinite-dimensional Banach space E, there exists a
compact convex set X such that (E ∗ + R)|X 6= Ac (X).
Exercise 2.113. Find a compact convex set X ⊂ R2 such that far X 6= exp X.
Hint. Let
X = {(x, y) ∈ R2 : |x|3 ≤ y ≤ 1, −1 ≤ x ≤ 1}.
Then the point (0, 0) is obviously exposed. However, (0, 0) ∈
/ far X as an easy geo-
metrical argument shows.
First hint. Let H be a Hilbert space, M be a dense proper subspace of H ∗ and let
Φ : H ∗ → H be the mapping assigning to each ϕ ∈ H ∗ a point y ∈ H such that
ϕ(h) = (h, y) for any h ∈ H. Let X := BH be equipped with the σ(H, M )-topology.
Since BH is w-compact and σ(H, M ) is weaker and Hausdorff, w = σ(H, M ) on
X. Choose x ∈ SH \ Φ(M ) and denote ψ := Φ−1 (x). Show that ψ is not σ(H, M )-
continuous. Further show that ψ|X ∈ Ac ((X, σ(H, M )) and that ψ exposes x. On the
other hand, no functional from (H, σ(H, M ))∗ = M exposes x. (This easily follows
from the fact that, given x, y ∈ SH , (x, y) = 1 if and only if y = x.)
Exercise 2.115. Keeping the notation of the first hint in Exercise 2.114, consider lo-
cally convex spaces E1 := (H, w) and E2 := (H, σ(H, M )). Prove that the compact
convex sets (BH , σ(H, M )) and (BH , w) are affinely homeomorphic (by the identity
mapping), and that the point x is exposed by a functional from E1∗ whereas it is not
exposed by a functional from E2∗ . (Compare with the proof of Theorem 2.51.)
Hint. For any x ∈ co K \ K there exists µ ∈ Mx (K) (see Proposition 2.39). Obvi-
ously, µ 6= εx , and thus x ∈
/ ext X by Theorem 2.40.
Exercise 2.117. Let F be a closed face of a compact convex set X and let U ⊂ X be
an open set containing F . Then F ∩ co(X \ U ) = ∅.
L := µ ∈ M1 (K) : µ ⊥ ν .
and [
Ln := L(n, G) : G ⊂ K open, ν(G) < 2−n .
T∞
Since each set L(n, G) is open in M1 (K) by Theorem A.85(b), the set n=1 Ln is a
Gδ set. The proof will be complete once we show that
∞
\
L= Ln .
n=1
∞
[
1 ≥ µ(G) = lim µ( Gn ) ≥ lim sup µ(Gm ) ≥ lim sup(1 − 2−m ) = 1,
m→∞ m→∞ m→∞
n=m
Exercise 2.121. Find an Fσ face that is not a countable union of closed faces.
Hint. Use the set F from Proposition 2.94. If F were a countable union of closed
faces, F would be measure extremal by Proposition 2.86. But this is not the case.
Exercise 2.122. If F is a closed face of X, then there exists a set A ⊂ ext X such that
F = co A. In particular, if F is a nonempty closed face of X, then F ∩ ext X 6= ∅
(cf. Theorem 2.20).
Exercise 2.123. Let X be a convex subset of a locally convex space E. Prove that the
boundary ∂X is an extremal subset of X.
Hint. If Int X = ∅, the assertion is obvious. Hence, assume that Int X 6= ∅ and
that ∂X is not extremal. In this case, there exist z ∈ ∂X, x ∈ X and y ∈ Int X so
that z = λx + (1 − λ)y for some λ ∈ (0, 1). Find a neighborhood V of 0 such that
y + V ⊂ Int X. For every ε > 0, there exists zε such that zε ∈ (z + εV ) \ X. Let
zε − λx
wε := .
1−λ
Show that we can choose ε small enough such that wε ∈ y + V . Since zε = λx +
(1 − λ)wε , this implies that zε ∈ X, which is a contradiction.
is convex.
Hint. The set D is obviously closed and extremal. Then use characterization (vi) of
Proposition 2.69.
Exercise 2.125. Find a continuous affine surjection of a compact convex set X onto a
compact convex set Y such that ext Y 6= ϕ(ext X).
Hint. Consider a triangle given as the convex hull of points (0, 0), (1, 0), ( 12 , 1) and
its projection onto the unit segment [0, 1].
48 2 Compact convex sets
G := µ ∈ M1 ([0, 1]) : µ = µc
of all continuous Radon measures on [0, 1] is measure convex (cf. also Proposition
2.93).
Hint. Let Ω ∈ M1 M1 ([0, 1]) , Ω(G) = 1 with r(Ω) = ω be given. For a bounded
Borel function f on [0, 1], define an affine function If on M1 ([0, 1]) by the formula
Now, use the Lebesgue dominated convergence theorem to show that equality (2.4)
holds for any bounded Borel function f on [0, 1]. Pick x ∈ [0, 1]. Then
Z
ω({x}) = Ic{x} (ω) = Ic{x} (µ) dΩ(µ) = 0.
G
and n fω o
A2 := : 0 ≤ f ≤ 1 Borel, ω(f ) > 0 .
ω(f )
Prove that
(a) face ω = A1 = A2 ,
(b) face ω = M1 ([0, 1]) if K = [0, 1] and ω is Lebesgue measure on [0, 1].
Hint. (a) Let µ ∈ face ω be given. Then there exist α ∈ (0, 1] and ν ∈ M1 (K) such
that ω = αµ + (1 − α)ν. Then 0 < αµ ≤ ω.
If µ ≤ ω, by the Radon–Nikodym theorem there exists a Borel measurable function
f such that µ = f ω. Then 0 ≤ f ≤ 1 ω-almost everywhere and ω(f ) = µ(K) > 0.
Hence
µ fω
=
µ(K) ω(f )
and µ ∈ A2 .
2.5 Notes and comments 49
fω
Finally, let µ = ω(f ) for some Borel function f with 0 ≤ f ≤ 1 and ω(f ) > 0. We
may assume that ω(f ) < 1. Then
fω (1 − f )ω
ω = ω(f ) + ω(1 − f ) ,
ω(f ) ω(1 − f )
and µ ∈ face ω.
(b) Assume that λ is Lebesgue measure on [0, 1]. Given a point x ∈ [0, 1], it is easy
to find a sequence of measures in face λ that converges to εx . Hence face λ is a convex
set containing all extreme points of M1 ([0, 1]), and thus face λ = M1 ([0, 1]).
In part, the material Subsections 2.3.B and 2.3.C concerning a more detailed study
of measure convex and measure extremal sets is taken from the paper [146] by
P. Dostál, J. Lukeš and J. Spurný. Theorem 2.75 and its proof are taken from
D. H. Fremlin and J. D. Pryce [185]. Alfsen’s example 2.81 is taken from [5], p. 130.
The proof of Proposition 2.92 can also be proved by means of a result of J. Saint
Raymond 10.75 without recourse to Proposition 2.80. Counterexamples contained in
Propositions 2.93 and 2.94 are just suitable modifications of the example by G. Cho-
quet in [106]. The examples of Propositions 2.95 and 2.96 partially use a construction
of H. v. Weizsäcker (see [463]). We also refer the reader to Lecture Notes [473] by
G. Winkler where a thorough investigation of measure convex sets is given.
The result of Exercise 2.103(a) is due to G. B. Price [377]. A characterization of
continuous affine functions on X (cf. Lemma 2.34 and Exercise 2.112) belonging to
E ∗ + R|X even for noncompact convex sets X is given in a paper [281] by M. Kraus.
Examples of Exercises 2.114 and 2.115 are due to M. Kraus and O. Kurka.
Chapter 3
Choquet theory of function spaces
This chapter lays the groundwork for the rest of the book by presenting the founda-
tions of the Choquet theory of function spaces. The central concept of a function
space is defined and its basic properties investigated in Section 3.1. We generalize
the framework of spaces of affine continuous functions on compact convex sets by
taking a subspace H of the space C(K) of continuous functions on a compact space
K such that H contains constants and separates points of K. Then we introduce
H-representing measures, H-affine and H-convex functions, and so on. A suitable
substitute for the set of extreme points turns out to be the Choquet boundary and we
show in Proposition 3.15 its nonemptiness and prove a minimum principle in Theo-
rem 3.16.
A crucial notion for obtaining integral representation theorems is the Choquet or-
dering introduced in Definition 3.19. This ordering somehow indicates how close to
the Choquet boundary a measure is situated. A key tool for handling function spaces
is Lemma 3.21 which serves as a substitute for the Hahn–Banach theorem. Several of
its applications are shown afterwards, along with Bauer’s characterization 3.24 of the
Choquet boundary.
These abstract results are then applied to a reexamination of Korovkin’s theorems;
Theorems 3.32, 3.34 and 3.36 are proved by means of the Choquet theory. After gen-
eralizing the concept of the barycenter mapping in Section 3.3, we turn our attention
to the Choquet representation theorem 3.45 for function spaces on metrizable compact
spaces. Our approach is to use the existence of a “strictly convex” function.
Next, Section 3.5 indicates how the Key lemma 3.21 enables us to prove analogues
of classical results on approximation of semicontinuous convex functions on compact
convex sets. More precisely, we show that semicontinuous H-convex functions can be
approximated by continuous H-convex functions (see Propositions 3.48 and 3.54). An
important corollary is the fact that the Choquet ordering of measures can be extended
to semicontinuous H-convex functions (see Proposition 3.56).
Measures maximal with respect to the Choquet ordering are investigated in Sec-
tion 3.6. First we prove Mokobodzki’s characterization in Theorem 3.58. Then we
show that the set of maximal measures is rich enough (see Theorem 3.65) and we fin-
ish the section with Theorem 3.70, which describes the space of boundary measures.
In order to prove the most important properties of maximal measures, namely, that
they are carried by any Baire set containing the Choquet boundary, we need some
kind of Fatou’s lemma (see Lemma 3.77). This task is accomplished in Section 3.7
by presenting the important Simons inequality 3.74 and a couple of its applications.
3.1 Function spaces 53
Then we can prove the integral representation theorem for nonmetrizable spaces in
Theorem 3.81.
The existence of representing measures “carried” by the Choquet boundary opens
the way for the proof of several variants of the minimum principle, as shown in Sec-
tion 3.9. The last section is devoted to a characterization of the fact that a pair of
measures µ, ν is related in the Choquet ordering. Theorem 3.92 shows that the impor-
tant notion of a dilation plays the key role here.
Examples 3.2. (a) Continuous functions. The space C(K) of all continuous functions
on a compact space K represents a simple example of a function space. Clearly, the
space C(K) separates points of K.
(b) Quadratic polynomials. The space P2 ([0, 1]) of all quadratic polynomials on the
interval [0, 1] is a further example of a function space. We considered this example in
Chapter 1.
(c) Convex case – affine functions. Let X be a compact convex subset of a locally
convex space E. The linear space Ac (X) consisting of all continuous affine functions
on X is a function space.
(d) Harmonic case – harmonic functions. Let U be a bounded open subset of the
Euclidean space Rd . The function space H(U ) consisting of all continuous functions
on U which are harmonic on U is another example of a function space.
More generally, we can consider a relatively compact open subset U of a Bauer
harmonic space (cf. Section A.8) and the function space H(U ), the linear subspace
of C(U ) of functions which are harmonic on U . We tacitly assume that constant
functions are harmonic and H(U ) separates points of U .
(e) If K is a compact subset of Rd , we define
[
H0 (K) := {H(U )|K : U is relatively compact open set, K ⊂ U }.
Generally, the function space H0 (K) is not a closed subspace of C(K) (see Exer-
cise 13.155).
(f) Further examples can be found in 3.47, 3.82, 3.83, 3.103, 3.106, 3.111, 3.119,
6.67, 6.76, 6.77, 6.94, 7.65, 8.11(a), 8.29, 8.80, 9.11, 9.56, 10.97, 10.98, 10.99 and
12.3.B.
54 3 Choquet theory of function spaces
Definition 3.3 (H-representing measures). Recall that M1 (K) denotes the set of all
probability Radon measures on K. We denote by Mx (H) the set of all H-represen-
ting measures for x ∈ K, that is,
Z
1
Mx (H) := {µ ∈ M (K) : f (x) = f dµ for any f ∈ H}.
K
Definition 3.4 (Choquet boundary). Define the Choquet boundary ChH (K) of a func-
tion space H as the set of those points x ∈ K for which the Dirac measure εx is the
only H-representing measure for x; that is,
Examples 3.5. We describe the Choquet boundary of our main examples from 3.2.
We postpone the proofs to later sections.
(a) Continuous functions. In the case when H = C(K) (K is a compact space), the
equality ChH (K) = K follows immediately from the definition.
(b) Quadratic polynomials. Let H = P2 ([0, 1]) be the function space of all quadratic
polynomials on the interval [0, 1]. Then ChH ([0, 1]) = [0, 1], as easily follows by
Proposition 3.7.
(c) Convex case – affine functions. If H is the linear space Ac (X) of all continuous
affine functions on a compact convex set X, then ChAc (X) (X) = ext X, by Bauer’s
characterization of ext X in Theorem 2.40.
(d) Harmonic case – harmonic functions. If H(U ) consists of all continuous func-
tions on U ⊂ Rd which are harmonic on U , then ChH(U ) U = ∂ reg U (cf. Theo-
rem 13.35).
In the general situation of an abstract Bauer harmonic space, the situation is more
delicate, cf. Theorem 13.41.
(e) The Choquet boundary of H0 (K) (see Example 3.2(e)) consists of stable points
of K (see Definition 13.4 and Theorem 13.35).
3.1 Function spaces 55
(f) Let
H := {f ∈ C([−1, 1]) : 2f (0) = f (−1) + f (1)}.
Then ChH ([−1, 1]) = [−1, 1] \ {0}.
Further examples of Choquet boundaries can be found in the examples of function
spaces mentioned in Examples 3.2(f).
Definition 3.6 (H-exposing functions and H-exposed points). Let x ∈ K. A function
h ∈ H such that 0 = h(x) < h(t) for any t ∈ K, t 6= x, is said to be an H-
exposing function for x. A point x ∈ K is called an H-exposed point if there exists
an H-exposing function for x.
The set of all H-exposed points of K will be denoted by expH (K).
Proposition 3.7. Any H-exposed point belongs to the Choquet boundary of H.
Proof. Let x ∈ K and let h ∈ H for which 0 = h(x) < h(t) for any t ∈ K \ {x}. If
µ ∈ Mx (H), then 0 = h(x) = µ(h). Hence spt µ ⊂ {x}, and therefore µ = εx . We
see that x ∈ ChH (K).
Definition 3.8 (H-affine, H-convex and H-concave functions). We define the family
A(H) of all H-affine functions as the family of all universally measurable functions
f : K → [−∞, ∞] such that µ(f ) exists for every µ ∈ Mx (H), x ∈ K, and the
following barycentric formula holds:
Z
f (x) = f dµ for each x ∈ K, µ ∈ Mx (H).
K
(b) The families W(H), S c (H), S usc (H) and S lsc (H) form min-stable convex cones.
(c) The space W(H) − W(H) is dense in C(K).
Proof. Since (a) is obvious, we proceed to the proof of (b). Let F be any of the
considered families. Obviously, F is a convex cone and the required topological
property is stable with respect to taking finite minima (for F = W(H) we use identity
n ^
^ k
f1 ∧ · · · ∧ fn + g1 ∧ · · · ∧ gk = (fi + gj ).)
i=1 j=1
For the proof of (c), we just use the lattice version of the Stone–Weierstrass theorem
from Proposition A.31.
Lemma 3.13. Let F be a closed H-extremal set and f ∈ S lsc (H). Then the set
H := {x ∈ F : f (x) = min f (F )}
is H-extremal.
Proposition 3.14. The family F of all closed H-extremal sets is stable under finite
unions and arbitrary intersection.
3.1 Function spaces 57
H ⊂ (K \ Fa1 ) ∪ · · · ∪ (K \ Fan ).
Since µ(K \ Faj ) = 0 for each j = 1, . . . , n, we get µ(H) = 0. From the inner
regularity of µ we obtain µ(K \ F ) = 0.
The proof of Proposition 3.15 follows along the lines to that of the Krein–Milman
theorem 2.22.
Proposition 3.15. The Choquet boundary ChH (K) intersects any nonempty closed
H-extremal set. In particular, the Choquet boundary ChH (K) is nonempty if K 6= ∅.
Proof. Let F be a nonempty closed H-extremal set. We partially order the family S
of all nonempty closed
T H-extremal subsets of F by the reverse inclusion. If Z is a
chain in S, then {C : C ∈ Z} is nonempty since it is the intersection of a down-
directed family of nonempty compact sets. Moreover, it is H-extremal according to
the previous Proposition 3.14. Zorn’s lemma now provides a maximal element H ∈
S. Assume that H contains two distinct points x and y. Since H is a function space,
there exists a function h ∈ H such that h(x) 6= h(y). According to Lemma 3.13, the
set
{z ∈ H : h(z) = min h(H)}
is a closed H-extremal set strictly contained in H, which contradicts the fact that H
is maximal. Hence, H = {x} for some x ∈ F . Since one-point H-extremal sets are
in ChH (K), it follows that x ∈ ChH (K).
Theorem 3.16 (Minimum principle for S lsc (H)). Let f ∈ S lsc (H) be a lower semi-
continuous H-concave function, f ≥ 0 on ChH (K). Then f ≥ 0 on K.
If x ∈ K and µ = εx , we define
F ∗
f (x) := Qεx ,F (f ) and F
f∗ (x) := −(F(−f )∗ (x)).
Lemma 3.18. Let µ ∈ M+ (K), F ⊃ ChH (K), and f be an upper bounded function
on F . Then
(a) Ff ∗ is an upper semicontinuous H-concave function,
(b) Qµ,F (f ) = µ(Ff ∗ ),
(c) if f is bounded, then −kf k ≤ Ff∗ ≤ Ff ∗ ≤ kf k,
(d) if f is upper semicontinuous on F and
f (x), x ∈ F,
g := lim sup f (y), x ∈ F \ F,
y→x,y∈F
Proof. The proof of (a) is obvious. To prove (b), let µ ∈ M+ (K) and an upper
bounded function f on F be given. By Lemma 3.11(b), the family
{k ∈ S c (H) : k ≥ f on F }
Proof. Given µ ∈ M+ (K), we assume first that f ∈ C(K). From Lemma 3.18(e)
we know that the mapping p : g 7→ Qµ,F (g) is a sublinear functional on C(F ).
We define a functional ϕ : span{f } → R as
ϕ(tf ) = tQµ,F (f ), t ∈ R.
Since
0 = Qµ,F (0) = Qµ,F (f − f ) ≤ Qµ,F (f ) + Qµ,F (−f ),
we get
ϕ(−f ) = −Qµ,F (f ) ≤ Qµ,F (−f ).
Hence ϕ ≤ p on span{f }.
The Hahn–Banach theorem provides a linear functional ν on C(F ) such that ν(f ) =
Qµ,F (f ) and ν ≤ p on C(F ). Since ν(g) ≤ p(g) ≤ 0 whenever g ∈ C(F ) and g ≤ 0,
we see that ν is, according to the Riesz representation theorem A.75, a Radon measure
on F . Let k ∈ S c (H). Then
for each ψ ∈ G.
Hence, using Theorem A.84 for verification of the equalities,
Proof. For the proof of (a), let f ∈ C(K) satisfy f∗ = f ∗ . Then f ∈ S usc (H) ∩
Klsc (H) = Ac (H). Conversely, if f ∈ Ac (H), then f∗ = f ∗ by definition.
For the proof of (b), let f be an upper bounded upper semicontinuous function on
ChH (K). Let g be the upper semicontinuous extension of f to ChH (K) defined as in
Lemma 3.18(d). By Lemma 3.22, for each x ∈ ChH (K) there exists µ ∈ Mx (H)
carried by ChH (K) such that
Proof. For the proof of (i) =⇒ (v), let x ∈ ChH (K) be given and f be an upper
semicontinuous function on K. By Lemma 3.21, there exists a measure µ ∈ Mx (H)
such that f ∗ (x) = µ(f ). Then µ = εx , and thus
Since (v) =⇒ (iv) =⇒ (iii) =⇒ (ii) are obvious, we close the chain of implications
by proving (ii) =⇒ (i). To this end, let a point x ∈ K satisfy f ∗ (x) = f (x) for each
f ∈ −W(H). Then for each µ ∈ Mx (H) and f ∈ −W(H),
(a) Then
F ∗
f = inf {g ∈ H : g ≥ f on F }
= inf {g ∈ Ac (H) : g ≥ f on F }
= inf {g ∈ W(H) : g ≥ f on F }
= inf {g ∈ S usc (H) : g ≥ f on F } .
(b) If f is upper semicontinuous, then
F ∗
f = inf k ∈ S lsc (H) : k ≥ f on F .
Proof. For the proof of (a), we first use Lemma 3.21 to conclude that
inf {g ∈ S usc (H) : g ≥ f on F } = inf {g ∈ S c (H) : g ≥ f on F } . (3.1)
Indeed, given g ∈ S usc (H) with g ≥ f on F , x ∈ K and ε > 0, by Lemma 3.21 there
exists a measure µ ∈ Mx (H) ∩ M+ (F ) such that Fg ∗ (x) = µ(g) ≤ g(x). Thus there
exists a function g 0 ∈ S c (H) with g 0 ≥ g on F such that
g 0 (x) ≤ g(x) + ε.
Hence (3.1) follows.
Further, we fix a point x ∈ K and define p : C(F ) → R as
p(g) = inf {g(x) : g ∈ H, g ≥ f on F } .
Then p is a sublinear functional on C(F ) and we may follow the beginning of the proof
of Lemma 3.21 to conclude that for any function g ∈ C(F ), there exists a measure µ
in Mx (H) ∩ M+ (F ) such that µ(g) = p(g).
Thus for every k ∈ S c (H),
p(k) = µ(k) ≤ k(x),
so that k(x) = inf{h(x) : h ∈ H, h ≥ f on F }. Hence
inf{h ∈ H : h ≥ f on F } = inf{k ∈ S c (H) : k ≥ f on F }.
Obviously,
inf{k ∈ S c (H) : k ≥ f on F } ≤ inf{k ∈ W(H) : k ≥ f on F }
≤ inf{k ∈ H : k ≥ f on F }.
Thus
F ∗
f ≤ inf{k ∈ S usc (H) : k ≥ f on F } ≤ inf{h ∈ Ac (H) : h ≥ f on F }
≤ inf{h ∈ H : h ≥ f on F } = inf{k ∈ W(H) : k ≥ f on F }
= inf{k ∈ S c (H) : k ≥ f on F }
= Ff ∗ .
3.1 Function spaces 63
Proof. Given a function k ∈ S c (H) and ε > 0, let h ∈ H be such that k ≤ h and
h(x) ≤ k(x) + ε (use Proposition 3.25(a)). Then
Hence εx ≺ µ, as needed.
Proof. Since the family S c (H) is min-stable and closed in C(K) and Ac (H) =
S c (H) ∩ (−S c (H)), necessity immediately follows.
As for the sufficiency, we show that S c (H) ⊂ W. Then
s = s∗ = inf {ϕ ∈ H : ϕ ≥ s} ≥ inf {ϕ ∈ W : ϕ ≥ s} ≥ s.
Kor(H) = Ac (H).
Proof. Pick f ∈ Ac (H) and ε > 0. By Proposition 3.25(a) and Corollary 3.23, for
any x ∈ K, there exists a function hx ∈ H and a neighborhood Ux of x such that
Ln f ≤ Ln hj ≤ hj + ε, j = 1, . . . , k,
and therefore
Ln f ≤ h1 ∧ · · · ∧ hk + ε = h + ε.
Given x ∈ K, we see that
Analogously, using lower envelopes we can prove the existence of n1 ∈ N such that
and define
gn (t) := 1 − nρ(x, t) ∨ 0, t ∈ K.
Then gn ∈ C(K) and
0 ≤ gn ≤ 1, gn (x) = 1 and gn = 0 on K \ Bn .
If
Ln : ϕ 7→ gn µ(ϕ) + (1 − gn )ϕ, ϕ ∈ C(K),
then Ln : C(K) → C(K) is a positive linear operator. Let h ∈ H and t ∈ K. Since
it follows that
kLn h − hk ≤ sup {|h(x) − h(t)| : t ∈ Bn } .
Since the function h is continuous at x, we get kLn h − hk → 0. We see that
the sequence {Ln } is H-admissible, and therefore kLn f − f k → 0. In particular,
Ln f (x) → f (x). But Ln f (x) = µ(f ), and the proof is complete.
Remark 3.30. Notice that in the proof of the inclusion Ac (H) ⊂ Kor(H) we did not
use linearity of Ln , while the metrizability condition was used only in the course of
the proof that Kor(H) ⊂ Ac (H).
Now, we are in a position to answer our query from Chapter 1, where we asked
under which conditions the equality Kor(H) = C(K) holds.
Proof. The preceding Proposition 3.29 shows that Kor(H) = Ac (H), and therefore
we only need to examine when Ac (H) = C(K). According to Bauer’s character-
ization 3.24 of the Choquet boundary, we know that x ∈ ChH (K) if and only if
f∗ (x) = f ∗ (x) for any f ∈ C(K). Hence, ChH (K) = K if and only if f∗ = f ∗ on
K for any f ∈ C(K); that is, if and only if Ac (H) = C(K), by Corollary 3.23.
3.322 (First Korovkin theorem revisited). Let H be the linear span of func-
Theorem
tions 1, id, id on [a, b]. Then Kor(H) = C([a, b]).
Proof. Let t ∈ [a, b]. Since x 7→ (t − x)2 , x ∈ [a, b], is an H-exposing function for t,
we may apply Proposition 3.7 to see that ChH ([a, b]) = [a, b]. Proposition 3.31 then
shows that Kor(H) = C([a, b]).
66 3 Choquet theory of function spaces
If P := {1, ϕ1 , . . . , ϕd }, then
Kor(P) = C(S).
Proof. For the function space H = Lin{1, ϕ1 , . . . , ϕd } we have ChH (S) = S. In-
deed, given w ∈ S, it is straightforward to check that the function
x 7→ 1 − x · w, x ∈ S,
Corollary 3.34 (Second Korovkin theorem). Let H be the linear span of {1, cos, sin}
on [0, 2π]. Then the Korovkin closure Kor(H) equals the Banach space of all contin-
uous functions f on [0, 2π] with f (0) = f (2π) (equipped with the sup-norm).
Proof. The circle S := {x ∈ R2 : |x| = 1} can be topologically identified with the
interval [0, 2π] after having identified the points 0 and 2π via the mapping ω : t 7→
(cos t, sin t), t ∈ [0, 2π]. Then the functions cos, sin correspond to the functions ϕ1 ,
ϕ2 from Theorem 3.33.
Remark 3.35. There is also a direct proof of this corollary following the lines of the
proof of Theorem 1.2, considering the function
t−x 1
x 7→ sin2 = (1 − cos x cos t − sin x sin t), x ∈ [0, 2π],
2 2
instead of the function x 7→ (x − t)2 .
Theorem 3.36 (Third Korovkin theorem). Let H be a linear span of {1, ϕ} on [a, b],
where ϕ is a continuous function on [a, b]. Then Kor(H) 6= C([a, b]).
Proof. The assertion is obvious if the function ϕ does not separate points
of [a, b].
Otherwise, there exists x ∈ (a, b) such that ϕ(x) = 21 ϕ(a) + ϕ(b) . Then the
measure 12 (εa + εb ) represents x, and therefore x does not belong to ChH ([a, b]). In
view of Proposition 3.31, Kor(H) 6= C([a, b]).
Remark 3.38. Note that in the “convex case”, M(Ac (X)) = M1 (X) (see Theo-
rem 2.29).
If H = C([0, 1]), then M(H) = {εx : x ∈ [0, 1]} 6= M1 ([0, 1]). See also
Exercise 3.106.
r : µ ∈ Mx (H) 7→ x, µ ∈ M(H).
Proof. Since the Banach space C(K) is separable, there exists a countable set {hn :
n ∈ N} that is dense in {h ∈ H : 0 ≤ h ≤ 1}.
We claim that the function h2n is H-convex for any n ∈ N. To this end, pick x ∈ K
and µ ∈ Mx (H). Then
Z 2 Z Z
h2n (x) = hn dµ ≤ 1 dµ h2n dµ = µ(h2n ). (3.2)
K K K
Proposition 3.43. Let H be a function space on a compact space K and let there exist
a continuous strictly H-convex function h on K. Then
Let x ∈ K \ ChH (K). There exists a measure µ ∈ Mx (H), µ 6= εx . Then, using the
strict H-convexity of h and Corollary 3.22, we have h(x) < µ(h) ≤ h∗ (x).
3.4 The Choquet representation theorem 69
Thus, we have
∞
∗
\
∗ 1
ChH (K) = {x ∈ K : h(x) = h (x)} = x ∈ K : h (x) − h(x) < .
n
n=1
for each n ∈ N. Thus ChH (K), being the intersection of these sets, is a Gδ set.
Remarks 3.44. (a) In the case of the “convex” function space Ac (X), there is an
alternative proof of Theorem 3.42, cf. Exercise 3.107 and Remark 3.9.
(b) If a compact convex set X admits a continuous strictly convex function, then X
is metrizable (see Exercise 4.40). This assertion is no longer true for general function
spaces; see Example 3.47.
Remark 3.46. By Theorem 3.42, any metrizable compact set admits a continuous
strictly H-convex function. Therefore, Proposition 3.43 and Theorem 3.45 hold under
the assumption that the compact space K is metrizable. On the other hand, there are
function spaces on nonmetrizable compact spaces which admit a continuous strictly
H-convex function. As a trivial example consider a nonmetrizable compact space K
and H = C(K). In this example, any continuous function on K is strictly H-convex.
We present the following slightly less trivial example.
70 3 Choquet theory of function spaces
K := {g ∈ S c (H) : g ≥ f on K}
and
L := {g ∈ S c (H) : g > f on K} ,
then K and L are down-directed families and f = inf K = inf L.
Proof. Since the cone S c (H) is min-stable (cf. Proposition 3.11), the sets K and L
are down-directed.
It suffices to show that f = inf L. To this end, choose x ∈ K and numbers
f (x) < c < d. By Proposition 3.25(a), f ∗ = f and there exists h ∈ H such that
Now, if g := h+d−c, then g ∈ H, g(x) < d and g > f . Since d is chosen arbitrarily,
f = inf L.
Proof. By Proposition 3.25(a), f ∗ = f . With the aid of Proposition 3.48, for any
x ∈ K we select a function gx ∈ S c (H) such that
Now appeal to the upper semicontinuity of f and the lower semicontinuity of g to find
an open neighborhood Ux of x such that
K = Ux1 ∪ · · · ∪ Uxn .
Then the function k := gx1 ∧ · · · ∧ gxn belongs to S c (H) and it is a routine argument
to show that f < k < g on K. Indeed, given y ∈ K, there exists j ∈ {1, . . . , n} such
that y ∈ Uxj . Then
f (y) < k(y) ≤ gxj (y) < g(y).
g ≤ uε ≤ vε ≤ f on K
and
vε − uε < ε on K.
This can be done quite easily. Namely, by Proposition 3.51, there exists a function
k ∈ S c (H) such that g < k < f + ε. It is easy to check that the functions
uε := (k − ε) ∨ g and vε := f ∧ k
do the job.
Now, using the above construction we can construct inductively two sequences
{un } and {vn } so that, for each n ∈ N, un is upper semicontinuous, vn ∈ S lsc (H),
g ≤ un ≤ un+1 ≤ vn+1 ≤ vn ≤ f
and
1
vn − un < .
2n
Then the sequence {vn } converges uniformly to an H-concave continuous function
k. The inequalities
1 1
g ≤ vn ≤ un + n ≤ f + n
2 2
ensure that g ≤ k ≤ f .
S := {g ∈ S c (H) : g < f on K}
and
T := {g ∈ S c (H) : g ≤ f on K} ,
then S and T are up-directed families and f = sup S = sup T.
Proof. It suffices to establish that f = sup S. Since
Proposition 3.51 asserts the existence of g ∈ S c (H) such that g1 ∨ g2 ≤ (g1 ∨ g2 )∗ <
g < f.
All that is left to be shown is that T is up-directed. To finish the proof, consider
g1 , g2 ∈ T. By Proposition 3.53, there exists g ∈ S c (H) such that g1 ∨g2 ≤ g ≤ f .
Proposition 3.55. Let f ∈ S c (H) and ε > 0. Then there exists h ∈ W(H) such that
f ≤ h ≤ f + ε.
Proof. By Proposition 3.25,
f = inf {h ∈ W(H) : h ≥ f } ,
Proposition 3.56. Let H be a function space on K. Then the following assertions are
equivalent for µ, ν ∈ M+ (K):
(i) µ ≺ ν,
(ii) µ(f ) ≤ ν(f ) for any f ∈ −W(H),
(iii) µ(f ) ≤ ν(f ) for any f ∈ Kusc (H),
(iv) µ(f ) ≤ ν(f ) for any f ∈ Klsc (H).
Proof. The implications (i) =⇒ (iii) and (i) =⇒ (iv) follow from Proposition 3.54,
Corollary 3.49 and Theorem A.84. Obviously, (i) =⇒ (ii), (iii) =⇒ (i) and (iv)
=⇒ (i). By Proposition 3.55, (ii) =⇒ (i).
Hence µ(f ) = ν(f ). Since the space W(H) − W(H) is uniformly dense in C(K),
we conclude that µ = ν. We see that µ is maximal.
Proof. The assertion follows immediately from Mokobodzki’s maximality test 3.58.
Corollary 3.60. If µ ∈ M+ (K) and µ(F ) = 0 for any closed F ⊂ K \ ChH (K),
then µ is a maximal.
we have Z Z
µ(v) = v dµ + v dµ = µ(v ∗ ),
A B
where B := {x ∈ K : v ∗ (x) = v(x)}. Thus (iii) =⇒ (i) by Mokobodzki’s maxi-
mality test 3.58.
Remark 3.63. We know from Corollary 3.62 that in the metrizable case any maximal
measure is carried by the Choquet boundary, which is a Borel set. This assertion
is no longer valid for nonmetrizable situations. Section 3.8 focuses on the Choquet
theory in the nonmetrizable setting. Nevertheless, in this case any maximal measure is
carried at least by the closure of the Choquet boundary; see the next Proposition 3.64.
On the other hand, a measure carried by ChH (K) need not be maximal. Indeed,
consider a function space H on a metrizable compact space K with a nonclosed Cho-
quet boundary (cf. Exercise 2.103(b)). For a point x ∈ ChH (K) \ ChH (K), the
Dirac measure at x is clearly carried by ChH (K) and it fails to be maximal in view of
Corollary 3.62.
A more striking example of this phenomenon is the Poulsen simplex (see Subsec-
tion 12.3.A), where the set of extreme points is dense in the whole set.
Proposition 3.64. Any maximal measure is carried by ChH (K).
Proof. Let µ be a maximal measure on K and C a compact subset of K \ ChH (K).
Urysohn’s lemma ensures the existence of a function f ∈ C(K), 0 ≤ f ≤ 1 on K,
such that
f = 0 on ChH (K) and f = 1 on C.
76 3 Choquet theory of function spaces
Theorem 3.65. For every measure µ ∈ M+ (K) there exists a maximal measure λ
such that µ ≺ λ.
Proof. Set M := {ν ∈ M+ (K) : µ ≺ ν}. Since for any ν ∈ M we have kνk = kµk,
it follows that M is contained in the compact set {η ∈ M+ (K) : kηk = kµk}. The
assertion will follow from Zorn’s lemma once it is shown that every chain in M has
an upper bound. Let R be such a chain. Then R is a net in a compact set (directed
by ≺) and hence there exists a subnet J of R which converges to some element
ν0 ∈ M+ (K). Since M is closed, ν0 ∈ M .
It remains to show that ν ≺ ν0 for any ν ∈ R. To this end, fix ν ∈ R, f ∈ Kc (H)
and ε > 0. There is η ∈ J such that ν ≺ η and |ν0 (f ) − η(f )| < ε. Then
ν0 (f ) ≥ η(f )−ε ≥ ν(f )−ε. Therefore ν ≺ ν0 since ε is an arbitrary positive number.
We conclude by Zorn’s lemma: the ordered set M has a maximal element.
Proof. Let µ ∈ M+ (K) be maximal and x ∈ K \ ChH (K). By Theorem 3.24, there
exists a function f ∈ −W(H) such that f ∗ (x) > f (x). Since µ({y ∈ K : f ∗ (y) >
f (y)}) = 0 (see Corollary 3.59), µ({x}) = 0.
Proof. By Theorem 3.58, (i) =⇒ (ii). Assume (ii), choose an upper semicontinuous
function f on K and ε > 0. Using Theorem 3.48, we can find a function g ∈ S c (H)
such that g > f ∗ and
Since f is upper semicontinuous, by Proposition A.50 and Theorem A.84 there exists
h ∈ C(K) such that h > f and
It follows,
and therefore,
−2ε = µ(ϕ∗ ) − µ(ϕ) − 2ε ≤ µ(f ∗ ) − µ(f )
≤ µ(ϕ∗ ) − µ(ϕ) + 2ε = 2ε.
We see that µ(f ) = µ(f ∗ ); hence (ii) =⇒ (iii).
It remains to show that (iii) =⇒ (i). We have to verify that |µ| = µ+ + µ− is
carried by the set Af := {x ∈ K : f ∗ (x) = f (x)} for any f ∈ C(K). To this end,
let f ∈ C(K) be given. Let again µ = µ+ − µ− be the decomposition of µ into its
positive and negative part, µ+ be carried by a set P and µ− be carried by a set N ,
where P ∩ N = ∅. It suffices to prove that µ+ (L) = 0 whenever L ⊂ P \ Af is a
compact set. So, fix such a set L and set
(
f on K \ L,
g := ∗
f on L.
We recall that the space M(K), as a dual space to the ordered Banach space C(K),
is an ordered Banach space. Moreover, it is a lattice. Theorem 3.70 below shows that
Mbnd (H) is a lattice with the structure inherited from M(K).
Hence ν(g) = ν(g ∗ ) for any g ∈ C(K), and thus ν is maximal by Theorem 3.58.
Theorem 3.70. The set Mbnd (H) is a norm closed subspace of M(K). Moreover, it
is a lattice considered with the order inherited from M(K).
Proof. We first show that Mbnd (H) is a norm closed subspace of M(K). To this
end, let µ1 , µ2 ∈ Mbnd (H) and g ∈ C(K) be given. Since |µ1 | + |µ2 | is carried by
the set {x ∈ K : g ∗ (x) = g(x)}, it is easy to see that |µ1 + µ2 | is carried by the set
{x ∈ K : g ∗ (x) = g(x)}. Thus µ1 + µ2 is a boundary measure by Theorem 3.58.
Analogously we show that cµ ∈ Mbnd (H) for any c ∈ R and µ ∈ Mbnd (H). If {µn }
is a sequence of boundary measures norm converging to µ ∈ M(K), we use again
Theorem 3.58 to show that µ is a boundary measure.
To verify the lattice property we use Proposition A.15. Obviously, any measure
µ ∈ Mbnd (H) can be written as a difference of positive maximal measures, and
hence
Mbnd (H) = (Mbnd (H))+ − (Mbnd (H))+ = Mmax (H) − Mmax (H).
Let µ1 , µ2 ∈ Mmax (H) be given. Since M(K) is a lattice, there exists a measure
ν ∈ M+ (K) such that ν = µ1 ∧ µ2 (with respect to the ordering of M(K)). Since
ν ≤ µ1 + µ2 , ν is absolutely continuous with respect to the maximal measure µ1 + µ2 .
By Proposition 3.69, ν ∈ Mmax (H) as well. Hence ν is a supremum of {µ1 , µ2 } in
Mbnd (H) and Mbnd (H) is a lattice.
Definition 3.73 (Boundary). Let K be a set and let F be a subset of `∞ (K). We say
that a nonempty set B in K is a boundary for F if every function from F attains its
maximum at a point of B.
Proof. Let {fn } be a bounded sequence on K. We start theP proof by fixing ε > 0
and a sequence {ρi } of strictly positive numbers such that ∞ i=1 ρP
i = 1. We set
σk := i=k ρi and find strictly positive numbers εk , k ∈ N, such that ∞
P∞ εk
k=1 σk+1 < ε.
Let Ck := coσ ({fn : n ≥ k}), k ∈ N.
For a bounded function f on K, we write p(f ) := sup f (B) and q(f ) := sup f (K).
We inductively find functions gi ∈ Ci , i ∈ N, such that
• p(ρ1 g1 ) ≤ infh∈C1 p(ρ1 h) + ε1 ,
Pk Pk−1
i=1 ρi gi ) ≤ infh∈Ck p( i=1 ρi gi + ρk h) + εk , k ≥ 2.
• p(
We set g := ∞
P
i=1 ρi gi .
P0
Proof of the claim. Let n ∈ N be given. Throughout the proof we interpret i=1 as
0. We pick a number k ∈ {1, . . . , n}. Since
∞ k−1
1 X X
ρi gi − ρi gi ∈ Ck ,
σk
i=1 i=1
we get
k k−1 ∞ k−1
!
X X 1 X X
p ρi g i ≤ p ρi gi + ρk ( ρi gi − ρi gi ) + εk
σk
i=1 i=1 i=1 i=1
k−1
ρk ρk X
≤ p(g) + 1 − p ρi gi + εk .
σk σk
i=1
Hence
k k−1
σk+1 X σk+1 X
p(g) 1 − ≥p ρi gi − p ρi g i − ε k . (3.4)
σk σk
i=1 i=1
80 3 Choquet theory of function spaces
We divide (3.4) by σk+1 and sum these inequalities for k = 1, . . . , n. Using p(0) = 0,
we have
n n
1 1 X X εk
p(g) −1 ≥ p ρk gk − .
σn+1 σn+1 σk+1
k=1 k=1
By the choice of the numbers εk , we get
n
X
p(g) ≥ p ρk gk + σn+1 p(g) − ε .
k=1
We apply the assumption for the function g, and find b ∈ B such that g(b) =
p(g) = q(g). Then, for n ∈ N, the claim yields
∞
X n
X
ρi gi (b) ≥ ρi gi (b) + σn+1 (q(g) − ε).
i=1 i=1
1 P∞
Since σn+1 ( i=n+1 ρi gi ) ∈ Cn+1 , we get
∞
1 X
q(g) − ε ≤ ρi gi (b)
σn+1
i=n+1
Corollary 3.75 (Simons lemma). Let {fn } and B ⊂ K be as in Lemma 3.74. Then
Proof. Let {fn } be as in Lemma 3.74. We again adopt the notation p(f ) = sup f (B)
and q(f ) = sup f (K) for each f ∈ `∞ (K). Assume that p(lim supn→∞ fn ) < c <
d < q(lim supn→∞ fn ). Let x ∈ K satisfy lim supn→∞ fn (x) > d. Let n1 < n2 <
. . . be natural numbers such that fnk (x) > d, k ∈ N. Clearly, B is a boundary for
coσ {fnk : k ∈ N}. By Lemma 3.74, there exists f ∈ coσ {fnk : k ∈ N} such that
Proof. (a) Let {fn } and B ⊂ K be as in the statement of the corollary. Note that B is
a boundary for coσ (span A). We again write p(f ) := sup f (B) for f ∈ `∞ (K). Let
{fn (b)} be convergent for each b ∈ B, and assume that there exists x ∈ K such that
{fn (x)} is not Cauchy. By choosing a subsequence if necessary, we may assume that
there exists a strictly increasing sequence {nk } of natural numbers and c > 0, such
that
c < fn2k (x) − fn2k−1 (x), k ∈ N.
Then the functions gk := fn2k − fn2k−1 , k ∈ N, satisfy p(lim supk→∞ gk ) = 0, yet
0 < c ≤ lim supk→∞ gk (x). This contradicts Corollary 3.75 and proves (a).
(b) Let fn → g on B. Then, for each x ∈ K, Corollary 3.75 gives
Proof. Let {fn } be as in the statement of the lemma. Without loss of generality we
may assume that {fn } is a bounded sequence, sup{kfn k : n ∈ N} ≤ C (otherwise we
would consider the functions −1∨fn ). Fix x ∈ K and ε > 0. Using Proposition 3.48,
we find functions kn ∈ Kc (H) such that kn ≤ fn and kn (x) ≥ fn (x) − ε. Without
loss of generality we may assume that sup{kkn k : n ∈ N} ≤ C. By Theorem 3.16,
each function in coσ ({kn : n ∈ N}), as an H-convex continuous function, attains its
maximum on ChH (K). Since
Proposition 3.80. Assume that ChH (K) is a Lindelöf set and µ ∈ M+ (K). Then the
following assertions are equivalent:
(i) µ∗ K \ ChH (K) = 0,
(ii) µ is maximal.
3.8 The Bishop–de Leeuw theorem 83
Proof. Since (ii) =⇒ (i) is proved in Theorem 3.79, it remains to show the reverse
implication. Let µ ∈ M+ (K) be a measure satisfying µ∗ (K \ ChH (K)) = 0 and let
f ∈ Kc (H) be fixed. By Theorem 3.24, {x ∈ K : f (x) < f ∗ (x)} is disjoint from
ChH (K). Since µ({x ∈ K : f (x) < f ∗ (x)}) = 0, by assumption, Corollary 3.59
concludes the proof.
Theorem 3.81 (Bishop–de Leeuw). For any point x ∈ K, there exists a measure
µ ∈ Mx (H) such that µ(A) = 1 for every K-analytic set A ⊃ ChH (K).
Example 3.82. There exist a function space H on a compact space K and a maximal
measure on K carried by a compact set disjoint from ChH (K).
Proof. Let K := [0, 1]×{−1, 0, 1} be equipped with the topology defined as follows:
points of [0, 1] × {−1, 1} are open (hence [0, 1] × {−1, 1} is discrete), and the base
of neighborhoods of (x, 0) ∈ [0, 1] × {0} are of type (U × {−1, 0, 1}) \ F , where
U ⊂ [0, 1] is a Euclidean neighborhood of x, and F is finite. If f is a continuous
function on this compact space K, we notice that, for each δ > 0, the set
Claim. If f ∈ Kc (H) and δ > 0, then {x ∈ [0, 1] : f ∗ (x, 0) − f (x, 0) > δ} is finite.
Proof of the claim. Let f ∈ Kc (H) and δ > 0 be given. If x ∈ [0, 1], by Lemma 3.21,
there exists a measure µ ∈ M(x,0) (H) such that f ∗ (x, 0) = µ(f ). Since f is H-
convex, µ(f ) ≤ ν(f ) ≤ f ∗ (x, 0) for any ν ∈ Mx (H) with µ ≺ ν. Hence we may
assume that µ is maximal. By Proposition 3.66, µ({(x, 0)}) = 0. As shown above,
µ = 21 (ε(x,−1) + ε(x,1) ). Thus
1
f ∗ (x, 0) = (f (x, −1) + f (x, 1)), x ∈ [0, 1].
2
By (3.5), the proof of the claim is finished.
84 3 Choquet theory of function spaces
Thus by Corollary 3.59, µ is maximal. On the other hand, µ is carried by the compact
set [0, 1] × {0}, which is disjoint from ChH (K).
Example 3.83. There exists a function space H on a compact space K and a measure
on K carried by any Baire set containing ChH (K) which is not maximal.
Proof. Let H be the function space from Example 3.47. Then ChH (K) = K \ {d∞ }
and εd∞ is not a maximal measure. On the other hand, if A ⊃ ChH (K) is a Baire set,
then A = K.
Indeed, any Baire set in K is Lindelöf (see Theorem A.111(i) and (d)) and hence
K \ {d∞ } is not a Baire set. Otherwise, as an open set it would be an Fσ set, which
it obviously is not.
Hence εd∞ (A) = 1 for each Baire set A containing ChH (K).
Theorem 3.85 (Minimum principle for S usc (H)). Let f ∈ S usc (H) be an upper se-
micontinuous H-concave function. Then f satisfies the minimum principle.
Theorem 3.86 (Minimum principle for Baire H-concave functions). Let f be a Baire
H-concave function. Then f satisfies the minimum principle.
3.9 Minimum principles 85
Proof. Assume that f ≥ 0 on ChH (K) and f (x) < 0 for some x ∈ K. The set
F := {y ∈ K : f (y) ≤ f (x)}
is a nonempty Baire subset of K, disjoint from ChH (K). Let µ ∈ Mx (H) be a max-
imal measure. By the Choquet–Bishop–de Leeuw theorem 3.81, µ(F ) = 0. Hence
Z Z
f (x) ≥ µ(f ) = f dµ > f (x) dµ(y) = f (x),
K\F K\F
which is a contradiction.
Proposition 3.87 (Strict minimum principles). Let f be a lower bounded H-concave
function on K. If f is lower or upper semicontinuous, or if f is a Baire function on
K, then f satisfies the strict minimum principle.
Proof. Let f ∈ S lsc (H), f > 0 on ChH (K). By Theorem 3.16, f ≥ 0 on K. If Z :=
{x ∈ K : f (x) = 0}, then Z is a closed H-extremal set. Since Z ∩ ChH (K) = ∅, it
follows from Proposition 3.15 that Z = ∅.
Assume now that f ∈ S(H) is a Baire function. The set Z := {x ∈ K : f (x) = 0}
is a Baire set disjoint from ChH (K). If Z contains a point x, choose a maximal
measure µ ∈ Mx (H). Then
Z
0 = f (x) ≥ µ(f ) = f dµ > 0,
K\Z
a contradiction.
Finally, let f ∈ S usc (H) be upper semicontinuous, and let f (x) = 0 for some
x ∈ K. By Proposition 3.48, there exists a sequence {fn } in S c (H) such that fn ≥
fn+1 ≥ f for each n ∈ N and fn (x) → 0. If h := inf {fn : n ∈ N}, then h is a Baire
H-concave function such that h > 0 on ChH (K). By the strict minimum principle
for H-concave Baire functions which we have just proved, h > 0 on K. However,
h(x) = 0, a contradiction that finishes the proof.
Proposition 3.88 (Minimum principle for semicontinuous functions). Assume that H
is a function space on a compact space K and f , g are bounded functions on K such
that f , −g are H-convex, f ≤ g on ChH (K) and f, g are semicontinuous. Then
f ≤ g on K.
Proof. Assume, for example, that both functions are upper semicontinuous. We fix
x ∈ K and ε > 0 and find a continuous H-concave function g 0 on K such that g ≤ g 0
and g 0 (x) ≤ g(x) + ε (see Proposition 3.48). Then g 0 − f ≥ 0 on ChH (K) and thus
g 0 − f ≥ 0 on K by Theorem 3.16. Hence
f (x) ≤ g 0 (x) ≤ g(x) + ε.
Since ε is arbitrary, the proof is complete.
The remaining cases can be dealt with by a similar argument.
86 3 Choquet theory of function spaces
(µ, ν) ∈ co M.
Assume that this is not the case. Then we use the Hahn–Banach theorem to find
continuous functions f1 ∈ C(F1 ), f2 ∈ C(F2 ) and c ∈ R such that
Hence
µ(f1 + F2f2∗ ) ≥ µ(f1 ) + ν(F2f2∗ ) ≥ µ(f1 ) + ν(f2 )
> c ≥ sup{εz (f1 ) + λ(f2 ) : (εz , λ) ∈ M }
≥ f1 (x) + F2f2∗ (x).
Thus
µ(f1 + F2f2∗ ) > c ≥ sup (f1 (x) + Ff2∗ (x)).
x∈F1
Proof. Let us first notice that (3.6) is clear for f1 , f2 ∈ C(K). To extend this formula
to universally measurable functions, we will first consider characteristic functions.
So, let A be the family of all universally measurable sets A ⊂ K such that (3.6) holds
for pairs of functions (cA , 0) and (0, cA ).
Step 1. We show that the family A contains all Borel subsets of K.
Clearly, K ∈ A. If G ⊂ K is open, Proposition A.50 yields
cG = sup{ϕ ∈ C(K) : 0 ≤ ϕ ≤ cG }.
By Theorem A.84,
are suprema of up-directed families of continuous functions. Hence we may use The-
orem A.84 again for the measure Λ to obtain validity of (3.6) for the pairs (cG , 0) and
(0, cG ). Hence A contains all open subsets of K.
Obviously K ∈ A and K \A ∈ A whenever A ∈ A. Since A is closed with respect
to the formation of unions of two pairwise disjoint elements of A, and the Lebesgue
monotone convergence theorem implies that the countable union of an increasing se-
quence of sets from A also belongs to A, we see that A is a Dynkin system. Since
A contains all open subsets of K, the Dynkin lemma A.68 yields that A contains any
Borel subset of K.
Step 2. To show that any universally measurable subset of K belongs to A, let A
be such a set.
We write
A = H ∪ N, H is an Fσ set, µ1 (N ) = µ2 (N ) = 0, H ∩ N = ∅.
Since we know that H ∈ A, to finish the argument we need to prove that the functions
Nq := {(λ1 , λ2 ) ∈ P : λ1 (N ) ≥ q}
∪ {(λ1 , λ2 ) ∈ P : λ2 (N ) ≥ q}
and
Nqn :={(λ1 , λ2 ) ∈ P : λ1 (Gn ) ≥ q}
∪ {(λ1 , λ2 ) ∈ P : λ2 (Gn ) ≥ q}.
Since the sets Gn belong to A, for any n ∈ N and q ∈ (0, 1) ∩ Q, we get
Z Z
n 1
Λ(Nq ) = 1 dΛ(λ1 , λ2 ) ≤ (λ1 (Gn ) + λ2 (Gn )) dΛ(λ1 , λ2 )
Nqn q P
1
= (µ1 + µ2 )(Gn ).
q
Since Nq ⊂ Nqn , n ∈ N, the choice of the sets Gn yields that Nq is Λ-measurable and
Λ(Nq ) = 0.
Thus
Λ ({(λ1 , λ2 ) ∈ P : λ1 (N ) + λ2 (N ) > 0})
[
= Λ(Nq ) = 0,
q∈Q∩(0,1)
T f (x) = Tx (f ).
Before stating our nextS theorem, we will recall and introduce some notation. Re-
member that M(H) := x∈K Mx (H) (cf. Proposition 3.37) and r : M(H) → K is
the H-barycenter mapping (cf. Definition 3.39). Let P := M1 (K) × M1 (K). We
will also consider the barycenter mapping rP : M(P ) → P , here the subscript P
serves only for easier distinguishing which barycenter mapping we are just operating
with.
We denote by πi : K × K → K the i-the projection, i = 1, 2. Then they induce
mappings
(πi )] : M1 (K × K) → M1 (K), i = 1, 2.
Analogously, let pi : P → M1 (K) be the i-th projection, i = 1, 2. As above, we
obtain mappings
M := {(εx , λ) ∈ P : εx ≺ λ, x ∈ K}.
surable and h ∈ H,
(iv) there exists a µ-dilation such that T µ = ν,
(v) there exists Λ ∈ M1 (M(H)) such that
90 3 Choquet theory of function spaces
R
• ν(f ) = M(H) λ(f ) dΛ(λ), for any bounded universally measurable func-
tion f on K,
• r] Λ = µ,
(vi) whenever µ = ni=1 µi withP µi ∈ M+ (K), there exist measures νi ∈ M+ (K),
P
i = 1, . . . , n, such that ν = ni=1 νi and µi − νi ∈ H⊥ .
If, moreover, K is metrizable, then (i) is equivalent with the following assertion:
(iv’) there exists a dilation T such that T µ = ν.
Proof. We already know from Proposition 3.89 that (i) ⇐⇒ (ii). We will prove (i)
=⇒ (iii) =⇒ (iv) =⇒ (i), (ii) =⇒ (v) =⇒ (i), (ii) =⇒ (vi) =⇒ (i) and (ii)
=⇒ (iv’) =⇒ (i) in the case when K is metrizable.
For the proof of (i) =⇒ (iii), assume that µ ≺ ν. For any functions f, g : K → R,
we denote by f ⊗ g : K × K → R,
V := {f ⊗ 1 + 1 ⊗ g : f, g ∈ C(K)}
We define σ : C(K × K) → R by
Z ∗
σ(f ) := (fx )∗ (x) dµ(x), f ∈ C(K × K).
K
R∗
( is the upper integral, see Definition A.69). Then σ is a sublinear functional on
C(K × K) satisfying
σ(f ) ≤ 0 for f ≤ 0. (3.7)
Further, ρ ≤ σ on V . Indeed, if f, g ∈ C(K) are given, we have
and µ(g ∗ ) ≥ ν(g ∗ ) (see Proposition 3.56 and Lemma 3.18). Thus
Z ∗
σ(f ⊗1 + 1 ⊗ g) = ((f ⊗ 1 + 1 ⊗ g)x )∗ (x) dµ(x)
K
Z
= (f (x) + g ∗ (x)) dµ(x) ≥ µ(f ) + ν(g ∗ )
K
≥ µ(f ) + ν(g) = ρ(f ⊗ 1 + 1 ⊗ g).
3.10 Orderings and dilations 91
By the Hahn–Banach theorem, there exists a functional λ ∈ (C(K × K))∗ such that
λ = ρ on V and λ ≤ σ on C(K × K). By (3.7), λ is a Radon measure on K × K.
If k ∈ S c (H) and g ∈ C(K) positive are given, then
Z ∗
λ(g ⊗ k) ≤ σ(g ⊗ k) = ((g ⊗ k)x )∗ (x) dµ(x)
K
Z Z
= g(x)k ∗ (x) dµ(x) = g(x)k(x) dµ(x) = λ(gk ⊗ 1).
K K
This gives
λ(gh ⊗ 1) = λ(g ⊗ h), g ∈ C(K), h ∈ H. (3.8)
(If g is not positive, let c ≥ 0 be such that g + c is positive, and apply (3.8) to
(g + c) − c.)
We claim that λ ∈ M+ (K ×K) possesses all the required properties. First, λ(1) =
µ(1) = 1, and hence λ ∈ M1 (K × K). Further, for f ∈ C(K), we have
(π1 )] λ(f ) = λ(f ◦ π1 ) = λ(f ⊗ 1) = ρ(f ⊗ 1) = µ(f ),
and
(π2 )] λ(f ) = λ(f ◦ π2 ) = λ(1 ⊗ f ) = ρ(1 ⊗ f ) = ν(f ).
Finally, (3.8) and standard measure-theoretic techniques yield
Z Z
g(x)h(x) dλ(x, y) = g(x)h(y) dλ(x, y)
K×K K×K
for h ∈ H and a bounded universally measurable function g on K. By setting g = cA ,
where A ⊂ K is universally measurable, we get
Z Z
h(x) dλ(x, y) = h(y) dλ(x, y), h ∈ H.
A×K A×K
To finish the proof of the implication, let h ∈ H be given. For any A ⊂ K Borel,
h := cA ⊗ h. Then e
we define e h is λ-measurable and using (iii) we obtain
Z Z Z Z
T h dµ = eh(x, y) dTx (y) dµ(x) = h(x, y) dλ(x, y)
e
A K K K×K
Z Z
= h(y) dλ(x, y) = h(x) dλ(x, y)
A×K A×K
Z Z
= h(x) d((π1 )] λ) = h(x) dµ(x).
A A
This shows that T h = h for µ-almost all x ∈ K.
We proceed with the proof of (iv) =⇒ (i). Let T = {Tx }x∈K : K → M1 (K) be
a µ-dilation such that T µ = ν. Let f ∈ −W(H) be given; that is, f = h1 ∨ · · · ∨ hn ,
h1 , . . . , hn ∈ H. We set
A0i := {x ∈ K : f (x) = hi (x)}, i = 1, . . . , n,
(3.10)
A1 := A01 , Ai := A0i \ (A1 ∪ · · · ∪ Ai−1 ), i = 2, . . . , n.
Let N ⊂ K be a set of µ-measure zero such that T hi (x) = hi (x) for each x ∈ K \ N
and i = 1, . . . , n. Then
Z Xn Z
µ(f ) = f dµ = hi dµ
K\N i=1 Ai \N
n Z
X n Z
X
= T hi dµ ≤ T f dµ
i=1 Ai \N i=1 Ai \N
Z
≤ T f dµ = T µ(f ) = ν(f ).
K
Hence µ ≺ ν by Proposition 3.56.
For the proof of (ii) =⇒ (v), let Ω ∈ M1 (M ) ⊂ M1 (P ) satisfy rP (Ω) = (µ, ν).
We claim that the measure Λ := (p2 )] Ω satisfies our requirements.
First, we observe that p2 (M ) ⊂ M(H), so that
Λ ∈ M1 (M(H)) ⊂ M1 (M1 (K)).
Let f be a bounded universally measurable function on K. Then Proposition 3.90
yields Z Z
ν(f ) = λ(f ) dΩ(εx , λ) = λ(f ) dΛ(λ)
M M(H)
and Z Z
(r] Λ)(f ) = f (r(λ)) dΛ(λ) = f (r(λ)) dΩ(εx , λ)
M(H) M
Z
= εx (f ) dΩ(εx , λ) = µ(f ).
M
Hence r] Λ = µ.
3.10 Orderings and dilations 93
Hence µ ≺ ν.
1 1
Pn=⇒ (vi), let Ω ∈ M (M ) ⊂ M (P ) satisfy rP (Ω) = (µ, ν).
For the proof of (ii)
Suppose that µ = i=1 µi , where Radon measures µi are nontrivial. According
to the Radon–Nikodym theorem, there exist positive Borel measurable functions fi ,
i = 1,P. . . , n, such that µi (ϕ) = µ(fi ϕ) for each bounded Borel function ϕ on K.
ThenP ni=1 fi = 1 µ-almost everywhere. Without loss of generality we may assume
that ni=1 fi = 1 on K.
For each i ∈ {1, . . . , n}, we define the function gi : M → R by
gi Ω
Ωi := .
Ω(gi )
Thus
µi = Ω(gi )λ1i , i = 1, . . . , n.
We set
νi := Ω(gi )λ2i , i = 1, . . . , n.
µi − νi ∈ H⊥ , i = 1, . . . , n.
94 3 Choquet theory of function spaces
Finally, for any ϕ ∈ C(K) we get (using (3.6) from Proposition 3.90)
n
X n
X Z n Z
X
νi (ϕ) = Ω(gi ) λ(ϕ) dΩi (εx , λ) = gi (εx , λ) λ(ϕ) dΩ(εx , λ)
i=1 i=1 M i=1 M
Xn Z Z n
X
= εx (fi )λ(ϕ) dΩ(εx , λ) = εx fi λ(ϕ) dΩ(εx , λ)
i=1 M M i=1
Z
= λ(ϕ) dΩ(εx , λ) = ν(ϕ).
M
µi := µ|Ai , i = 1, . . . , n,
n Z
X n
X n
X n
X
µ(f ) = hi dµ = µi (hi ) = νi (hi ) ≤ νi (f ) = ν(f ).
i=1 Ai i=1 i=1 i=1
Ψ(εx , λ) := x, (εx , λ) ∈ M.
Then ψ ∈ Ac (P ) and
Z
Sx (ψ) = ψ(λ̃, λ) dSx (λ̃, λ) = ψ(rP (Sx ))
M
= p2 (rP (Sx ))(f ) = Tx (f ), x ∈ K.
Hence
Z Z
T µ(f ) = Tx (f ) dµ(x) = Sx (ψ) dµ(x) = Ω(ψ) = ψ(rP (Ω))
K K
= ψ(µ, ν) = ν(f ).
To show that (iv’) =⇒ (i), let T = {Tx }x∈K be a dilation such that T µ = ν. Then,
for f ∈ Kc (H),
Z Z
ν(f ) = (T µ)(f ) = Tx f dµ(x) ≥ f (x) dµ(x)
K K
3.11 Exercises
Exercise 3.93 (Bauer functionals). For a measure µ ∈ M+ (K) and an upper bounded
real-valued function f on K denote
B µ (f ) := inf {µ(h) : h ∈ H, h ≥ f } .
96 3 Choquet theory of function spaces
Hint. The proof is almost the same as the proof of Lemma 3.21.
Hint. Let H be the space of all affine continuous functions on the closed interval
[−1, 1]. Then S c (H) equals the set of all continuous concave functions on [−1, 1].
Show that for
1
f (x) := −|x| + 1 and µ := (ε−1 + ε1 )
2
we get
Qµ (f ) = 0 and B µ (f ) = 1.
Exercise 3.96. Prove that, for any upper bounded function f on K, the inequalities
“≥” in the families of functions in Proposition 3.25(a) may be replaced by “>”.
kFf ∗ − Fg ∗ k ≤ kf − gk`∞ (F )
Thus
F ∗
f (x) − Fg ∗ (x) ≤ F(f − g)∗ (x) ≤ kf − gk,
and the assertion easily follows.
Exercise 3.99. Let {fα }α∈A be a down-directed net of upper semicontinuous func-
tions on K, fα & f . Prove that fα∗ & f ∗ .
Hint. It is clear that {fα∗ }α∈A is a down-directed net of functions satisfying f ∗ ≤ fα∗
for every α ∈ A. Since by Exercise 3.96
f ∗ = inf {h ∈ H : h > f } ,
it is enough to verify that for any h > f , h ∈ H, there exists α ∈ A such that fα ≤ h.
Since fα & f , we can use the upper semicontinuity of the functions {fα } and the
compactness of K to find α1 , . . . , αn ∈ A such that
Exercise 3.100. Let {xα } be a net of points converging to x ∈ ChH (K). Let µα ∈
Mxα (H). Then µα → εx .
Hint. If we suppose the contrary, then there exist a measure µ 6= εx and a subnet {µβ }
so that µβ → µ. It is straightforward to verify that µ is an H-representing measure
for x. Since µ is not the Dirac measure at x, we have arrived at a contradiction with
the assumption that x ∈ ChH (K).
Exercise 3.101. Prove that a closed set F ⊂ K is H-extremal if and only if cK\F ∈
S lsc (H).
Exercise 3.102. Let X be a compact convex subset of a locally convex space and
B b (X) be the family of all bounded Borel functions on X.
98 3 Choquet theory of function spaces
Exercise 3.103. Let H be the linear span of 1, id, id3 on [−1, 1]. Prove that Kor(H)
6= C([−1, 1]). Find a continuous function on [−1, 1] which does not belong to Kor(H).
Exercise 3.104. Let H be the linear span of {1, id, ϕ} on an interval [a, b] , where ϕ
is a continuous function on [a, b]. Prove that Kor(H) = C([a, b]) if and only if ϕ is
either strictly convex or strictly concave on [a, b].
3.11 Exercises 99
Hint. If ϕ is strictly convex and a ∈ [0, 1], then the function ϕ − h exposes a for each
affine function h supporting ϕ at a.
For the converse, set
s0 (f ) + s1 (f ) + · · · + sn−1 (f )
Tn f := ,
n
then the sequence Tn f converges uniformly to f on R.
Hint. Since the operators Tn are obviously linear and positive in view of
2
2π sin( nt
2 )
Z
1
Tn f (x) = f (x + t) dt, x ∈ [0, 2π],
2πn 0 sin( 2t )
it suffices to use the second Korovkin theorem 3.34 to show that the sequence {Tn } is
H-admissible if H equals the linear span of {1, cos, sin}.
Exercise 3.106. If Z 1
H := f ∈ C([0, 1]) : f (0) = f
0
Hint. According to Proposition 2.45, we may assume that X is a norm compact con-
vex subset of `2 . Then x 7→ kxk2 , x ∈ X, is the required strictly convex function.
100 3 Choquet theory of function spaces
Exercise 3.108 (Rosenthal’s proof). In the following, supply the details of Rosenthal’s
proof of the Choquet representation theorem in the setting of metrizable compact
convex sets. More precisely, prove that given a metrizable compact convex subset
X of a locally convex space E and x ∈ X, then there exists a probability measure
µ ∈ Mx (X) such that µ is carried by extreme points of X.
Hint. Exercise 3.107 provides a strictly convex function Q on X. For n ∈ N, define
1 1
Fn := x ∈ X : there are y, z ∈ X, 2x = y + z, (Q(y) + Q(z)) ≥ Q(x) + .
2 n
Obviously, each Fn is closed. First, prove the following assertion.
Claim. Let ρ ∈ M1 (Fn ). Then there exists a measure σ ∈ M1 (X) such that
Z Z Z Z
1
f dσ = f dρ, f ∈ E ∗ , and Q dσ ≥ Q dρ + .
X Fn X Fn n
Proof of the claim. First, the claim holds in the case when ρ = εx for some x ∈
Fn . It suffices to put σ = 21 (εy + εz ), where y, z are as in the definition of the set
Fn . The assertion then easily follows for any molecular measure. In general, use
approximation by molecular measures (cf. Proposition 2.27).
Proof of the Choquet theorem. Using the compactness of M1 (X) (cf. again Propo-
sition 2.27) show that there exists a representing measure µ ∈ Mx (X) such that
Z nZ o
Q dµ = sup Q dν : ν ∈ Mx (X) =: m.
X K
find n ∈ N so that r := µ(Fn ) > 0. By the claim, there exists a measure σ for
ρ := r1 µ|Fn . Denote
κ := rσ + µ|X\Fn
and show that κ ∈ Mx (X). Then deduce that
Z Z
r r
m≥ Q dκ ≥ Q dµ + = m + ,
X X n n
which is an obvious contradiction.
F := {f − k + ε : k ∈ S c (H), k ≥ g} .
exposes the points rn and 1 + rn , these points belong to the Choquet boundary
ChH (K). Also the points 0 and 1 are exposed, as the functions f (x) := x and
1,
x ∈ [0, 21 ] ∩ K,
f (x) := 3n−12n2
, x = 1 − rn , n ≥ 2,
x − 1, x ∈ [1, 2] ∩ K,
Any measure from M1−rn (H) is carried by the set {rn , 1 − rn , 1 + rn }. Hence
the function (
1, x = 0,
f :=
0 elsewhere
is H-concave. It is clear that f is also upper semicontinuous. Let g be a lower semi-
continuous function defined as
(
1, on K ∩ [0, 12 ],
g :=
0 elsewhere.
Exercise 3.112. Prove that Mmax (H) is a norm closed cone in M(K).
Exercise 3.113. Prove that the set of all probability maximal measures is a norm
closed face of M1 (K).
Exercise 3.116. Let {xn } be a sequence in a Banach space E that converges weakly
Then for every ε > 0 there exist λi > 0, i = 1, . . . , n, ni=1 λi = 1,
P
to x ∈ E. P
such that k ni=1 λi xi − xk < ε. Prove this without the recourse to Mazur’s lemma
(cf. W. Rudin [403], Theorem 3.13).
Hint. By the uniform boundedness principle (see Theorem 2.5 in W. Rudin [403]),
the sequence {xn } is bounded. If we view elements of E as continuous functions on
the dual unit ball BE ∗ endowed with the w∗ -topology, then the assertion follows from
Exercise 3.115.
for any h ∈ H.
A measure µ ∈ M+ (K) is called l-maximal if given ν ∈ M+ (K), µ l ν, then
µ(h2 ) = ν(h2 ) for any h ∈ H.
Prove the following assertions hold.
(a) The relation l is reflexive and transitive.
(b) If µ ≺ ν, then µ l ν.
(c) For any measure µ ∈ M+ (K) there exists a l-maximal measure λ ∈ M+ (K)
such that µ l λ.
(d) Any l-maximal measure is ≺-maximal.
Hint. The verification of reflexivity and transitivity in assertion (a) is straightforward.
For the proof of (b), we just note that h2 ∈ Kc (H) for any h ∈ H, and assertion (c)
can be proved by an analogous argument to that in Theorem 3.65.
For the proof of (d), let µ be a l-maximal measure. According to Theorem 3.58,
it is enough to show that µ(f ) = µ(f ∗ ) for any f ∈ −W(H). Let ε > 0 and
f ∈ −W(H) be given; that is,
f = h1 ∨ · · · ∨ hn , h1 , . . . , hn ∈ H.
Let
H := {x ∈ K : f ∗ (x) − f (x) ≥ ε}
104 3 Choquet theory of function spaces
Then r(µ) = r(ν), ν is ≺-maximal, νlµ, but there exists h ∈ H with ν(h2 ) < µ(h2 ).
Obviously, ν is ≺-maximal because ν is carried by extreme points of K. The
verification of ν l µ can be done by a straightforward computation. Finally, if we
define
h(x, y) := x + y, (x, y) ∈ K,
3
then ν(h2 ) < 2 = µ(h2 ).
In [109], G. Choquet discloses the following three sources of motivation for his in-
vestigation on integral representation in compact convex sets: study of certain capaci-
ties (alternating set functions of infinite order), R.S. Martin’s representation of positive
harmonic functions in a Euclidean domain and R. Godement’s result on positive op-
erators on Hilbert spaces. G. Choquet completed his proof on integral representation
Theorem 3.45 (the metrizable case) in July 1956. The result, examples and an applica-
tion to a new proof of the ergodic theorem were presented in G. Choquet [111]. New
proofs appeared soon afterwards (M. Hervé [223], L. H. Loomis [311], F. F. Bonsall
[75]). Theorem 3.42 was proved for the convex case by M. Hervé in [223].
The characterization of maximal measures contained in Theorem 3.58 is due to
G. Mokobodzki [347].
The Simons inequality in 3.74 and the Simons lemma in 3.75 go back to S. Simons’
paper [417]. Proof of Lemma 3.74 follows the lines of a proof given by M. Ruiz Galán
and S. Simons [404]. We use the Simons lemma 3.75 for the proof of Lemma 3.77.
Another proof of this lemma can be found in E. M. Alfsen [5], Proposition 1.4.10. (see
also Exercise 4.45). For generalizations and applications of the Simons inequality, as
well as for its simplified proofs, we refer the reader to E. Oja [367], K. Kiviso and
E. Oja [270], G. Godefroy [199] and R. Deville and C. Finet [143]. For a related
lemma by V. Pták we refer the reader to [380].
The approach proposed by E. Bishop and K. de Leeuw in [58] turned out to be
particularly important. It relies on introducing partial ordering comparing measures
by means of squares of continuous affine functions. Maximal measures, which in the
metrizable case are carried by the set of extreme points, then play a role of desired
representing measures. Using a different ordering on measures induced by all con-
tinuous convex functions, G. Choquet obtained a similar representation theorem for
nonmetrizable case in [105]. Example 3.82 is due to E. Bishop and K. de Leeuw [58].
Example 3.83 is a variant of an example due to G. Mokobodzki (see [374], Chapter 10,
Example).
The characterization given in (vi) of Theorem 3.92 was taken as the definition of the
strong ordering of measures by L. H. Loomis in [311]. The equivalence of his strong
ordering with the Choquet ordering was later established in P. Cartier, J. M. G. Fell and
P. -A. Meyer [97]. Theorem 3.92 follows their result and Theorem 1.3.6 in G. Winkler
[473] dealing with noncompact case. Conditions of Theorem 3.92 and the proof of
their equivalence follows R. D. Bourgin [82], Chapter 6, G. A. Edgar [152], [153],
J. L. Doob [144], J. Hoffmann-Jørgensen [227], H. v. Weizsäcker [462] and the paper
H. v. Weizsäcker and G. Winkler [464].
Exercise 3.102 is taken from J. D. Maitland Wright [476]. For details of Rosen-
thal’s proof of the Choquet representation theorem indicated in Exercise 3.108, see
the paper H. P. Rosenthal [397]. Exercises 3.114 and 3.115 use techniques from
S. Simons [418], Exercise 3.117 is taken from E. Bishop and K. de Leeuw [58] and
Exercise 3.119 is taken from the paper [311] by L. H. Loomis.
Chapter 4
Affine functions on compact convex sets
The preceding chapter deals with abstract function spaces on compact spaces. Here
we focus on a particular example of a function space, namely, on the space Ac (X)
of continuous affine functions on a compact convex set X, and show how classical
concepts from convex analysis fit into the general framework of function spaces. We
first establish an easy, but important, Lemma 4.3 on approximation of probability
measures on X by molecular measures with the same barycenter. Next we introduce
a notion of strongly affine functions, that is, the functions satisfying the barycentric
formula. Corollary 4.8 then identifies semicontinuous convex and affine functions
on X with semicontinuous H-convex and H-affine functions. As a consequence, we
obtain classical results on approximation of semicontinuous convex functions (see
Propositions 4.9, 4.10 and 4.12).
Next Section 4.2 characterizes strongly affine functions. The main result contained
in Theorem 4.19 shows that strongly affine functions can be characterized by means
of the Haydon condition and a Lusin type property. A list of conditions characteriz-
ing strongly affine functions is further enriched by Theorem 4.34. As a corollary of
Theorem 4.19 we get Choquet’s barycentric theorem 4.22 and Mokobodzki’s approx-
imation theorem 4.24.
Section 4.3 introduces the state space S(H) of a function space H. This provides an
efficient link between the theory of function spaces and the theory of compact convex
sets. In a series of propositions we describe basic facts that will enable us later on to
transfer properties of function spaces to the framework of compact convex sets and
vice versa. First glimpses of this kind of reasoning can be observed in a reformulation
of Goldstine’s lemma 4.33 and in Theorem 4.35.
As an application of Mokobodzki’s approximation theorem, we prove in Section 4.4
that affine Baire-one functions on the dual unit ball of a Banach space E are pointwise
limits of a sequence of elements from E. This also shows that w∗ -sequential closure
of E in its bidual is norm closed.
for each x, y ∈ X and for each λ ∈ (0, 1). We denote by A(X) the set of all affine
functions on X, by Ab (X) the space of all bounded affine functions on X, and by
Ac (X) the set of all continuous functions from A(X). It is obvious that Ac (X) is a
function space.
We also recall that the Choquet boundary of Ac (X) coincides with the set ext X of
all extreme points of X.
More generally, a function f is called concave if f is upper or lower finite and
The collection of all concave functions on X will be denoted by S(X). Further, let
Sc (X), Susc (X) and Slsc (X) denote the families of all continuous, upper semicon-
tinuous and lower semicontinuous concave functions on X, respectively.
A function f is convex if −f is concave. We write K(X) for the family of all
convex functions on X. The notation Kc (X), Kusc (X) and Klsc (X) should be clear.
Recall that we abbreviated Mx (Ac (X)) to Mx (X) and notice that µ ∈ Mx (X)
means exactly that x is a barycenter of µ; for short, r(µ) = x.
Analogously we write Mmax (X) and Mbnd (X) instead of Mmax (Ac (X)) and
Mbnd (Ac (X)).
Proposition 4.3. Let x ∈ X and µ ∈ Mx (X). Then there exists a net {µα } of
molecular measures in Mx (X) such that µα → µ.
4.1 Affine functions and the barycentric formula 109
|µ(fi ) − ν(fi )| ≤ ε, i = 1, . . . , n.
Let J be the set of indices for which µ(Aj ) 6= 0. Let λj := µ(Aj ) for j = 1, . . . , m
and
µj := λ−1
j µ|Aj , j ∈ J.
z − xj = (z − x0j ) + (x0j − xj ),
and so by (4.1)
|fi (z) − fi (xj )| ≤ 2ε, i = 1, . . . , n.
Since µj is a probability measure carried by X ∩ Wj , we obtain
≤ 2ε, i = 1, . . . , n.
Claim. Let x ∈ X and c > f (x) be given. Then there exists n ∈ N such that
where
Proof of the claim. Note that both sets Fn and Hn are compact convex sets in E × R,
and hence
co (Fn ∪ Hn ) = λu + (1 − λ)v : u ∈ Fn , v ∈ Hn , λ ∈ [0, 1] .
Suppose that the claim does not hold. Then, for any n ∈ N, there exists tn ∈ R such
that c ≤ tn and (x, tn ) ∈ co (Fn ∪ Hn ) . Hence there exist yn , zn ∈ X, λn ∈ [0, 1],
sn ∈ R, satisfying
Using compactness we find subnets {yni }, {zni } and {λni } of {yn }, {zn } and {λn },
respectively, and points y, z ∈ X and λ ∈ [0, 1] such that yni → y, zni → z and
λni → λ. Then x = λy + (1 − λ)z and, for each index i,
f (x) < c ≤ tni = λni sni + (1 − λni )(−ni ) ≤ λni f (yni ) + (1 − λni )(−ni ). (4.3)
If λ = 1, we get from (4.3) f (x) < c ≤ f (x). It does not matter whether f (x) is
finite or infinite; in both cases we obtain a contradiction.
If λ < 1, we get c ≤ −∞, which again is a contradiction with the choice of c. The
proof of the claim is thus completed.
Now, the proof of the proposition can be finished as in the case of a real-valued
upper semicontinuous functions. Let x ∈ X, µ ∈ Mx (X) and c > f (x) be given.
Find an integer n ∈ N with the property guaranteed by the claim. Then the point (x, c)
can be separated from the set co (Fn ∪ Hn ); in other words, there exists a continuous
affine function h on X such that f ≤ h and h(x) < c. Then µ(f ) ≤ µ(h) = h(x) < c,
and the proof is finished.
Slsc (X) = S lsc (Ac (X)), Susc (X) = S usc (Ac (X)),
Ausc (X) = Ausc (Ac (X)), Alsc (X) = Alsc (Ac (X)),
and Sc (X) = S c (Ac (X)).
Proof. The necessity of the condition is obvious. Concerning the sufficiency, let f, −g
be upper semicontinuous functions on X such that µ(f ) < ν(g) for any µ, ν ∈
M1 (X) satisfying r(µ) = r(ν). We claim that f ∗ < g∗ . Indeed, pick x ∈ X.
By Lemma 3.21, there exist measures µ, ν ∈ Mx (X) such that
J1 := {(x, t) ∈ E × R : m ≤ t ≤ f1 (x)}
and
J2 := {(x, t) ∈ E × R : g1 (x) ≤ t ≤ M }.
4.2 Barycentric theorem and strongly affine functions 113
A := {g ∈ Ac (X) : g < f on X} .
Definition 4.14 (Convex inner measure). For a Radon measure µ on X, let the convex
inner measure µ for A ⊂ X be defined as
We note that the convex inner measure µ need not be a measure at all.
Lemma 4.15. Let µ ∈ M1 (X) and {An } be a decreasing sequence of convex subsets
of X. Then
\∞
µ An = lim µ(An ).
n→∞
n=1
Proof. Since µ(A) ≤ µ(B) whenever A ⊂ B, we obtain that limn→∞ µ(An ) exists
and
∞
\
µ( An ) ≤ lim µ(An ).
n→∞
n=1
For the converse, fix ε > 0. By induction, we will construct a decreasing sequence of
compact convex sets Kn ⊂ An such that
First find a compact convex set K1 ⊂ A1 such that µ(K1 ) > µ(A1 ) − 2ε . If the sets
K1 , . . . , Kn−1 satisfying the required condition have been constructed, select a real
number δ such that 0 < δ < ε and
There exists a compact convex set L ⊂ An such that µ(L) > µ(An ) − δ. If we set
Kn := L ∩ Kn−1 , we obtain the desired compact convex set. Indeed, since
we get
Lemma 4.16. Assume that a bounded affine function f on X satisfies the Haydon
condition. Then for any µ ∈ M1 (X) and any b ∈ R,
Proof. Let µ be a Radon probability measure on X. We will prove that for any pair
of real numbers a < b, the inequality
µ(A) + µ(B) ≥ 1
λ := µ|X\S∞
n=1 (Ln ∪Kn )
is a nonzero Radon measure. Hence we may apply the Haydon condition in order to
obtain a compact convex set L such that λ(L) > 0 and such that L is contained either
in A or in B. Suppose that L ⊂ A. Since A is convex, co(L ∪ Ln ) is contained in A
for any n ∈ N. Thus we obtain
Thus using the monotonicity of a convex inner measure (Lemma 4.15), we obtain
Lemma 4.18. Let f be a function on a compact convex set X such that X is saturated
with respect to f . Then f is universally measurable.
Proof. Let µ ∈ M1 (X) be given. We fix a sequence {εk } of strictly positive numbers
converging to 0. For each k ∈ N, we find a sequence {Knk }∞ of compact convex
n=1 S
sets in X such that diam f (Kn ) ≤ εk for each n ∈ N and µ(X \ ∞
k k
n=1 µ(Kn )) = 0.
Sn−1 k
We set An := Kn \ i=1 Ki , n ∈ N, and pick numbers an ∈ f (An ) if the set Akn is
k k k k
nonempty. We define
(
akn , if x ∈ Akn , n ∈ N,
fk (x) :=
if x ∈ X \ ∞ k
S
0, n=1 An .
Theorem 4.19. Let f be a bounded affine function on X. Then the following condi-
tions are equivalent:
(i) f satisfies the barycentric formula,
(ii) for any µ ∈ M+ (X) and any ε > 0, there exists a compact convex set L ⊂ X
such that µ(L) > µ(X) − ε and f |L is continuous,
(iii) f satisfies the Haydon condition,
(iv) X is saturated with respect to f.
Proof. (i) =⇒ (ii): Suppose that f satisfies the barycentric formula and µ is a
Radon measure on X. Clearly we may assume that µ is a probability measure. For
given ε > 0, we use Lusin’s theorem A.76 to find a compact set F ⊂ X such that
f |F is continuous and µ(F ) ≥ 1 − ε. If we define L := coF, we obtain the required
compact convex set.
Indeed, we want to prove that f |L is continuous. Let x be a point in L and {xα }α∈A
be a net of points of L converging to x. By the Milman theorem 2.43, ext L is a
subset of F. The Integral representation theorem 2.31 ensures the existence of a net of
probability measures {µα }α∈A carried by ext L with barycenters xα . Without loss of
generality, we may assume that {µα }α∈A converges to a probability measure ν. Since
the net {xα }α∈A converges to x, the measure ν represents x. Since f is continuous
on ext L and µα and ν are carried by ext L, we get µα (f ) → ν(f ). Now we apply the
assumption to obtain
continuous and µ(L) > 0. We select a point x ∈ spt µ|L . Without loss of generality
we may assume that f (x) < b. By the continuity of f |L we can find a closed convex
neighborhood V of x such that f < b on V ∩L. Then V ∩L satisfies our requirements.
For the proof of the implication (iii) =⇒ (iv) we will use Lemma 4.16. Let
a Radon probability measure µ on X and ε > 0 be given. Fix δ > 0. Since f is
bounded, we may suppose that 0 ≤ f < 1. Choose N ∈ N such that N1 < ε and define
Fn := {x ∈ X : n−1 n
4.16, N
P
N ≤ f < N }, n = 1, . . . , N . By Lemma P n=1 µ(Fn ) = 1.
N
Hence there exist compact convex sets Kn ⊂ Fn such that n=1 µ(Kn ) > 1 − δ.
Then
1
diam f (Kn ) ≤ diam f (Fn ) ≤ < ε.
N
If we put together all compact convex sets obtained by choosing δl = 1l , l ∈ N, we
get the required sequence.
It remains to prove (iv) =⇒ (i). Let a Radon probability measure µ on X with the
barycenter x and ε > 0 be given. By condition (iv), there exists a sequence {Kn } of
closed convex subsets of X such that
∞
[
diam f (Kn ) < ε and µ X\ Kn = 0.
n=1
λn := µ(En ), n = 0, . . . , N.
Obviously
N
X N
X N
X
λn = 1, λn xn = x and λn µn = µ. (4.7)
n=0 n=0 n=0
118 4 Affine functions on compact convex sets
Lemma 4.20. Let f be a convex real-valued function on a compact convex set X that
is lower bounded on some nonempty open subset of X. Then f is lower bounded on
X.
In particular, if f has a point of continuity, then f is lower bounded on X.
Proof. Given f as in the lemma, let U be a nonempty open subset of X such that
f ≥ C on U for some C ∈ R. Assume that f is not lower bounded on X; that is, that
there exist points xn ∈ X, n ∈ N, such that f (xn ) → −∞. Let x be a cluster point of
the sequence {xn } and let y ∈ U be chosen arbitrarily. Then there exists α ∈ (0, 1)
such that αx + (1 − α)y ∈ U . Since x is a cluster point of {xn }, for infinitely many
indices n ∈ N we have αxn + (1 − α)y ∈ U . Then for these indices we get
Theorem 4.21. Let f be an affine function on a compact convex set X such that f is
fragmented. Then f satisfies the barycentric formula.
Proof. First, we note that f is bounded by Lemma 4.20 and Theorem A.121.
We will show that X is saturated with respect to f . To this end, let µ ∈ M+ (X)
and ε > 0 be given. Let
Proof. By Theorem A.124 and Theorem A.121, any Baire-one function on X is frag-
mented. Thus the assertion follows from Theorem 4.21.
Proof. Choquet’s barycentric theorem 4.22 asserts that any affine Baire-one function
is Ac (X)-affine. Thus the affine B1 -minimum principle is a particular case of Theo-
rem 3.86. The second assertion is then an immediate consequence.
fn % f and gn & f.
120 4 Affine functions on compact convex sets
1 1
fn − < hn < gn + .
n n
Then {hn } is a bounded sequence of continuous affine functions such that hn → f
on X.
Definition 4.25 (State space). Let H be a function space on a compact space K and
S(H), the state space of H, be defined as
S(H) := {ϕ ∈ H∗ : ϕ ≥ 0, ϕ(1) = 1} .
φ : x 7→ φx , x ∈ K,
where
φx : h 7→ h(x), h ∈ H.
Let Φ : H → Ac (S(H)) be the mapping defined for h ∈ H as
(c) r(φ] µ) = π(µ) for any µ ∈ M1 (K); in particular, r(φ] µ) = φ(x) for µ ∈
Mx (H),
(d) φ(ChH (K)) = ext S(H),
(e) Φ is a positive linear mapping and kΦ(h)k = khk for each h ∈ H,
(f) Φ is surjective if and only if H is closed,
(g) if Φ is surjective, then the inverse mapping is realized by
The Hahn–Banach theorem provides a Radon measure µ on K such that kµk = ksk =
1 and s(g) = µ(g) for each g ∈ H. Since µ(1) = s(1) = 1, the measure µ belongs to
M1 (K). Thus s(h) = µ(h) ∈ co h(K). But this contradicts (4.11). Thus S(H) =
co(φ(K)).
Lemma 2.34 yields (b). Indeed, Lemma 2.34 ensures that ((H ∗ , w∗ )∗ + R)|S(H)
is dense in Ac (S(H)). Hence the space Φ(H) + R is dense in Ac (S(H)). Since H
contains constant functions, Φ(H) + R = Φ(H).
For the proof of (c), let µ ∈ M1 (K) be given. By (b), it is enough to verify that
We proceed with the proof of (d). Pick s ∈ ext S(H). Since S(H) = co(φ(K)),
the Milman theorem 2.43 implies that
give (e).
If Φ is onto, the function space H is closed because Ac (S(H)) is a complete space
and Φ is an isometric mapping. Conversely, suppose that H is a closed space. Since
Φ(H) is a dense subspace of Ac (S(H)), H is complete and Φ is an isometry, the space
Φ(H) is closed in Ac (S(H)). Hence
Proposition 4.28. Let H be a function space on a compact space K and S(H) its
state space. Then the following assertions hold.
(a) If µ ∈ Mφ(x) (Ac (S(H))) with spt µ ⊂ φ(K), then (φ−1 )] µ ∈ Mx (H).
(b) π(M1 (F )) = co φ(F ) for any closed set F ⊂ K.
(c) If µ, ν ∈ M+ (K), then µ ≺ ν if and only if φ] µ ≺ φ] ν.
(d) A measure Λ ∈ M+ (S(H)) is Ac (S(H))-maximal if and only if Λ = φ] λ for
some H-maximal measure λ ∈ M+ (K).
(e) A measure Λ ∈ M(S(H)) is a boundary measure if and only if Λ = φ] λ for some
H-boundary measure λ ∈ M(K).
4.3 State space and representation of affine functions 123
Proof. Assertion (a) is a direct consequence of the definitions and the density of Φ(H)
in Ac (S(H)).
For the proof of (b), if s ∈ co φ(F ), then there exists a measure µ ∈ M1 (φ(F ))
which represents s (see Proposition 2.39). By assertion (a),
Then we obtain
µ(h) = s(h) > sup t(h) ≥ sup t(h) = sup h(F ) ≥ µ(h),
t∈co φ(F ) t∈φ(F )
and hence φ] µ ≺ φ] ν.
Conversely, assume that φ] µ ≺ φ] ν. By Proposition 3.56, it is enough to show that
µ(f ) ≤ ν(f ) for any f = h1 ∨ · · · ∨ hn , where h1 , . . . , hn ∈ H. By the assumption
and the definition of the image of a measure, we get
Z Z
h1 ∨ · · · ∨ hn dµ = Φ(h1 ) ∨ · · · ∨ Φ(hn ) d(φ] µ)
K φ(K)
Z Z
≤ Φ(h1 ) ∨ · · · ∨ Φ(hn ) d(φ] ν) = h1 ∨ · · · ∨ hn dν,
φ(K) K
hand, if ϕSis a nonzero positive functional on H and c = kϕk, then ϕ = c(c−1 ϕ), and
thus ϕ ∈ c≥0 cX. Obviously, (H∗ )+ ∩ −(H∗ )+ = {0}.
If ϕ ∈ H∗ is given, extend ϕ to a signed measure µ ∈ M(K) with µ = ϕ
on H. If we decompose µ into its positive and negative part µ+ and µ− and set
ϕ+ := µ+ |H and ϕ− := µ− |H , we obtain the desired decomposition of ϕ. Thus
H∗ = (H∗ )+ − (H∗ )+ .
For the third part of the proof, pick F ∈ Ac (X). Since H is closed, F = Φ(f ) for
some f ∈ H by Proposition 4.26(f). Setting
Fb(ϕ) := ϕ(f ), ϕ ∈ H∗ ,
(e) For every affine continuous function f on X there exists a unique function fb ∈
(E ∗ , w∗ )∗ such that f = fb ◦ φ on X.
(f) The mapping id : (Ac (X), τX ) → (E, w) is a homeomorphism.
4.3 State space and representation of affine functions 125
Theorem 4.33 (Goldstine’s lemma). If X is a compact convex set, then BAc (X) is
τX -dense in BAb (X) .
Proof. The assertion is just a reformulation of Goldstine’s lemma (see Theorem A.4).
Indeed, if E stands for the Banach space Ac (X), then the unit ball BE is w∗ -dense
in BE ∗∗ . Let ε : E → E ∗∗ be the canonical embedding and T : E ∗∗ → Ab (X) be
the mapping from Proposition 4.32(b). If I : Ac (X) → Ab (X) denotes the identity
mapping, then I = T ◦ ε. From this the assertion follows.
Theorem 4.34. For a real-valued function f on a compact convex set X, the following
assertions are equivalent:
(i) f is strongly affine,
(ii) f is bounded and for any measure µ ∈ M1 (X) there exists a sequence {fn } of
continuous affine functions on X such that kfn k ≤ kf k, n ∈ N, and fn → f
µ-almost everywhere.
Proof. To prove (i) =⇒ (ii), let f be a strongly affine function. Since f is bounded
by Lemma 4.5, we may assume without loss of generality that kf k = 1. Given a
measure µ ∈ M1 (X), Theorem 4.19 implies the existence of compact convex sets
Kn ⊂ X, n ∈ N, such that µ(Kn ) → 1 and f |Kn is continuous. For a fixed natural
number n ∈ N, consider the restriction operator R : Ab (X) → Ab (Kn ) defined as
Rf := f |Kn , f ∈ Ab (X).
By Goldstine’s lemma 4.33, BAc (X) is τX -dense in BAb (X) . Since the operator R is
(τX –τKn )-continuous, R(BAc (X) ) is τKn -dense in R(BAb (X) ). Hence we have Rf ∈
4.4 Affine Baire-one functions on dual unit balls 127
τK
R(BAc (X) ) n . Since Rf is continuous on Kn and τKn coincides with the weak
topology on Ac (Kn ) (see Proposition 4.31(f)), Mazur’s theorem (see Theorem A.2 or
Exercise 3.116) yields
k·kAc (Kn ) k·kAc (Kn )
Rf ∈ co R(BE ) = R(BE ) .
It follows that there exists a function g ∈ BAc (X) that is uniformly close to f on Kn
as much as we wish. Applying this argument to each Kn , we construct the required
sequence {fn } in BAc (X) .
To show (ii) =⇒ (i), assume that f satisfies (ii). First we notice that the assump-
tion ensures that given finitely many measures µ1 , . . . , µk in M+ (X), we can find a
sequence {fn } of continuous affine functions on X such that kfn k ≤ kf k, n ∈ N, and
fn → f µi -almost everywhere for each i = 1, . . . , k (we just apply the assumption to
a multiple of µ1 + · · · + µk ).
Let µ ∈ M1 (X) with x = r(µ) be given. We find a sequence {fn } of continuous
affine functions on X such that kfn k ≤ kf k, n ∈ N, fn → f µ-almost everywhere
and fn (x) → f (x) (use the preceding argument for the measures µ and εx ). Then
µ(f ) = lim µ(fn ) = lim fn (x) = f (x),
n→∞ n→∞
and the assertion follows since M is norm closed (see Proposition A.126).
4.5 Exercises 129
4.5 Exercises
Exercise 4.39. Let f be a continuous function on a compact convex set X and x ∈ X.
Then f ∗ (x) = sup{µ(f ) : µ ∈ Mx (X), µ is molecular}.
Hint. Use Lemma 3.21 and Proposition 4.3.
Exercise 4.40. If a compact convex set X in a locally convex space E admits a strictly
convex continuous function, then X is metrizable.
Hint. In order to show that X is metrizable, it suffices to find a countable family
B ⊂ C(K) such that B separates points on X (see Lemma 10.45).
So, suppose that f is a continuous strictly convex function on X. By Corollary 4.8
and Proposition 3.55, there exist kn ∈ N, n ∈ N, and functions fn = hn1 ∨ · · · ∨ hnkn ,
hni ∈ Ac (X), i = 1, . . . , kn , such that fn < f < fn + n1 , n ∈ N. We want to show
that the family
B := {hni : i = 1, . . . , kn , n ∈ N}
separates points on X.
Assume that there are distinct x, y ∈ X such that a(x) = a(y) for each a ∈ B.
Let z := 12 (x + y). For ε > 0, find a1 , a2 ∈ B such that f (x) < a1 (x) + ε and
f (y) < a2 (y) + ε. Then
1 1
f (z) ≤ (f (x) + f (y)) ≤ (a1 (x) + a2 (y)) + ε
2 2
1
≤ (f (z) + f (z)) + ε.
2
1
Hence f (z) = 2 (f (x) + f (y)) and f is not strictly convex.
Exercise 4.41. Prove that there exists a strictly convex function f on a compact convex
set X that is not Ac (X)-strictly convex.
Hint. Consider the compact convex set M1 ([0, 1]) and a strictly convex continuous
function f on M1 ([0, 1]), 0 ≤ f ≤ 21 . If
g : µ 7→ µd ([0, 1]), µ ∈ M1 ([0, 1]),
where µd denotes the discrete part of µ, then g is a Borel (in fact, by Proposition 2.63,
of the second Baire class) affine function on M1 ([0, 1]). Show that f − g is a strictly
convex function on M1 ([0, 1]) which is not strictly Ac (M1 ([0, 1]))-convex.
By Proposition 2.27, the mapping ε : x 7→ εx is a homeomorphism of [0, 1] onto
ext M1 ([0, 1]). Let Λ := ε] λ be the image of Lebesgue measure λ by ε. By Proposi-
tion 2.54, λ is a barycenter of Λ. Moreover,
1
Λ(f − g) = Λ(f ) − Λ(g) ≤ − 1 < 0 ≤ f (λ) = f (λ) − g(λ)
2
= (f − g)(λ).
130 4 Affine functions on compact convex sets
for any µ, ν ∈ M1 (X) with r(µ) = r(ν). Thus we may apply Lemma 4.11.
Exercise 4.45. Let X be a compact convex set. Prove Lemma 3.77 for this particular
case without the Simons lemma 3.75.
Hint. Let {fn } be an upper bounded sequence of lower semicontinuous convex func-
tions on X satisfying lim supn→∞ fn ≤ 0 on ext X. Without loss of generality we
may assume that {fn } is a bounded sequence. Given x ∈ X and ε > 0, we find con-
tinuous affine functions hn , n ∈ N, such that hn ≤ fn and hn (x) > fn (x) − ε (use
Corollary 3.23(b) and Proposition 3.25(a)). We define a continuous affine mapping
ϕ : X → RN by
ϕ(x) = {hn (x)}n∈N , x ∈ X,
and let Y := ϕ(X). Then Y is a metrizable compact convex set and ϕ(ext X) ⊃
ext Y (see Proposition 2.72(c)). Find a maximal measure µ ∈ Mϕ(x) (Y ) and let
h0n ∈ Ac (Y ) be such that h0n ◦ ϕ = hn , n ∈ N. Then lim supn→∞ h0n ≤ 0 on ext Y .
By Corollary 3.61 and Fatou’s lemma,
lim sup hn (x) = lim sup h0n (ϕ(x)) = lim sup µ(h0n )
n→∞ n→∞ n→∞
≤ µ(lim sup fn ) ≤ 0.
n→∞
Hence
lim sup fn (x) ≤ ε + lim sup hn (x) ≤ ε.
n→∞ n→∞
This concludes the proof.
4.5 Exercises 131
and
∞ ∞
X X 1
2−j ,
sn := yn := r j
εxj , n ∈ N.
sn 2
j=n+1 j=n+1
Then
n
X 1
x= x j + s n yn . (4.13)
2j
j=1
Exercise 4.47. Prove that there exists a function f on a compact convex set X sat-
isfying the barycentric formula such that f |F has no point of continuity for some
nonempty closed F ⊂ X.
Exercise 4.49. Let X be a compact convex set and A ⊂ Ac (X) be a vector space
containing the constant functions and separating points of X. Then A is dense in
Ac (X).
Hint. By the Hahn–Banach theorem it is enough to show that ϕ = 0 whenever ϕ ∈
(Ac (X))∗ satisfies ϕ = 0 on A. By Proposition 4.31(b) there exist numbers a1 , a2 ≥ 0
and x1 , x2 ∈ X so that ϕ(f ) = a1 f (x1 ) − a2 f (x2 ), f ∈ Ac (X). By the assumption,
0 = ϕ(1) = a1 − a2 . Hence f (x1 ) = f (x2 ) for each f ∈ A. Since A separates points
of X, x1 = x2 and ϕ = 0.
Exercise 4.50. Let X be a compact convex set and A ⊂ Ac (X) be convex. Prove that
τ
f ∈ Ac (X) is in the norm closure of A if and only if f ∈ A X .
Hint. This follows from Theorem A.2 and Proposition 4.31(f).
Exercise 4.51. Let H be a function space on a compact space K and let {1, f1 , . . . , fd }
be a basis of H. Let ψ : K → Rd be defined as
ψ(x) = (f1 (x), . . . , fd (x)), x ∈ K.
Prove that S(H) is affinely homeomorphic to co ψ(K).
Hint. Let ϕ : S(H) → Rd be defined by
ϕ(s) := (s(f1 ), . . . , s(fd )), s ∈ S(H).
Then ϕ is a homeomorphism of S(H) into Rd such that ϕ ◦ φ = ψ on K. By
Proposition 4.26(a),
ϕ(S(H)) = ϕ(co φ(K)) = co(ϕ ◦ φ(K)) = co ψ(K).
Since co ψ(K) = co ψ(K) by Corollary 2.8, S(H) is affinely homeomorphic to
co ψ(K).
4.6 Notes and comments 133
Exercise 4.52. Find a compact convex set X and a point x ∈ X such that face x is not
a face.
1 1
H := {f ∈ C(K) : f ( ) = (f (−1) + f (2))}.
2 2
Indeed, the inclusion face s ⊂ π(face λ) follows from the fact that π −1 (face s) is a
face containing λ (see Proposition 2.72).
For the converse inclusion, let t ∈ face s be given. Then there exist α ∈ (0, 1] and
t0 ∈ X such that s = αt + (1 − α)t0 . Let µ, µ0 ∈ M1 (K) satisfy π(µ) = t and
π(µ0 ) = t0 . We set A := {−1, 12 , 2} and find a sequence {fn } of functions from H
such that 0 ≤ fn ≤ 1, n ∈ N, and fn → cA . Then the equality
Hence µ ∈ face λ by Exercise 2.127(a) and t ∈ π(face λ). This proves (4.14).
By (4.14), Proposition A.40 and Exercise 2.127(b),
Hence φ 1 ∈ face s, φ 1 = 21 (φ−1 + φ2 ) and φ−1 , φ2 are not contained in face s. Thus
2 2
face s is not a face.
Section 4.2 follows mainly the paper R. Haydon [220]. Choquet’s barycentric the-
orem 4.22 is proved by G. Choquet in [106] (see also E. M. Alfsen [5], Theorem I.2.6,
or R. R. Phelps [374]). A slightly simplified proof can be found in J. Lukeš, J. Malý,
I. Netuka, M. Smrčka and J. Spurný [319]. G. Choquet himself also showed that the
barycentric formula is no longer valid for affine functions of the second Baire class
(see Proposition 2.63). Theorem 4.24 due to G. Mokobodzki can be found in M. Ro-
galski [392].
Results on the state space contained in Section 4.3 are standard; they are variants
of results in E. M. Alfsen [5], Chapter 2, §1 and L. Asimow and A. J. Ellis [24],
Chapter 2. Theorem 4.34 is taken from S. Teleman [449].
Section 4.4 is a well-known application of G. Mokobodzki’s theorem 4.24; it can
be found for example in S. A. Argyros, G. Godefroy and H. P. Rosenthal [19] and
E. Odell and H. P. Rosenthal [366].
Exercise 4.40 is proved in M. Hervé [223]. Exercise 4.45 is the classical way of
proving Lemma 3.77, see E. M. Alfsen [5], Proposition I.4.10. In A. Goullet de Rugy
[200], Exercise 4.46 is ascribed to G. Mokobodzki.
Chapter 5
Perfect classes of functions and representation of
affine functions
The aim of this chapter is to introduce a hierarchy of Borel sets in topological spaces,
show its relation to inductive generation of Baire and Borel functions, and apply these
facts to get a link between topological properties of function spaces and compact
convex sets.
The first part is devoted to an abstract generation of classes of sets and functions
from a given family of sets. It turns out that a suitable starting point for this procedure
is an algebra of sets, because then the standard properties of Borel sets hold (see
Proposition 5.4) as well as the characterization of Baire classes of functions via level
sets (see Theorem 5.9).
The next section employs an abstract procedure for creating classes of Baire and
Borel sets in topological spaces. The main result is Theorem 5.16, showing that Borel
classes are “quotient” with respect to perfect mappings. The key ingredient of the
proof is elementary Lemma 5.15.
Abstract results obtained so far are applied in Section 5.3 to get a characterization
of Baire and Borel mappings via their measurability (see Corollary 5.22). As another
consequence we get Theorem 5.23 characterizing Baire-one functions on compact
spaces. The nice behavior of our Borel classes with respect to perfect mappings is
pointed out by Theorem 5.26, which is a result widely used throughout the book.
The next section begins to describe the interplay between topological results and
properties of functions on compact convex sets. To emphasize crucial properties of
affine functions, we introduce in Definition 5.28 a notion of “affinely perfect func-
tions”. Theorem 5.31 and Corollary 5.32 show to what extent the behavior of an
affine function is determined by its properties on the closure of extreme points and
Theorem 5.33 provides examples of affinely perfect classes.
The last section is crucial for the transfer of properties of function spaces to the
context of compact convex sets. A precise description of this transfer is contained in
Theorem 5.40 and Corollary 5.41.
First we recall several definitions.
and [
Πα (F) := Σβ (F) .
δ
β<α
Elements of the family Σα (F) are termed the sets of abstract additive class α and
elements of the family Πα (F) are called the sets of abstract multiplicative class α.
The sets in ∆α (F) = Σα (F) ∩ Πα (F) are the sets of abstract ambiguous class α.
Remark 5.3. The definition and notation of abstract Borel classes is chosen in such
a way that it coincides with the definition of Borel sets in metrizable spaces; see
Definition A.113, Proposition 5.14 or Section 11.B of [262].
Proposition 5.4. Let F be an algebra of sets in a set X. Then the following assertions
hold:
(a) Σα (F) ∪ Πα (F) ⊂ ∆β (F), 0 < α < β < ω1 ,
(b) a set A ⊂ X is in Σα (F) if and only if X \ A ∈ Πα (F), α ∈ (0, ω1 ),
(c) the family Σα (F) is stable with respect to countable unions and finite intersec-
tions, Πα (F) is stable with respect to finite unions and countable intersections
and ∆α (F) is an algebra, α ∈ (1, ω1 ),
S S
(d) α<ω1 Σα (F) = α<ω1 Πα (F) is the σ-algebra generated by F,
(e) Σα (F) = (∆α (F))σ , α ∈ (1, ω1 ),
(f) for each α ∈ (0, ω1 ), Σα (F) has the generalized reduction property; that is, if
An ∈ Σα (F), n ∈ N, then pairwise disjoint sets A0n ∈ Σα (F), n ∈ N,
there existS
such that An ⊂ An and n=1 An = ∞
0
S∞ 0
n=1 An ,
(g) if α ∈ (0, ω1 ) and A ∈ ∆α+1 (F), then there exists a sequence {An } of sets such
that An ∈ ∆α (F), n ∈ N, and cA = limn→∞ cAn ,
(h) if α ∈ (1, ω1 ) is a limit ordinal
S and A ∈ ∆α+1 (F), then there exists a sequence
{An } of sets such that An ∈ β<α ∆β (F), n ∈ N, and cA = limn→∞ cAn .
5.1 Generation of sets and functions 137
Proof. Assertions (a), (b) and (c) follow by a straightforward transfinite induction, (d)
follows from (a) and (c), and (e) follows from (a).
For the proof of (f), let An ∈ Σα (F), n ∈ N, be given. If α = 1, we use the fact
that F Sis an algebra to get the required sets A0n , n ∈ N. Otherwise, using (e), we write
An = ∞ k=1 Ank , where Ank ∈ ∆α (F). We enumerate the family {Ank : n, k ∈ N}
0
as {Bj } and define
Bn ⊂ An ⊂ Cn .
Without loss of generality we may assume that the sequences {Bn }, {Cnk }∞ k=1 are
increasing and the sequences {Cn }, {Bnk }∞
k=1 are decreasing. Then the sets
n
[ p
\
An = (Bpn ∩ Cjn ), n ∈ N,
p=1 j=1
are contained in ∆αn (F) for some αn < α and limn→∞ cAn = cA .
Indeed, let x ∈ A be given. First we show that x is contained in An for all n ∈ N
sufficiently large. To this end, let w ∈ N be chosen in such a way that x ∈ Bw . Then
x ∈ Bwp for all p ∈ N. Let r ∈ N be such that
Then
w
\
x ∈ Bwn ∩ Cjn , if n > r.
j=1
Remark 5.6. Later on (see Theorem 5.9 and Definition 5.18), it will be sometimes
convenient to denote the starting family of the inductive definition by Φ1 . More pre-
cisely, we start from a family
S denoted by Φ1 and then Φα consists of all pointwise
limits of sequences from 1≤β<α Φβ , α ∈ (1, ω1 ). The purpose of this convention
is that we want to start the generation of mappings between topological spaces from
“Baire-one” mappings.
g(x) = yi , x ∈ Ai , i = 1, . . . , n.
The proof of Theorem 5.9 below closely follows the proof of Theorem 24.3 in
[262]; however, we include the following lemma.
Lemma 5.8. Let F be an algebra of sets in a set X and let 0 < ε < ε0 . Let f, g :
X → Y be mappings from a set X to a separable metric space (Y, ρ) such that
ρ(f (x), g(x)) < ε, x ∈ X. Let fn , gn : X → Y , n ∈ N, be Σγn (F)-measurable
mappings such that fn → f , gn → g and γn ∈ (1, ω1 ).
Then there exist Σγn (F)-measurable mappings hn : X → Y , n ∈ N, so that
hn → g and ρ(fn (x), hn (x)) < ε0 for each x ∈ X and n ∈ N.
Proof. Given the objects as in the lemma, for n ∈ N we set
A1n := {x ∈ X : ρ(fn (x), gn (x)) < ε0 }, A2n := {x ∈ X : ρ(fn (x), gn (x)) > ε}.
By the separability of the space Y we get that the sets A1n , A2n are in Σγn (F).
Indeed, we select countably many open sets Uk , Vk , k ∈ N, in Y such that
∞
[
(Uk × Vk ) = {(y1 , y2 ) ∈ Y × Y : ρ(y1 , y2 ) < ε0 }.
k=1
Hence
∞
[
A1n = (fn−1 (Uk ) ∩ gn−1 (Vk ))
k=1
Theorem 5.9. Let F be an algebra of sets in a set X and let Y be a separable metriz-
able space. Let Φ1 stand for the family of all Σ2 (F)-measurable mappings from X to
Y and, for α ∈ (1, ω1 ), let Φα be defined from Φ1 as in Definition 5.5.
Then, for each α ∈ (0, ω1 ) and f : X → Y , the following assertions are equiva-
lent:
140 5 Perfect classes of functions and representation of affine functions
(i) f ∈ Φα ,
(ii) f is Σα+1 (F)-measurable.
and the sets fk−1 (Um ) ∈ Παk +1 (F) by the inductive assumption. Thus f −1 (U ) ∈
Σα+1 (F).
For the proof of (ii) =⇒ (i) we use transfinite induction again. The case α = 1
follows from the definition. Assume that α ∈ (1, ω1 ) and the assertion holds for all
β < α. Let f : X → Y be a Σα+1 (F)-measurable function.
We assume first that f : X → Y has only finitely many values; that is, there
exist pairwise disjoint sets Ai ⊂ X and points yi ∈ Y , i = 1, . . . , n, such that
Ai ∈ ∆α+1 (F) and f = yi on Ai , i = 1, . . . , n.
Assume first that α is isolated, that is, α = α0 + 1 for some α0 ∈ [1, ω1 ). According
to Proposition 5.4(g), there exist sequences {Aki }∞ k=1 of sets in ∆α (F), i = 1, . . . , n,
such that cAk → cAi . We may assume that {Ak1 , . . . , Akn } is a disjoint family for each
i
k ∈ N. (Otherwise we would take B1k = Ak1 and Bjk = Akj \ (Ak1 ∪ · · · ∪ Akj−1 ),
j = 2, . . . , n.)
Then
fk (x) = yi , x ∈ Aki , i = 1, . . . , n, k ∈ N,
If x ∈ X, from ρ(f k (x), f (x)) < 2−k it follows that ρ(f k (x), f k+1 (x)) < 2−k+1 .
Now we employ Lemma 5.8 to replace {fnk } by {hkn } in such a way that
• h1 = f 1 , n ∈ N,
n n
ρ(hkn (x), hk+1 −k+2 , x ∈ X, n, k ∈ N,
n (x)) < 2
•
h1n := fn1 , n ∈ N,
Then the mappings hk are Σγkk +1 (F)-measurable, where γkk < α, and hk → f . By
the inductive assumption, hk ∈ Φγkk , k ∈ N. Hence f ∈ Φα , as required.
Theorem 5.10. Let F be an algebra of sets in a set X and let Φ1 stand for the family
of all Σ2 (F)-measurable functions from X to R and, for α ∈ (1, ω1 ), let Φα be the
abstract Baire class defined from Φ1 . Let α ∈ (0, ω1 ).
Then Φα is a vector space containing constant functions that is closed with respect
to uniform convergence, finite minima, multiplications and inversions (if defined).
Proof. Let α ∈ (0, ω1 ) be given. By Theorem 5.9, it is enough to show that the family
M of all Σα+1 (F)-measurable functions on X possesses all the required properties.
Obviously, M contains constant functions. Further, if f, g are Σα+1 (F)-measur-
able functions on X, the mapping ψ : X → R2 defined as
is Σα+1 (F)-measurable. (This follows from the fact that any open set in R2 is a
countable union of open rectangles.) If ϕ : R2 → R is any of the following mappings
where the union on the right-hand side is disjoint. Indeed, let I, J ⊂ {1, . . . , n} be
distinct sets. Assume that j ∈ I \ J. If
\ [ \ [
x ∈ Hi \ Fi ∩ Hi \ Fi ,
i∈I i∈{1,...,n}\I i∈J i∈{1,...,n}\J
5.2 Baire and Borel sets 143
then [
x ∈ Hj ⊂ Fj and x∈
/ Fi .
i∈{1,...,n}\J
where the last set is a finite disjoint union of differences of sets from F. Thus inter-
section of any pair of sets from B is again in B, and, consequently, B is stable with
respect to finite intersections.
Thus B is an algebra and A ⊂ B.
Definition 5.13 (Baire and Borel sets). We consider the following families of subsets
of X.
(a) The algebra Bas(X) generated by zero sets. By Lemma 5.12,
n
[
Bas(X) = (Fi \ Hi ) : Fi , Hi is a zero set in X, n ∈ N .
i=1
(c) If we start the Borel hierarchy as defined in Section 5.1 from the sublattice G(X)
of all open subsets of X, for metrizable spaces we get the standard Borel classes
Σ0α (X) and Π0α (X) as defined in Definition A.113. We denote as Σ0α (G(X)) and
Π0α (G(X)) the families obtained by this procedure. We show below its relation to the
families defined in (a), (b). We just mention that a set A belongs to Σ02 (G(X)) if and
only if A = F ∪ G, where F is Fσ and G is open.
For both families (a) and (b) above we consider the abstract Borel classes defined in
Section 5.1 and call them the sets of additive, or multiplicative, Borel class α and the
sets of additive, or multiplicative, Baire class α for α < ω1 , respectively.
144 5 Perfect classes of functions and representation of affine functions
Lemma 5.14. Let X be a topological space. Then the following assertions hold.
(a) Σα (Bas(X)) ⊂ Σα (Bos(X)), α ∈ (0, ω1 ).
S
(b) The family
S α<ω1 Σα (Bas(X)) is the σ-algebra of all Baire sets in X and the
family α<ω1 Σα (Bos(X)) is the σ-algebra of all Borel sets in X.
(c) If A is a subset of a normal space, then A ∈ ∆2 (Bas(X)) if and only if A is both
Fσ and Gδ .
(d) If X is metrizable, then
(d1) Σα (Bas(X)) = Σα (Bos(X)), α ∈ (0, ω1 ),
(d2) Σα (Bos(X)) = Σ0α (G(X)) = Σ0α (X), α ∈ (1, ω1 ).
Proof. Assertion (a) follows from the fact that Bas(X) ⊂ Bos(X) and (b) is obvious.
For the proof of (c), notice that any set in ∆S2 (Bas(X)) T
is both Fσ and Gδ in any
topological space. If X is normal and A = ∞ n=1 Fn = ∞
n=1 Gn , where Fn ⊂ X
closed, Gn ⊂ X open, n ∈ N, then there exist continuous functions fn : X → [0, 1]
so that (
1 on Fn ,
fn = n ∈ N.
0 on X \ Gn ,
Then fn → cA , and hence
∞ \
∞ n ∞ ∞
[ 1o \ [ n 1o
A= x ∈ X : fn (x) ≥ = x ∈ X : fn (x) >
2 2
k=1 n=k k=1 n=k
is in ∆2 (Bas(X)).
Assertions (d1) and (d2) follow from the fact that any closed subset of a metrizable
space is a zero set. This concludes the proof.
Φ−1 (F ) = {y ∈ Y : Φ(y) ∩ F 6= ∅}
is closed in Y for each closed set F in X. Let {Fn } be a sequence of closed subsets
of X.
Then there exists a mapping φ : Y → X such that
(a) φ(y) ∈ Φ(y) for every y ∈ Y ,
(b) φ−1 (Fn ) ∩ φ−1 (X \ Fn ) = ∅ for each n ∈ N, and
(c) φ−1 (Fn ) ∈ Bos(Y ) for each n ∈ N.
condition (b) is satisfied and φ−1 (Fn ) = (Φn )−1 (Fn ). Thus (c) follows from condi-
tion (f).
is a closed set in Y for each closed F ⊂ X and hence the assumptions of Lemma 5.15
are satisfied.)
Since B = φ−1 (A),
Theorem 5.17. Let A be a Baire subset of a compact space X and α ∈ (0, ω1 ). Then
A ∈ Σα (Bas(X)) if and only if A ∈ Σα (Bos(X)).
The following theorem justifies the term “measurability” in the definition above.
(iii) f is fragmented,
(v) f ∈ B 1 (X).
Proof. It follows from Theorem A.121 that (i) =⇒ (ii) ⇐⇒ (iii) ⇐⇒ (iv). By
Theorem A.124, (i) ⇐⇒ (v). To finish the proof we show that (iii) =⇒ (v). Since
f −1 (U ) is a Baire set in X for each U ⊂ R open, Corollary 5.22 gives that f ∈
B α (X) for some α < ω1 . It is easy to see that there exists a family Φ := {fn :
n ∈ N} of continuous functions on X such that f ∈ Φα (see Definition 5.5 and cf.
Proposition A.48). If ϕ : X → RN is defined as
V := H \ ϕ(F \ U )
is a nonempty relatively open set in H such that diam g(V ) < ε. Hence g is frag-
mented.
By Theorem A.127, g is a Baire-one function on Y , and thus f is a Baire-one
function on X. This finishes the proof.
Proof. For the proof of the nontrivial implication (ii) =⇒ (i), notice that A :=
f −1 ((c, ∞)) is an open Baire set in X for each c ∈ R. Since any Baire subset of a
compact space is Lindelöf, A is a cozero set. By Proposition A.53, f is a limit of an
increasing sequence of continuous functions.
5.4 Perfect classes of functions 149
are continuous images of Baire sets in a compact space, they are K-analytic (see
Theorem A.111(a)) and hence g −1 (U ) is a Baire set by the separation principle (see
Theorem A.112). By Theorem 5.17, g −1 (U ) ∈ Σα+1 (Bas(Y )).
The proof of (d) follows the same line of argument as the proof of (c); in fact, it is
enough to use Theorem 5.16 together with Theorem 5.19.
It remains to prove the last assertion. Similarly to the previous paragraph, we need
to prove that A ⊂ Y is universally measurable if ϕ−1 (A) is universally measurable
in X. Let ν be a probability measure on Y . Since the mapping ϕ] : M1 (X) →
150 5 Perfect classes of functions and representation of affine functions
Thus ν(ϕ(N )) = 0 and A is a disjoint union of an Fσ set ϕ(H) and ν-null set ϕ(N ).
Hence A is ν-measurable and the proof is finished.
Corollary 5.27. The following classes of functions on topological spaces are perfect:
(a) increasing limits of continuous functions,
(b) lower semicontinuous and upper semicontinuous functions,
(c) Baire functions of class α and Baire functions,
(d) Borel measurable functions of class α and Borel functions,
(e) universally measurable functions.
Proof. Let µ ∈ M1 (X). Proposition A.92 yields that for every bounded universally
measurable function h on Y we have
(ϕ] µ)(h) = µ(h ◦ ϕ). (5.3)
We observe that r(ϕ] µ) = ϕ(r(µ)). Indeed, for any h ∈ Ac (Y ) we get
h(r(ϕ] µ)) = (ϕ] µ)(h) = µ(h ◦ ϕ) = (h ◦ ϕ)(r(µ)) = h(ϕ(r(µ))).
Assume now that g ∈ Abar (Y ). Then, for any µ ∈ M1 (X), we obtain
µ(g ◦ ϕ) = (ϕ] µ)(g) = g(r(ϕ] µ)) = (g ◦ ϕ)(r(µ)).
Thus g ◦ ϕ ∈ Abar (X).
Conversely, let g ◦ ϕ be in Abar (X). By Theorem 5.26(e), the function g is univer-
sally measurable on Y . It remains to prove the validity of the barycentric formula for
g.
To this end, let ν be a probability measure on Y . We find a probability measure
µ ∈ M1 (X) with ϕ] µ = ν.
By the assumption, the function f := g ◦ ϕ satisfies the barycentric formula on X.
Then (5.3) gives
ν(g) = (ϕ] µ)(g) = µ(g ◦ ϕ) = µ(f ) = f (r(µ)) = g(ϕ(r(µ))) = g(r(ν)),
which concludes the proof.
is Σ2 (Bos(M1 (K)))-measurable. S
If A ∈ ∆2 (Bos(K)), then A = ∞ n=1 An , where A1 ⊂ A2 ⊂ · · · and An ∈
Bos(K), n ∈ N. Then, for c ∈ R, the set
∞
[
{µ ∈ M1 (K) : IcA (µ) > c} = {µ ∈ M1 (K) : µ(An ) > c}
n=1
Proof. Let r : M1 (ChH K) → K be the barycentric mapping (see Section 3.3). Then
r is a surjective continuous mapping. The function
ν1 (f ) = µ1 (f ) = µ2 (f ) = ν2 (f ),
Corollary 5.32. Let X be a compact convex set in a locally convex space and f be a
bounded function on X such that, for any µ ∈ M1 (ext X), f is µ-measurable and
µ(f ) = f (r(µ)). Then f satisfies the barycentric formula on X.
Moreover, such a function f belongs to a class listed in Corollary 5.27 if f |ext X
belongs to this class.
154 5 Perfect classes of functions and representation of affine functions
Theorem 5.33. The classes of strongly affine Baire-α functions and the class of
strongly affine Baire functions are affinely perfect.
Proof. The property (AP1) follows from Proposition 5.29, (AP2) is obvious, (AP3)
follows from Proposition 5.30 and (AP4) from Theorem 3.86.
Remarks 5.34. (a) We recall that a function f on a topological space X has the Baire
property if f is measurable with respect to the σ-algebra of the sets with the Baire
property. A function f on a topological space X has the restricted Baire property if,
for any closed F ⊂ X, f |F has the Baire property. A set A ⊂ X is perfectly meager
if A ∩ P is meager in P for every perfect P ⊂ X.
To see that neither class of functions is perfect, we recall an example of W. Sier-
pinski who constructed in Théorème of [416] a perfectly meager set A ⊂ [0, 1] and
a continuous function ϕ : [0, 1] → [0, 1] such that ϕ−1 (A) does not have the Baire
property. Hence the characteristic function cA has the restricted Baire property and
cA ◦ ϕ does not have even the Baire property.
(b) Later on we will present several examples of classes of functions that are not
affinely perfect, see Exercise 9.59, Subsections 12.3.B and 12.3.C.
We note that an affine function f on a symmetric compact convex set is odd if and
only if f (0) = 0.
Lemma 5.39. Let F1 , F2 be closed convex subsets of a compact convex set X such that
co(F1 ∪ F2 ) = X. Let f be an affine function on X. Then the following assertions
hold.
(a) The function f is universally measurable (or of any class listed in Corollary 5.27,
respectively) on X if and only if f is universally measurable (or of any class listed
in Corollary 5.27, respectively) on F1 ∪ F2 .
(b) The function f belongs to Abar (X) if and only if f |F1 ∈ Abar (F1 ) and f |F2 ∈
Abar (F2 ).
Then fb inherits topological properties of f |F1 ∪F2 and fb = f ◦ ϕ. Thus Corollary 5.27
yields (a).
156 5 Perfect classes of functions and representation of affine functions
For the proof of the nontrivial part of (b), let f be an affine function on X such that
f |Fi ∈ Abar (Fi ), i = 1, 2. By (a), f is universally measurable on X.
Let x ∈ X and µ ∈ M1 (ext X) with r(µ) = x be given. Since X ⊂ co(F1 ∪ F2 ),
the Milman theorem 2.43 yields ext X ⊂ F1 ∪ F2 . We set
µ1 := a−1
1 µ|F1 and µ2 := a−1
2 µ|F2 \F1 .
Then r(µi ) ∈ Fi , i = 1, 2,
By the hypotheses,
Thus f satisfies the barycentric formula with respect to any measure carried by ext X.
By Corollary 5.32, f ∈ Abar (X). This concludes the proof.
= sup |f (x)| = kf k
x∈K
If
b (µ) = µ(f ), µ ∈ BM(K) .
If (s) = µ(f ) = a1 µ1 (f ) − a2 µ2 (f )
= a1 F (π(µ1 )) − a2 F (π(µ2 )) = F (π(µ))
= F (s).
Thus I −1 (F ) = F ◦ φ and we get the inclusions “⊃” in (a), (c) and (d).
Using the Lebesgue dominated convergence theorem, we get (e).
By transfinite induction and (e), I(Hα,b ) = Aodd (X) ∩ Aα (X) for α ∈ [0, ω1 ),
which finishes the proof.
s 7→ µ(f ), µ ∈ π −1 (s), s ∈ X.
Then I is an isometric isomorphism between U(K) ∩ H⊥⊥ and Abar (X) such that
I = Φ on H and I preserves natural order of functions. Its inverse is given by
Further, for any class of functions C listed in Corollary 5.27, If ∈ C(X) if and
only if f ∈ C(K).
158 5 Perfect classes of functions and representation of affine functions
Proposition 5.42. Let H be a function space on a compact space K. Then the follow-
ing assertions hold:
(a) U(K) ∩ H⊥⊥ ∩ B 1 (K) = H1,b ,
(b) If f ∈ U(K) ∩ H⊥⊥ is lower semicontinuous, then f = sup{h ∈ H : h < f }
and this family is up-directed.
Proof. Let X be the state space of H, Φ : H → Ac (X) be the isometry into and I be
as in Corollary 5.41.
For the proof of (a) we notice that If ∈ B 1 (X) ∩ Abar (X). By Theorem 4.24,
If ∈ A1 (X); that is, there exists a bounded sequence {hn } ⊂ Ac (X) pointwise
converging to If . Since Φ(H) is dense in Ac (X), we may assume that hn ∈ Φ(H),
n ∈ N. Then Φ−1 hn → f , and f ∈ H1 .
Similarly we verify (b). Indeed, If is lower semicontinuous and affine provided
f ∈ H⊥⊥ is lower semicontinuous. By Proposition 4.12,
where this family is up-directed. Thus the set {h ∈ H : h < f } is up-directed as well
and its supremum equals f .
Proof. The implication (i) =⇒ (ii) is obvious. To show the converse implication, let
f, −g be upper semicontinuous functions on K satisfying (ii). Then f < g. Define
functions F and G on the state space X := S(H) as follows:
(
f ◦ φ−1 on φ(K),
F :=
min f (K) on S(H) \ φ(K),
and (
g ◦ φ−1 on φ(K),
G :=
max g(K) on S(H) \ φ(K).
Then F , −G are upper semicontinuous functions on X with F < G. Let µ1 , µ2 ∈
M1 (X) with r(µ1 ) = r(µ2 ) be given. Let νi ∈ M+ (X) be maximal such that
µi |X\φ(K) ≺ νi , i = 1, 2.
Then
λi := µi |φ(K) + νi , i = 1, 2,
5.7 Exercises 159
(φ−1 )] λ1 − (φ−1 )] λ2 ∈ H⊥ .
By the assumption,
H ∈ ((H∗ , w∗ )∗ + R) |X = Φ(H)|X
5.7 Exercises
Exercise 5.44. For every α ∈ (1, ω1 ) there exists a Baire subset A of a normal space
Xα such that A is of additive Borel class 1 and A is not of additive Baire class α + 1.
Hint. If denotes the Euclidean topology of the real line R, let A ⊂ R be a set in
Πα (Bas(R, )) \ Σα (Bas(R, )) (see [262], Theorem 22.4). Let τ be the topology on
R such that its open sets are of the form U ∪ V , where U is -open and V is any subset
of R \ A. It is easy to see that (R, τ ) is a Hausdorff regular space. It is even a normal
space.
Indeed, let U be an open cover of (R, τ ). Since (A, τ ) = (A, ), we can find a
countable open family U 1 ⊂ U covering A. Then
V = U 1 ∪ {{x} : x ∈ R \ A}
Exercise 5.45. There exists a normal space X such that for each α ∈ (1, ω1 ) there
exists a Baire set A ⊂ X such that A ∈ Σ1 (Bos(X)) \ Σα+1 (Bas(X)).
Hint. For every α ∈ (1, ω1 ), we construct Xα as in Exercise 5.44. Then the topolog-
ical sum of these spaces is the space with the required properties.
160 5 Perfect classes of functions and representation of affine functions
Exercise 5.46. Let X be a normal space and F1 , . . . , Fn be its closed subsets. Then
βX βX βX
F1 ∩ · · · ∩ Fn = F1 ∩ · · · ∩ Fn .
βX
Hint. First we show the following claim: If F1 , F2 ⊂ X are closed and F1 ∩
βX
F2 6= ∅, then F1 ∩ F2 6= ∅.
To show this assume that F1 ∩ F2 = ∅. By Tietze’s theorem A.26, there exists a
bounded continuous function f on X such that f = 0 on F1 and f = 1 on F2 . Let
fb ∈ C(βX) be the continuous extension of f . Then
βX βX
F1 ⊂ {x ∈ βX : fb(x) = 0} and F2 ⊂ {x ∈ βX : fb(x) = 1},
βX βX
and thus F1 ∩ F2 = ∅, a contradiction. This proves the claim.
βX βX
Now let F1 , F2 ⊂ X be closed with x ∈ F1 ∩ F2 . Let U 3 x be an open set.
Then
βX βX
βX βX
x ∈ F1 ∩ (X ∩ U ) ∩ F2 ∩ (X ∩ U ) .
By the claim,
βX
F1 ∩ F2 ∩ (X ∩ U ) 6= ∅.
βX
Since U is arbitrary, x ∈ F1 ∩ F2 .
βX
By induction we prove the assertion for finitely many sets in X. If x ∈ F1 ∩
βX
· · · ∩ Fn+1 , where F1 , . . . , Fn+1 are closed in X, then
βX βX
x ∈ F1 ∩ · · · ∩ Fn ∩ Fn+1 ,
and thus
βX
x ∈ F1 ∩ · · · ∩ Fn+1 .
Exercise 5.47. Let (NN , ) denote the space NN endowed with the usual product topol-
ogy and let X = NN ∪ {p} equipped with the following topology τ : the points of NN
are discrete and a base of neighborhoods of the point p is formed by the sets
and
Is = ϕ(Us ) = {σ ∈ NN : σ||s| = s} ∪ {p}, s ∈ N<N .
If s, t ∈ N<N are different sequences of length n, then Exercise 5.46 yields
βX βX βX
Is ∩ It = Is ∩ It = {p}. (5.8)
We claim that
∞ [
\ βX
X= Is . (5.9)
n=1 |s|=n
Exercise 5.48. Let X be a completely regular space and α ∈ (1, ω1 ). Prove that the
following assertions are equivalent:
(i) X ∈ Σα (Bos(cX)) for any compactification cX of X,
(ii) X ∈ Σα (Bos(cX)) for some compactification cX of X,
(iii) X ∈ Σα (Bos(βX)) for the Čech–Stone compactification βX of X.
Exercise 5.49. Let X be a completely regular space and α ∈ (0, ω1 ). Prove that the
following assertions are equivalent:
(i) X ∈ Σα (Bas(cX)) for any compactification cX of X,
(ii) X ∈ Σα (Bas(cX)) for some compactification cX of X,
(iii) X ∈ Σα (Bas(βX)) for the Čech–Stone compactification βX of X.
Hint. Follow the proof of Exercise 5.48 along with Theorem 5.26(c).
Exercise 5.50. We recall that a compact space K is scattered if every subset of K has
an isolated point. Verify the following assertions:
(a) A compact space is scattered if and only if every closed subset of K has an iso-
lated point.
(b) A continuous image of a scattered compact space is scattered.
(c) If K is not scattered, then there exists a perfect subset P ⊂ K and a continuous
surjection ϕ : P → {0, 1}N .
(d) If K is a metrizable compact scattered space, then it is countable.
Exercise 5.51. Prove that a compact space K is scattered if and only if every f ∈
C(K) has countable range.
Exercise 5.52. Any scattered compact space K has a base consisting of clopen sets.
and
{z ∈ K : f (z) > r} = {z ∈ K : f (z) ≥ r}
are clopen sets separating x and y. From this the assertion follows.
164 5 Perfect classes of functions and representation of affine functions
Hint. If B ⊂ F is a zero set, by Tietze’s theorem there exists a zero set A ⊂ K with
A ∩ F = B. If F denote the family of all Baire sets in F that are traces of Baire
sets in K, it is easy to verify that F is a σ-algebra. Hence F contains all Baire sets in
F.
Exercise 5.54. Prove that for a compact space K the following assertions are equiva-
lent:
(i) K is not scattered,
(ii) for any α < ω1 there exists a Baire set A ∈
/ Σα (Bas(K)).
Hint. Assume that K is scattered and A ⊂ K is a Baire set. Using Proposition A.48
we find a countable family {fn : n ∈ N} of continuous functions such that A is
contained in the σ-algebra generated by {fn−1 ((0, 1]) : n ∈ N}. Then ϕ : K → RN
defined as
ϕ(x) := {fn (x)}n∈N , x ∈ K,
is a continuous surjection of K on a metrizable compact space ϕ(K). By Exer-
cise 5.50(b), (d), ϕ(K) is a countable compact space. Since A = ϕ−1 (ϕ(A)) and
Exercise 5.55. If K is a compact space that is not scattered, then for any α < ω1 there
exists a Borel set A ∈
/ Σα (Bos(K)).
Hint. If K is not scattered and α < ω1 , Exercise 5.54 provides a Baire set A ⊂ K not
contained in Σα (Bas(K)). By Theorem 5.17, A is not contained in Σα (Bos(K)).
Φ0 (y) := Φ(y), y ∈ Y.
Let
∞
\
φ(y) := Φn (y), y ∈ Y.
n=0
x1 ∈ φ(y) ⊂ Φn (y),
Φn (y) ∩ Fn+1 6= ∅. Hence Φn+1 (y) ⊂ Fn+1 , and thus x2 ∈ / Φn+1 (y). Since
φ(y) ⊂ Φn+1 (y), we get a contradiction.
Further, φ(y) ⊂ Φ(y) for each y ∈ Y . This follows from the first step of the
construction.
Finally, since
φ−1 (Fn+1 ) = (Φn )−1 (Fn+1 ) ∈ Bos(Y ),
and any open set in X is a countable union of elements from {Fn : n ∈ N}, φ is
Σ2 (Bos(Y ))-measurable mapping. In other words, φ ∈ Bof1 (X, Y ).
166 5 Perfect classes of functions and representation of affine functions
Exercise 5.58. Let E be a Banach space and X := BE ∗ with the w∗ -topology. We set
E0 := E ⊂ E ∗∗ and inductively define
[
Eα := {x∗∗ |X : x∗∗ is a w∗ -limit of a sequence from Eβ }, α ∈ (0, ω1 ).
β<α
simplices by means of their facial structure are described in Theorem 6.69. Fakhoury’s
characterization of simplices is contained in Theorem 6.70.
We conclude the chapter with a simple result in Section 6.9 on restrictions of func-
tion spaces.
As usually, H will stand for a function space on a compact space K.
Lemma 6.4. If H is simplicial, then f ∗ (x) = δx (f ) for any x ∈ K and f ∈ Kusc (H).
Proof. Let f ∈ Kusc (H) and x ∈ K be given. Key Lemma 3.21 provides a measure
µ ∈ Mx (H) so that µ(f ) = f ∗ (x). By Lemma 6.3(a), µ ≺ δx . Hence
f ∗ (x) = µ(f ) ≤ δx (f ) ≤ f ∗ (x),
which concludes the proof.
Proof. For the proof of (i) =⇒ (iv), let f ∈ Kusc (H) be given. We fix x ∈ K and
µ ∈ Mx (H). By Lemma 6.3(a), µ ≺ δx . Using Lemma 3.18(b), Lemma 6.4, and
Lemma 3.18(a) we get
Hence f ∗ is H-affine.
Obviously, (iv) =⇒ (iii) =⇒ (ii). To close the chain of implications, assume
that (ii) holds. Let x ∈ K and µ, ν ∈ Mx (H) be maximal measures. Then, for each
f ∈ −W(H), Theorem 3.58 yields
F := {f − g + ε : s ≤ f ≤ g ≤ t, f, −g ∈ Kc (H)} .
Since the cone S c (H) is min-stable, by Proposition 3.53 there exists a function g ∈
S c (H) such that s∗ ≤ g ≤ k ∧ t. By Theorem 6.5, s∗ is an H-affine function. Now,
by an analogous reasoning, there exists f ∈ Kc (H) such that s∗ ≤ f ≤ g ≤ t. Then
and therefore
µ(f − g + ε) = εµ(K) + µ(f − g) > 0.
6.1 Basic properties of simplicial spaces 171
With this in hand, it is now possible to construct sequences {fn } in Kc (H) and {gn }
in S c (H) so that
1
s ≤ fn ≤ fn+1 ≤ gn+1 ≤ gn ≤ t and 0 < gn − fn <
2n
for each n ∈ N. Both sequences converge uniformly on K to an H-affine function h
satisfying s ≤ h ≤ t.
Since obviously (iii) =⇒ (ii), all that remains to be proved is that (ii) =⇒ (i).
Pick x ∈ K and consider maximal measures µ, ν ∈ Mx (H). If f ∈ Kc (H), then by
assumption (ii), it follows that the set {h ∈ Ac (H) : h ≥ f } is down-directed. Hence,
using Proposition 3.25(a) and Theorem A.84,
Since the space Kc (H)−Kc (H) is uniformly dense in C(K), it follows that µ = ν.
Proof. For the proof of (a), let f ∈ C(K) be given. By Proposition 3.11(c), there
exist sequences {fn }, {gn } of H-convex continuous functions such that fn − gn → f
uniformly. Then T (fn −gn ) → T f uniformly, and so Lemma 6.4 concludes the proof
of (a) (note that semicontinuous functions on metrizable spaces are Baire-one and
Baire-one functions are stable with respect to uniform convergence, see Section A.5).
To verify (b), let B consist of all bounded Baire functions f on K, for which T f
is Borel. It is easy to check that B is vector space that is stable with respect to tak-
ing pointwise limits of bounded sequences. Since C(K) ⊂ B by (a), B contains all
bounded Baire functions by Proposition A.47.
For the proof of (c), we fix f ∈ Kc (H). Let µ, ν ∈ M+ (K) satisfy µ − ν ∈
(Ac (H))⊥ . By Theorem 6.6, the family inf{h ∈ Ac (H) : h ≥ f } is down-directed.
Using Proposition 3.25(a) and Theorem A.84,
Proposition 6.9. Let H be a function space on a compact space K. Then the following
conditions are equivalent:
(i) H is simplicial,
(ii) if µ, ν ∈ Mmax (H) with µ − ν ∈ (Ac (H))⊥ , then µ = ν,
(iii) (Ac (H))⊥ ∩ Mbnd (H) = {0}.
By Proposition 3.11(c), µ = ν.
Obviously, (ii) =⇒ (iii). Assuming (iii), let x ∈ K and µ, ν ∈ Mx (H) be
maximal. Since µ − ν ∈ (Ac (H))⊥ , µ = ν by the assumption. Hence (iii) =⇒ (i).
Remark 6.10. Example 3.2(b) shows that a closed function space H can be simplicial
even though H⊥ ∩ Mbnd (H) 6= {0}.
6.1 Basic properties of simplicial spaces 173
imply that T µ is H-maximal (see Theorem 3.58). Finally, for such a function f ∈
Kc (H) we get
Z Z
µ(f ) = f (x)dµ(x) ≤ δx (f )dµ(x) = T µ(f ).
K K
We use Example 3.82 as a starting point for a general construction leading to var-
ious examples of simplicial function spaces. In particular, we get that the space con-
structed in Example 3.82 is a simplicial function space.
Proof. For a point x in some Lb and a relatively compact open set U ⊂ Lb containing
x, let fb : Lb → [0, 1] be a continuous function such that fb (x) = 0 and fb = 1 on
Lb \ U . Then the function
(
fb on Lb ,
fU :=
1 on K \ Lb ,
belongs to H, and thus shows that the support of any measure H-representing x is
contained in U . Hence Mx (H) = {εx } and K \ L ⊂ ChH (K).
6.1 Basic properties of simplicial spaces 175
belongs to H,Sand shows that the support of any measure H-representing x is con-
tained in U ∪ b∈B∩U Lb . Hence L \ B ⊂ ChH (K). Obviously, B ⊂ K \ ChH (K)
and hence ChH (K) = K \ B.
Let b ∈ B be arbitrary. Similar functions as in (6.1) show that any measure µ ∈
Mb (H) is carried by {b} ∪ Lb . Let µ ∈ Mb (H) satisfy µ({b}) = 0. Let g1 : Lb →
[0, 1] be a continuous function with compact support and ε > 0. We find a continuous
function g2 : {b} ∪ Lb → [0, 1] such that |g2 (b) − µb (g1 )| < ε and and g1 = g2 on the
support of g1 . Then the function
(
g2 (x), x ∈ {b} ∪ Lb ,
f (x) :=
g2 (b), x ∈ K \ ({b} ∪ Lb ),
belongs to H. Thus
µb (f ) = f (b) = µ(f ).
Example 6.15. There exists a simplicial function space on a compact space K and a
bounded Borel function f on K such that T f is not universally measurable.
T f |L = cL\B .
f1 ∨ f2 < f1 ∨ f2 + ε ≤ h ≤ g1 ∧ g2 − ε < g1 ∧ g2 .
f1 ∨ · · · ∨ fm < h < g1 ∧ · · · ∧ gn .
6.2 Characterizations of simplicial spaces 177
For the proof of (v) =⇒ (iii), let f, −g ∈ Kc (H) with f < g be given. Find ε > 0
g ∈ W(Ac (H))
so that f + ε < g − ε. By Proposition 3.55, there exist functions fb, −b
with
f − ε ≤ fb ≤ f and g ≤ gb ≤ g + ε.
f ≤ fb + ε < h < gb − ε ≤ g,
Since the latter family is down-directed by our assumption, we can employ Theo-
rem A.84 and deduce that f ∗ is H-affine. Hence (iii) =⇒ (i), which concludes the
proof.
Remark 6.17. The conditions from Theorem 6.16(vi) and (vii) express that the space
Ac (H) has the Riesz interpolation property and the weak Riesz interpolation property,
respectively.
Proof. By Proposition A.11 we know that condition (ii) is equivalent to condition (vi)
of Theorem 6.16.
Remark 6.19. If the function space Ac (H) satisfies the requirements of Theorem
6.18(ii), we say that Ac (H) has the Riesz decomposition property.
Proof. Assume that H is a simplicial function space and {B(hi , ri )}ni=1 is a finite
family of closed balls in Ac (H) with centers hi and radii ri , i = 1, . . . , n, such that
B(hi , ri ) ∩ B(hj , rj ) 6= ∅, i, j = 1, . . . , n. Then the functions fi = hi − ri and
gi = hi + ri , i = 1, . . . , n, satisfy
f1 ∨ · · · ∨ fn ≤ g1 ∧ · · · ∧ gn .
f1 ∨ · · · ∨ fn ≤ h ≤ g1 ∧ · · · ∧ gn .
For the proof of (ii) =⇒ (i), let f, g1 , g2 ∈ Ac (H) be positive functions such that
f ≤ g1 + g2 . Without loss of generality we may assume that g1 + g2 ≤ 1 on K. Then
the closed balls
0 ∈ B1 ∩ B2 ∩ B3 , f ∈ B1 ∩ B3 ∩ B4 and g1 ∈ B2 ∩ B4 .
Remark 6.21. If the function space Ac (H) satisfies the requirements of Theorem
6.20(ii), we say that Ac (H) has the finite binary intersection property.
Proof. The proof follows from Theorem 6.20, because the finite binary intersection
property of a Banach space is preserved by an isometric isomorphism.
Proof. Let R : Mbnd (H) → (Ac (H))∗ be the restriction mapping. By Proposi-
tion 6.9, R is injective. Using Lemma 6.23 applied to Ac (H) and Proposition 3.67,
we infer that R is a surjective isometry. It is easy to check that Rµ is a positive func-
tional on Ac (H) if and only if µ ∈ Mbnd (H) is a positive measure (use Lemma 6.23).
Hence R preserves order, which concludes the proof.
Proof. By Proposition 6.24 it is enough to show that Mbnd (H) is order isometric to
L1 (X, Σ, σ) for a suitable measure space (X, Σ, σ). We use throughout the proof the
identification of Radon measures on K with finite measures defined on Borel sets that
are inner regular with respect to compact sets (see Proposition A.73).
Using Zorn’s lemma, we find a set {µi : i ∈ I} of pairwise singular probability
measures in Mmax (H) that is maximal with respect to inclusion. For each i ∈ I we
consider the measure space (K, Σi , σi ), where Σi denotes the σ-algebra of Borel sets
180 6 Simplicial function spaces
(T ν)|Ki := hi , i ∈ I.
If, moreover, K is metrizable, then H is simplicial if and only if the weak Dirichlet
problem for H is solvable.
Proof. The proof of the first part of the theorem is an immediate consequence of the
Edwards in-between theorem 6.6. Indeed, assume that H is simplicial, D ⊂ ChH (K)
compact and f ∈ C(D). Set
( (
min f (D), x ∈ K \ D, max f (D), x ∈ K \ D,
s(x) := t(x) :=
f (x), x ∈ D, f (x), x ∈ D.
(µ + ν)(K \ D) < ε.
Since the weak Dirichlet problem for H is solvable, there exists h ∈ Ac (H) such that
Then
|µ(f ) − ν(f )| ≤ |µ(f ) − h(x)| + |h(x) − ν(g)|
= |µ(f ) − µ(h)| + |ν(h) − ν(f )|
Z Z Z
≤ |f − h| dµ + |h − f | dν = |f − h| d(µ + ν)
K K K
Z
= |f − h| d(µ + ν) ≤ 2kf kε.
K\D
0 ≤ hy ≤ 1, hy (x) = 0, hy (y) = 1.
Since ChH (K) \ {x} is Lindelöf, there are yn ∈ ChH (K), n ∈ N, such that
∞
[
ChH (K) \ {x} ⊂ {t ∈ K : hyn (t) > 0} .
n=1
Set
∞
X 1
h := hy .
2n n
n=1
Then h ∈ Ac (H),
h ≥ 0 on K and h(x) = 0. It remains to show that h > 0 on
K \ {x}. Let z ∈ K satisfy h(z) = 0. Since h > 0 on ChH (K) \ {x} and the
maximal measure δz at z is carried by ChH (K), it follows that spt δz ⊂ {x}. Thus
z = x.
We remark that ChH (K)f∗,α ≤ ChH (K)f ∗,α due to the Minimum principle 3.86.
Theorem 6.31. Let α ∈ [0, ω1 ), f be a bounded function on ChH (K) and let g :=
ChH (K)f ∗,α . Then the following assertions are equivalent:
(ii) There exists a unique function h ∈ A(H)∩B bα (K) such that h = f on ChH (K).
(iii) There exists a function h ∈ A(H) ∩ B bα (K) such that h = f on ChH (K).
Proof. Let a bounded function f on ChH (K) be given. To start the proof we notice
that (ii) =⇒ (iii) is obvious. For the proof of (iii) =⇒ (i), let h ∈ B bα (K) ∩ A(H)
h ∈ B bα (K) ∩ A(H) be any function such that f ≤ b
extend f . Let b h on ChH (K). By
the Minimum principle 3.86, h ≤ h on K. Hence h ≤
b Ch H (K) ∗,α
f . Analogously we
obtain ChH (K)f∗,α ≤ h.
On the other hand,
h ≤ ChH (K)f∗,α ≤ ChH (K)f ∗,α ≤ h
by the definition. Hence h = ChH (K)f ∗,α = ChH (K)f∗,α on ChH (K) and property (i1)
follows.
Since (i2) and (i3) are obvious, the proof of (iii) =⇒ (i) is finished.
To close the chain of implications, we have to verify that (i) =⇒ (ii). It follows
from the Minimum principle 3.86 that the extension is unique, provided it exists. Let
f and g be as in (i). We denote Z := ChH (K).
Step 1: For any z ∈ Z and µ ∈ Mz (H) ∩ M1 (Z), we have µ(g) = g(z).
Indeed, let z ∈ Z and µ ∈ Mz (H) ∩ M1 (Z) be given. Then
µ(g) = µ inf{h ∈ A(H) ∩ B bα (K) : h ≥ f on ChH (K)}
It follows from Step 3 that h is well defined. Obviously, h = f on ChH (K). Finally,
Theorem 5.31 yields that h ∈ B bα (K) ∩ A(H). This finishes the proof.
Theorem 6.32. Let f be a bounded function on ChH (K) and let g := ChH (K)f ∗ . Then
the following assertions are equivalent:
(i) The function f satisfies
(i1) ChH (K)f ∗ = ChH (K)f∗ on ChH (K),
(i3) µ(g) = 0 for any boundary measure µ ∈ H⊥ .
(ii) There exists a unique function h ∈ H such that h = f on ChH (K).
(iii) There exists a function h ∈ H such that h = f on ChH (K).
Proof. The proof follows from Theorem 6.31 once we realize that (i1) implies the
continuity of g on ChH (K). Hence we can find a function h ∈ Ac (H) extending g.
To finish the proof we have to verify that h ∈ H. Since µ(f ) = 0 for any boundary
µ ∈ H⊥ , Theorem 5.31 gives h ∈ H⊥⊥ . Hence h ∈ H by the Hahn–Banach theorem.
This concludes the proof.
We recall that H1,b stands for the space of all functions of H-affine class 1 (see
Definition 5.35).
Theorem 6.33. Let f be a bounded function on ChH (K) and let g := ChH (K)f ∗,1 .
Proof. The proof again follows from Theorem 6.31. The only fact we need to verify
is that the obtained extension h is contained in H1,b . We already know that h ∈
B b1 (K) ∩ A(H). Using assumption (i3) we get from Theorem 5.31(b) that h ∈ H⊥⊥ .
It follows from Proposition 5.42(a) that h ∈ H1,b . This concludes the proof.
Remark 6.34. Inspection of the previous proofs shows that the whole procedure can
be done for any affinely perfect class C of functions on compact convex sets (see
Section 5.5). We define envelopes similarly to above, namely, for a bounded function
f defined at least on the set ext X, we set
ext X ∗,C
f := inf{h ∈ C(X) : h ≥ f on ext X} and
ext X
f∗,C := −ext X(−f )∗,C .
Thus we obtain the following result.
Theorem 6.35. Let X be a compact convex set, C be an affinely perfect class of func-
tions, f be a bounded function on ext X and let g := ext Xf ∗,C . Then the following
assertions are equivalent:
(i) The function f satisfies
(i1) ext Xf ∗,C = ext Xf∗,C on ext X,
(i2) the function µ 7→ µ(g), µ ∈ M1 (ext X), is contained in C(M1 (ext X)),
(i3) µ(g) = 0 for any boundary measure µ ∈ A(X)⊥ .
(ii) There exists a unique function h ∈ C(X) such that h = f on ext X.
(iii) There exists a function h ∈ C(X) such that h = f on ext X.
Proof. For the proof that (i) =⇒ (ii), suppose that x ∈ K and µ ∈ Mx (H) with
spt µ ⊂ ChH (K) are given. If f ∈ C(K), then by Bauer’s characterization of the
Choquet boundary 3.24,
Hence
µ {x ∈ K : f (x) < f ∗ (x)} ≤ µ K \ ChH (K)) = 0.
This obvious contradiction yields the required conclusion, and therefore (v) =⇒
(vi).
Obviously, (vi) =⇒ (vii).
Suppose that (vii) holds and select h1 , h2 ∈ Ac (H). By (vii), there is h ∈ Ac (H)
such that h = h1 ∨ h2 on ChH (K). If k ∈ Ac (H), k ≥ h1 ∨ h2 on K, then k ≥ h on
ChH (K). Using the Minimum principle 3.85 again, we get k ≥ h on K. We see that
every pair of elements of Ac (H) has a least upper bound, in other words, Ac (H) is a
lattice in its natural ordering, hence (viii) holds.
The proof will be completed once it is shown that (viii) =⇒ (i). To this end, let
h1 , . . . , hn ∈ Ac (H) and let h ∈ Ac (H) be their least upper bound in Ac (H). By
Proposition 3.25(a), h = (h1 ∨ · · · ∨ hn )∗ . We see that the function (h1 ∨ · · · ∨ hn )∗
is H-affine, and thus Theorem 6.5 yields that H is simplicial.
It remains to prove that the Choquet boundary ChH (K) is closed. Assume that
x ∈ ChH (K) \ ChH (K). With the aid of Bauer’s characterization of ChH (K) (see
Theorem 3.24) we can find functions h1 , . . . , hn ∈ Ac (H) such that f := h1 ∨· · ·∨hn
satisfies f (x) < f ∗ (x).
As above, f ∗ ∈ Ac (H) is the least upper bound of h1 , . . . , hn in Ac (H) and f ∗ = f
on ChH (K). The continuity of f and f ∗ yields the equality f (x) = f ∗ (x), which is
an obvious contradiction.
Proposition 6.38. The set M1 (K) of all probability Radon measures on a compact
space K is a Bauer simplex.
Proof. We use condition (vi) of Theorem 6.37 as a characterization of Bauer sim-
plices. So assume that F ∈ C(ext M1 (K)) = C({εx : x ∈ K}) is given. We set
Proposition 6.39. Let X be a Bauer simplex. Then there exists a compact space K
such that X is affinely homeomorphic to M1 (K).
Proof. Choose K as ext X and consider the mapping x 7→ δx , x ∈ K, where δx is
the unique maximal measure representing x.
188 6 Simplicial function spaces
(see Proposition 6.9), and that simplicial function spaces H having a closed Choquet
boundary ChH (K) were labeled as Bauer simplicial spaces. The equivalence (i) ⇐⇒
(v) of Theorem 6.37 says that H is a Bauer simplicial space if and only if
HT := {f ∈ C(K) : T f = f } ,
Φ : f 7→ T f (x), f ∈ C(K),
Lemma 6.41. Let T be a Markov projection on K such that the space HT separates
points of K. Then
ChHT (K) = x ∈ K : µTx = εx .
Theorem 6.42. Let H be a function space on K. The following assertions are equiv-
alent:
(i) there exists a Markov projection on K such that H = HT ,
(ii) H is a Bauer simplicial space and H = Ac (H),
(iii) H|ChH (K) = C(ChH (K)).
Proof. If (i) holds and f ∈ C(ChH (K)) = C(ChHT (K)) = C(ChHT (K)) is given,
there exists g ∈ C(K) such that g = f on ChH (K). Then T g ∈ HT = H and
T g = g = f on ChH (K), so (i) =⇒ (iii).
Suppose now that (iii) holds. According to (i) ⇐⇒ (v) of Theorem 6.37, H
is a Bauer simplicial space. It remains to show that H = Ac (H). To this end, let
f ∈ Ac (H). There exists a function h ∈ H so that h = f on ChH (K). Since
h − f ∈ Ac (H), the Minimum principle 3.85 yields the equality h = f on K. Hence,
f ∈ H. This shows that (iii) =⇒ (ii).
Finally, suppose that H is as in (ii). Given x ∈ K, let δx be the unique maximal
measure in Mx (H). By Theorem 6.37, the mapping x 7→ δx is continuous, hence the
function T f defined as T f (x) = δx (f ), x ∈ K, is continuous for each f ∈ C(K).
Since T f is an H-affine function (cf. Theorem 6.8(c)), T is a Markov projection on
K. To show that H = HT , let h ∈ H. Then T h(x) = δx (h) = h(x) for each x ∈ K.
Thus, H ⊂ HT .
Conversely, pick g ∈ HT . Since the function T g is in Ac (H), we get as above
g = T g ∈ Ac (H) = H which shows that HT ⊂ H. This proves that (ii) =⇒ (i) and
finishes the proof.
Definition 6.43 (Markov simplicial spaces). Function spaces for which one of the
equivalent conditions of Theorem 6.42 holds are labeled as Markov simplicial spaces.
Corollary 6.44. If H is a Markov simplicial space on K, then the function space H is
closed and H⊥ ∩ Mbnd (H) = {0}.
Proof. The assertion easily follows from the facts that H = Ac (H) and that Ac (H)
is closed (see Proposition 3.11(a)).
190 6 Simplicial function spaces
Theorem 6.45. Let H be a simplicial function space on a compact space K such that
ChH (K) is Lindelöf.
(a) Let a bounded function f ∈ Kusc (H) be a pointwise limit of a decreasing se-
quence of continuous functions. Then f ∗ is a pointwise limit of a decreasing
sequence of continuous H-affine functions.
(b) Any function f ∈ C b (ChH (K)) can be extended to a Baire-one H-affine function
on K. Moreover, this extension is unique.
Proof. Let f ∈ Kusc (H) be as in the statement of the theorem. Due to the Edwards in-
between theorem (Theorem 6.6), the set {h ∈ Ac (H) : h ≥ f } is down-directed and
its pointwise infimum is f ∗ (see Corollary 3.25(a)). For x ∈ ChH (K), f ∗ (x) = f (x)
and so, by Lemma A.54, there is a decreasing bounded sequence {fn } of continuous
H-affine functions converging pointwise to f on ChH (K).
The set B := {x ∈ K : fn (x) → f (x)} is a Baire set containing ChH (K), and
therefore carries any maximal measure on K by Theorem 3.79(a). Hence, for all
y ∈ K,
Z Z
∗
f (y) = f (x) dδy (x) = lim fn (x) dδy (x) = lim fn (y).
B n→∞ B n→∞
Thus the sequence {fn } decreases to f ∗ on K and the first part of the proof is finished.
Let f be a bounded continuous function on ChH (K). We extend f to an upper
semicontinuous function g on X := ChH (K) by the formula
lim sup f (y), x ∈ X \ ChH (K),
g(x) := y→x,y∈ChH (K)
f (x), x ∈ ChH (K).
By Lemma 3.18(d),
X ∗
g (x) = g(x) = f (x), x ∈ ChH (K).
Since the latter set is down-directed, it follows from Lemma A.54 that there exist
a decreasing sequence of continuous H-concave functions {gn } and an increasing
sequence of continuous H-convex functions {hn } such that
h := lim fn
n→∞
Proof. We start the proof by noticing that the implications (i) =⇒ (ii) =⇒ (iii) are
obvious. To check (iii) =⇒ (iv), let f be a continuous H-convex function on K. By
Theorem 6.5, f ∗ is H-affine. By Lemma 6.4, f ∗ = T f , hence by the hypothesis, T f
is a Baire function.
To verify (iv) =⇒ (i), notice that H is simplicial by Theorem 6.5. Since Kc (H) −
Kc (H) is uniformly dense in C(K), T (Kc (H)) ⊂ B 1 (K) and the space of Baire-one
functions is closed with respect to the uniform convergence (see Proposition A.126),
we have T (C(K)) ⊂ B 1 (K). By Theorem 6.8(c), T (C(K)) ⊂ (Ac (H))⊥⊥ . By
Proposition 5.42(a), T (C(K)) ⊂ (Ac (H))1,b .
We proceed with the proof by showing (iii) =⇒ (v). We notice that a straight-
forward transfinite induction gives that T f is in F for any bounded Baire function f
on K. Given functions f1 , f2 ∈ F, the function T (f1 ∨ f2 ) ≥ f1 ∨ f2 by the Mini-
mum principle 3.86 and T (f1 ∨ f2 ) = f1 ∨ f2 on ChH (K). Again by the Minimum
principle 3.86, T (f1 ∨ f2 ) is the supremum of f1 ∨ f2 in the natural ordering.
We conclude the proof by showing (v) =⇒ (iii). Let f ∈ −W(H) be given, that
is, f = h1 ∨ · · · ∨ hn , h1 , . . . , hn ∈ H. By the assumption, there exists a function
h ∈ F such that h is the supremum of h1 , . . . , hn in the natural ordering. If g ∈ H is
a function with g ≥ f , then h ≤ g. Hence h ≤ f ∗ .
192 6 Simplicial function spaces
On the other hand, given x ∈ K, we find µ ∈ Mx (H) such that µ(f ) = f ∗ (x)
(see Lemma 3.21). Then
f ∗ (x) = µ(f ) ≤ µ(h) = h(x).
Hence f ∗ = h is H-affine and H is simplicial by virtue of Theorem 6.5. Thus we have
proved that f ∗ = T f is a Baire function for each f ∈ −W(H). Again we use the
density of W(H)−W(H) to conclude that T f is a Baire function for each f ∈ C(K).
This concludes the proof.
Remark 6.47. A function space satisfying condition (i) of Theorem 6.46 are some-
times termed analytic simplicial spaces or Lion simplicial spaces. By Theorem 6.45,
any simplicial function space with a Lindelöf Choquet boundary is Lion. Exam-
ple 6.48 below shows that there exists a Lion simplicial space such that its Choquet
boundary is not Lindelöf.
Example 6.48. There exists a Lion simplicial space such that ChH (K) is not Lindelöf.
Proof. Let L := [0, 1], B := L, Lb := {0, 1} and µb := 21 (ε0 + ε1 ) for b ∈ B, and
let H be the Stacey simplicial function space on the compact space K from Defini-
tion 6.13. Then ChH (K) is not Lindelöf, because it is an uncountable discrete space.
On the other hand, let f ∈ Kc (H) be given. For any δ > 0, the set
{x ∈ L : f ∗ (x) − f (x) > δ}
is finite. Hence it follows that the sets
{x ∈ L : f ∗ (x) − f (x) > c} and {x ∈ L : f ∗ (x) − f (x) ≥ c}
are finite for each c ∈ R. Since the points of K are Gδ , the function f ∗ − f is
Baire-one by Theorem A.124. Hence f ∗ is Baire-one as well. By Theorem 6.8(a),
T (C(K)) ⊂ B 1 (K), and thus H is Lion by Theorem 6.46.
Proof. For the proof of the implication (i) =⇒ (ii), suppose S∞ that H is simplicial
and ChH (K) is an Fσ set. Thus we can write ChH (K) = n=1 Fn where {Fn } is
an increasing sequence of compact sets. Let f be a bounded Baire-one function on
ChH (K) and {fn } be a sequence of continuous functions on ChH (K) converging
pointwise to f . We may assume that kf k, kfn k are all bounded by a positive number
M . By Theorem 6.27, there exists a sequence {hn } of H-affine continuous functions
on K such that hn = fn on Fn and khn k = kfn k.
The proof will be completed by showing that the sequence {hn } converges point-
wise to the function h := T f . Then T f is Baire-one and H-affine by the Lebesgue
dominated convergence theorem. Notice that the definition is meaningful since maxi-
mal measures are carried by ChH (K) due to Theorem 3.81. R
For a fixed point x ∈ K and ε > 0 we find n0 ∈ N such that ChH (K) |f −fn | dδx <
ε and δx (Fn ) > 1 − ε for all n ≥ n0 . Then, for n ≥ n0 , we have
Z
|T f (x) − hn (x)| = (f − hn ) dδx
K
Z Z
≤ |f − fn | dδx + 2M dδx
Fn0 K\Fn0
≤ ε + ε2M,
which proves the required statement and concludes the proof of (i) =⇒ (ii).
Since the implication (ii) =⇒ (iii) is obvious, we proceed to the proof of (iii)
=⇒ (iv). Let H be a function space on a compact space K satisfying condition
(iii). First we verify that H is simplicial. Indeed, for a given continuous H-convex
function f on K we find an H-affine Baire-one function h with h = f on ChH (K).
194 6 Simplicial function spaces
(Indeed, otherwise the set C ∩ ChH (K) would be an Fσ set, which is absurd.)
Another use of Theorem A.115 provides a homeomorphic copy D ⊂ C \ ∞
S
n=1 Kn
of the Cantor set {0, 1}N such that
Remark 6.50. We remark that only the proof of (v) =⇒ (i) in the preceding theorem
used the assumption of metrizability of K.
Example 6.51. There exists a simplicial space H on a metrizable space K such that
B b1 (K) ∩ A(H) is a lattice in the natural ordering but ChH (K) is not an Fσ set.
Proof. We take a suitable Stacey simplicial function space from Definition 6.13. Na-
mely, we take L := [0, 1] and B := Q ∩ [0, 1] and K as in Definition 6.13. We
enumerate B as {qn : n ∈ N} and take Lqn := {0, 1} with µqn := n1 ε0 + (1 − n1 )ε1 ,
n ∈ N.
By Lemma 6.14, H is simplicial and ChH (K) = K \ B. Hence ChH (K) is not an
Fσ set. Further we know from Lemma 6.14 that δqn = µqn . For brevity we write qn+
for the point 1 in Lqn and qn− for the point 0 in Lqn .
In order to check that A(H)∩B b1 (K) is a lattice in the natural ordering, it is enough
to prove that T (f ∨g) is a Baire-one function for every pair f and g of H-affine Baire-
one functions. Let f and g be such functions with values in [0, 1] and set h := f ∨ g.
We claim that
2
|h(qn+ ) − h(qn )| ≤ for every n ∈ N. (6.5)
n
Indeed, for a fixed n ∈ N we have
1 1
|f (qn+ ) − f (qn )| = |f (qn+ ) −f (qn− ) − (1 − )f (qn+ )|
n n
1 2
≤ |f (qn+ ) − f (qn− )| ≤ .
n n
By the same argument, |g(qn+ ) − g(qn )| ≤ n2 . We need to check this inequality for the
function h. The only nontrivial case is when h(qn+ ) = f (qn+ ) and h(qn ) = g(qn ) (or
196 6 Simplicial function spaces
is finite for every ε > 0. By Lemma A.128, T h is a Baire-one function and the space
A(H) ∩ B b1 (K) is a lattice in the natural ordering.
kI + Sk = 1 + kSk. (6.6)
(Here, I stands for the identity operator.) Afterwards, Banach spaces and operators
satisfying the equality (6.6) were studied. A Banach space E satisfying (6.6) for any
compact operator S on E is said to satisfy the Daugavet property, or, for short, the
Daugavet space. Among other properties, it was shown that a Banach space satisfies
the equality (6.6) for any compact operator S if and only if the equality (6.6) is satis-
fied for the class of all rank-one operators, and that the space C(K) has the Daugavet
property if and only if the compact space K has no isolated points (see, for example,
V. M. Kadets, R. V. Shvidkoy, G. G. Sirotkin and D. Werner [255]).
6.7 The Daugavet property of simplicial spaces 197
Definition 6.52 (The Daugavet property). A Banach space E has the Daugavet prop-
erty if
kI + F k = 1 + kF k
Theorem 6.53. Let H be a simplicial function space on a compact space K such that
ChH (K) has no isolated points. Then Ac (H) has the Daugavet property.
Proof. Let a rank-one operator F : Ac (H) → Ac (H) be given. There exists a func-
tional ϕ ∈ Ac (H)∗ and a function h0 ∈ Ac (H) such that
Fix ε > 0. By Theorem 3.85, there exists a point y0 ∈ ChH (K) (= ChAc (H) (K) by
Exercise 3.95) such that h0 (y0 ) = kh0 k. Let W be an open neighborhood of y0 such
that h0 > kh0 k − ε on W . Since ChH (K) has no isolated points, there exists a point
x0 ∈ W ∩ ChH (K) and an open neighborhood U of x0 such that |µ|(U ) < ε. Let f
be a continuous function on K such that kf k = 1 and µ(f ) > kµk − ε. According to
Tietze’s theorem, there exists a continuous function g on K such that
Then µ(g) > kµk − 2ε. Since µ is a boundary measure, µ(g) = µ(g ∗ ). For the
positive part µ+ of µ we find by Theorem A.84 a continuous H-concave function k +
such that
g ≤ k+ ≤ 1 and µ+ (k + ) − µ+ (g ∗ ) < ε.
g ≤ k− ≤ 1 and µ− (k − ) − µ− (g ∗ ) < ε.
(recall that k∗ (x0 ) = k(x0 ) by Theorem 3.24). Combining all the estimates together
we get
Theorem 6.54. For a function space H, the following assertions are equivalent.
(i) H is simplicial.
(ii) The ordered Banach space Ac (H)∗ is a vector lattice.
(iii) The compact convex set S(Ac (H)) is a simplex.
Proof. We start the proof by noticing that (i) =⇒ (ii) due to Theorem 6.18 and
Theorem A.24.
The next step is showing that (ii) =⇒ (iii). Let X := S(Ac (H)). We want to prove
that Ac (X) is simplicial, which means, by virtue of Theorem 6.5 and Corollary 4.8, to
check that f ∗ is affine for any continuous convex function f on X. Since f ∗ is always
concave, we need to check its convexity.
To this end, let s := α1 s1 + α2 s2 be a nontrivial convex combination of points
s1 , s2 ∈ X. We fix ε > 0 P and use Lemma 3.21 along with Proposition 4.3 to find
a convex combination s = nj=1 βj tj , t1 , . . . , tn ∈ X, β1 , . . . , βn strictly positive,
such that
Xn
βj f (tj ) > f ∗ (s) − ε.
j=1
6.8 Choquet simplices 199
Let γij ≥ 0 and uij ∈ X satisfy u0ij = γij uij , i = 1, 2, j = 1, . . . , n. Then we get
the following convex combinations
n
X γij
si = uij , i = 1, 2, and
αi
j=1
γ1j γ2j
tj = u1j + u2j , j = 1, . . . , n.
βj βj
We use the convexity of f , concavity of f ∗ and preceding equalities to get
n
∗
X γ1j γ2j
f (s) − ε < βj f ( u1j + u2j )
βj βj
j=1
n
X
≤ γ1j f (u1j ) + γ2j f (u2j )
j=1
n n
X γ1j X γ2j
= α1 f (u1j ) + α2 f (u2j )
α1 α2
j=1 j=1
n n
X γ1j X γ2j
≤ α1 f ∗ (u1j ) + α2 f ∗ (u2j )
α1 α2
j=1 j=1
≤ α1 f ∗ (s1 ) + α2 f ∗ (s2 ).
Since ε is arbitrary, f ∗ (s) = α1 f ∗ (s1 ) + α2 f ∗ (s2 ), and f ∗ is affine.
Now we verify (iii) =⇒ (i). Let φ : K → X be the embedding from Defini-
tion 4.25. Given a point x ∈ K, let µ, ν ∈ Mx (Ac (H)) be Ac (H)-maximal mea-
sures. Then φ] µ and φ] ν are maximal measures on X (see Proposition 4.28(d)) that
represent the point φ(x) (use Proposition 4.26(c)). Hence φ] µ = φ] ν, and µ = ν.
Thus Ac (H) is simplicial. By Lemma 6.3(b), H is simplicial.
Remark 6.55. We remark that the original definition of Choquet simplices is different
from ours. We recall that X is affinely homeomorphic to the state space S(Ac (X)),
which is contained in the dual space E = (Ac (X))∗ . This identification is called
a regular embedding of X (see p. 82 in E. M. Alfsen [5]). A regularly embedded
compact convex set X is called a simplex in Chapter II, §3 of [5] if E is a vector
lattice with the ordering given by the cone E + . Thus from Theorem 6.54 we get the
following classical definition of Choquet simplices.
200 6 Simplicial function spaces
Theorem 6.56. Let X be a compact convex set that is regularly embedded in the
locally convex space E = (Ac (X))∗ . Then X is a Choquet simplex if and only if E is
a vector lattice.
Definition 6.58 (Exposed and relatively exposed sets). A set F in a compact convex
set X is called relatively exposed if for any x ∈ X \ F there exists f ∈ Ac (X) so that
f ≥ 0 on X, f = 0 on F and f (x) > 0 (cf. Definition 8.47).
A set F ⊂ X is said to be exposed if there is a positive function f ∈ Ac (X) such
that f (x) = 0 if and only if x ∈ F .
We remark that any exposed set is relatively exposed and any relatively exposed set
is a closed face (cf. Proposition 8.48).
Definition 6.59 (Prime compact convex sets). A compact convex set X is said to
be prime if whenever F and G are closed relatively exposed faces of X such that
co(F ∪ G) = X, then either F = X or G = X.
Theorem 6.60. A compact convex set X is prime if and only if the function space
Ac (X) is an antilattice.
we get that f = f ∧ g almost everywhere with respect to δx . Since ChH (K) ⊂ spt δx ,
the equality f = f ∧ g holds on ChH (K). In other words, f ≤ g on ChH (K). The
Minimum principle 3.85 yields the desired conclusion that f ≤ g on K.
Lemma 6.66. Let X be a compact convex set such that ext X = X. Then X is prime.
Example 6.67. There exists a simplicial function space H on a compact space K such
that ChH (K) = K and H is not prime.
Proof. Let K := [0, 1] ∪ {2} ∪ {3} and H := {f ∈ C(K) : f (0) = 21 (f (2) + f (3))}.
Then H = Ac (H), H is simplicial and ChH (K) = K \ {0}. Nevertheless, H is not
an antilattice.
Indeed, if (
x(1 − x) x ∈ [0, 1],
f (x) :=
0 otherwise,
and (
1
4x x ∈ [0, 1],
g(x) :=
0 otherwise,
which yields y, z ∈ co H.
Proof. For the proof of (i) =⇒ (ii), let X be a Bauer simplex and F ⊂ X be a
face. Since F is convex, we only have to show that F is extremal, that is, our aim
is to prove that spt δx ⊂ F for every x ∈ F (see Exercise 6.82). First, suppose that
x ∈ F and that δx is not carried by F . Let V be a closed convex neighborhood not
intersecting F such that δx (V ) > 0. If δx (V ) = 1, then x ∈ V and thus V ∩ F 6= ∅,
which is impossible. Setting c := δx (V ) ∈ (0, 1), we have
hence
x = cr(δx |V ) + (1 − c)r(δx |X\V ).
for any f ∈ Kc (X). By Corollary 4.8 and Lemma 6.4, the previous equality can be
rewritten as
f ∗ (αx + (1 − α)y) = αf ∗ (x) + (1 − α)f ∗ (y),
which holds because the function f ∗ is affine by Theorem 6.5.
For the proof of the reverse implication, let T be an affine mapping provided by
condition (ii). We write Tx for the measure T (x). We need to prove that f ∗ is affine
for any f ∈ Kc (X). Since T is assumed to be affine, it is enough to verify that
Tx (f ) = f ∗ (x) for any f ∈ Kc (X) and x ∈ X.
Let f and x be as above and ε > 0. By Lemma 3.22 and Proposition 4.3, there
∗
Pn a molecular measure µ ∈ Mx (X) such
exists Pnthat µ(f ) ≥ fP(x)n
− ε. Then µ =
i=1 αi εxi where xi ∈ X, αi > 0 such that i=1 αi = 1 and i=1 αi xi = x. Since
f is convex and T is affine, Proposition 3.20(b) gives
n
X n
X
∗
f (x) − ε ≤ µ(f ) = αi εxi (f ) ≤ αi Txi (f )
i=1 i=1
= TPni=1 αi xi (f ) = Tx (f ) ≤ f ∗ (x).
Proof. For the proof of (a), we notice that for µ, ν ∈ M+ (L), µ ≺H ν if and only
µ(f ) ≤ ν(f ) for each f ∈ −W(H) (see Proposition 3.56), and this is the case if and
only if µ(f ) ≤ ν(f ) for any f ∈ −W(G). Further, (b) and (c) are obvious.
6.10 Exercises 205
6.10 Exercises
Exercise 6.73. Let H be a simplicial space and x ∈ K. Prove that δx ∈ ext(Mx (H)).
and
n
X d+1
X
x := c−1 ai xi = −c−1 ai xi .
i=0 i=n+1
Exercise 6.75. Let X be a compact convex set in Rd . Prove that X is a simplex if and
only if X is an n-simplex for some n ≤ d (see Definition 2.2).
ChH ([0, 1]) = ChH ([0, 1]) = ChC([0,1]) ([0, 1]) = [0, 1].
It follows that H is simplicial and {0} ∈ ChH ([0, 1]). Nevertheless, {0} is not H-
exposed.
and that a function f ∈ C(K) belongs to Ac (H) if and only if f is affine on each
segment contained in K. Further show that H is simplicial and that the point (1, 1) is
Ac (H)-exposed but it is not H-exposed.
Hint. In order to show that the point (1, 1) is Ac (H)-exposed, consider, for example,
the function
(x, y) 7→ | − x − y + 2| + |1 − y|, (x, y) ∈ K.
Hint. If f ∈ Ac (H) and ε > 0, for any x ∈ K we use Corollary 3.23(a) and
Proposition 3.25(a) to find a function hx ∈ H such that f < hx and hx (x) <
f (x) + ε. A simple compactness argument yields the existence of finitely many func-
tions h1 , . . . , hn ∈ H such that
f < h1 ∧ · · · ∧ hn < f + ε.
f > g1 ∨ · · · ∨ gm > f − ε.
6.10 Exercises 207
By the weak Riesz interpolation property, there exists a function h ∈ H such that
g1 ∨ · · · ∨ gm < h < f1 ∧ · · · ∧ fn .
Exercise 6.79. Find an example of a simplicial function space whose state space is
not a simplex.
Hint. Consider the function space H := P 2 ([0, 1]) from Example 3.2(b). Then
ChH ([0, 1]) = [0, 1] (see Example 3.5(b)), and thus H is simplicial. On the other
hand, if ψ : [0, 1] → R2 is defined by
S(H) is affinely homeomorphic to co ψ([0, 1]) (use Exercise 4.51). Hence S(H) is
not a simplex.
Hint. If z ∈ spt µ \ co spt ν, find a continuous affine function f on X such that
f (z) = −1 and inf f co spt ν ≥ 0. If k := f ∧ 0, then k is continuous concave on
X, k = 0 on spt ν and ν(k) ≤ µ(k) < 0.
Hint. The characteristic function cX\F is concave (cf. Proposition 2.68) and lower
semicontinuous. By Proposition 3.56, ν(cX\F ) ≤ µ(cX\F ).
Exercise 6.82. Let F be a closed convex subset of a compact convex set X. Prove
that F is extremal if and only if spt ν ⊂ F for any x ∈ F and any maximal measure
ν ∈ Mx (X).
Hint. If F is a closed convex extremal set, it carries any maximal measure with
barycenter in F by Proposition 2.69. Conversely, let x ∈ F and µ ∈ Mx (X).
Let ν ∈ M1 (X) be maximal with µ ≺ ν. Then ν ∈ Mx (X). By Exercise 6.80,
spt µ ⊂ co spt ν ⊂ F . Hence spt µ ⊂ F and F is extremal.
Exercise 6.83. Let F be a closed face in a compact convex set X. Prove that Ac (F )-
maximal measures coincide with Ac (X)-maximal measures carried by F .
208 6 Simplicial function spaces
Exercise 6.88. Verify the details of the following example to provide an example of a
face in a simplex whose closure is not a face.
Hint. Let
P := {xn } ∈ c0 : xn ≥ 0 for all n ∈ N
and
Q := {n−1 en : n ∈ N}
and X := co {−e1 } ∪ Q ,
where en are the standard unit vectors in c0 . Show that X is a simplex and
ext X = {−e1 } ∪ Q.
Exercise 6.89. Prove that there exist metrizable Bauer simplices X, Y and affine con-
tinuous surjections ϕ : X → Y and ψ : Y → X such that X, Y are not affinely
homeomorphic.
Hint. Let K := [0, 1]. There exist continuous surjections f : K → K 2 and g : K 2 →
K. Let X := M1 (K) and Y := M1 (K 2 ). Then f] is an affine surjection of X onto
Y and g] is an affine surjection of Y onto X.
The sets X and Y are metrizable Bauer simplices (Proposition 6.38). If we assume
that X and Y are affinely homeomorphic, the sets ext X and ext Y are homeomorphic.
Then, by Proposition 2.27, K and K 2 are homeomorphic, a contradiction.
(Ac (X))∗ is a lattice by Theorem A.24. By Proposition A.17 and Theorem A.24,
(Ac (X))∗∗ is a lattice (we remark that (Ac (X))∗∗ can be identified with Ab (X) by
Proposition 4.32, and thus satisfies the condition
To show (b), notice that Ab (X) can be identified with Ac (Y ) via the canonical
embedding of Ab (X) into (Ab (X))∗∗ . Thus assuming the existence of s ∈ Y \ φ(X),
we can find a function f ∈ Ab (X) such that s(f ) > sup f (X). Without loss of
generality we may suppose that f is positive. Then
a contradiction.
The last assertion (c) follows now from (a) and Theorem 6.37. Indeed, Ac (Y ), as
well as Ab (X), is a lattice, and thus (viii) of Theorem 6.37 applies.
S := {s ∈ A∗ : s ≥ 0, ksk = 1}
and ϕ : X → RF be defined as
ϕ(x) = {f (x)}f ∈F , x ∈ X.
For the proof of (e), let T ∗ : (A)∗ → (Ac (Y ))∗ be the dual operator of T from
(d). Then T ∗ is an affine homeomorphism of S onto S(Ac (Y )), and thus S is affinely
homeomorphic to Y by Proposition 4.31(a).
If A has the weak Riesz interpolation property, then so does A. Since T from
(d) preserves order, Ac (Y ) possesses the weak Riesz interpolation property as well.
Consequently, Y is a simplex by Theorem 6.16.
Analogously we get that Ac (Y ) is a lattice provided A is. In this case, Y is a Bauer
simplex by Theorem 6.37.
Exercise 6.92 (The ESP property). A function space H on K has the equal support
property, for short ESP, if for any x ∈ K and any two maximal measures µ1 , µ2 ∈
Mx (H) representing the point x we have spt µ1 = spt µ2 .
It is clear that any simplicial function space has the ESP. There are examples of
function spaces having the ESP without being simplicial. Construct a compact convex
set X with ext X closed such that X is not a simplex and X has the ESP.
Hint. Let µ be any nonzero signed Radon measure on [0, 1] such that
The space H separates points of [0, 1]: Given x, y ∈ [0, 1], assume that h(x) = h(y)
for any h ∈ H. Then εx −εy ∈ H⊥ and, consequently, εx −εy = αµ for some α ∈ R.
Since spt µ+ = spt µ− = [0, 1], we get x = y.
We see that H is a (closed) function space. Moreover, ChH ([0, 1]) = [0, 1]. To
show this, select any x ∈ [0, 1] and ν ∈ Mx (H). Then ν − εx ∈ H⊥ , so that
ν − εx = αµ for some α ∈ R. Hence ν = εx , and therefore x ∈ ChH ([0, 1]).
Let X be the state space of H. Since s := π(µ+ ) = π(µ− ) ∈ X and
by Proposition 4.28(a), we see that the point s has two different representing measures
φ] µ+ and φ] µ− carried by ChH ([0, 1]) (hence maximal by Corollary 3.61). Therefore,
X is not a simplex.
On the other hand, X has the ESP property: Suppose that Λ1 and Λ2 are maximal
probability measures on X representing a point x ∈ X. By Proposition 4.28(d), there
exist H-maximal measures λ1 , λ2 ∈ M1 ([0, 1]) such that Λ1 = φ] λ1 and Λ2 = φ] λ2 .
Then λ1 −λ2 ∈ H⊥ . Hence, there exists α ∈ R such that λ1 −λ2 = αµ = αµ+ −αµ− .
212 6 Simplicial function spaces
ext X.
Exercise 6.93. Let X be a compact convex set as given by Exercise 6.92. Recall that
X is a metrizable set with ext X closed such that X is not a simplex and such that
X has the ESP property. Prove that there exists x ∈ X such that, if F is a compact
convex set, x ∈ F ⊂ X and ext F ⊂ ext X, then F is not a simplex. (Compare with
Remark 2.14.)
Hint. Keep the notation of Exercise 6.92. Let x := π(µ+ ) and let F ⊂ X be a
compact convex set containing x such that ext F ⊂ ext X. There exists Λ ∈ Mx (F )
carried by ext F . Since Λ is carried by (a closed set) ext X, it is maximal. Hence,
spt Λ = ext X (see the hint of Exercise 6.92). It follows that ext F ⊃ ext X, and
therefore F = X.
Prove that
(a) ChH (K) = [0, 1], Ac (H) = H and H is not simplicial,
(b) H has the ESP,
(c) the state space S(H) has not the ESP.
Hint. Assertion (a) is easy to verify and (b) follows by a similar argument as to that
in Exercise 6.92. The only difference is that
and
Hint. If kI + T k = 1 + kT k, then
Subsection 6.6.A follows the paper by H. Bauer [38], see also E. M. Alfsen [5],
Theorem II.4.1.
The definition of a Markov simplex appeared in a paper [419] by R. Sine.
Theorem 6.45 is taken from F. Jellett [250] (see also H. Fakhoury [175]), Theo-
rem 6.46 from M. Rogalski [392] and A. Goullet de Rugy, C. Schol-Cancelier and
B. Taylor-MacGibbon [201]. We mention the following problem on the converse im-
plication in Theorem 6.45.
Problem 6.96. Let X be a compact convex set such that any f ∈ C b (ext X) has an
affine Baire-one extension. Is then ext X a Lindelöf space?
Subsection 6.6.D follows the papers by J. Spurný [424] and [427]. The easy impli-
cation of Theorem 6.49 can be found in M. Rogalski [392] and F. Jellett [250]. The
case of nonmetrizable spaces was solved in the paper J. Spurný and O. Kalenda [432]
by proving the following theorem.
For a compact convex set X, the following assertions are equivalent:
• X is a simplex and ext X is a Lindelöf resolvable set,
was given by E. Størmer in [440] (see also A. K. Roy [399] and [400]). Note that
E. StørmerSin [440] proved that a compact convex set X is a Bauer simplex if and
only if co( α Fα ) is a closed split face of X whenever {Fα } is any family of closed
split faces of X (Størmer’s axiom). Theorem 6.70 is from H. Fakhoury [174]
The example of Exercise 6.77 is taken from Bauer’s paper [44]. It represents a
slight modification of an earlier Bauer’s example from [38] and is attributed to S. Pa-
padopoulou. Exercise 6.88 is due to M. Kraus. Exercise 6.90 is taken from C. H. Chu
and B. Cohen [119] and P. Harmond and Å. Lima [215], Example 3.3(b).
The ESP property in Exercise 6.92 was introduced by I. Feinberg in his thesis
(cf. J. N. McDonald [336]). In the same paper, J. N. McDonald investigated the ESP
property and constructed an example of the simplex which fails the ESP property
(Exercise 6.92 follows this example; see also J. N. McDonald [338]). Exercises 6.80
and 6.81 can also be found in [336]. S. Alpay proved in [12] that if X is a compact
convex set and x ∈ X, then all maximal measures on X with barycenter x have the
same support if and only if x ∈ ext(co spt µ) whenever µ is a maximal measure on X
with barycenter x. The ESP property is also studied in M. W. Grossman [204].
Note that J. Köhn [274] localized the notion of a simplex in the convex setting and
gave several statements that a point has a unique maximal representing measure. We
also refer the reader to C. Cho-Ho [118] for similar results. Examples and a study of
points of simpliciality in the context of function spaces is presented in the paper [27]
by M. Bačák.
Chapter 7
Choquet theory of function cones
The notion of a function cone generalizes the concept of a function space. Even
though we focus mainly on function spaces, the Choquet theory of function cones is an
indispensable tool for us, since later on we investigate typical function cones arising
in potential theory. Another motivation is the selection theorem from Section 11.5 and
a description of boundary measures contained in Section 8.5. Since the basic results
of the Choquet theory of function cones very often use the same techniques as their
analogues from the theory of function spaces, we present them in Sections 7.1– 7.6 in
a brief but, we hope, comprehensible way.
We point out that Theorem 7.27 provides a measure on the Choquet boundary in-
duced by a maximal measure. This concept will be strengthened considerably in
Chapter 9. Other results not encountered before are Theorem 7.38 and Theorem 7.40,
which clear up the relation between simplicial function cones and function spaces
generated by them.
An important concept of ordered compact convex sets is studied in Section 7.5.
Main facts are summarized in Theorem 7.54 and an application of Theorem 7.27
is presented in Theorem 7.55. We show in Theorem 7.58 how an ordered compact
convex set can be recovered from the set of maximal elements by a method imitating
the Krein–Milman theorem. The Douglas characterization of simplicial measures is
given in Theorem 7.60.
W(S) = {s1 ∧ · · · ∧ sn : s1 , . . . , sn ∈ S, n ∈ N} = S.
Examples 7.2. (a) Any function space can serve as an example of a function cone.
(b) Let H be a function space on a compact space K. Then W(H) is a function cone
on K. Also the family S c (H) of all continuous H-concave functions is an example of
a function cone.
7.1 Function cones 217
Proof. Assertions (a) and (b) are obvious, the first part of (c) follows by the same
argument as in Proposition 3.11(b) and the second part from (b). Assertion (d) follows
from the definition, (e), (f) are consequences of (d) and (g) follows from (e). Finally,
(h) follows from the lattice version of the Stone–Weierstrass theorem contained in
Proposition A.31.
Qµ (f ) := inf{µ(s) : s ∈ S c (S), s ≥ f }.
Proof. Assertion (a) is obvious, (b) follows from Theorem A.84, (c) holds because S
contains constant functions and (d) follows by a straightforward verification.
p : g 7→ Qµ (g), g ∈ C(K).
G := {g ∈ C(K) : g ≥ f }
and use the first part of the proof to find a measure νg ∈ M+ (K) such that µ ≺ νg
and νg (g) = Qµ (g), g ∈ G. Given h ∈ G, let
Mh := {νg : g ∈ G, g ≤ h}.
Since Mh ⊂ {λ ∈ M+ (K) : kλk = kµk} and the family {Mh : h ∈ G} has T the finite
intersection property, a compactness argument yields the existence of ν ∈ {M h :
h ∈ G}. Then µ ≺ ν and
inf λ(h) : λ ∈ Mh = inf λ(h) : λ ∈ M h ≤ ν(h)
for each h ∈ G.
220 7 Choquet theory of function cones
Hence
Qµ (f ) ≤ inf {Qµ (g) : g ∈ G} = inf {νg (g) : g ∈ G}
≤ inf {inf {νg (h) : g ∈ G, g ≤ h} : h ∈ G}
≤ inf {ν(h) : h ∈ G}
= ν(f )
≤ inf {ν(k) : k ∈ S c (S), k ≥ f }
≤ inf {µ(k) : k ∈ S c (S), k ≥ f }
= Qµ (f ),
which yields the required equality.
The second part of the assertion follows by taking µ = εx and observing that
ν ∈ Mx (S) provided εx ≺ ν.
The inequality “≥” being obvious, we proceed with the proof of “≤”. Let x ∈ K and
g ∈ C(K) be fixed. The formula
p(s) = s(x), x ∈ K.
Hence
inf{g ∈ S : g ≥ f } = inf{g ∈ S c (S) : g ≥ f }. (7.2)
Further,
inf{g ∈ S c (S) : g ≥ f } = inf{g ∈ S usc (S) : g ≥ f } (7.3)
by (a).
Combining (7.2) and (7.3) we get
f ∗ = inf{g ∈ S c (S) : g ≥ f }
= inf{g ∈ S usc (S) : g ≥ f }
≤ inf{g ∈ W(S) : g ≥ f }
≤ inf{g ∈ S : g ≥ f }
= inf{g ∈ S c (S) : g ≥ f }.
To verify (c), let f be upper semicontinuous and s ∈ S lsc (S). Let x ∈ K and
µ ∈ Mx (S) be such that f ∗ (x) = µ(f ). Then
which yields “≤” in (c). The reverse inequality being obvious, the proof is finished.
Proof. Assertion (a) follows from inequalities µ(1) ≤ ν(1) and µ(−1) ≤ ν(−1). To
show (b), εx ≺ µ clearly implies µ ∈ Mx (S). Conversely, if µ ∈ Mx (S), s ∈ S c (S)
and ε > 0, let t ∈ S be such that s ≤ t and t(x) < s(x) + ε (use Proposition 7.11(a),
(b)). Then
µ(s) ≤ µ(t) ≤ t(x) ≤ s(x) + ε.
Hence εx ≺ µ, and the proof is complete.
Proposition 7.13. (a) If f ∈ S usc (S) and g is lower semicontinuous with f < g, then
there exists s ∈ W(S) such that f < s < g.
(b) If f ∈ S usc (S), then f = inf{g ∈ S c (S) : g > f } and this family is down-
directed.
222 7 Choquet theory of function cones
(c) If f ∈ S lsc (S) and g is upper semicontinuous with g < f , then there exists
s ∈ W(S) such that g < s < f .
(d) If f ∈ S lsc (S), then f = sup{g ∈ S c (S) : g < f } and this family is up-directed.
(e) W(S) is dense in S c (S).
Proof. To show (a), let f < g as in (a) be given. For any x ∈ K we use Proposi-
tion 7.11(a),(b) to find a function sx ∈ S such that f < s and f (x) < sx (x) < g(x).
By the semicontinuity of f and g there exists an open set Ux 3 x such that f < sx < g
on Ux . By the compactness of K, we can select finitely many points x1 , . . . , xn such
that f < sx1 ∧ · · · ∧ sxn < g. Hence s := sx1 ∧ · · · ∧ sxn is the required function.
Since (b) follows immediately from Proposition 7.11(a),(b) and Proposition 7.6(c),
we proceed to the proof of (c). Let f ∈ S lsc (S) and upper semicontinuous g with
g < f be given. For a fixed x ∈ K, let µ ∈ Mx (S) be such that g ∗ (x) = µ(g) (use
Lemma 7.10). Then
g ∗ (x) = µ(g) < µ(f ) ≤ f (x).
Hence g ∗ < f , and we may use (a) to find a function s ∈ W(S) with g ∗ < s < f .
Since
f = sup{g ∈ C(K) : g < f }
(see Proposition A.50(ii)), assertion (d) follows from (c).
Finally, let f ∈ S c (S) be given. Combining (a) with (b) we see that f = inf{s ∈
W(S) : s > f } and the latter family is down-directed. By Dini’s theorem, for any
ε > 0 there exists s ∈ W(S) such that f < s < f + ε. This concludes the proof.
Theorem 7.15. (a) The family of all closed S-extremal sets is stable with respect to
finite unions and arbitrary intersections.
(b) A closed set F ⊂ K is S-extremal if and only if cK\F ∈ S lsc (S).
(c) If f ∈ S lsc (S), then F := {x ∈ K : f (x) = min f (K)} is S-extremal.
(d) A point x ∈ K is in ChS (K) if and only if the set {x} is S-extremal.
(e) Any nonempty closed S-extremal set intersects ChS (K); in particular, ChS (K)
is nonempty.
(f) If f ∈ S lsc (S) is positive on ChS (K), then f ≥ 0 on K.
7.2 Maximal measures 223
Proof. The proof of (a) is analogous to the proof of Proposition 3.14 and (b), (c)
and (d) follows by a straightforward verification (cf. Lemma 3.13). To show (e) we
use Zorn’s lemma again (cf. Proposition 3.15 and Theorem 2.22). Given a closed S-
extremal set F , we consider the family Z of all nonempty closed S-extremal sets con-
tained in F endowed with the partial ordering given by the reverse inclusion. Zorn’s
lemma provides a maximal element H ∈ Z. If x1 , x2 ∈ H are distinct points, let
s ∈ S be such that s(x1 ) 6= s(x2 ). Then it is easy to see that
is a closed S-extremal set strictly smaller than H, a contradiction with the maximality
of H. Hence H is a singleton, say H = {x}, and thus x ∈ ChS (K) by (d). This
proves (e).
Finally, let f ∈ S lsc (S) be positive on ChS (K). Then F := {x ∈ K : f (x) =
min f (K)} is a closed S-extremal set, and thus F ∩ ChS (K) is nonempty by (e).
Hence f ≥ 0 on K, which concludes the proof.
Proposition 7.16. For measures µ, ν ∈ M+ (K), the following assertions are equiv-
alent:
(i) µ ≺ ν,
(ii) ν(s) ≤ µ(s) for all s ∈ W(S),
(iii) ν(s) ≤ µ(s) for all s ∈ S lsc (S),
(iv) ν(s) ≤ µ(s) for all s ∈ S usc (S).
Proof. Obviously, (i) =⇒ (ii), (iii) =⇒ (i), (iv) =⇒ (i). Proposition 7.13(b),(d)
yields (i) =⇒ (iii) and (i) =⇒ (iv). Finally, Proposition 7.13(e) yields (ii) =⇒ (i).
This concludes the proof.
Proof. Since S = W(S) by the assumption, the proof follows from Proposition 7.16.
Theorem 7.18. For every measure µ ∈ M+ (K) there exists a maximal measure λ
such that µ ≺ λ.
Theorem 7.19. For a measure µ ∈ M+ (K), the following assertions are equivalent:
(i) µ is maximal,
(ii) µ(f ) = µ(f ∗ ) for any f ∈ −W(S),
(iii) µ(f ) = µ(f ∗ ) for any f ∈ Kc (S),
224 7 Choquet theory of function cones
Proof. Obviously, (v) =⇒ (iv) =⇒ (iii) =⇒ (ii). To verify (i) =⇒ (v), let f
be an upper semicontinuous function. Using Lemma 7.10 we find ν ∈ M+ (K) such
that µ ≺ ν and Qµ (f ) = ν(f ). Since µ is maximal, µ = ν. Hence
by Lemma 7.8(b).
For the proof of (ii) =⇒ (i), let ν ∈ M+ (K) satisfy µ ≺ ν. For any f ∈ −W(S)
we get, using Theorem A.84, that
Since µ(f ) ≤ ν(f ), µ(f ) = ν(f ). Thus µ = ν on W(S) − W(S), which yields
µ = ν by Proposition 7.6(h).
Corollary 7.25. Let S be a function cone on a compact space K. Then for every
x ∈ K there exists µ ∈ Mx (S) such that µ(B) = 1 for any K-analytic set containing
ChS (K).
Proof. Combine Theorem 7.18 with Theorem 7.24(c).
226 7 Choquet theory of function cones
Theorem 7.26. Let S be a function cone on a metrizable compact space K. Then the
following assertions hold:
(a) ChS (K) is of type Gδ .
(b) A measure µ ∈ M+ (K) is maximal if and only if µ(K \ ChS (K)) = 0.
(c) For any x ∈ K there exists µ ∈ Mx (S) such that µ(K \ ChS (K)) = 0.
Proof. We select a dense set {kn : n ∈ N} in Kc (S). To show (a), we verify that
∞
\
ChS (K) = {x ∈ K : kn∗ (x) = kn (x)}. (7.4)
n=1
Indeed, the inclusion “⊂” follows from Theorem 7.21. Conversely, if x is contained
in the right-hand side of (7.4), f ∗ (x) = f (x) for any f ∈ Kc (S) due to the density
of {kn : n ∈ N} and Lemma 7.8(c). Thus Theorem 7.21 again implies x ∈ ChS (K).
Now (7.4) yields assertion (a).
To verify (b), we first notice that any measure µ ∈ M+ (K) satisfying µ(K \
ChS (K)) = 0 is maximal due to Theorem 7.19 and Theorem 7.21. Conversely, any
maximal measure µ ∈ M+ (K) satisfies µ(f ) = µ(f ∗ ) for all f ∈ Kc (S), and thus
ChS (K), being the intersection of countably many sets {x ∈ K : kn∗ (x) = kn (x)},
n ∈ N, is of full measure µ.
The last assertion (c) follows from Theorem 7.18 and (b).
It follows from Theorem 7.24 that the value µ0 (A0 ) does not depend on the choice of
A. It is easy to verify that µ0 is a measure on the σ-algebra Σ.
If f is a Baire function on K, let f 0 := f |ChS (K) . Given an open set U ⊂ R,
We want to use Lemma A.88 in order to find a strictly positive function in F. To this
end, choose a nonzero measure µ ∈ M+ (K). By Theorem A.84 and Lemma 7.8(b)
there exists a function k ∈ S c (S) such that
s∗ < g < k ∧ t.
s∗ < f < g.
Then
µ(g − f ) ≤ µ(g − s∗ ) < µ(k − s∗ ) < εµ(K),
7.4 Simplicial cones 229
and thus
µ(f − g + ε) = εµ(K) + µ(f − g) > 0.
This proves the existence of a strictly positive element in F and thus we are able to
find a pair of functions f, −g ∈ Kc (S) such that s < f < g < t and g − f < ε.
Now we construct inductively functions fn , −gn ∈ Kc (S), n ∈ N, such that
• s < f < f
n n+1 < gn+1 < gn < t,
• gn − fn < 2−n .
Both sequences {fn } and {gn } converge uniformly to an S-affine continuous function
h satisfying s < h < t.
To verify (iv) =⇒ (iii), let s, −t ∈ Kc (S) satisfy s ≤ t. We inductively construct
a sequence {hn }∞ n=0 of S-affine functions such that
• s − 1 < h < t + 1,
0
• (hn−1 − 2−n ) ∨ (s − 2−n ) < hn < (hn−1 + 2−n ) ∧ (t + 2−n ), n ∈ N.
To start the construction, we use (iv) to find h0 ∈ Ac (S) with s − 1 < h0 < t + 1.
Assume that we have constructed functions h0 , . . . , hn−1 . From
s − 2−(n−1) < hn−1 < t + 2−(n−1)
we get
(hn−1 − 2−n ) ∨ (s − 2−n ) < (hn−1 + 2−n ) ∧ (t + 2−n ).
By Lemma 7.29(c),
((hn−1 − 2−n ) ∨ (s − 2−n ))∗ < ((hn−1 + 2−n ) ∧ (t + 2−n ))∗ .
Moreover, both functions are S-affine by Theorem 7.30. Hence we can use (iv) again
to find a function hn ∈ Ac (S) satisfying
(hn−1 − 2−n ) ∨ (s − 2−n ) < hn < (hn−1 + 2−n ) ∧ (t + 2−n ).
This completes the construction.
It is obvious that {hn } is a Cauchy sequence converging to h ∈ Ac (S) that satisfies
s ≤ h ≤ t.
To finish the proof we have to verify (ii) =⇒ (i). Pick x ∈ K and consider a
maximal measure µ ∈ Mx (S). If f ∈ Kc (S), it follows from (ii) that
f ∗ = inf{h ∈ Ac (S) : h ≥ f }
and the latter family is down-directed. Hence
µ(f ∗ ) = inf{µ(h) : h ∈ Ac (S), h ≥ f }
= inf{h(x) : h ∈ Ac (S), h ≥ f }
= f ∗ (x).
Thus f ∗ is S-affine, which concludes the proof by Theorem 7.30.
230 7 Choquet theory of function cones
f ∗ = inf{h ∈ Ac (S) : h ≥ f }
Proof. The first part follows from Theorem 7.31(iii) and the second part is an imme-
diate consequence of the first one.
Proof. The proof is identical with the proof of Theorem 6.8; we only have to use
Lemma 7.29(c) and Corollary 7.32.
Proof. The implication (i) =⇒ (ii) follows from Corollary 7.32 and (ii) =⇒ (i) is
obvious.
Theorem 7.36. Let S be a simplicial cone and T : K → M1 (K) be the kernel from
Definition 7.33. Then the following assertions hold:
(a) For any µ ∈ M+ (K),
7.4 Simplicial cones 231
(a1) T µ ∈ M+ (K),
(a2) kT µk = kµk,
(a3) T µ is S-maximal,
(a4) µ ≺ T µ.
(b) T µ is a boundary measure for any µ ∈ M(K).
(c) T εx = δx , x ∈ K.
Proof. We start the proof of (a) by noticing that H is indeed a closed function space.
Inclusion Mx (S) ⊂ Mx (H) follows by the definition, inclusions S c (H) ⊂ S c (S),
Mbnd (H) ⊂ Mbnd (S) and ChH (K) ⊂ ChS (K) are its consequences and Ac (H) =
H follows from Theorem 3.27.
Now assume that S is simplicial. Let s, −t ∈ Kc (H) with s ≤ t be given. Since
K (H) ⊂ Kc (S), Theorem 7.31 yields the existence of a function h ∈ H such that
c
Hence µ = ν and µ is H-maximal. Thus Mbnd (S) ⊂ Mbnd (H), which gives
Mbnd (S) = Mbnd (H) by (a). Since the equality ChS (K) = ChH (K) is then an
immediate consequence, the proof is complete.
Lemma 7.39. If f ∈ Kusc (S) and g ∈ S usc (S) with f ≤ g on ChS (K), then f ≤ g
on K.
232 7 Choquet theory of function cones
Theorem 7.40. Suppose that Ac (S) separates points of K. If Ac (S) is simplicial and
ChAc (S) (K) = ChS (K), then S is simplicial.
Proof. Let H stand for the space Ac (S). The aim of the proof is to show that any S-
maximal measure is H-maximal. To this end, let f ∈ Kc (H) ⊂ Kc (S) be given. We
temporarily write f ∗,H and f ∗,S for the respective envelopes. We claim that f ∗,H =
f ∗,S on K.
Indeed, f ∗,H = f on ChH (K) by Theorem 3.24 and f ∗,S = f on ChS (K) by
Theorem 7.21. Since H is simplicial, f ∗,H is H-affine (see Theorem 6.5) and thus
also S-affine. Since f ∗,S ∈ S usc (S) and f ∗,H = f ∗,S on ChS (K), Lemma 7.39
yields f ∗,H ≤ f ∗,S on K. Since f ∗,H ≥ f ∗,S by the definitions and Theorem 7.38(a),
the claim is proved.
Let µ ∈ Mmax (S) and f ∈ Kc (H) be given. Then µ(f ) = µ(f ∗,S ) by Theo-
rem 7.19, and thus µ(f ) = µ(f ∗,H ) by the previous step. Theorem 3.58 gives that µ
is H-maximal.
Since Mx (S) ⊂ Mx (H) for all x ∈ K, the simpliciality of S follows.
Proof. The first part is nothing else than Lemma A.20. If x, y ∈ X are distinct points,
we may assume that x y. By the first part, f (y) < f (x) for a suitable f ∈ Lc (X).
This concludes the proof.
7.5 Ordered compact convex sets and simplicial measures 233
Example 7.43. Let H be a function space on a compact space K. Consider the locally
convex space M(K) ordered by the cone
Then the ordering induced by this cone on M+ (K) is nothing else than the Choquet
ordering ≺H .
The sets M1 (K), Mx (H) where x ∈ K, Mµ (H) := {ν ∈ M+ (K) : µ − ν ∈
H⊥ } and {ν ∈ M+ (K) : µ ≺ ν ∈ H⊥ } where µ ∈ M+ (K) are ordered compact
convex sets.
In these examples, the cone of affine isotone continuous functions coincides with
the set of all H-convex continuous functions on K, where we view C(K) as a subspace
of (M(K))∗ .
Notation 7.44 (Maximal elements). If X is an ordered compact convex set, the set of
all points of X, which are maximal in the given ordering, is denoted by Xmax .
Proposition 7.45. Let X be an ordered compact convex set. Then the following as-
sertions hold:
(a) For every point x ∈ X, there exists a maximal element z ∈ X so that x ≤ z.
(b) The set Xmax is an extremal subset of X and ext Xmax = Xmax ∩ ext X.
(c) If Xmax is convex, it is a face of X.
Proof. Given y ∈ Ymax , we select x ∈ ϕ−1 (y). Using Proposition 7.45(a) we find
z ∈ Xmax with x ≤ z. Since ϕ−1 (y) is hereditary upwards due to the assumption on
ϕ, z ∈ ϕ−1 (y), and the proof is complete.
Proposition 7.50. Let X be an ordered compact convex set. Then −Lc (X) is a closed
function cone.
Theorem 7.54. Let X be an ordered compact convex set and −Lc (X) be considered
as a function cone on X. Then the following assertions hold:
(a) If µ, ν ∈ M1 (X) satisfy µ ≺−Lc (X) ν, then r(µ) ≤ r(ν).
(b) The function −fe is isotone for any f ∈ C(X).
(c) A point x is in Xmax if and only if fe(x) = f (x) for any f ∈ Ac (X).
(d) Ch−Lc (X) (X) = ext Xmax .
(e) For every x ∈ X there exists a −Lc (X)-maximal measure µ ∈ Mx (−Lc (X)).
Conversely, let x ∈ X \ Xmax be given. It means that there exists y ∈ X such that
x < y. Using Proposition 7.42, we find f ∈ Lc (X) such that f (x) < f (y). Then
by Proposition 7.45(b).
Since (e) is a consequence of Theorem 7.18, the proof is complete.
Theorem 7.55. Let X be an ordered compact convex set and −Lc (X) be considered
as a function cone on X. If
then for every x ∈ Xmax there exists a measure µ0 on the σ-algebra Σ such that
Z
f (x) = f (t) dµ0 (t), f ∈ Ac (X).
ext Xmax
Proof. Given the objects as in the hypothesis, let µ ∈ Mx (−Lc (X)) be a −Lc (X)-
maximal measure (use Theorem 7.18) and let µ0 be the induced measure on Σ given
by Theorem 7.27. Since εx ≺−Lc (X) µ, x ≤ r(µ) by Theorem 7.54(a). Since x is a
maximal element of X, x = r(µ). Hence Theorem 7.27 gives
Z Z
f (t) dµ0 (t) = f (t) dµ0 (t)
ext Xmax Ch−Lc (X)
on E generated by D.
Example 7.57. If X is any of the ordered compact convex sets from Example 7.43,
then D is the linear span of functions of the form
Theorem 7.58. Let X 6= ∅ be an ordered compact convex set such that the topology τ
from Definition 7.56 is well defined. Then Xmax is convex and Xmax = coτ ext Xmax .
Proof. It follows from Theorem 7.54(c) that
\ \
Xmax = {x ∈ X : h|
g X (x) = h(x)} = {x ∈ X : ϕh (x) = h(x)}.
h∈E ∗ h∈E ∗
Thus Xmax is convex and it is a closed set in the topology τ . Hence we have Xmax ⊃
coτ ext Xmax . We know from Proposition 7.47 that ext Xmax is a nonempty set. As-
sume that z ∈ Xmax \ coτ ext Xmax .
By the Hahn–Banach separation argument, there exists a function f ∈ D such that
i |X on X. Thus
linear functionals on E such that ϕhi = hg
n
X
f = h0 + i |X
ci hg
i=1
Pn
on X. By Theorem 7.54(c), f = h0 + i=1 ci hi on Xmax . Thus
n
X
h := h0 + ci hi
i=1
^
Since z ∈ Xmax , −h| X (z) = −h(z) by Theorem 7.54(c). Hence there exists an
isotone function l ∈ Lc (X) so that
Then
F := {x ∈ X : l(x) = max l(X)}
is a nonempty closed face which is hereditary upwards. By Proposition 7.47, F inter-
sects ext Xmax , say at a point y. Then
Proof. Notice that H is a subspace of L1 (ν) and that H is dense in L1 (ν) if and only
if
F ∈ (L1 (ν))∗ : F (h) = 0 for all h ∈ H = {0} .
If H is not dense in L1 (ν), then there exists f ∈ L∞ (ν) such that kf kL∞ (ν) = 1 and
Z
hf dν = 0 for all h ∈ H.
K
ν1 := (1 − f )ν and ν2 := (1 + f )ν.
ν1 + ν2
ν= and ν1 6= ν 6= ν2 ,
2
the measure ν is not simplicial.
7.5 Ordered compact convex sets and simplicial measures 239
For the converse, assume that H is dense in L1 (ν) and that ν ∈ Mµ (H) satisfies
ν = ν1 +ν
2
2
where ν1 , ν2 ∈ Mµ (H). Since 0 ≤ ν1 ≤ 2ν and 0 ≤ ν2 ≤ 2ν, the Radon–
Nikodym theorem yields positive functions f1 , f2 ∈ L∞ (ν) such that f1 + f2 = 2
ν-almost everywhere and νj = fj ν for j = 1, 2. For any h ∈ H we have
Z Z
f1 h dν = ν1 (h) = µ(h) = ν2 (h) = f2 h dν.
K K
Since H is dense in L1 (ν), the same equalities hold for any h ∈ L1 (ν). We see that
f1 = f2 = 1 ν-almost everywhere. Therefore ν = ν1 = ν2 .
Theorem 7.62. Let H be a function space on a compact space K. Then for any
x ∈ K and f ∈ Kc (H) there exists a simplicial measure µ ∈ Mx (H) such that
µ(f ) = f ∗ (x).
Proof. Let X denote the ordered compact convex set Mx (H). Then
F := {µ ∈ X : µ(f ) = f ∗ (x)}
7.6 Exercises
Exercise 7.63. Let X be a compact convex set and S := Sc (X) (see Definition 4.1).
Prove that
• ≺ c +
A (X) coincides with ≺S on M (X),
• S-maximal measures coincide with Ac (X)-maximal measures,
Further,
Ac (S) = {f ∈ C(K) : f (xn ) = (1 − n−1 )f (yn ) + n−1 µn (f ),
f (xn ) = (1 − n−1 )f (yn ) + n−1 νn (f ), n ∈ N}.
/ ChAc (S) (K). Indeed, for any function f ∈ Ac (S) and
is a function space and z ∈
n ∈ N we have
(1 − n−1 )f (yn ) + n−1 µn (f ) = (1 − n−1 )f (yn ) + n−1 νn (f ).
Hence
µn (f ) = νn (f ), n ∈ N,
which in turn yields Z
f (t) dt = f (z).
I
This concludes the proof.
Exercise 7.66. Find an example of an ordered compact convex set X such that Xmax
is not convex.
Hint. Consider E := R2 ordered by the cone E + = {(x, y) ∈ E : x ≥ 0, y ≥ 0},
and let X := {(x, y) ∈ E : x2 + y 2 ≤ 1}. Then
Xmax = {(x, y) ∈ X : x ≥ 0, y ≥ 0, x2 + y 2 = 1}
is not convex.
Exercise 7.67. If X is an ordered compact convex set, x ∈ Xmax and µ ∈ Mx (X),
then µ(fe) = µ(f ) for any f ∈ Ac (X).
Hint. Given f ∈ Ac (X), −fe is upper semicontinuous and concave. Using Corol-
lary 4.8 and Theorem 7.54(c), we get
f (x) = µ(f ) ≤ µ(fe) ≤ fe(x) = f (x).
Exercise 7.68. If X is a metrizable ordered compact convex set, then Xmax is a mea-
sure extremal subset of X.
Hint. Let {fn : n ∈ N} be a dense countable subset in BAc (X) . From Lemma 7.8(c)
and Theorem 7.54(c) we deduce that
∞
\
Xmax = {x ∈ X : fe(x) = f (x)}.
n=1
Exercise 7.69. Let X be a metrizable compact convex set and f be a bounded uni-
versally measurable function on X such that, for any maximal µ ∈ M1 (X), µ(f ) =
f (r(µ)). Prove that f is strongly affine.
Hint. Given a measure µ ∈ Mx (X), let ν ∈ M1 (X) be maximal with µ ≺ ν. Then
ν ∈ Mx (X) as well. We consider Y := M1 (X) as an ordered compact convex set
and let πi : Y × Y → Y be the projections, i = 1, 2. Let
M := {(εx , λ) ∈ Y × Y : εx ≺ λ}.
((π2 )] Ω)(Ymax ) = 1.
By assumption, ν(f ) = f (x). We apply Proposition 3.90 and the assumption to the
pairs (f, 0) and (0, f ) to get
Z Z
µ(f ) = λ1 (f ) dΩ(λ1 , λ2 ) = λ1 (f ) dΩ(λ1 , λ2 )
M M ∩(Y ×Ymax )
Z
= λ1 (f ) dΩ(λ1 , λ2 )
{(εx ,λ)∈Y ×Y :εx ≺λ,λ∈Ymax }
Z
= λ2 (f ) dΩ(λ1 , λ2 )
{(εx ,λ)∈Y ×Y :εx ≺λ,λ∈Ymax }
Z
= λ2 (f ) dΩ(λ1 , λ2 )
M
= ν(f ) = f (x).
and cok·k ext Xmax consists of discrete probability measures on K. Hence Xmax 6=
cok·k ext Xmax .
Problem 7.72. We do not know whether the assertion of Exercise 7.69 holds without
the metrizability assumption.
Chapter 8
Choquet-like sets
The aim of this chapter is to transfer the concept of faces of compact convex sets to
the framework of general function spaces. The most important examples are P -sets
and M -sets, which are analogues of parallel and split faces.
We start by recalling in Theorems 8.5 and 8.7 measure theoretic characterizations
of split and parallel faces that guide us to the general definition. In order to be able to
handle “faces” in function spaces as in the framework of compact convex sets, we need
several analogues of classical tools from convex analysis. This task is accomplished in
Section 8.2, where we define H-convex and H-extremal sets and prove results imitat-
ing the Krein–Milman theorem, the Milman theorem and the Hahn–Banach separation
theorem (see Corollary 8.19, Corollary 8.20 and Proposition 8.23).
In Section 8.3, we present analogues of faces in function spaces termed Choquet
sets. Particular examples of P -sets and M -sets are defined there and their basic char-
acterizations are proved (see Theorem 8.39 and Theorem 8.44). An important result
due to C. J. K. Batty on characterization of H-maximal measures is given in Theo-
rem 8.32. H-exposed sets are investigated in Section 8.4.
The next section provides preparatory results on boundary measures that are nec-
essary in Section 8.6. The central result of this chapter is Theorem 8.60 and its con-
sequences contained in Theorem 8.62 and Corollary 8.63. They provide characteriza-
tion of simplicial function spaces by means of Choquet sets. In the context of compact
convex sets we get a characterization of simplices by means of split and parallel faces.
Throughout this chapter, we frequently use the notation and results from Sec-
tion 4.3. Hence H is a function space on a compact space K, φ is the evaluation
mapping from K to the state space S(H) and X denotes a compact convex subset of
a locally convex space.
Definition 8.1 (Complementary sets). Let F ⊂ X. The union of all faces of X disjoint
from F is called the complementary set of F and it is denoted by F 0 .
8.1 Split and parallel faces 245
x = λy + (1 − λ)z, (8.1)
Remarks 8.3. (a) Since (cF )∗ is an upper semicontinuous function, the complemen-
tary set F 0 of a closed face F is a Gδ set.
(b) Notice that
{x ∈ X : cF (x) = c∗F (x)} ⊂ F ∪ F 0 .
In general, the decomposition x = λy + (1 − λ)z in (8.1) is not unique. The question
of uniqueness of the decomposition leads to notions of parallel and split faces and is
investigated next.
Definition 8.4 (Split faces). A face F of X is said to be a split face if its complemen-
tary set F 0 is convex and if any point in x ∈ X \ (F ∪ F 0 ) can be uniquely represented
as a convex combination
Theorem 8.5. The following statements about a closed face F of X are equivalent:
(i) F is a split face,
(ii) µ|F ∈ (Ac (X))⊥ for any boundary measure µ ∈ (Ac (X))⊥ .
Definition 8.6 (Parallel faces). A face F is called parallel if its complementary set F 0
is convex, X = co(F ∪ F 0 ), and if for every point x ∈ X \ (F ∪ F 0 ) the barycentric
coefficient λ in the convex combination
is uniquely determined.
246 8 Choquet-like sets
As in the case of split faces we have the following measure theoretic characteriza-
tion of parallel faces.
Theorem 8.7. The following statements about a closed face F of X are equivalent:
(i) F is a parallel face,
(ii) µ(F ) = 0 for any boundary measure µ ∈ (Ac (X))⊥ .
Proof. See B. Hirsberg [226], Theorem 2.12 or L. Asimow and A. J. Ellis [24], The-
orem 2.10.10.
and
1
H := f ∈ C(K) : f (2) = (λ1 (f ) + λ2 (f )) ,
2
then K is an H-extremal set. On the other hand, sj = π(λj ), j = 1, 2, are elements
of S(H),
1
φ(2) = (s1 + s2 ) ∈ φ(K),
2
while s1 and s2 do not belong to φ(K) since φ(K) = {π(εx ) : x ∈ K}.
8.2 H-extremal and H-convex sets 247
(b) Example 8.29 illustrates the fact that there is an H-extremal set H in K such that
co φ(H) is not an extremal subset of S(H).
F 0 := {x ∈ K : (cF )∗ (x) = 0} .
Definition 8.15 (closed H-convex hull). Let F be a subset of K. The closed H-convex
hull of F is the set
\
coH F := {C : C ⊃ F, C is closed and H-convex} .
and
Fb := x ∈ X : there is µ ∈ Mx (H) such that spt µ ⊂ F .
The set Fe is closed and contains F . Moreover, Fe is H-convex. Indeed, given x ∈ K
and µ ∈ Mx (H) such that spt µ ⊂ Fe, we have
Z Z
|h(x)| = |µ(h)| = h dµ ≤ |h| dµ ≤ sup {|h(t)| : t ∈ F }
Fe Fe
Corollary 8.20 (Milman type theorem). Let F be a subset of K such that K = coH F .
Then ChH (K) ⊂ F .
Proof. Take x ∈ ChH (K). Since x ∈ coH F , by Proposition 8.18 there exists µ ∈
Mx (H) such that spt µ ⊂ F . Since x ∈ ChH (K), µ = εx , which implies that
x ∈ F.
Remarks 8.21. (a) Let F be a subset of a compact convex set X and H = Ac (X).
Then coH F is nothing else than the closed convex hull of F .
(b) It is not true that, for a Borel subset F of a compact convex set X and H = Ac (X),
Consider the compact convex set X := M1 ([0, 1]) and F := {εx : x ∈ S} where
S is a countable dense subset of [0, 1]. Then co F = X while for any measure Λ ∈
M1 (X) representing Lebesgue measure λ ∈ X we have Λ(F ) = 0.
Proof. Let F be a closed H-convex set and let x be in K such that φ(x) ∈ co φ(F ).
By Proposition 4.28(b), there exists a measure µ ∈ M1 (φ(F )) representing the point
φ(x). Then (φ−1 )] µ ∈ Mx (H) and spt µ ⊂ F . Thus x ∈ coH F = F .
Conversely, let F = φ−1 (co F ∩ φ(K)) and let µ be a measure representing a point
x ∈ K such that spt µ ⊂ F . Then spt φ] µ ⊂ φ(F ), and hence φ(x) ∈ co φ(F ). Due
to the assumption, φ(x) ∈ φ(F ).
Proof. Let µ ≺ ν and x ∈ spt µ \ coH spt ν. By Proposition 8.23, there exists a
function h ∈ H such that h(x) > max h(coH spt ν). By adding a suitable constant,
we may assume that max h(coH spt ν) = 0. Then h ∨ 0 is a continuous H-convex
function and
µ(h ∨ 0) > 0 ≥ ν(h ∨ 0)
which contradicts the assumption µ ≺ ν.
Lemma 8.25. Let F be a closed set in K. Then F ∩ ChH (K) = (coH F ) ∩ ChH (K).
250 8 Choquet-like sets
Proof. The inclusion “⊂” being obvious, we need to verify the reverse one. Let
x ∈ (coH F ) ∩ ChH (K) be given. By Proposition 8.18 there exists a measure
µ ∈ Mx (H) ∩ M1 (F ). Since x ∈ ChH (K), µ = εx and x ∈ F , which completes
the proof.
Proof. Let L be a compact subset of coH F \ F . Given ε > 0, there exists a compact
set M ⊂ K \ coH F such that
Hence, a closed set F is a Choquet set if and only if the following two conditions
are satisfied:
(a) given x ∈ F and µ ∈ Mx (H), then spt µ ⊂ F ,
(b) if x ∈ K \ F and µ ∈ Mx (H), then spt µ is not contained in F .
Example 8.28 (Convex case). Let X be a compact convex subset of a locally convex
space. A closed set F ⊂ X is a Choquet set if and only if F is a closed face of X.
This assertion follows by the fact that a closed set F in X is convex or extremal if and
only if F is Ac (X)-convex or Ac (X)-extremal, respectively.
8.3 Choquet sets, M -sets and P -sets 251
Example 8.29 (State space). Let H be a closed face of S(H). Then the set F :=
φ−1 (H ∩ φ(K)) is a closed Choquet set. Indeed, F is H-convex due to Proposi-
tion 8.22 and H-extremal by Lemma 8.10.
The converse implication is false in general. Let K := [0, 1] ∪ [3, 4]. Denote by λ1
and λ2 restrictions of Lebesgue measure on [0, 1] and [3, 4], respectively. Let
H := {f ∈ C(K) : λ1 (f ) = λ2 (f )}.
Then F := [0, 1] is a Choquet set but the set H := co φ(F ) is not Ac (S(H))-extremal
in S(H). Indeed, the state s := π(λ1 ) belongs to H, the measure φ] λ2 represents s
and (φ] λ2 )(H) = 0.
Proposition 8.30. (a) If F is a closed H-extremal set, then F ⊂ coH (F ∩ ChH (K)).
(b) For any closed Choquet set F ⊂ K we have F = coH (F ∩ ChH (K)).
(c) The family of closed Choquet sets is stable with respect to arbitrary intersections.
Proof. For the proof of (a), let F be a closed H-extremal set and let x ∈ F be given.
First, we show that
h(x) ≤ sup h(F ∩ ChH (K)) (8.2)
holds for any strictly positive h ∈ H. Let h ∈ H be a strictly positive function. Then
f := hcF is an upper semicontinuous H-convex function, and hence it attains its
maximum at some point y ∈ ChH (K) (see Lemma 3.13 and Proposition 3.15). Since
h is strictly positive, y ∈ F ∩ ChH (K), and (8.2) follows.
If h ∈ H is arbitrary, we apply the previous argument to h + c for a suitable c ∈ R
to obtain (8.2) for any h ∈ H. Given h ∈ H, (8.2) applied to h and −h yields
|h(x)| ≤ sup |h|(F ∩ ChH (K)). By Proposition 8.18, x ∈ coH (F ∩ ChH (K)).
To verify (b), we first notice that obviously F ⊃ coH (F ∩ ChH (K)). Since the
reverse inclusion follows by (a), the proof of (b) is complete.
By combining Propositions 3.14 and 8.17 we get (c).
Proposition 8.31. If H is a simplicial function space, then coH F is a Choquet set for
any compact F ⊂ ChH (K).
(ii) there exists a set S ⊂ C(K) separating points of K such that each function in S
is constant on Fx (H) for µ-almost all x ∈ K,
(iii) any continuous function on K is constant on Fx (H) for µ-almost all x ∈ K.
Proof. We start the proof by showing (i) =⇒ (ii). Let µ be a maximal measure and
let h ∈ H be arbitrary. Then µ((h2 )∗ ) = µ(h2 ) by Theorem 3.58, and thus (h2 )∗ = h2
on a set A ⊂ K with µ(K \ A) = 0. Let x ∈ A be arbitrary and ν ∈ Mx (H). Then,
by the Hölder inequality,
Thus Z Z Z
2 2
(h − h(x)) dν = h dν − 2h(x) h dν + h2 (x) = 0.
K K K
Hence h = h(x) on spt ν, and h is constant on Fx (H). Thus we can take H to be the
required family S of continuous functions.
For the proof of (ii) =⇒ (iii), let S ⊂ C(K) be as in (ii). Then the family
is a closed linear space stable with respect to pointwise multiplication and maxima.
Since S ⊂ F, F = C(K) by the Stone–Weierstrass theorem.
Assume now that (iii) holds. Given a function f ∈ C(K), let x ∈ K be a point such
that f is constant on Fx (H). We find a measure ν ∈ Mx (H) such that ν(f ) = f ∗ (x)
(see Lemma 3.21). Since x ∈ Fx (H) and spt ν ⊂ Fx (H), f = f (x) on Fx (H). Thus
Thus f (x) = f ∗ (x) for µ-almost all x ∈ K, and thus µ(f ) = µ(f ∗ ). By Theo-
rem 3.58, µ is maximal.
Definition 8.33 (Conditions (M) and (P)). We say that a closed set F ⊂ K satisfies
condition (M) if
Proposition 8.34. Let F ⊂ K be a closed set. Then F satisfies condition (M) or (P)
if and only if the set coH F satisfies (M) or (P), respectively.
Proof. If µ ∈ H⊥ ∩ Mbnd (H) is given, Lemma 8.26 asserts that µ|F = µ|coH F , from
which the assertion readily follows.
8.3 Choquet sets, M -sets and P -sets 253
Example 8.35 (State space). If a closed set F satisfies (M), then φ(F ) satisfies (M)
in the state space as well. Indeed, given µ ∈ Ac (S(H))⊥ ∩ Mbnd (S(H)), we have
(φ−1 )] µ ∈ H⊥ ∩ Mbnd (H) which yields (φ−1 )] µ|F ∈ H⊥ . Then µ|φ(F ) ∈
Ac (S(H))⊥ .
Similarly, we see that φ(F ) satisfies (P) provided F satisfies (P).
Definition 8.36 (P-sets). We say that a set F is a P -set if F is a closed Choquet set
satisfying condition (P).
Example 8.37 (Convex case). If F is a closed face in a compact convex set X, then
F is a P -set if and only if F is a parallel face. This follows from a measure theoretic
characterization of parallel faces stated in Theorem 8.7.
Proposition 8.38 (State space). The following assertions hold:
(a) If H is a closed parallel face of S(H), then the set F := φ−1 (H ∩ φ(K)) is a
P -set.
(b) If F is a P -set, then co φ(F ) is a parallel face of S(H).
Proof. (a) We already know from Example 8.29 that F is a Choquet set. Let µ ∈
H⊥ ∩ Mbnd (H) be given. Then φ] µ ∈ Ac (S(H))⊥ ∩ Mbnd (S(H)). Thus we get
which gives the desired conclusion that co φ(F ) is extremal and consequently a face.
Hence, it is a parallel face by Example 8.37.
254 8 Choquet-like sets
Theorem 8.39. Let F be a closed H-convex subset of K. Then the following asser-
tions are equivalent:
(i) F is a P -set,
(ii) µ(c∗F ) = ν(c∗F ) for any probability measures µ and ν with µ − ν ∈ H⊥ ,
(iii) the set {h ∈ H : h > cF } is down-directed,
(iv) the function c∗F belongs to H⊥⊥ ,
(v) there exists a parallel face H ∈ S(H) such that F = φ−1 (H ∩ φ(K)).
Proof. The equivalence (i) ⇐⇒ (v) follows from Proposition 8.38 and Proposition
8.22.
Assume now that F is a P -set and choose µ, ν ∈ M1 (K) such that µ − ν ∈ H⊥ .
Lemma 3.21 provides a measure µ0 ∈ M1 (K) so that µ ≺ µ0 and Qµ (cF ) = µ0 (cF ).
Let µm be a maximal measure satisfying µ0 ≺ µm . Since F is a closed Choquet set,
cF is upper semicontinuous and H-convex, and thus µ0 (cF ) ≤ µm (cF ). Let ν m be a
maximal measure with ν ≺ ν m . Then µm − ν m ∈ H⊥ .
Then, using Lemma 3.18, Proposition 3.56, Theorem 3.58, and condition (P) we
get
µ(c∗F ) = Qµ (cF ) = µ0 (cF )
≤ µm (cF ) = µm (F ) = ν m (F )
= ν m (cF ) = ν m (c∗F ) ≤ ν(c∗F ).
By interchanging µ with ν we get that µ(c∗F ) = ν(c∗F ). Hence, (i) =⇒ (ii).
Suppose next that (ii) holds and that h1 , h2 ∈ H, h1 > cF , h2 > cF . Then
cF < h1 ∧ h2 . Let µ, ν ∈ M1 (K) with µ − ν ∈ H⊥ be given. Since
∗
µm = µm |F + µm |G and ν m = ν m |F + ν m |G .
By condition (M),
µm |F − ν m |F ∈ H⊥ .
256 8 Choquet-like sets
It follows that
µm |G − ν m |G ∈ H⊥ .
Then
µ(fe) ≤ µm (fe) = µm |F (f0 ) + µm |G (f ) = ν m |F (f0 ) + ν m |G (f )
< ν m |F (g0 ) + ν m |G (g) = ν m (e
g ) ≤ ν(e
g ).
Using Lemma 5.43 we can find a function h ∈ H such that fe < h < ge. This function
possesses all the required properties.
Theorem 8.44. Let F be a closed H-convex subset of K. Then the following asser-
tions are equivalent:
(i) F is an M -set,
(ii) µ((f cF )∗ ) = ν((f cF )∗ ) for any measures µ, ν with µ−ν ∈ H⊥ and any f ∈ H
satisfying f |F ≥ 0,
(iii) the set {h ∈ H : h > f cF } is down-directed for any f ∈ H satisfying f |F ≥ 0,
(iv) for any f ∈ H satisfying f |F ≥ 0, we have (f cF )∗ ∈ H⊥⊥ ,
(v) there exists a closed split face H ∈ S(H) such that F = φ−1 (H ∩ φ(K)).
Proof. As in the proof of Theorem 8.39 we notice that (i) ⇐⇒ (v) by Proposi-
tion 8.42 and Proposition 8.22. For the proof of (i) =⇒ (ii), let f ∈ H with
f |F ≥ 0 be given and let µ, ν ∈ M1 (K) satisfy µ − ν ∈ H⊥ . We proceed sim-
ilarly as in Theorem 8.39. Lemma 3.21 provides a measure µ0 ∈ M1 (K) so that
Qµ (f cF ) = µ0 (f cF ) and µ ≺ µ0 . Let µm be a maximal measure satisfying µ0 ≺ µm .
Since F is a closed Choquet set, f cF is upper semicontinuous and H-convex, and
thus µ0 (f cF ) ≤ µm (f cF ). Let ν m be a maximal measure with ν ≺ ν m .
Then, using Lemma 3.18, Proposition 3.56 and Theorem 3.58, we get
µ((f cF )∗ ) = Qµ (f cF ) = µ0 (f cF )
≤ µm (f cF ) = ν m (f cF )
= ν m ((f cF )∗ ) ≤ ν((f cF )∗ ).
Definition 8.45 (Archimedean sets). A closed set F is called Archimedean if for each
g ∈ H with g ≥ 0 on F there exists a function h ∈ H+ such that h = g on F and
h ≥ g on K.
Proposition 8.46. Let H be a closed function space on a compact space K. Then any
M -set is Archimedean.
Proof. Let F ⊂ K be an M -set, g ∈ H, g ≥ 0 on F . By Lemma 8.43, there is
h1 ∈ H such that
Since khn − hn−1 k < 2−n , there is a positive function h ∈ H such that hn → h
uniformly on K. Obviously, h = g on F and h ≥ g on K.
where, for any x ∈ K \ F , hx is a function from H+ such that hx (x) > 0 and hx = 0
on F . Conversely, any intersection of H-exposed sets is relatively H-exposed.
Since any H-exposed set is clearly closed, a relatively H-exposed set F (being an
intersection of H-exposed sets) is itself closed.
Let µ ∈ M1 (F ) be a measure representing x ∈ K. If x ∈ / F , then there exists a
function hx ∈ H+ with hx (x) > 0 and hx = 0 on F . Then
Corollary 8.50. Let H be a closed function space on a compact space K. Then any
M -set is relatively H-exposed.
Proof. This follows immediately from Propositions 8.46 and 8.49.
Proof. Suppose that (µ, ν) ∈ ext Mmax is a pair of nontrivial measures. First we
realize that µ, ν ∈ M1 (K). Indeed, since (µ, ν) ∈ M, then µ(K) = ν(K) ≤ 1.
Suppose that 0 < µ(K) = ν(K) < 1. Then
µ ν
(µ, ν) = µ(K) , + (1 − µ(K))(0, 0),
µ(K) µ(K)
2 > kµ − νk = kη + k + kη − k = 2kη + k.
η−
+
+ η λ λ
(µ, ν) = kη k , + kλk , ,
kη + k kη + k kλk kλk
H⊥ ⊥
2 := H ∩ {µ ∈ M(K) : kµk ≤ 2}
2 = kηk = kη + k + kη − k.
(η + , η − ) = α(µ1 , µ2 ) + (1 − α)(ν1 , ν2 )
η = α(µ1 − µ2 ) + (1 − α)(ν1 − ν2 ).
Since
µ − ν = η = α(η1+ − η1− ) + (1 − α)(η2+ − η2− )
= αη1+ + (1 − α)η2+ − (αη1− + (1 − α)η2− ),
we get
µ ≤ αη1+ + (1 − α)η2+ and η ≤ αη1− + (1 − α)η2− .
Since all these measures are probability,
Since (µ, ν) ∈ ext Mmax , we get η1+ = η2+ and η1− = η2− . Thus η1 = η2 , and hence
µ − ν ∈ ext(H⊥
2 ∩ Mbnd (H)).
Lemma 8.55. If H⊥ ⊥
2 ∩ Mbnd (H) 6= {0}, then ext(H2 ∩ Mbnd (H)) 6= ∅.
8.5 Weak topology on boundary measures 261
Proof. By Lemma 8.52 and Lemma 8.54 it suffices to show that the set
is nonempty.
Let D be the linear span of functionals on M(K) × M(K) of the form
and
(µ, ν) 7→ µ(f ∗ ) + ν(g ∗ ), f, g ∈ C(K).
Consider the locally convex topology σ on M(K) × M(K) generated by functionals
from D. It follows from Theorem 7.58 that
Now the proof of the lemma follows easily. If the set A were empty, then every
pair (µ, ν) ∈ Mmax would satisfy µ = ν. But our hypothesis assures that there is
a nonzero measure µ ∈ H⊥ +
2 ∩ Mbnd (H). Then its positive and negative parts µ ,
− + − + −
µ satisfy (µ , µ ) ∈ Mmax and µ 6= µ , a contradiction. This concludes the
proof.
Definition 8.56 (Weak topology generated by upper envelopes). Let D be the linear
span of the family C(K)∪{f ∗ : f ∈ C(K)} in the space of all bounded Borel functions
on K. We denote by τ the (locally convex) topology on M(K) generated by the
functionals
µ 7→ µ(f ), f ∈ D.
η ∈ H⊥ τ ⊥
2 ∩ Mbnd (H) \ co (ext(H2 ∩ Mbnd (H)).
262 8 Choquet-like sets
where the latter equality follows by Lemma 8.54 and Lemma 8.52. Since (µ, ν) ∈
ext Mmax if and only if (ν, µ) ∈ ext Mmax , it follows that s ≥ 0.
By the definition of τ there are continuous functions f0 , . . . , fn on K such that
n
X
f = f0 + fi∗ .
i=1
Pn
Set g := f0 + i=1 fi and let η + and η − be the positive and the negative part of η.
Since η is a boundary measure, we obtain
As each set Fn+ is H-exposed, it satisfies condition (P), and so by Lemma 8.58 we
have
Z
+ +
|µ(Fn ) − µn (Fn )| ≤ sup h d(µ − µn )
{h∈H:khk≤2} K
Thus |µ(Fn+ ) − µn (Fn+ )| → 0. Since Fn+ is an H-exposed set, µ(Fn+ ) = 0 due to the
assumption. Recall that spt µ+ + − +
n ⊂ Fn and spt µn ∩ Fn = ∅. Hence,
−
kµ+ + + + + + +
n k = µn (Fn ) = µn (Fn ) − µn (Fn ) = µn (Fn )
= µ(Fn+ ) + µn (Fn+ ) − µ(Fn+ ) ≤ |µn (Fn+ ) − µ(Fn+ )|,
−
1 = kLk = lim kLn k = lim kµn k = lim kµ+
n→∞ n→∞ n k + kµn k = 0,
n→∞
Theorem 8.60. Let H be a closed function space on a compact space K. Then the
following assertions are equivalent:
(i) H⊥ ∩ Mbnd (H) = {0},
(ii) any closed Choquet subset of K is an M -set,
(iii) any closed Choquet subset of K is a P -set,
(iv) any H-exposed subset of K is a P -set.
Proof. The implications (i) =⇒ (ii) =⇒ (iii) =⇒ (iv) are obvious. It remains to
prove that (iv) =⇒ (i).
By Theorem 8.57 it is enough to show that there are no nonzero measures in
ext(H⊥ ⊥
2 ∩ Mbnd (H)). Suppose that µ is a nonzero element of ext(H2 ∩ Mbnd (H)).
Let µ = µ+ − µ− be the decomposition of µ into the positive and negative part.
Since spt µ+ cannot be a singleton, we can find a Borel subset B of K so that
µ+ (K) 6= µ+ (B) 6= 0. Set
µ+ (B) −
λ1 := µ+ |B − µ and λ2 := µ − λ1 .
µ− (K)
Then kµk = kλ1 k + kλ2 k and both λ1 , λ2 are nontrivial boundary measures. If
λ1 ∈ H⊥ , then both λ1 and λ2 are contained in H⊥
2 ∩ Mbnd (H) and µ is a nontrivial
convex combination of suitable multiples of λ1 and λ2 . Since this contradicts the
/ H⊥ . Set
extremality of µ, we have λ1 ∈
Z
L(h) := h dλ1 , h ∈ H.
K
Put
spt µ+ +
n ⊂ Fn and spt µ− −
n ⊂ Fn .
Proof of Claim 1. Suppose the contrary. Hence there exists an increasing sequence
{nk } in N so that spt µ ⊂ Fn+k . As in the proof of Lemma 8.59 we obtain that
|µ− − − − −
nk (Fnk ) − λ1 (Fnk )| = |µnk (Fnk ) − λ1 (Fnk )| → 0.
Since we assume that spt µ ⊂ Fn+k , we have spt λ1 ⊂ Fn+k as well. Thus we get
lim kµ−
nk k = 0.
k→∞
= lim µ+
nk (1) = lim kµnk k = lim kLnk k = kLk
k→∞ k→∞ k→∞
Similarly we deduce that spt µ * Fn− for all but finitely many n ∈ N.
lim kµ+ + +
n k = lim µn (Fn ) = 0.
n→∞ n→∞
−
kLk = lim kLn k = lim kµn k = lim kµ+
n→∞ n→∞ n→∞ n k + kµn k = 0,
Remark 8.61. Before proceeding, we have a need of a short note. Given a function
space H on K, the space Ac (H) of all continuous H-affine functions on K can be
considerable larger than H. Although a lot of objects determined by H and by Ac (H)
coincide (for example, representing measures, boundary measures, Choquet sets), this
coincidence is no longer valid for M -sets, P -sets and exposed sets. In order to obtain
next Theorem 8.62 in a full generality, we must specify whether we think over M -
sets and P -sets with respect to H or with respect to Ac (H). In order to avoid an
inaccuracy, we point out that conditions (iii)–(v) of the next theorem are laid down
for the function space Ac (H). More precisely, for example, condition (iv) means that
any closed Choquet set F ⊂ K satisfies µ(F ) = 0 for any measure µ ∈ (Ac (H))⊥ ∩
Mbnd (H).
With these preliminaries out of the way, we may state the main result of this chapter.
Theorem 8.62. Let H be a function space on a compact space K. Then the following
assertions are equivalent:
(i) H is simplicial,
(ii) (Ac (H))⊥ ∩ Mbnd (H) = {0} ,
(iii) any closed Choquet subset of K is an M -set with respect to Ac (H),
(iv) any closed Choquet subset of K is a P -set with respect to Ac (H),
(v) any Ac (H)-exposed subset of K is a P -set with respect to Ac (H).
8.6 Characterizations of simpliciality by Choquet sets 267
Proof. We already know from Theorem 8.60 that assertions (ii)–(v) are equivalent
(note that the space Ac (H) is closed). Since (i) ⇐⇒ (ii) by Proposition 6.9, the proof
is finished.
Corollary 8.63. For a compact convex set X the following assertions are equivalent:
(i) X is a simplex,
(ii) any closed face of X is a split face,
(iii) any closed face of X is a parallel face,
(iv) any exposed subset of X is a parallel face.
Proof. This follows from Theorem 8.62, if we use the characterization of split and
parallel faces from Example 8.41 and Example 8.37.
Corollary 8.64. Let H be a simplicial function space. Then closed Choquet sets co-
incide with relatively Ac (H)-exposed sets.
Proof. If F is a closed Choquet set, then by Theorem 8.62, F is an M -set with respect
to Ac (H). Corollary 8.50 asserts that F is relatively Ac (H)-exposed.
The converse implication follows by Proposition 8.48 and the observation that a set
F is a Choquet set with respect to H if and only if F is a Choquet set with respect to
Ac (H) (in virtue of the equality Mx (H) = Mx (Ac (H)) for any x ∈ K).
There are simplicial function spaces where supports of maximal measures are big
enough. In this case we are able to give a better characterization of both closed Cho-
quet sets and M -sets.
Proposition 8.66. Let H be a simplicial function space so that spt δx = ChH (K) for
any x ∈ K \ ChH (K). Then a closed set F ⊂ K is Choquet if and only if either
F = K or F is a proper subset of ChH (K).
Proof. Assume first that F is a Choquet set which is not a proper subset of ChH (K).
We show that ChH (K) ⊂ F . If z ∈ F \ ChH (K), then spt δz = ChH (K) by the
assumption. Since F is an H-extremal set, we have spt δz ⊂ F . Hence ChH (K) ⊂ F .
By Corollary 8.19,
K = coH (ChH (K)) ⊂ coH F = F.
Conversely, assume that F is a closed proper subset of ChH (K). By Example 8.9,
F is H-extremal. We now wish to show that F is H-convex. To this end, fix z ∈ K
268 8 Choquet-like sets
Proposition 8.67. Let H be a simplicial space such that H = Ac (H) and spt δx =
ChH (K) for any x ∈ K \ ChH (K). Then a closed set F ⊂ K is an M -set if and only
if either F is a proper subset of ChH (K) or equals K.
8.7 Exercises
Exercise 8.68. Let F ⊂ K be a closed H-extremal set and µ a Radon measure carried
by F . If ν is a Radon measure, µ ≺ ν, then ν is carried by F as well.
Hence ν is carried by F .
Exercise 8.69. Let F ⊂ K and let µ be a Radon measure carried by the complemen-
tary set F 0 . If ν is a Radon measure such that µ ≺ ν, then ν is carried by F 0 .
Hint. Since the function c∗F is upper semicontinuous and H-concave, we have by
Proposition 3.56
0 ≤ ν(c∗F ) ≤ µ(c∗F ) = 0.
It follows that ν(c∗F ) = 0, which yields the conclusion that ν is carried by F 0 .
Exercise 8.71. Let F be a P -set. Then c∗F ∈ H if and only if the complementary set
F 0 is closed.
8.7 Exercises 269
Hint. Clearly, if c∗F ∈ H, then the complementary set F 0 is closed since c∗F is contin-
uous and F 0 = {x ∈ K : c∗F (x) = 0}.
Conversely, suppose that F 0 is closed. Choose ε > 0. By (i) =⇒ (iii) of Theo-
rem 8.39 and Dini’s theorem, there exists a function h ∈ H so that h > cF and
h<1+ε on F, h < ε on F 0 .
Exercise 8.72. Find an example of a compact set F ⊂ ext X such that co F is not a
face.
Hint. Consider the unit square in R2 .
Exercise 8.73. Let H be a closed function space. If F is a P -set for which its com-
plementary set F 0 is closed, then F is Archimedean.
Hint. Let g ∈ H, g ≥ 0 on F . Consider the function u := 1 − c∗F . Then u ∈ H by
Exercise 8.71, 0 ≤ u ≤ 1, and
Exercise 8.74. Clearly, any H-exposed set F is a Gδ subset of K. Prove the following
converse assertions.
(a) Let H be a closed function space. Then any relatively H-exposed Gδ set F is
H-exposed.
(b) Relatively H-exposed sets are H-exposed provided the compact space K is
metrizable.
270 8 Choquet-like sets
Exercise 8.76. Let X be a compact convex set, E := Ac (X) and φ : X → S(Ac (X))
be the evaluation mapping. Prove that φ(X) is a parallel face of BE ∗ .
Hint. It is easy to show that φ(X) and −φ(X) are closed faces. By Proposition
4.31(c), BE ∗ = co(φ(X) ∪ −φ(X)). Hence, φ(X) and −φ(X) are complementary
faces. They are parallel by Proposition 4.31(b).
To verify (b), we use Theorem 8.5. Let µ ∈ (Ac (X))⊥ be a boundary measure of
norm 1 and assume that µ = cεx + ν, where c 6= 0 and |ν|({x}) = 0. We select
ε ∈ (0, 1) such that |c|(1 − ε) − 2ε > 0 and a closed neighborhood U of x with
|ν|(U ) < ε. Let h ∈ Ac (X) be chosen for U and ε. Then |h| ≤ ε on ext X \ U , and
hence Z Z Z
|h| d|ν| = |h| d|ν| + |h| d|ν| ≤ ε + ε.
X U ext X\U
Thus
0 = |µ(h)| ≥ |cεx |(h) − |ν|(|h|) ≥ |c|(1 − ε) − 2ε > 0,
a contradiction.
For the proof of (c), let {x} be a parallel face, ext X closed and U 3 x along with
ε > 0 be given. We choose η > 0 with (1 + η)−1 > 1 − ε.
By Theorem 8.39, (c{x} )∗ is an affine function, the family
is down-directed and its infimum equals (c{x} )∗ . Since (c{x} )∗ = 0 on ext X \ {x}
and ext X is a closed set, Dini’s theorem yields the existence of a function h ∈ Ac (X)
such that c{x} < h < 1 + η and h < η on ext X \ U . Then (1 + η)−1 h is the sought
function witnessing that x is a weak peak point.
Exercise 8.78. Find an example of an extreme point of a simplex that is not a weak
peak point.
Hint. Consider an extreme point x of the Poulsen simplex S (see Definition 12.56).
Let x0 ∈ S be arbitrary and y = 13 x + 23 x0 . If U is a closed convex neighborhood of
x not containing y, then there is no h ∈ Ac (S) such that khk ≤ 1, h(x) > 1 − 71 and
|h| < 17 on ext S \ U .
Indeed, assume that h is such a function. Since x0 ∈ / U and ext S is dense in S,
h(x0 ) ≤ 71 . Then
1 2 4
h(y) = h(x) + h(x0 ) ≥ .
3 3 21
Let {xn } be a sequence in ext S converging to y. Since all but finitely many points xn
are in X \ U , h(y) ≤ 71 . Thus we get a contradiction.
Exercise 8.79. Let X ⊂ Rd be a compact convex set such that {x} is a split face for
each x ∈ ext X. Prove that X is a simplex.
272 8 Choquet-like sets
n
X
x = a1 x1 + a a−1 ai xi .
i=2
Since {x1 } is a split face and ni=2 a−1 ai xi is contained in the complementary face
P
{x1 }0 , there is only one possibility how to decompose x as a convex combination of
x1 and an element from {x1 }0 . Proceeding inductively we get that for any x ∈ X there
is only one possibility how to write it as a convex combination of extreme points.
Further we show that X has at most d + 1 extreme points. Assume that this is
not the case and {x0 , . . . , xd+1 } are distinct extreme points of X. Then they are not
affinely
Pd+1 independent, and thus there exist nonzero real numbers a0 , . . . , ad+1 with
Pd+1
i=0 ai = 0 such that i=0 ai xi = 0. We assume that a0 , . . . , aj are positive and
aj+1 , . . . , ad+1 are negative. If a := ji=0 ai , then
P
j
X d+1
X
a−1 ai xi = a−1 ai xi
i=0 i=j+1
are different convex combinations of extreme points expressing the same point. Hence
we have arrived to a contradiction with the first part of the proof.
Combining both parts together we get that ext X is finite and any x ∈ X can be
written uniquely as a convex combination of ext X. Thus X is a simplex.
Exercise 8.80. Find an example of a compact convex set X such that each extreme
point of X is a weak peak point and X is not a simplex.
Hint. Consider the compact space K := [0, 1] ∪ [2, 3] and the function space
n Z 1 Z 3 o
H := f ∈ C(K) : f (t) dt = f (t) dt ,
0 2
and let X := S(H). We denote by λ1 and λ2 Lebesgue measure on [0, 1] and [2, 3], re-
spectively. The set X is not a simplex since the point π(λ1 ) has two different measures
carried by ext X = φ(ChH (K)), namely φ] λ1 and φ] λ2 (here we use the notation and
results from Section 4.3).
On the other hand, if x ∈ K, ε > 0 and an open set U 3 x are given, it is not
difficult to construct a function f ∈ H such that f (x) = 0, 0 ≤ f ≤ 1 and f > 1 − ε
on K \ U . Hence any extreme point of X is a weak peak point.
8.8 Notes and comments 273
Section 8.5 is based upon E. Briem’s paper [92]. The central Section 8.6 combines
some techniques of A. J. Ellis and A. K. Roy [167] (see also Theorem 3.2.5 in [24]),
E. Briem [94] and E. G. Effros [163], [164]. We also mention another characterization
of simplices given by C. J. K Batty [33] which reads as follows:
A compact convex set X is a simplex if and only if there exists C > 0 such that,
whenever F ⊂ X is a closed extremal set, each positive f ∈ Ac (coF ) has a positive
extension g ∈ Ac (X) with kgk ≤ Ckf k.
Corollary 8.64 is a slightly more general version of Theorem 2.5 in R. E. Atalla
[25], where it is proved that for Markov simplicial function spaces (see Definition
6.43) the class of relatively H-exposed sets coincides with the class of closed Choquet
sets.
Exercises 8.77 and 8.78 are taken from C. H. Chu and B. Cohen [119].
Chapter 9
Topologies on boundaries
In this chapter we further explore families of H-extremal and Choquet sets and show
that they generate compact topologies on Choquet boundaries. The fact that these
topologies are typically non Hausdorff turns out to be a minor inconvenience. The first
section introduces the topologies σext and σmax generated by closed H-extremal and
H-maximally extremal sets. Theorem 9.10 shows that these topologies are Hausdorff
if and only if the Choquet boundary is closed.
Theorem 9.12 provides a key tool for introducing measures on Choquet boundaries
induced by maximal measures. This procedure is clarified by Theorem 9.19 which
shows that the induced measures are regular in some sense. As a consequence we
get that a bounded H-affine Baire function satisfies the barycentric calculus with re-
spect to the induced measures (see Corollary 9.20). Next we characterize functions
continuous with respect to the topologies σext (see Theorem 9.25) and σmax (see The-
orem 9.29).
Section 9.4 is devoted to a study of H-strongly universally measurable functions.
The main result, Theorem 9.36, proves that upper semicontinuous affine functions on
compact convex sets are measurable with respect to the completion of the induced
measure from Theorem 9.19 and that they satisfy the barycentric calculus with re-
spect to these measures. As a consequence we obtain that H-strongly universally
measurable functions satisfy the barycentric calculus with respect to these measures
(see Theorem 9.38).
The last section describes properties of the so-called facial topology generated by
M -sets. It turns out in Theorem 9.48 that facially continuous functions on Choquet
boundaries can be identified with the center of H (see Definition 9.46). In the case of
prime simplicial function spaces, there are no nonconstant facially continuous func-
tions (see Corollary 9.50).
Examples 9.2. (a) If M = {εx : x ∈ K}, then any universally measurable subset is
M-extremal.
(b) If M = M(H), then M-extremal sets are exactly H-extremal sets (see Defini-
tion 3.12).
(c) If M = M(H) ∩ Mmax (H), we get the family of H-maximally extremal sets.
Proposition 9.3. If M ⊂ M(H), then the family of all closed M-extremal sets is
stable with respect to arbitrary intersections and finite unions.
Proof. The proof follows by an easy verification from the definitions.
Examples 9.6. (a) If M = {εx : x ∈ K}, then the topology τM is the original
topology on ChH (K).
(b) If M = M(H), we denote the resulting topology as σext .
(c) If M = M(H) ∩ Mmax (H), we write σmax for the resulting topology.
Proposition 9.7. Both σext and σmax are compact topologies on ChH (K) such that
one-point subsets of ChH (K) are closed.
Proof. Let M stand for M(H) or M(H) ∩ Mmax (H) and let σ stand for σext or
σmax . Since any point of ChH (K) is M-extremal, it is a σ-closed set. To check
the compactness of (ChH (K), σ), let {Hi : i ∈ I} be a family of σ-closed sets
in ChH (K) having the finite intersection property. Let Fi , i ∈ I, be closed M-
extremal sets with Hi = Fi ∩ ChHT(K), i ∈ I. Then {Fi } has the finite intersection
property as well, and hence F := i∈I Fi is a nonempty closed M-extremal set (see
Proposition 9.3). Thus
\ \
Hi = Fi ∩ ChH (K) = F ∩ ChH (K)
i∈I i∈I
f ∗ (x) = ν1 (f ) ≤ ν2 (f ) ≤ f ∗ (x).
To verify (b), let x ∈ ChG (L). Then εx is G-maximal, and thus H-maximal by
(a). It follows that x ∈ ChH (K). Since the reverse inclusion follows from Proposi-
tion 6.72(c), assertion (b) is proved.
Since (c) follows immediately from (a), the proof is finished.
Proposition 9.9. A set H ⊂ ChH (K) is σmax -closed if and only if H is relatively
closed in ChH (K) and H is H-maximally extremal.
Proof. Let τ stand for the original topology on ChH (K). Obviously we have
We prove (i) =⇒ (iv) =⇒ (ii) =⇒ (iii) =⇒ (i) and (iv) =⇒ (v) =⇒ (iii).
If ChH (K) is closed, any τ -closed subset of ChH (K) is H-extremal and hence
σext -closed. Hence τ = σext , and thus (i) =⇒ (iv). Obviously, (iv) =⇒ (ii) and (ii)
=⇒ (iii) because of (9.2).
To verify (iii) =⇒ (i), assume that there exists x ∈ ChH (K) \ ChH (K). Let F
be the smallest closed H-maximally extremal set containing x (such a set exists by
virtue of Proposition 9.3). Then H := F ∩ ChH (K) is σmax -closed and nonempty by
Proposition 9.4(b). Moreover, H contains two different points, say x1 , x2 . We claim
that they cannot be separated by disjoint σext -open sets.
Indeed, assume that U1 , U2 are such sets. Let Fi , i = 1, 2, be closed H-maximally
extremal sets such that ChH (K) \ Ui = Fi , i = 1, 2. Then ChH (K) ⊂ F1 ∪ F2 .
We suppose that x ∈ F1 . By the minimality of F , F ⊂ F1 and thus x1 ∈ / F.
278 9 Topologies on boundaries
Example 9.11. There exists a compact convex set X such that σext is strictly weaker
than σmax .
Proof. Let x0 , x2 , x2 be distinct points in [0, 1]. We identify (x0 , 0) and (x2 , 0) in
[0, 1]×[0, 21 ] and let K := p([0, 1]×[0, 12 ]), where p is the quotient mapping. We write
ω0 for the point p(x0 , 0) and identify K \ {ω0 } with [0, 1] × [0, 21 ] \ {(x0 , 0), (x2 , 0)}.
Let
H := {f ∈ C(K) : f (ω0 ) = tf (x1 , t) + (1 − t)f (x2 , t), t ∈ (0, 12 ],
Z
f (x1 , 0) + f (ω0 ) = 4 f (x, t) dx dt}.
K
Then H is a closed function space on K whose Choquet boundary equals K \ {ω0 }.
Let X stand for the state space of H and let φ : K → X be the homeomorphic
embedding from Definition 4.25. Let
C contains both φ(x1 , t) and φ(x2 , t) for each t ∈ (0, 12 ]. Since D is a face, we get
1
(φ(x1 , t) + φ(x2 , t)) ∈ D, t ∈ (0, 21 ],
2
as well. Thus 12 (φ(x1 , 0) + φ(ω0 )) ∈ D because D is closed. If µ = p] (2λ), where
λ is Lebesgue measure on [0, 1] × [0, 21 ], then 21 (φ(x1 , 0) + φ(ω0 )) is the barycenter
of φ] µ. Thus D, and consequently C, contains the support of φ] µ. Hence C ⊃ φ(K)
and F is not σext -closed.
Baire subsets of K and F ⊂ C(K) be a norm separable set. Then there exist a
metrizable compact space K 0 , a function space H0 on K 0 , a family F 0 ⊂ C(K 0 ), and
a continuous surjection ϕ : K → K 0 such that
(a) h ◦ ϕ ∈ H for every h ∈ H0 ,
(b) ϕ] µn is H0 -maximal, n ∈ N,
(c) ϕ(An ) is a Baire subset of K 0 and ϕ−1 (ϕ(An )) = An , n ∈ N,
(d) for every f ∈ F there exists f 0 ∈ F 0 so that f = f 0 ◦ ϕ.
Moreover, if H is simplicial, H0 can be chosen to be simplicial as well and then
Ac (H0 ) = H0 .
Proof. Given the objects as in the hypotheses, we find a dense countable subset D of
F. Next we select a countable family F 1 ⊂ C(K) such that each An is contained in
the σ-algebra generated by {f −1 (U ) : U ⊂ R open, f ∈ F 1 } (see Proposition A.48).
Without loss of generality we may assume that 1 ∈ F 1 . For each function f ∈ D ∪F 1
we find a countable family Hf ⊂ H such that f ∈ W(Hf ) − W(Hf ). Let
[
F 2 := Hf .
f ∈D∪F 1
and ϕ : K → RN as
x 7→ {f (x)}f ∈F ∞ .
Then K 0 := ϕ(K) is a metrizable compact space and for any function f ∈ F ∞ there
exists a function gf ∈ C(K 0 ) such that f = gf ◦ ϕ. Let
H0 := {gf : f ∈ F ∞ }.
It follows from (f), that F ∞ is a linear space, hence H0 is a linear space as well.
Obviously, it contains constant functions and separates points of K 0 , so it is a function
space. Since F ∞ ⊂ H, h ◦ ϕ ∈ H for every h ∈ H0 , and (a) therefore holds.
280 9 Topologies on boundaries
Hence ϕ] µn is H0 -maximal.
To verify (c) and (d), we notice that, by the construction, any function from D ∪ F 1
is contained in the closure of the family {g ◦ ϕ : g ∈ W(H0 ) − W(H0 )}. Hence any
function from D ∪ F 1 can be expressed as g ◦ ϕ for a suitable function g ∈ C(K 0 ).
From this observation the existence of the required family F 0 ⊂ C(K 0 ) as well as (c)
and (d) follow.
Moreover, if H is simplicial, we adjust the construction of the families F k , k ≥ 3,
in the following way. We moreover assure in the inductive step that the family F k+1
possesses the following property:
(h) for each f1 , −f2 ∈ −W(F k ) with f1 ≤ f2 there exists f ∈ F k+1 with f1 ≤ f ≤
f2 .
Then the resulting family F ∞ , and consequently H0 , possesses the weak Riesz in-
terpolation property, which yields that H0 is simplicial and that Ac (H0 ) = H0 (see
Theorem 6.16 and Exercise 6.78). This concludes the proof.
Proof. Given a maximal measure µ ∈ M+ (K), a Baire set B ⊂ K and ε > 0, let
(K 0 , H0 ) and ϕ : K → K 0 be constructed as in Theorem 9.12. Then ϕ] µ, as an
H0 -maximal measure on a metrizable compact space K 0 , satisfies
Using the regularity of ϕ] µ, we find a compact set F 0 ⊂ ϕ(B) ∩ ChH0 (K 0 ) such that
For ε > 0 we find closed H-extremal sets Fn ⊂ Gn of type Gδ such that µ(Gn \Fn ) ≤
ε2−n . Then F := ∞
T
n=1 n is a closed H-extremal set of type Gδ with F ⊂ G, and
F
∞
[
µ(G \ F ) ≤ µ( (Gn \ Fn )) ≤ ε.
n=1
Notation 9.15. Let Σ stand for the σ-algebra generated by the family consisting of all
Baire sets and closed H-extremal sets, and let Σ0 := {A ∩ ChH (K) : A ∈ Σ}.
Let
T := {A ∈ Σ : µ1 (A) = µ(A), µ1 (K \ A) = µ(K \ A)}.
It follows from Proposition 9.14(a) that T contains all Baire sets. Further, any closed
H-extremal set is in T , by Proposition 9.14(b).
We show that T is an σ-algebra. Clearly, T is closed with respect to complements.
Let {An } be a sequence of sets in T and let A := ∞
S
n=1 An . For ε > 0 we find k ∈ N
such that µ(A \ kn=1 An ) < ε. For each n ∈ N we select closed H-extremal sets
S
Fn , Fn0 ⊂ K such that
Fn ⊂ An , Fn0 ⊂ K \ An and µ(An \ Fn ) + µ(K \ (An ∪ Fn0 )) < ε2−n .
Then
k
[ ∞
\
F := Fn and F 0 := Fn0
n=1 n=1
are closed H-extremal sets. Further, F ⊂ A, F0 ⊂ K \ A and
µ(A \ F ) + µ(K \ (A ∪ F 0 )) < 2ε + ε.
Hence A ∈ T , and T is a σ-algebra.
It follows that Σ = T , which proves the first equality in (9.3). To prove the second
equality, we have
µ2 (A) ≥ µ1 (A) = µ(A), A ∈ Σ. (9.4)
Let A ∈ Σ and F, F 0 be closed H-extremal set with
F ∩ ChH (K) ⊂ A and F 0 ∩ ChH (K) ⊂ K \ A.
We claim that F ∩ F 0 = ∅.
Indeed, assume that x ∈ F ∩ F 0 . Consider the function space G := {h|F ∩F 0 : h ∈
H}. Let λ be a G-maximal measure in Mx (G). By Propositions 3.64 and 9.8(b), λ is
carried by
ChG (F ∩ F 0 ) = ChH (K) ∩ F ∩ F 0 = ∅,
which is impossible. Hence F ∩ F 0 = ∅.
Given A ∈ Σ, let F be a closed H-extremal set with F ∩ ChH (K) ⊂ A. We fix
ε > 0 and find a closed H-extremal set F 0 ⊂ K \ A such that
µ(K \ A) − µ(F 0 ) < ε.
By the argument above, F ∩ F 0 = ∅. Thus
µ(F ) ≤ µ(K \ F 0 ) = µ(K) − µ(F 0 )
≤ µ(K) − µ(K \ A) + ε = µ(A) + ε.
Hence µ2 (A) ≤ µ(A) for each A ∈ Σ, which, in conjunction with (9.4), completes
the proof.
9.2 Induced measures on Choquet boundaries 283
Corollary 9.17. If µ is maximal and A ∈ Σ disjoint from ChH (K), then µ(A) = 0.
Then µ is H-maximal.
Proof. We verify condition (b) of Theorem 8.32; more precisely, we show that any
function
S h ∈ H is constant on Fx (H) for µ-almost all x ∈ K (we recall that Fx (H) =
ν∈Mx (H) spt ν). Let h ∈ H be arbitrary and let [a, b] be a closed interval in R.
BySthe assumption, there exist closed H-extremal sets Fn ⊂ h−1 ([a, b]) such that
∞ ∞
µ( n=1 Fn ) = µ(h−1 ([a, b])). For a point x ∈ n=1 Fn and ν ∈ Mx (H),
S
∞
[
spt ν ⊂ Fn ⊂ h−1 ([a, b]).
n=1
Hence, Fx (H) ⊂ h−1 ([a, b]) for µ-almost all x ∈ h−1 ([a, b]).
Thus there exists a set K 0 ⊂ K satisfying
• µ(K \ K 0 ) = 0,
• if x ∈ K 0 and [a, b] is a closed interval with rational endpoints such that x ∈
h−1 ([a, b]), then Fx (H) ⊂ h−1 ([a, b]).
It follows that, for any x ∈ K 0 , the function h is constant on Fx (H). Thus µ is
maximal.
Moreover,
and
by Theorem 9.19.
Hence we get from Fubini’s theorem that
Z Z 1
f (x) = f (y) dµ(y) = µ(Ft ) dt
K 0
Z 1 Z
0
= µ (Ft ∩ ChH (K))) dt = f (y) dµ0 (y).
0 ChH (K)
Then σext = τext |ChH (K) and σmax = τmax |ChH (K) .
9.3 Functions continuous in σext and σmax topologies 285
Lemma 9.22. Let f : ChH (K) → [0, 1] be a σext -upper semicontinuous function.
Then there exist a closed H-extremal set K 0 ⊃ ChH (K) and a τext -upper semicontin-
uous function g : K 0 → [0, 1] that extends f .
Proof. Let f : ChH (K) → [0, 1] be σext -upper semicontinuous. We fix n ∈ N and
find closed H-extremal sets Fin , i = 0, . . . , 2n − 1, such that
Lemma 9.23. Let K 0 be a closed H-extremal set containing ChH (K) and g : K 0 → R
a τext -upper semicontinuous function such that g = 0 on ChH (K). Then g ≤ 0 on
K 0 and there exists a closed H-extremal set K 00 ⊂ K 0 containing ChH (K) such that
g = 0 on K 00 .
Proof. Assume that sup g(K 0 ) > 0. Since g is upper semicontinuous, g attains its
maximum on K 0 . Then there exists a closed H-extremal set F ⊂ K such that
g −1 ([−n−1 , ∞)) = Fn ∩ K 0 , n ∈ N.
Then F := ∞ 00
T
n=1 Fn is a closed H-extremal set containing ChH (K), and thus K :=
0
K ∩ F is the required set.
Theorem 9.25. Let f : ChH (K) → R. Then the following assertions are equivalent:
(i) f is σext -continuous.
(ii) there exist a closed H-extremal set K 0 ⊃ ChH (K) and a τext -continuous func-
tion g : K 0 → R extending f .
Proof. Let f : ChH (K) → R be σext -continuous. By the compactness of ChH (K)
in σext , f is bounded. Using Lemma 9.22, we find closed H-extremal sets K1 , K2
containing ChH (K), a τext -lower semicontinuous function g1 : K1 → R and a τext -
upper semicontinuous function g2 : K2 → R that extend f .
Then K1 ∩ K2 is a closed H-extremal set containing ChH (K). By Lemma 9.23,
there exists a closed H-extremal set K 0 ⊂ K1 ∩ K2 containing ChH (K) such that
g1 = g2 on K 0 . Then g := g1 |K 0 is the sought extension. This proves (i) =⇒ (ii).
Conversely, if g : K 0 → R is a τext -continuous function on a closed H-extremal set
0
K containing ChH (K), then f := g|ChH (K) is σext -continuous. Indeed,
Lemma 9.26. Let f : ChH (K) → [0, 1] be a σmax -upper continuous function. Then
there exists a τmax -upper semicontinuous function g : ChH (K) → [0, 1] extending f .
9.3 Functions continuous in σext and σmax topologies 287
Proof. The proof is similar to that of Lemma 9.22. Let f : ChH (K) → [0, 1] be
σmax -upper semicontinuous function. We fix n ∈ N and define
By Proposition 9.9, each Fin is an H-maximally extremal set. We define gn : ChH (K)
→ [0, 1] as
gn := sup (i + 1)2−n cFin .
i=0,...,2n −1
Lemma 9.27. Let g : ChH (K) → [0, ∞) be a τmax -upper semicontinuous function
such that g = 0 on ChH (K). Then g = 0 on ChH (K).
Proof. Let g : ChH (K) → [0, ∞) be τmax -upper semicontinuous and assume that
g(x) > δ > 0 at some point x ∈ ChH (K). Then
is a nonempty closed H-maximally extremal set that does not intersect ChH (K), a
contradiction with Proposition 9.4(b). Since g ≥ 0 by upper semicontinuity, the proof
is finished.
Proof. Let g : ChH (K) → R be a τmax -continuous function and H be a closed set
in R. Then g −1 (H) is a closed H-maximally extremal set. If x ∈ ChH (K) and
µ ∈ Mx (H) ∩ Mmax (H), then µ is carried by
Theorem 9.29. Let f : ChH (K) → R. Then the following assertions are equivalent:
(i) f is σmax -continuous,
(ii) there exists a continuous function g : ChH (K) → R extending f such that
for every x ∈ ChH (K) and µ ∈ Mx (H) ∩ Mmax (H), g = g(x) µ-almost
everywhere.
R
Moreover, if g is the extension of f , then h(g(x)) = K h(g(y)) dµ(y) for each x ∈
ChH (K), µ ∈ Mx (H) ∩ Mmax (H) and h : R → R continuous.
Proof. To show (i) =⇒ (ii), let f : ChH (K) → R be a σmax -continuous function.
Since ChH (K) is σmax -compact, f is bounded. By Lemma 9.26, there exist bounded
functions g1 , g2 : ChH (K) → R such that
• g1 = f = g2 on ChH (K),
• g1 , −g2 are τmax -lower semicontinuous.
Since τmax is weaker than the original topology, g1 , −g2 are lower semicontinuous.
Hence g1 ≤ g2 on ChH (K). By Lemma 9.27, g := g1 = g2 on ChH (K) is a τmax -con-
tinuous extension of f . By Proposition 9.28, g possesses the property required in (ii).
Conversely, let g : ChH (K) → R be as in (ii). We show that F := g −1 ([a, b])
is H-maximally extremal for each closed interval [a, b] ⊂ R. For x ∈ F and µ ∈
Mx (H) ∩ Mmax (H), µ is carried by the set {y ∈ ChH (K) : g(y) = g(x)}. Hence µ
is carried by F and F is H-maximally extremal. Thus g|ChH (K) is σmax -continuous.
The last part of theorem follows from the fact that a function g satisfying condition
(ii) is τmax -continuous, and hence Proposition 9.28 applies.
Proposition 9.31. Any H-strongly universally measurable function is in H⊥⊥ and any
bounded semicontinuous function from H⊥⊥ is H-strongly universally measurable.
Proof. If f is H-strongly universally measurable, then f is obviously universally mea-
surable and bounded. Given µ, ν ∈ M1 (K) with µ − ν ∈ H⊥ and ε > 0, let g, h be
as in Definition 9.30. Then the inequalities
as required.
From the considerations above it also follows that µ(f ) = f (r(µ)), and f is
strongly affine.
Lemma 9.35. Let h1 , h2 be bounded affine functions on a compact convex set X with
h1 ≤ h2 . Let Γ := {(x, t) ∈ X × R : h1 (x) ≤ t ≤ h2 (x)} and let π : X × R → X
be the projection.
(a) We have
(b) If h1 , −h2 are lower semicontinuous, then for each x ∈ X, µ ∈ Mx (X) and
t ∈ [h1 (x), h2 (x)] there exists ν ∈ M1 (Γ) such that π] ν = µ and ν ∈ M(x,t) (Γ).
(c) If h1 , −h2 are lower semicontinuous, then for each x ∈ X, a maximal measure
µ ∈ Mx (X) and t ∈ [h1 (x), h2 (x)] there exists a maximal measure ν ∈ M1 (Γ)
such that π] ν = µ and ν ∈ M(x,t) (Γ).
Then Γ ⊂ Γf1 ,f2 and thus we may use the first part to deduce that
µ ∈ π] (M(x,t) (Γf1 ,f2 )).
Since \
Γ= {Γf1 ,f2 : f1 ∈ F 1 , f2 ∈ F 2 }
and the latter family is down-directed,
\
M(x,t) (Γ) = {M(x,t) (Γf1 ,f2 ) : f1 ∈ F 1 , f2 ∈ F 2 }
and the latter family is down-directed as well.
By compactness and the continuity of π] , Proposition A.40 yields
\
π] (M(x,t) (Γ)) = {π] (M(x,t) (Γf1 ,f2 )) : f1 ∈ F 1 , f2 ∈ F 2 },
and thus
µ ∈ π] (M(x,t) (Γ)).
To verify (c), let µ ∈ Mx (X) be a maximal measure. Using (b) we find λ ∈
M(x,t) (Γ) with π] λ = µ. Let ν ∈ M1 (Γ) be a maximal measure with λ ≺ ν. Then
ν ∈ M(x,t) (Γ) and µ ≺ π] ν.
Indeed, if k is a continuous convex function on X, k ◦ π is a continuous convex
function on Γ. Hence
µ(k) = π] λ(k) = λ(k ◦ π) ≤ ν(k ◦ π) = π] ν(k).
Since µ is maximal, π] ν = µ. This finishes the proof.
Proof. Since f is bounded (see Proposition 4.7 and Lemma 4.5), we may assume that
f is a lower semicontinuous function on X with 0 ≤ f < 1. Let
Γ := {(x, t) ∈ X × R : f (x) ≤ t ≤ 1}.
Using Lemma 9.35(c), we find a maximal measure ν ∈ M1 (Γ) such that
π] ν = µ and r(ν) = (x, f (x)).
We claim that the set gr f is a Gδ face of Γ and gr 1 = {(x, 1) : x ∈ X} is a closed
face of Γ. To verify this, we first notice that both sets are obviously faces of Γ and
gr 1 is closed. Since
∞ n
\ 1o
gr f = (x, t) ∈ Γ : f (x) ≤ t < f (x) +
n
n=1
292 9 Topologies on boundaries
and the sets on the right-hand side are open in Γ by the lower semicontinuity of f ,
gr f is of type Gδ in Γ.
Further, ν(gr f ) = 1. To see this, we consider a positive continuous affine function
F on Γ defined as
F (x, t) := t − f (x), (x, t) ∈ Γ.
Then Z
0 = F (x, f (x)) = F (r(ν)) = F (y, s) dν(y, s).
Γ
Hence F = 0 ν-almost everywhere, and thus
ν(gr f ) = 1. (9.6)
Our aim is to prove that both E 0 and F 0 are µ00 -measurable. Let (µ0 )∗ and (µ0 )∗ denote
the outer and inner measure on ext X induced by µ0 , respectively.
We first notice that Proposition 9.14(b) and Theorem 9.19 applied to the open set
F give
Proof of the claim. Obviously we may assume that 0 ≤ a < 1. Since gr f is a Gδ set
in Γ, Proposition 9.14 yields the existence of a closed extremal set D1 ⊂ Γ such that
Let
G := {(x, t) ∈ Γ : t ≤ a}, G0 := G ∩ ext Γ.
Then
G0 = G ∩ ext gr f
and
E = π(G), E 0 = π(G0 ).
9.4 Strongly universally measurable functions 293
Since G is a closed Gδ set, Theorem 9.19 provides a closed extremal set D2 ⊂ Γ such
that
Then
D := D1 ∩ D2
is a closed extremal set in Γ contained in gr f such that
This implies
µ(E) = ν(π −1 (E)) = ν(π −1 (π(G))) = ν 0 (G0 ). (9.10)
From (9.9) and (9.10) we get
Hence
(µ0 )∗ (E 0 ) ≥ µ(E)
as required.
It follows from the claim that the set F 0 is µ00 -measurable. Indeed, the claim, along
with (9.7), yields
Thus f is µ00 -measurable and to finish the proof we need to verify the barycentric
formula. To this end, let
Fa := {y ∈ X : f (y) > a}, Fa0 := Fa ∩ ext X, a ∈ R.
By the first part of the proof,
µ00 (Fa0 ) = µ(Fa ), a ∈ R.
Since µ(f ) = f (x) by Proposition 4.7, Fubini’s theorem gives
Z Z 1
f (x) = f (y) dµ(y) = µ(Fa ) da
X 0
Z 1 Z
= µ00 (Fa0 ) da = f (y) dµ00 (y).
0 ext X
This concludes the proof.
Proof. Given the objects as in the statement of the lemma, let Σ(K) and Σ0 (ChH (K))
stand for the σ-algebras from Notation 9.15, where K indicates their relevance to the
space K. Analogously, we write Σ(X) and Σ0 (ext X) for the respective σ-algebras on
X. Now (ChH (K), Σ00 (ChH (K)), µ00 ) denotes the completion of the measure space
(ChH (K), Σ0 (ChH (K), µ0 )), and (ext X, Σ00 (ext X), (φ] µ)00 ) denotes the completion
of (ext X, Σ0 (ext X), (φ] µ)0 ).
First we notice that
(φ0 )−1 (A ∩ ext X) = φ−1 (A) ∩ ChH (K), A ⊂ X. (9.11)
Clearly, φ−1 (A) is a Baire set in K for any Baire set A ⊂ X and φ−1 (A) is a closed
H-extremal set for any closed extremal set A ⊂ X (see Lemma 8.10). By (9.11),
(φ0 )−1 (A0 ) ∈ Σ0 (ChH (K)), A ∈ Σ0 (ext X).
From this it follows that
φ0 : (ChH (K), Σ0 (ChH (K))) → (ext X, Σ0 (ext X))
is measurable.
9.4 Strongly universally measurable functions 295
Further, for any A ∈ Σ(X), we get from Theorem 9.19 and (9.11) that
Hence
((φ0 )] µ00 )(A) = (φ] µ)00 (A), A ∈ Σ00 (ext X),
which is nothing else than the desired result
We define
an := h1 ∨ · · · ∨ hn , bn := g1 ∧ · · · ∧ gn , n ∈ N.
296 9 Topologies on boundaries
Then
sup an ≤ f ≤ inf bn ,
n∈N n∈N
= 0.
hn (r(φ] µ)) = (φ] µ)00 (hn |ext X ) ≤ (φ] µ)00 (If |ext X )
≤ (φ] µ)00 (gn |ext X ) = gn (r(φ] µ)), n ∈ N,
gives
(φ] µ)00 (If |ext X ) = If (r(φ] µ)) = µ(f ).
Let φ0 : ChH (K) → ext X be the restriction of φ. By the first part of the proof, If
is (φ] µ)00 -measurable. Lemma 9.37 yields
Proposition 9.39. The family of all M -sets is closed with respect to taking arbitrary
intersections and H-convex hulls of finite unions.
by Theorem A.84. Hence it is enough to show that F possesses condition (M) when
F is finite. By a simple induction, it follows that we need to verify that F1 ∩ F2
satisfies condition (M) whenever both F1 and F2 do. But this is obvious, because
µ|F1 ∈ H⊥ ∩ Mbnd (H) and hence
Let F1 , F2 be M -sets and let F := coH (F1 ∪ F2 ). We first verify that F satisfies
condition (M). To this end, let µ ∈ H⊥ ∩ Mbnd (H) and h ∈ H be given. Then
µ|F1 ∪F2 (h) = µ|F1 (h) + µ|F2 (h) − µ|F1 ∩F2 (h) = 0,
Definition 9.40 (Facial topology). It follows from Propositions 9.39 and 8.30(b) that
the family
{F ∩ ChH (K) : F is an M -set}
defines the family of closed sets for a topology on ChH (K). We call it the facial
topology and denote it σfac .
Proof. Let {Fi }i∈I be a family of M -sets such that {Fi ∩ ChH (K)}i∈I has the finite
intersectionTproperty. Then {Fi }i∈I has the finite intersection property as well, and
thus F := i∈I Fi is a nonempty M -set (see Proposition 9.39). Since F ∩ChH (K) 6=
∅ (see Proposition 9.4(b)), we get
\
(Fi ∩ ChH (K)) = F ∩ ChH (K) 6= ∅.
i∈I
Theorem 9.42. Let f : ChH (K) → R be a bounded σfac -upper semicontinuous func-
tion and a ∈ H be a positive function. Then there exists a unique upper semicontinu-
ous function h ∈ H⊥⊥ such that
Proof. Assume first that f is as in the hypothesis and f (ChH (K)) ⊂ [0, 1]. We fix
n ∈ N and find for i = 0, . . . , n an M -set Fi ⊂ K such that
i
{x ∈ ChH (K) : f (x) ≥ } = Fi ∩ ChH (K).
n
Theorem 9.43. Let f : ChH (K) → R be a σfac -continuous function and a ∈ H. Then
there exists a unique h ∈ H such that
Proof. Given f and a as in the theorem, we first notice that f is bounded by Proposi-
tion 9.41. Let m > 0 satisfy a(K) ⊂ [−m, m]. By Theorem 9.42, there exist upper
semicontinuous functions h1 , h2 ∈ H⊥⊥ such that
Thus h is the sought function. Since the uniqueness follows trivially from the mini-
mum principle, the proof is finished.
Corollary 9.44. Let f : ChH (K) → R be a σfac -continuous function. Then there
exists a unique h ∈ H such that h = f on ChH (K).
Proof. We start by proving (ii) =⇒ (iii). First we show that ChH (K) is a closed
set. Assume that there exists x ∈ ChH (K) \ ChH (K). Let F be the smallest M -set
containing x (see Proposition 9.39). Since x ∈ / ChH (K), F intersects ChH (K) in at
least two distinct points, say x1 and x2 (see Proposition 3.15 and Proposition 8.30(b)).
For i = 1, 2, let Ui be disjoint σfac -open sets containing xi , respectively. By the
definition of σfac , there exist M -sets Fi such that
Since U1 ∩ U2 = ∅, then
x ∈ ChH (K) ⊂ F1 ∪ F2 = F1 ∪ F2 .
Definition 9.46 (Center of H). We say that a function f ∈ H belongs to the center of
H, if for every a ∈ H there exists h ∈ H such that
Proposition 9.47. If H is a closed function space, then the center Z(H) is a com-
mutative Banach algebra and a vector lattice, where the multiplication and lattice
operations are considered pointwise on ChH (K).
Proof. First we notice that Z(H) is a linear subspace of H containing constant func-
tions.
Step 1. Next we show that Z(H) is closed in H. Let {fn } be a sequence of
functions from Z(H) uniformly convergent to f ∈ H. For any a ∈ H and n ∈ N,
9.5 Facial topology generated by M -sets 301
h = f1 f2 on ChH (K).
We claim that h ∈ Z(H) as well. Indeed, for a given a ∈ H we can find h1 ∈ H such
that h1 = f2 a on ChH (K). Further we find h2 ∈ H with h2 = f1 h1 on ChH (K).
Then
Hence h ∈ Z(H).
Step 3. If f ∈ Z(H), then f = (kf k + f ) − kf k and thus Z(H) = (Z(H))+ −
(Z(H))+ .
We show that for any f ∈ Z(H) there exists h ∈ Z(H) such that h = f ∨ 0 on
ChH (K). To this end, let {pn } be a sequence of polynomials on R that converges to
the function t 7→ t∨0, t ∈ R, uniformly on f (K). By the second step of the proof, we
find functions hn ∈ Z(H), n ∈ N, satisfying hn = pn ◦ f on ChH (K). Again by the
minimum principle, we deduce that {hn } converges uniformly on K to an element
h ∈ Z(H). It easily follows that h = f ∨ 0 on ChH (K). Proposition A.15 yields that
Z(H) is a vector lattice, which completes the proof.
Theorem 9.48. The mapping R : Z(H) → C(ChH (K), σfac ) defined as the restric-
tion to ChH (K) is an isometric isomorphism preserving multiplication and lattice
operations.
Proof. We first show that h|ChH (K) is σfac -continuous for each h ∈ Z(H). Since
Z(H) is a linear space, to this end it is enough to show that the set
is σfac -closed for any h ∈ Z(H) with h(K) ⊂ [0, 1] and α ∈ [0, 1].
For any n ∈ N we use the algebraic structure of Z(H) to find a function hn ∈
Z(H) such that
are closed Choquet sets with ChH (K) ⊂ F1 ∪ F2 . Since H is simplicial, they are
M -sets with respect to Ac (H) (see Theorem 8.62). Then
Ui := ChH (K) \ Fi , i = 1, 2,
are σfac -open disjoint sets in ChH (K). Since neither F1 nor F2 covers ChH (K) (oth-
erwise h = h1 or h = h2 ), the sets U1 , U2 are nonempty. This concludes the proof.
Corollary 9.50. Let H be a prime simplicial space. Then the center of Ac (H) consists
of constant functions.
Proof. The assertion follows from Proposition 9.49 and Theorem 9.48.
9.6 Exercises
Exercise 9.51. Let X be a compact convex set, X 0 a minimal closed extremal set
containing ext X and g : X 0 → R τext -continuous. Prove that g is constant on any
segment and on any face contained in X 0 .
Hint. Let x, y ∈ X 0 be distinct points such that the segment [x, y] ⊂ X 0 and g(x) <
g(y). Then
1
F1 := {z ∈ X 0 : g(z) ≤ (g(x) + g(y))}
2
and
1
F2 := {z ∈ X 0 : g(z) ≥ (g(x) + g(y))}
2
are closed extremal subsets of X with X 0 ⊂ F1 ∪ F2 . On the other hand, if z =
1
2 (x + y), then it follows from the extremality of F1 and F2 that z ∈/ F1 ∪ F2 . This
contradiction shows that g is constant on the segment joining points x and y.
If F ⊂ X 0 is a face, then g is constant on any segment contained in F . Hence g is
constant on F .
Hint. Consider H := C(K) on a compact space K (see Example 3.2(a)). Then any
continuous function on K is τext -continuous and any closed set F ⊂ K is a closed
Choquet set.
Exercise 9.53. Let Σ0 be the σ-algebra on ChH (K) generated by the family
Exercise 9.54. Let Σ0 and µ0 be as in the previous Exercise 9.53. Let f : ChH (K) →
R be σmax -continuous and g : ChH (K) → R be the extension provided by Theo-
rem 9.29. Let h : R → R be continuous, x ∈ ChH (K) and µ ∈ Mx (H) ∩ Mmax (H).
Prove that Z
h(g(x)) = h(f (y)) dµ0 (y).
ChH (K)
Hint. By Proposition 9.28, g = g(x) µ-almost everywhere and by the proof of Theo-
rem 9.29, g is τmax -continuous. Hence f is µ0 -measurable and
is a τmax -closed set which carries µ. Thus µ0 (F ∩ ChH (K)) = µ(F ) = 1 and
Z Z
h(f (y)) dµ0 (y) = h(f (y)) dµ0 (y)
ChH (K) F ∩ChH (K)
Z
= h(g(x)) dµ0 (y)
F ∩ChH (K)
= h(g(x)).
Exercise 9.55. Prove that any H-strongly affine function satisfies the minimum prin-
ciple.
Hint. Use either Theorem 9.38 or the definition along with the Minimum princi-
ple 3.16.
Exercise 9.56. Let φ0 : ChH (K) → ext S(H) be the mapping from Lemma 9.37.
Prove that
(a) φ0 : (ChH (K), σext ) → (ext S(H), σext ) is continuous,
(b) φ0 : (ChH (K), σmax ) → (ext S(H), σmax ) is a homeomorphism,
(c) φ0 : (ChH (K), σext ) → (ext S(H), σext ) need not be a homeomorphism.
9.6 Exercises 305
Hint. For the σext -continuity of φ0 use Lemma 8.10. To show that φ0 is a home-
omorphism with respect to σmax -topologies, it is enough to show that F ⊂ K is
H-maximally extremal if and only if φ(F ) is S(H)-maximally extremal. But this
easily follows from the facts that µ ∈ M+ (K) is H-maximal if and only if φ] µ is
Ac (S(H))-maximal (see Proposition 4.28(d)) and that Ac (S(H))-maximal measures
on S(H) are carried by φ(K) (see Proposition 3.64).
To find an example required by (c), consider the following compact space K ⊂ R2 .
Let
In := [−2, −1] × {n−1 }, Jn := [1, 2] × {n−1 }, n ∈ N,
I0 := [−2, −1] × {0}, J0 := [1, 2] × {0},
I−1 := [−2, −1] × {−1},
and let µn , n ∈ N ∪ {0, −1} and νn , n ∈ N ∪ {0}, be copies of one-dimensional
Lebesgue measure on the sets In and Jn , respectively. Let z := (1, 1) ∈ R2 . We
define [ [
K := In ∪ Jn
n∈N∪{0,−1} n∈N∪{0}
and n 1 1
H := f ∈ C(K) : f (z) = µn (f ) + (1 − )νn (f ), n ∈ N,
n n
o
µ0 (f ) = µ−1 (f ) .
Then H is a function space on K and it can be easily seen that ChH (K) = K \ {z}.
We claim that K \ I−1 is H-extremal.
To this end, we consider functions hn defined as
on I−1 ∪ I0 ∪ ∞
S
1
k=n Ik ,
−1
S∞
hn := −(k − 1) on k=n Jk , n ≥ 2.
0 otherwise,
φ0 (F ) ⊂ H ∩ ext S(H).
306 9 Topologies on boundaries
Then φ(z) ∪ φ(I0 ) ⊂ H. Let π : M1 (K) → S(H) be the restriction mapping and
denote
sn := π(µn ), n ∈ N ∪ {0, −1}, tn := π(νn ), n ∈ N.
Then
1 1
sn + (1 − )tn , n ∈ N,
φ(z) =
n n
and thus sn , tn ∈ H, n ∈ N. Since H is closed, s0 ∈ H as well. We get that
the measure φ] µ−1 ∈ Ms0 (S(H)), and hence its support is contained in H (see
Proposition 2.86). Thus
φ(I−1 ) ⊂ H ∩ ext(S(H)).
Hence φ0 (F ) is strictly smaller than ext(S(H)) ∩ H for any closed extremal set H in
S(H) satisfying φ0 (F ) ⊂ ext(S(H)) ∩ H.
Fn ⊂ f −1 (Un ) and Hn ⊂ f −1 (R \ Un ), n ∈ N,
such that
Then
∞
\
F := (Fn ∪ Hn )
n=1
is a σext -closed set such that µ00 (ChH (K) \ F ) < ε and f |F is σext -continuous.
Exercise 9.58. Find a compact convex set X, a maximal measure µ ∈ M1 (X) and a
strongly affine function f on X such that f |ext X is not µ00 -measurable.
Hint. Let H be the function space from Example 3.82, X := S(H) be the state
space of H and let f := c[0,1]×{1} − c[0,1]×{−1} . Let λ be Lebesgue measure on
[0, 1] × {0} and λ00 be the induced measure on ChH (K) from Notation 9.34. Since H
is simplicial, H = Ac (H) (see Lemma 6.14 and its proof) and f is Borel, f ∈ H⊥⊥
(see Corollary 6.12). Thus If is a strongly affine function on X (see Corollary 5.41).
By Lemma 9.37, (φ] λ)00 = (φ0 )] λ00 . Further,
and hence (φ] λ)00 -measurability of If |ext X implies λ00 -measurability of f |ChH (K) .
Thus it suffices to show that f |ChH (K) is not λ00 -measurable. Since Σ00 is the com-
pletion of Σ0 with respect to the measure λ0 , we know from Theorem 9.19 that
Thus to disprove that f |ChH (K) is µ00 -measurable it is enough to show that λ00 (F ) = 0
for any σext -closed set F ⊂ [0, 1] × {1} (and analogously for [0, 1] × {−1}). If F is
such a set, then there exists a closed H-extremal set H ⊂ K such that
H ∩ ChH (K) = F.
Exercise 9.59. Prove that the class of strongly universally measurable functions is not
affinely perfect. More precisely, find compact convex sets X, Y , an affine continuous
surjection π : X → Y and a function f : Y → R such that f ◦π is strongly universally
measurable and f is not strongly universally measurable.
Hint. Consider the function space H from Example 3.82 and set Y := S(H), X :=
M1 (K). Let π : M1 (K) → S(H) be the restriction mapping and let f and λ be as
in Exercise 9.58. If I is the mapping from Corollary 5.41, the function If |ext Y is not
(φ] λ)00 -measurable. The function g : X → R defined as g(µ) = µ(f ), µ ∈ X, is
strongly universally measurable on X.
Indeed, given ε > 0 and ν ∈ X, let h1 , −h2 be upper semicontinuous bounded
functions on K such that h1 ≤ f ≤ h2 and ν(h2 ) − ν(h1 ) < ε. Then the functions
hi (µ) := µ(hi ), µ ∈ X, i = 1, 2, witness that g is strongly universally measurable.
b
Finally, the formula
If (π(µ)) = g(µ), µ ∈ X,
Hint. Assertions (a), (b) and (c) are easy from the definitions. To prove (d), let F be
a closed semiextremal set. We consider the family
F := {F ∩ H : H closed extremal, F ∩ H 6= ∅}
Take any point x ∈ F and a maximal measure µ ∈ Mx (X). Then µ(F ) > 0
(otherwise x ∈ X \ F by Proposition 2.76). Thus there exists a maximal measure ν
carried by F . If ν = εy for some y ∈ X, y is an extreme point contained in F and the
proof is finished.
Now consider the case when ν is not a Dirac measure. Then we can find a pair of
disjoint Baire sets A1 , A2 ⊂ K such that ν(Ai ) > 0, i = 1, 2. Using Proposition 9.14
we find closed extremal sets Fi ⊂ Ai such that ν(Fi ) > 0, i = 1, 2. In particular, the
sets Fi intersect F . It follows from Proposition 2.69 that there exists a pair of closed
faces Hi ⊂ X such that
∅ 6= Hi ∩ F ⊂ Fi ∩ F, i = 1, 2.
The goal of this chapter is to present several important results on function spaces and
compact convex sets. The first section is devoted to boundaries and to applications of
this notion in Banach spaces. The first result is Theorem 10.2, proving the existence of
the smallest closed boundary (so-called Shilov boundary) under a rather mild condi-
tion. Theorem 10.3 recalls the classical fact that the closure of the Choquet boundary
is the Shilov boundary for a function space.
Next we prove several results on James’ boundaries which is a concept particularly
useful in the theory of Banach spaces. We present an approach based upon the no-
tion of I-envelopes of sets in duals of Banach spaces. This leads us to Theorem 10.7
by V. Fonf and J. Lindenstrauss, which opens a way for the proof of Rode’s result
on norm separability of dual spaces (see Theorem 10.8). As a consequence of the
Simons lemma, we prove James’ theorem characterizing weakly compact sets in sep-
arable Banach spaces (see Theorem 10.9). Then we present Grothendieck’s result on
angelicity of spaces of continuous functions on compact spaces when they are en-
dowed with the pointwise topology. As its application we show Khurana’s proof of a
theorem by J. Bourgain and M. Talagrand on angelicity of bounded sets in a Banach
space with respect to the weak topology generated by extreme points of the dual unit
ball (see Theorem 10.12).
The next section contains the proof of Lazar’s generalization of the Banach–Stone
theorem (see Theorems 10.17 and 10.18).
Then we turn our attention to results on automatic boundedness and continuity of
affine functions. First we need several important auxiliary facts on the Cantor set, such
as the 0–1 laws of Theorem 10.24. The main tool needed later is Theorem 10.26 on
countable additivity of finitely additive measures on the Cantor set. Then we present
Christensen’s results on strongly linearly independent sequences in locally convex
spaces and functionals on L∞ (µ) (see Theorems 10.30 and 10.32). Results on auto-
matic boundedness of affine and convex functions on compact convex sets are proved
in Theorems 10.31 and 10.37.
Haydon’s characterization of Banach spaces not containing `1 by means of strongly
affine functions is contained in Theorem 10.42.
Metrizability of compact convex sets is investigated in Section 10.5. The first result
in Theorem 10.51 shows that a compact convex set X is metrizable provided it has
10.1 Boundaries 311
10.1 Boundaries
10.1.A Shilov boundary
Definition 10.1 (Shilov sets). Let K be a nonempty compact space and F be a non-
empty family of lower semicontinuous functions on K. A closed set S ⊂ K is called
a Shilov set for F if for every u ∈ F there exists x ∈ S such that
u(x) = inf u(K).
If there exists a Shilov set, which is contained in any other Shilov set, it is called
the Shilov boundary.
Obviously, K is a Shilov set for F. Notice that if F contains the constants and
u + c ∈ F whenever u ∈ F and c ∈ R, S is a Shilov set if and only if for every u ∈ F
with u ≥ 0 on S we have u ≥ 0 on K.
Consider the following simple example of the space K := {x1 , x2 , x3 } with the
discrete topology. We identify functions on K with vectors in R3 and denote
F := {(1, 0, 0), (0, 1, 0), (0, 0, 1)} .
Then every two-point subset of K is a Shilov set for F. Consequently, there exists no
smallest Shilov set for F.
312 10 Deeper results on function spaces and compact convex sets
Proof. A moment’s reflection shows that we may suppose that both F and S are min-
stable (otherwise we would consider a new family W(F) and a new convex cone
W(S) formed by finite minima of elements of F and S, respectively).
Claim. The system of Shilov sets is closed with respect to finite intersections.
Proof of the claim. Let S and T be Shilov sets. Suppose that u ∈ F and u ≥ 0 on
S ∩ T . Fix ε > 0 and define
U := {x ∈ K : u(x) + ε > 0} .
s − t ≥ 1 ≥ (1 − a)s on V,
B := {b ∈ [0, ∞) : v + bs ≥ 0 on K} .
10.1 Boundaries 313
Fix b ∈ B. Then
0 ≤ v + bs = v + bt on W
0 ≤ v + bt ≤ v + bas on V.
0 ≤ v + bas on U ∪ V ⊃ S.
for every n ∈ N. Since a ∈ (0, 1), we get v ≥ 0 on K. This establishes the claim.
Define now
Σ := {S ⊂ K : S is a Shilov set for F}
and denote
\
S0 := {S : S ∈ Σ}.
U := {x ∈ K : u(x) + ε > 0} .
Proof. By Theorem 10.2, there exists the Shilov boundary S for H. By the Minimum
principle 3.16, ChH (K) is a Shilov set for H. To show its minimality, assume that
there exists x ∈ ChH (K) \ S. Then x ∈ / coH S (see Proposition 8.18), and thus there
exists a function h ∈ H such that h(x) > max h(S) (see Proposition 8.23). But this
contradicts the fact that S is a Shilov set.
314 10 Deeper results on function spaces and compact convex sets
Proof. Let x∗ ∈
/ I-env(B), that is, there exist d > 0 and sets Bn ⊂ B, n ∈ N, such
that
∞
!
w∗
[
∗
B(x , d) ∩ co co Bn = ∅.
n=1
By replacing the sets Bn by Bn ∩ kB SE , n, k ∗∈ N, if necessary, we may assume that
∗
the sets Bn are bounded. Since co( nk=1 cow (Bk )) is a w∗ -compact convex set for
each n ∈ N, the set
n ∞
! !
w∗ w∗ w∗
[ [
co (B1 ∪ · · · ∪ Bn ) ⊂ co co Bk ⊂ co co Bn
k=1 n=1
Proof. For the proof of (ii) =⇒ (i), let x∗ and {xn } be as in (ii). We find c1 , c2 ∈ R
satisfying
inf x∗ (xn ) > c1 > c2 > sup lim sup b∗ (xn ).
n∈N b∗ ∈B n→∞
10.1 Boundaries 315
By setting
∞
\
Bn := {b∗ ∈ E ∗ : b∗ (xk ) ≤ c2 }, n ∈ N,
k=n
a contradiction.
Hence x∗ is not contained in the norm closure of the convex hull of ∞
S
n=1 Bn , and
∗
therefore x ∈ / I-env(B).
For the proof of the converse implication (i) =⇒ (ii), assume that x∗ ∈
/ I-env(B).
Without loss of generality we may assume that x∗ = 0. By the assumption and
Lemma 10.5, there exist d > 0 and an increasing sequence {Bn } such that
∞ ∞
!
∗
[ [
B= Bn and B(0, d) ∩ co cow Bn = ∅.
n=1 n=1
Without loss of generality we may assume that all points xn are of norm 1. Hence
Theorem 10.7. Let B be a subset of a w∗ -compact convex set X ⊂ E ∗ such that each
element x ∈ E attains its maximum on X at some point of B. Then X = I-env(B).
Proof. Assume that there exists x∗ ∈ X \ I-env(B). By Proposition 10.6, there exist
a sequence {xn } in BE and c1 , c2 ∈ R such that
P Simons inequality 3.74, there exists a sequence {λn } of positive numbers such
By the
that ∞ n=1 λn = 1 and
X∞
sup y ∗
λn xn < c2 .
y ∗ ∈X n=1
316 10 Deeper results on function spaces and compact convex sets
n=1
Hence the closed convex hull of {b∗n : n ∈ N} is norm dense in X, from which the
theorem follows.
Theorem 10.9 (James’ theorem for separable spaces). Let B ⊂ E be a closed con-
vex subset of a separable Banach space E such that each element of E ∗ attains its
supremum on B. Then B is weakly compact.
Proof. If B is as in the hypothesis, we first notice that B is bounded since it is weakly
bounded (see the uniform boundedness principle, Theorem 3.15 in [173]) and weakly
w∗
closed by the Mazur theorem A.2. Thus we need to show that B ⊂ E ∗∗ is a subset
w∗
of E. Indeed, if this is the case, B is a w∗ -compact subset of E considered as a
w w∗
subset of E ∗∗ , and thus B = B = B is weakly compact.
w∗
To this end, let ϕ ∈ B be given. By Theorem A.7, it is enough to show that ϕ
is w∗ -continuous on BE ∗ . Since E is separable, it suffices to check the w∗ -sequential
continuity of ϕ|BE∗ . Let {x∗n } be a sequence in BE ∗ w∗ -converging to 0.
Assume that {ϕ(x∗n )} does not converge to 0. Without loss of generality we may
assume that there exists c > 0 such that ϕ(x∗n ) > c for all n ∈ N. If we consider
w∗ w∗
points of E ∗ as functions on B , the assumption ensures that B ⊂ B is a boundary
for E ∗ . Hence B is a boundary for any x∗ ∈ coσ {x∗n : n ∈ N}. Since x∗n (b) → 0 for
each b ∈ B, Corollary 3.74 gives
0 = sup lim sup x∗n (b) = sup lim sup ψ(x∗n ) ≥ lim sup ϕ(x∗n ) ≥ c,
b∈B n→∞ w∗ n→∞ n→∞
ψ∈B
a contradiction.
Thus ϕ(x∗n ) → 0, which is the desired conclusion.
10.1 Boundaries 317
Definition 10.10 (Relatively countably compact and angelic spaces). We recall that
a subset A of a topological space X is relatively countably compact if any sequence
{xn } of elements of A has a cluster point in X.
A topological space X is angelic if each relatively countably compact set A ⊂ X
is relatively compact and any point x ∈ A can be obtained as a limit of a sequence of
points from A.
(e) for any I ⊂ B ∪ {g} finite, ε > 0 and x ∈ X there exists y ∈ Y such that
|f (y) − f (x)| ≤ ε for each f ∈ I,
(f) for any I ⊂ Y finite and ε > 0 there exists f ∈ B such that |f (x) − g(x)| ≤ ε
for every x ∈ I.
The construction will be done inductively. First we notice that, for any finite set
I ⊂ C(X), the set {(f (x))f ∈I : x ∈ X} is a separable subset of RI considered
with the supremum norm. Hence there exists a countable set YI ⊂ X such that
{(f (y))f ∈I : y ∈ Y } is dense in {(f (x))f ∈I : x ∈ X}. If I ⊂ X is finite, we can
choose a sequence {fj(I) } from A pointwise converging to g on I. Let BI consists of
the elements of the sequence {fj(I) }.
We set inductively
[n n−1
[ o [n n−1
[ o
Yn := YI : I ⊂ {g} ∪ Bi and Bn := BI : I ⊂ Yi , n ∈ N.
i=1 i=1
S∞ S∞
At the end of the construction we set Y := n=1 Yn and B := n=1 Bn . Then the
requirements (e) and (f) are satisfied.
By (f), there exists a sequence {fj } of functions from A pointwise converging to g
on Y . We claim that fj (y) → g(y) for all y ∈ X.
Assume that this is not the case. Then there exist y ∈ X and ε > 0 such that the
set {j ∈ N : |fj (y) − g(y)| ≥ ε} is infinite. For I = {f1 , . . . , fm , g} we use (e) to
find ym ∈ Y satisfying
a contradiction.
Hence fj → g as required, and the proof is finished.
Theorem 10.12. Let X be a compact convex set. Then (BAc (X) , τext X ) is an angelic
space. In particular, any norm bounded relatively τext X -countably compact set A is
relatively τX -compact.
Proof. Let A ⊂ BAc (X) be relatively τext X -countably compact. Since (C(X), τX )
is an angelic space, to conclude that A is relatively τX -compact it suffices to show
that any sequence {fn } in A has a τX -cluster point. We select a τext X -cluster point
10.1 Boundaries 319
f0 of {fn } and show that f0 is also a τX -cluster point of {fn }. First we notice that
kf0 k ≤ 1 by the minimum principle of Corollary 2.24.
If we assume a contrary, after omitting finitely many elements from {fn } we may
assume that there exist x1 , . . . , xk ∈ X and ε > 0 such that
Then for i = 1, . . . , k,
Z Z
|fnk0 (xi ) − f0 (xi )| = gnk0 (t) dνi (t) − g0 (t) dνi (t)
ext Y ext Y
Z Z
≤ gnk0 (t) dνi (t) − g0 (t) dνi (t)
K K
Z
+ |gnk0 (t) − g0 (t)| dνi (t)
ext Y \K
ε ε
< + 2 = ε.
2 4
But this contradicts (10.2) and shows that A is relatively τX -compact.
τ
Thus A X is τX -compact, and hence also τext X -compact; in particular τext X -closed.
Thus
τ τ τ
A ext X ⊂ A X ⊂ A ext X .
τ τ
It follows that the identity mapping id : (A ext X , τX ) → (A ext X , τext X ), as a continu-
ous injective mapping, is a homeomorphism. Since the topology τX is angelic, τext X
τ
is angelic on A ext X , which concludes the proof.
320 10 Deeper results on function spaces and compact convex sets
Proof. Let A be a norm bounded relatively σ(E, ext BE ∗ )-countably compact set in
a Banach space E. Then A can be viewed as a relatively τext BE∗ -countably compact
subset of the space Ac (BE ∗ ) (here BE ∗ is endowed with the w∗ -topology). Moreover,
the weak topology coincide with the topology τBE∗ (see Proposition 4.31(f)). Hence
the assertion follows from Theorem 10.12.
Proof. If T is as in the hypothesis, the dual operator T ∗ : (Ac (Y ))∗ → (Ac (X))∗ is
a positive isometry as well and, moreover, it is a (w∗ –w∗ )-homeomorphism. Hence
T ∗ is an affine homeomorphism of S(Ac (Y )) to S(Ac (X)). To see its surjectivity, let
s ∈ S(Ac (X)) be arbitrary. Then t : g 7→ s(T −1 g), g ∈ Ac (Y ), is in S(Ac (Y )) and
T ∗ t = s.
Using the identification of X with S(Ac (X)) and Y with S(Ac (Y ) (see Proposi-
tion 4.31(a)) we finish the proof.
Lemma 10.15. Let X1 , X2 be compact convex sets and, for i = 1, 2, let Fi be closed
faces of Xi such that
• Fi is a split face of Xi ,
• the complementary face Fi0 is closed.
Let ϕ : F1 ∪ F10 → F2 ∪ F20 be a continuous mapping that is affine on F1 and F10 and
maps F1 to F2 and F10 to F20 .
Then there exists a unique continuous affine mapping ψ : X1 → X2 extending ϕ.
Moreover, if ϕ is injective or surjective, then ψ is injective or surjective, respectively.
Proof. Given x ∈ X1 , let ψ(x) be defined as follows. Let x = αx1 + (1 − α)x2 be the
unique decomposition of x with α ∈ [0, 1] and x1 ∈ F1 , x01 ∈ F10 (see Definition 8.4).
We set
ψ(x) := αϕ(x1 ) + (1 − α)ϕ(x2 ).
It is a matter of routine verification that ψ : X1 → X2 is a well-defined affine map-
ping.
To show its continuity, pick g ∈ Ac (X2 ). Then g ◦ ψ is an affine function that
is continuous on F1 ∪ F10 . By Lemma 5.39, g ◦ ψ is continuous on X1 . Hence ψ
is continuous with respect to the weak topology on X2 . Since X2 is compact, ψ is
continuous.
10.2 Isometries of spaces of affine continuous functions 321
Assume that ϕ is injective on F1 ∪ F10 and let ψ(x) = ψ(y) for some x, y ∈ X1 .
Choose α, β ∈ [0, 1], x1 , y1 ∈ F1 and x2 , y2 ∈ F10 such that
Then
αϕ(x1 ) + (1 − α)ϕ(x2 ) = βϕ(y1 ) + (1 − β)ϕ(y2 ).
Since F2 is a split face, this decomposition yields α = β and ϕ(xi ) = ϕ(yi ), i = 1, 2.
Hence xi = yi , i = 1, 2, and x = y.
If ϕ is surjective, then X2 = co(F2 ∪ F20 ) gives ψ(X1 ) = X2 .
are closed faces such that F1 is parallel and F2 is its complementary face.
Proof. Since kT 1k = 1, both F1 , F2 are closed faces. We use the identifications from
Proposition 4.31, namely, that span X = Ac (X)∗ and span Y = Ac (Y )∗ . Then T ∗
is an isometry, and thus T ∗ maps extreme points of BAc (Y )∗ onto ext BAc (X)∗ . Hence
for any y ∈ ext Y there exists x ∈ ext X such that T ∗ y = x or T ∗ y = −x. Then
either
Hence
Y = co ext Y = co(F1 ∪ F2 ).
From this it follows that F10 = F2 (see Exercise 10.93).
Let y = αy1 + (1 − α)y2 for yi ∈ Fi , i = 1, 2, and α ∈ [0, 1]. Then
Theorem 10.17. Let X, Y be compact convex sets with the following property:
• if F1 , F2 are closed faces such that F1 is parallel and F2 = F10 , then F1 is a split
face.
Let T : Ac (X) → Ac (Y ) be an isometric isomorphism. Then there exists an affine
homeomorphism ϕ : Y → X and a function h ∈ Ac (Y ) such that
and
By Lemma 10.16, G1 is a closed parallel face of Y such that the closed face G2 is its
complementary face. Similarly, F1 is a closed parallel face of X such that F2 = F10 .
By the assumption, both F1 and G1 are split faces.
We use the identification of span X with Ac (X)∗ and span Y with Ac (Y )∗ (see
Proposition 4.31). We claim that T ∗ (G1 ) = F1 and T ∗ (G2 ) = −F2 .
Indeed, for any y ∈ G1 we see from
T ∗ y(1) = T 1(y) = 1,
Thus T f (y) = h(y)f (ϕ(y)) for all y ∈ ext Y , and the proof is complete.
10.3 Baire measurability and boundedness of affine functions 323
Proof. The proof follows from Theorem 10.17 and Corollary 8.63.
Further, {0, 1}N can be as a set identified with the family of all subsets of N as
follows. A set A ⊂ N is mapped to its characteristic function cA ∈ {0, 1}N . Thus the
group operation + on {0, 1}N is nothing else than the symmetric difference of sets.
Further, the mapping
copy of the measure on {0, 1} assigning to both {0} and {1} measure 12 . We point
out the fact that λ, as a Haar measure, is invariant with respect to the operation +.
Further,
λ(Us ) = 2−|s| , s ∈ {0, 1}<N .
Lemma 10.20. The following mappings from {0, 1}N × {0, 1}N → {0, 1}N defined as
v = σ + τ ∈ (σ + Q) ∩ A ∩ V.
Definition 10.22 (Q-invariance and tail sets). We say that a function f : {0, 1}N → R
is Q-invariant, if f (σ) = f (τ ) whenever σ differs from τ at finitely many coordinates,
σ, τ ∈ {0, 1}N .
A set A ⊂ {0, 1}N is a tail set if cA is Q-invariant.
10.3 Baire measurability and boundedness of affine functions 325
Corollary 10.23. Let f : {0, 1}N → R have the Baire property and be Q-invariant.
Then f is constant on a comeager subset of {0, 1}N .
Proof. The proof follows from Lemma 10.21(b), because the set Q := {σ ∈ {0, 1}N :
σ has finite support} is a countable dense set.
Proof. For the proof of (a), we notice that f := cA is Q-invariant because A is a tail
set. By Corollary 10.23, f is constant on a comeager set. From this the assertion
follows.
For the proof of (b), assume that A is a λ-measurable tail set. We set
= λ(B)λ(A).
Since any Qλ-measurable set B ⊂ {0,Q1}N can be approximated by the sets of the
form F × ∞ k=n+1 {0, 1}, where F ⊂
n
k=1 {0, 1} (see Proposition A.98), we get the
required conclusion.
Now, since A ∈ A, we obtain λ(A) = λ(A)λ(A), which concludes the proof.
Theorem 10.26. Let µ : {0, 1}N → R be finitely additive and have the Baire property
in the restricted sense. Then µ is a continuous function and countably additive.
326 10 Deeper results on function spaces and compact convex sets
Proof. We divide the proof in several steps. Let Q stand for the set {σ ∈ {0, 1}N :
σ has finite support}.
Step 1: The function µ|Q is uniformly continuous; that is, for every ε > 0 there
exists k ∈ N such that |µ(α) − µ(β)| < ε whenever α|k = β|k , α, β ∈ Q.
For the proof, we use the Baire property of µ to find a comeager set A ⊂ {0, 1}N
such that f |A is continuous. Since
\
A∩ (A − τ )
τ ∈Q
We finish the proof by showing µ = ν on {0, 1}N . Fix σ0 ∈ {0, 1}N and let N ⊂ N
be the support of σ0 . We identify
We have to verify that µ(σ0 ) = ν(σ0 ). To this end we prove the following assertion.
Step 3: We have µ = ν on {0, 1}N .
Let ω := µ − ν. According to the assumption, ω is a finitely additive function with
the Baire property in the restricted sense on {0, 1}N that equals 0 on Q ∩ {0, 1}N .
Hence ω is a Q-invariant function on {0, 1}N , and thus there exist a comeager set A ⊂
{0, 1}N and a ∈ R such that ω = a on A (see Lemma 10.21(b)). By Lemma 10.20,
the set
is comeager in {0, 1}N × {0, 1}N . In particular, we can pick a point (σ, τ ) ∈ B. Then
yields a = 0. Hence
It follows from Zorn’s lemma that every filter is contained in an ultrafilter and it is
easy to see that there are free ultrafilters on N.
Proof. Consider the Cantor set {0, 1}N as a subset of (B`∞ , w∗ ) and let λ stand for the
Haar measure on {0, 1}N . We select a free ultrafilter U on N and define f : `∞ → R
as f (x) := limU x, x ∈ `∞ (equivalently, f (x) = x b(U), where x b is the continuous
extension of x to the Čech–Stone compactification of N).
328 10 Deeper results on function spaces and compact convex sets
For σ ∈ {0, 1}N \ Q, we assign a sequence {pσ (k)}, where pσ (k) is the position of
the k-th 1 in σ.
For each n ∈ N, we define
n
Mn := σ ∈ {0, 1}N \ Q : there exists x∗ ∈ E ∗ such that for each k ∈ N
1o
we have |un (k) − x∗ (xpσ (k) )| ≤ .
n
10.3 Baire measurability and boundedness of affine functions 329
Proof of the claim. First we check that each Mn is open in {0, 1}N \ Q. We pick
σ ∈ Mn and find x∗ ∈ E ∗ such that for each k ∈ N we have |un (k)−x∗ (xpσ (k) )| ≤ n1 .
We find an index j0 ∈ N such that
• un (j) = 0 for j ≥ j0 ,
• |x∗ (xj )| ≤ 1
n for j ≥ j0 .
Let U be the open set in {0, 1}N \ Q consisting of all elements of {0, 1}N \ Q that
coincide with σ on the first pσ (j0 ) coordinates.
Let τ ∈ {0, 1}N \ Q be such a sequence. Let k ∈ N be given. If k ≤ j0 , then
pσ (k) = pτ (k) and thus
1
|un (k) − x∗ (xpτ (k) )| = |un (k) − x∗ (xpσ (k) )| ≤ .
n
If k > j0 , un (k) = 0 and pτ (k) ≥ k > j0 . Hence
1
|un (k) − x∗ (xpτ (k) )| = |x∗ (xpτ (k) )| ≤ .
n
Thus the chosen set U is contained in Mn .
Next we prove that Mn is dense in {0, 1}N \ Q. Let σ ∈ {0, 1}N \ Q be given and
let U be a neighborhood of σ consisting of all elements τ ∈ {0, 1}N \ Q that coincide
with σ on the first m coordinates. We may assume that m is so large that un (k) = 0
for k ≥ m.
We claim that there exists x∗ ∈ E ∗ such that
This means
j
X
lim λi x∗ (xpσ (i) ) = 0, x∗ ∈ E ∗ ,
j→∞
i=1
or equivalently,
∞
X
λi ui = 0, u ∈ U.
i=1
Since U is norm dense in c0 and (c0 )∗ = `1 , λ = 0. This concludes the proof.
10.3 Baire measurability and boundedness of affine functions 331
is a well-defined homeomorphism,
(b) the function h : {0, 1}N → R defined as
where m ∈ N is the least index such that σm 6= τm . Thus x∗ ◦ϕ is continuous for each
x∗ ∈ E ∗ which gives that ϕ : {0, 1}N → X is continuous with respect to the weak
topology of E. Since the weak topology of E coincides with the original topology on
the compact set X, ϕ is continuous.
Further, we check the injectivity of ϕ. Let σ, τ ∈ {0, 1}N satisfy ϕ(σ) = ϕ(τ ), that
is, X X
σk 2−nk yk = τk 2−nk yk . (10.5)
k∈N k∈N
Since the sequence {yk } is strongly independent, (10.5) yields that σ = τ . Hence, ϕ
is injective and thus a homeomorphism.
For the proof of (b), we have to verify that h has the Baire property on each compact
set K ⊂ {0, 1}N . But this is obvious, since ϕ : K → X is a homeomorphic mapping.
This concludes the proof of the claim.
Thus
∞
[ [
ν An = f (cSn∈N An ) = f (cSn∈M An ) = ν An and
n=1 n∈M
X X
ν(An ) = ν(An ).
n∈N n∈M
Hence we may assume that all the sets An are of strictly positive measure.
Then the mapping ϕ : {0, 1}N → BE ∗ defined as
∞
X
ϕ : σ 7→ σn cAn , σ ∈ {0, 1}N ,
n=1
Notation 10.33 (Positive cone of `1 ). Let P stand for the positive cone in `1 , that is,
P := {x ∈ `1 : x ≥ 0}.
Apart from its usual norm topology, we consider on P a finer topology τ whose basis
of neighborhoods for x ∈ P consists of the sets of the form
V (x, ε) := {y ∈ P : y ≥ x, ky − xk ≤ ε}.
Lemma 10.34. The cone P is a Baire space in the topology τ .
Proof. Let {Vn } be a sequence of dense open sets in (P, τ ), and let V0 be a given
nonempty τ -open set. Inductively we find points xn ∈ P and εn ∈ (0, 2−n ), n ≥ 0,
such that V (x0 , ε0 ) ⊂ V0 and
V (xn+1 , εn+1 ) ⊂ Vn+1 ∩ V (xn , εn ), n ≥ 0.
Then the sequence {xn } satisfies xn ≤ xn+1 , n ≥ 0, and kxn − xn+1 k ≤ 2−n . It
follows that the sequence {xn } converges to some x ∈ P . Since {xn } is increasing,
∞
\
x ∈ V0 ∩ Vn ,
n=1
x
Since the mapping x 7→ (kxk, kxk ) is a homeomorphism of P \{0} onto (0, ∞)×{λ ∈
P : kλk = 1}, the function g is Borel on P . Moreover, g is sublinear. Hence the
assertion follows from Lemma 10.35.
Theorem 10.37. Let f be a Borel convex function on a compact convex set X. Then
f is lower bounded.
10.4 Embedding of `1
Definition 10.38. If X is a set and {fn } a bounded sequence in `∞ (X), we say that
{fn } is equivalent to `1 -basis (with constant C) if there exists C > 0 such that
n
X n
X
ci fi `∞ (X)
≥C |ci |
i=1 i=1
Lemma 10.40. Let X be a set, α < β and {fn } be a bounded sequence of functions
on X that satisfies the following condition: for every pair M, N of finite disjoint sets
in N there exists x ∈ X satisfying
\ \
x∈ {y ∈ X : fi (y) < α} ∩ {y ∈ X : fi (y) > β}.
i∈M i∈N
fi (x) > β, i ∈ M,
fi (x) < α, i ∈ N.
Then
n
X X X
ci fi (x) = ci fi (x) + ci fi (x)
i=1 i∈M i∈N
X X
≥β ci + α ci
i∈M i∈N
n n
β+αX β−αX
= ci + |ci |
2 2
i=1 i=1
n
β−α X
≥ |ci |.
2
i=1
Hence
n n n
X X β−αX
ci fi `∞ (X)
≥ ci fi (x) ≥ |ci |.
2
i=1 i=1 i=1
Pn
If (α + β) i=1 ci < 0, we use the previous estimate for the numbers −ci . In both
cases we get
n n
X β−αX
ci fi `∞ (X)
≥ |ci |,
2
i=1 i=1
are nonempty.
Assume now that the construction has been completed up to the n-th stage, that is,
we have x1 , . . . , xn ∈ BE such that the set
\ \
UM,N := {z ∗ ∈ K : z ∗ (xi ) > β} ∩ {z ∗ ∈ K : z ∗ (xi ) < α}
i∈M i∈N
intersects K for each pair of disjoint nonempty finite sets M, N in {1, . . . , n}.
∗
For every such pair we select a point zM,N ∈ K ∩ UM,N and use the assumption
w∗
∗ ∗
to find xM,N , yM,N ∈ co (K ∩ UM,N ) such that f (x∗M,N ) > β and f (yM,N ∗ ) < α.
We find xn+1 ∈ BE satisfying
Then
K ∩ UM,N ∩ {z ∗ ∈ K : z ∗ (xn+1 ) > β} 6= ∅ and
K ∩ UM,N ∩ {z ∗ ∈ K : z ∗ (xn+1 ) < α} 6= ∅.
It follows that
\ \
K∩ {z ∗ ∈ K : z ∗ (xi ) > β} ∩ {z ∗ ∈ K : z ∗ (xi ) < α} 6= ∅
i∈M i∈N
for each pair of disjoint nonempty subsets of {1, . . . , n + 1}. This finishes the con-
struction.
By (10.6), {xn |K } satisfies the assumptions of Lemma 10.40. If C > 0 satisfies
K ⊂ CBE ∗ , for any numbers c1 , . . . , cn we get
n n n
X X β−αX
C ci xi E
≥ ci xi |K ≥ |ci |.
`∞ (K) 2
i=1 i=1 i=1
338 10 Deeper results on function spaces and compact convex sets
Thus
n n
β−αX X
|ci | ≤ ci xi E
,
2C
i=1 i=1
which shows that the sequence {xn } is equivalent to `1 -basis and the proof is finished.
Theorem 10.42. For a Banach space E the following assertions are equivalent:
(i) E does not contain an isomorphic copy of `1 ,
(ii) f |BE∗ is universally measurable for any f ∈ E ∗∗ ,
(iii) for any f ∈ BE ∗∗ and µ ∈ M1 (BE ∗ ) there exists a sequence {xn } in BE such
that xn → f on BE ∗ µ-almost everywhere,
(iv) f |BE∗ satisfies the barycentric formula for any f ∈ E ∗∗ .
Proof. For the proof of (i) =⇒ (iv), let f be an element of E ∗∗ . For a measure
µ ∈ M1 (BE ∗ ) given, pick any numbers β > α. We set K := spt µ and find β 0 > α0
such that [α0 , β 0 ] ⊂ (α, β). By Lemma 10.41, there exists a w∗ -open set U intersecting
K such that
∗
cow (U ∩ K) ⊂ {z ∗ ∈ K : f (z ∗ ) ≥ α0 } ⊂ {z ∗ ∈ K : f (z ∗ ) > α} or
w∗ ∗ ∗ 0 ∗ ∗
co (U ∩ K) ⊂ {z ∈ K : f (z ) ≤ β } ⊂ {z ∈ K : f (z ) < β}.
In both cases we get a w∗ -compact convex set L with strictly positive measure such
that either f > α on L or f < β on L. Hence f satisfies the Haydon condition
(Definition 4.13) and, by Theorem 4.19, f is strongly affine on BE ∗ .
For the proof of (iv) =⇒ (iii) use Theorem 4.34. Since the implication (iii)
=⇒ (ii) is obvious, we proceed to the last implication (ii) =⇒ (i). Assume that
T : `1 → E is an isomorphic embedding satisfying kT −1 k ≤ 1. Then the dual
operator T ∗ : E ∗ → `∞ is surjective and T ∗ (BE ∗ ) ⊃ B`∞ .
Let g ∈ (`∞ )∗ be an element whose restriction to B`∞ is not universally measurable
(see Lemma 10.28). Then f := g ◦ T ∗ is in E ∗∗ and it is not universally measurable
by Corollary 5.27. This concludes the proof.
Then the family {fxn ,yn : n ∈ N} separates points of X. This proves (i) =⇒ (ii).
For the proof of (ii) =⇒ (iii), let {fn : n ∈ N} be a countable family of continuous
functions on X separating points of X. Then
[
(X × X) \ ∆ = {fn−1 (U ) × fn−1 (V ) :
U, V ⊂ R are disjoint closed intervals with rational endpoints}
is an Fσ set. Thus ∆ is Gδ and the proof is complete.
Since
−1
{x ∈ X : πX (x) ⊂ G1 × G2 } = {x ∈ X : ϕ(x) ⊂ G2 } ∩ G1
is an open subset of X,
−1
πX (F ) = πX ((X × Y ) \ U ) = X \ {x ∈ X : πX (x) ⊂ U }
Proof. Let N be a network for a regular space X. Since for each x ∈ X and open set
U containing x there exists an open set V such that x ∈ V ⊂ V ⊂ U , assertion (a)
follows.
For the proof of (b), we start with the assumption that X is determined. Let
f : Y → X be a continuous surjection from a separable metric space Y . If B is
a countable base of open sets in Y , then it easily follows that
{f (B) : B ∈ B}
We claim that Nσ is either empty or a singleton for each σ ∈ {0, 1}N . Indeed, let
σ ∈ {0, 1}N and x, y ∈ X be distinct points. We find a neighborhood V of x such
that y ∈
/ V and n ∈ N such that x ∈ Mn ⊂ V . Then, x ∈ / Nσ if σn = 0, whereas
y∈/ Nσ if σn = 1. Hence Nσ cannot contain both points x, y. We set
Proof. Let X be a determined compact space. By Lemma 10.48, there exists a count-
able network N for X consisting of closed sets. For any pair N, N 0 ∈ N of disjoint
sets we choose a continuous function f(N,N 0 ) ∈ C(X) such that
(
0 on N,
f(N,N 0 ) =
1 on N 0 .
10.5 Metrizability of compact convex sets 343
Then
{f(N,N 0 ) : N, N 0 ∈ N , N ∩ N 0 = ∅}
is a countable family of continuous functions separating points of X. Thus X is
metrizable by Lemma 10.45.
Let K(X) denote the family of all compact subsets of X endowed with the following
metric:
0
if A = B = ∅,
dH (A, B) = 1 if exactly one of A, B is empty,
δ(A, B) ∨ δ(B, A) if both A, B are nonempty.
By Chapter II, Section 4.F in A.S. Kechris [262], dH is a metric on K(X) that turns
it into a compact metric space. Moreover, it is compatible with the so-called Vietoris
topology.
Proof. Let N be a countable network of B (see Lemma 10.48) and let ε ∈ (0, 1) be
fixed. For each finite family F ⊂ N , we set
Let K([−1, 1]) denote the space of all compact subsets of [−1, 1] endowed with the
Hausdorff metric dH (see Definition 10.50) and, for n ∈ N, let (K([−1, 1]))n be
the product space of n copies of K([−1, 1]) with the maximum metric. If F =
{F1 , . . . , Fn }, we define a mapping ϕF : AF → (K([−1, 1]))n as
Since (K([−1, 1]))n is a compact metric space, ϕF (AF ) is its separable subspace.
We select a countable set DF ⊂ AF such that ϕF (DF ) is dense in ϕF (AF ).
Let [
D := {DF : F ⊂ N finite}.
Step 1: For each f ∈ BAc (X) there exists a sequence {fn } of functions from D
such that
lim sup |fn (x) − f (x)| ≤ 2ε, x ∈ B.
n→∞
344 10 Deeper results on function spaces and compact convex sets
Given a function f ∈ BAc (X) , we find a finite open cover U of X such that
diam f (U ) ≤ ε for each U ∈ U. We set
F n := {F1 , . . . , Fn }, n ∈ N,
|fk (x) − f (x)| ≤ |fk (x) − fk (y)| + |fk (y) − f (x)| ≤ 2ε.
Hence lim supn→∞ |fn (x) − f (x)| ≤ 2ε for each x ∈ B, concluding the proof.
Step 2: For each f ∈ BAc (X) there existP
a sequence {fn } of functions from D and
a sequence {λn } of positive numbers with ∞ n=1 λn = 1 such that
∞
X
λn |fn (x) − f (x)| ≤ 3ε, x ∈ X.
n=1
Given f ∈ BAc (X) , by Step 1 there exists a sequence {fn } in D such that
−2ε ≤ lim inf(fn (x) − f (x)) ≤ lim sup(fn (x) − f (x)) ≤ 2ε, x ∈ B.
n→∞ n→∞
−2ε ≤ lim inf(fn (x) − f (x)) ≤ lim sup(fn (x) − f (x)) ≤ 2ε, x ∈ X.
n→∞ n→∞
In other words,
lim sup |fn (x) − f (x)| ≤ 2ε, x ∈ X.
n→∞
A usePof the Simons inequality 3.74 provides a sequence {λn } of positive numbers
with ∞ n=1 λn = 1 such that
∞
X
λn |fn (x) − f (x)| ≤ 3ε, x ∈ X.
n=1
10.5 Metrizability of compact convex sets 345
Step 3: For each f ∈ BAc (X) there Pnexist a finite family {f1 , . . . , fn } ⊂ D and a
positive numbers {c1 , . . . , cn } with k=1 ck = 1 such that
n
X
ck |fk (x) − f (x)| ≤ 6ε, x ∈ X.
k=1
If f ∈ BAc (X) , let {fk } and {λk } be as in Step 2. We find n ∈ N such that
λ := ∞ −1 < 1 + ε and set
P
k=n λk satisfies (1 − λ)
ck := (1 − λ)−1 λk , k = 1, . . . , n.
Then
n
X n
X
ck |fk (x) − f (x)| = (1 − λ)−1 λk |fk (x) − f (x)| ≤ 3ε(1 + ε) ≤ 6ε, x ∈ X,
k=1 k=1
• X is distinguishable.
Then X is separable metrizable.
Proof. Let {gn : n ∈ N} be a family of continuous functions on βX that distinguishes
X in βX. Let {fn : n ∈ N} be a family of continuous functions on X that separates
points of X. We may assume that all functions fn are bounded. We extend the
functions fn (and denote them likewise) to the whole βX.
Let B be a countable base of open sets in R and let
Thus
n
\
x∈ Uyi ⊂ V,
i=1
and {U ∩ X : U ∈ U} is a subbase of X.
As any regular space with a countable base is metrizable (see R. Engelking [169],
Theorem 4.4.7) and clearly separable, the proof is finished.
Lemma 10.53. Let X be completely regular space. Then the following assertions are
equivalent.
(i) X is determined,
Proof. Assume first that X is determined, that is, there exist a separable metric space
M and a continuous surjective mapping f : M → X. Then f ×f : M ×M → X ×X
defined as
f (m, n) := (f (m), f (n)), (m, n) ∈ M × M,
(X × X) \ ∆ = (f × f ) (f × f )−1 ((X × X) \ ∆) ,
separates points of gr ϕ.
10.5 Metrizability of compact convex sets 347
bM (β(gr ϕ) \ gr ϕ) ⊂ M
π c \ M.
Proof. Given F as in the statement of the lemma, we may assume without loss of gen-
erality that 1 ∈ F. Then G := W(span F) − W(span F) is the desired space (recall
that W(H) is the smallest min-stable cone generated by H; see Definition 3.10).
Then there exists a family F 2 ⊂ C b (ext Y ) so that f1 ∈ F 1 if and only if there exists
f2 ∈ F 2 satisfying f1 = f2 ◦ ϕ on ϕ−1 (ext Y ) ∩ ext X.
Thus f2 is both upper and lower semicontinuous, and hence continuous on ext Y .
Obviously, f2 is bounded.
If x ∈ ext X ∩ ϕ−1 (ext Y ), then
Hence F 2 consisting of all such functions constructed for each f1 ∈ F 1 is the required
family.
Theorem 10.56. Let X be a compact convex set. Then the following assertions are
equivalent.
(i) X admits a continuous strictly convex function,
(ii) ext X is a Baire subset of X,
(iii) ext X is distinguishable in X,
(iv) ext X is K-countably determined and countably many continuous functions on
ext X separate points,
(v) ext X is Lindelöf and countable many continuous functions on ext X separate
points,
(vi) X is metrizable.
Proof. We will prove (i) ⇐⇒ (vi) =⇒ (ii) =⇒ (iii) =⇒ (iv) =⇒ (v) =⇒ (iii)
and (iv) =⇒ (vi).
For the proof of (i) =⇒ (vi), let f be a strictly convex continuous function on X.
Then ϕ : X × X → [0, ∞) defined as
1 x + y
ϕ(x, y) := (f (x) + f (y)) − f , (x, y) ∈ X × X,
2 2
is a positive continuous function and
∞ n
\ 1o
∆= (x, y) ∈ X × X : ϕ(x, y) < .
n
n=1
Assume now that (iii) is satisfied. By Lemma 10.46, ext X is a K-countably deter-
mined space. Define ϕ : X × X → X by the formula
1
ϕ(x, y) := (x + y), (x, y) ∈ X × X.
2
It is easy to see that both
Let G 1 consist of the elements of all these sequences for f running through F 2 .
Claim. For every x1 ∈ ext X and x2 ∈ X \ ext X there exists g ∈ G 1 so that g(x1 ) 6=
g(x2 ).
Proof of the claim. We assume that g(x1 ) = g(x2 ) for each g ∈ G 1 . We find a Baire
set A ⊂ X with x1 ∈ A ⊂ X \ {x2 } and a maximal measure µ ∈ Mx2 (X). We use
Theorem 9.12, for maximal measures εx1 , µ, a Baire set A and a family G 1 , to get a
continuous affine surjection ϕ : X → Y of X onto a metrizable compact convex set
Y and a set G 2 ⊂ C(Y ) such that
(a) ϕ] εx1 and ϕ] µ are maximal on Y ,
350 10 Deeper results on function spaces and compact convex sets
= f3 (y1 ).
Analogously we get (ϕ] µ)(f3 ) ≤ f3 (y1 ).
Since ϕ] µ is maximal, it is carried by ext Y . HenceSwe may find an increasing
sequence {Kn } of compact sets in ext Y such that (ϕ] µ)( ∞n=1 Kn ) = 1 and y1 ∈ K1 .
Now we may use the Stone–Weierstrass theorem for the restriction of B to each Kn
to conclude that B = C b (ext Y ).
Indeed, let f ∈ C b (ext Y ) and ε > 0 be given. We find Kn so that (ϕ] µ)(Kn ) >
1 − ε and a function g ∈ F 3 so that
sup |f (y) − g(y)| < ε.
y∈Kn
The claim yields that ext X is distinguishable in ext X by means of functions from
G 1 . This concludes the proof of (v) =⇒ (iii).
To finish the proof we verify (iv) =⇒ (vi). If (iv) holds, Lemma 10.52 yields that
ext X is a separable metrizable space. By Theorem 10.51, X is metrizable.
Proof. We first notice that the implications (ii) =⇒ (i) and (ii’) =⇒ (i’) are
obvious. We start the proof of the converse implications by showing (i) =⇒ (ii). To
this end, let µ ∈ M1max (X) be given. We fix an arbitrary closed set F ⊂ Y \ ext Y .
Since ext Y is Lindelöf, there exists a countable family of cozero sets {Un : n ∈ N}
in Y such that
∞
[
ext Y ⊂ Un ⊂ Y \ F.
n=1
S∞
Then G := ϕ−1 ( n=1 Un ) is an Fσ set. By the assumptions, ext X ⊂ G and hence
µ(G) = 1. Thus
∞
[
(ϕ] µ) Un = µ(G) = 1,
n=1
Indeed, given y ∈ ext Y ∩ ϕ(X), the set ϕ−1 (y) is a closed face. Since
the assumption yields that ϕ−1 (y) is a singleton. Hence (10.10) follows.
Let µ ∈ M1max (X) be given. For any set F ⊂ X \ ext X, (10.10) gives
ϕ(F ) ⊂ Y \ ext Y.
This along with Proposition 3.80 and the first part of the proof yields
Hence
and thus
µ(ϕ−1 (ϕ(F ))) = µ(F ), F ⊂ X, F closed. (10.11)
If µ, ν ∈ M1max (X) are measures with ϕ] µ = ϕ] ν, then (10.11) yields
it is not difficult to realize that ϕ(F ∩ ext X) = ϕ(F ) ∩ ext Y for any F ⊂ X. Hence
ϕ : ext X → ext Y is a closed mapping (that is, ϕ maps closed sets to closed sets),
and thus a homeomorphism on ext X.
Hence we obtain that ext X is a Lindelöf space if ext Y is Lindelöf and ϕ is a
surjective mapping with the properties as above.
Theorem 10.59. Let ϕ : X → Y be a continuous affine mapping of a compact convex
set X to a simplex Y .
(a) Then the following assertions are equivalent:
10.6 Continuous affine images 353
Proof. For the proof of (a), we first verify (i) =⇒ (ii). To this end, let µ be a maximal
probability measure on X. To show that ϕ] µ is maximal on Y , we use Mokobodzki’s
maximality test 3.58.
Let g be a continuous convex function on Y . Since Y is a simplex, g ∗ is an affine
function (see Theorem 6.5). By the assumption ,
g ∗ ◦ ϕ = (g ◦ ϕ)∗ on ext X.
Hence g ∗ ◦ ϕ = (g ◦ ϕ)∗ on X.
Thus the equality
We start the proof of (b) by showing (i’) =⇒ (iii’). We know from (a) that ϕ(X)
is a face of Y and hence a simplex. Since ext ϕ(X) = ϕ(X) ∩ ext Y , we may assume
from now on that ϕ is a surjective mapping onto a simplex Y . By the assumption,
a contradiction.
Step 2: The mapping ϕ is injective.
Suppose that there exists a point y ∈ Y such that the fiber ϕ−1 (y) contains two
different points, say x1 and x2 . Let f ∈ Ac (X) satisfy f (x1 ) > f (x2 ). Let g :
Y → R be defined as above and let h := g ◦ ϕ − f . By the first step, h is upper
semicontinuous and affine. Further, h = 0 on ext X by the assumption. By the
Minimum principles 2.24 and 3.85, h = 0 on X. But h(x1 ) < 0, a contradiction.
This finishes the proof of injectivity of ϕ, and thus ϕ is a homeomorphism.
Since the remaining implications are obvious, the proof is finished.
phic embedding from Definition 4.25. If ϕ : S(H) → S(Ac (F)) is the restriction
mapping, then
φ2 = ϕ ◦ φ1 . (10.12)
By the assumption and Proposition 4.26(d),
and ϕ is injective on ext S(H) by (10.12). Since S(Ac (F)) is a simplex (see Theo-
rem 6.54), Theorem 10.59(b) implies that ϕ is a surjective homeomorphism. Hence
Ac (F) = H, which finishes the proof.
Proof. Assertion (a) follows easily from the definition (see the proof of Theorem
10.59). To show (b), let h ∈ Ac (Y ) satisfy h ◦ ϕ ≥ f . Then
[ d
X
= y∈Y: sup λi fi (x) ≤ 0
λ∈S x∈ϕ−1 (y) i=1
[ d
X
= y∈Y: λi fi ϕ
(y) ≤ 0 .
λ∈S i=1
10.6 Continuous affine images 357
d
X
C := (y, λ) ∈ Y × S : λi fi ϕ (y) ≤ 0 .
i=1
Lemma 10.64. If X is a Bauer simplex, then the sets of the form {x ∈ ext X : f (x) >
a}, f ∈ Ac (X), a ∈ R, form a base of open sets in ext X.
Proof. If X is a Bauer simplex, ext X is a compact set. Thus the sets ni=1 {x ∈
T
ext X : fi (x) > 0}, where n ∈ N and the functions fi are continuous on ext X, form
a base of the topology of ext X. Given functions f1 , . . . , fn in Ac (X), let f ∈ Ac (X)
be such that f1 ∧ · · · ∧ fn = f on ext X (see Theorem 6.37). Then
n
\
{x ∈ ext X : fi (x) > 0} = {x ∈ ext X : f (x) > 0},
i=1
Proof. Given a function f ∈ Ac (X) and y ∈ Y , let δy be the unique maximal measure
in My (Y ). Let x ∈ ϕ−1 (y) be such that f (x) = fϕ (y) and let µx ∈ Mx (X) be
a maximal measure. By Theorem 10.59(a), ϕ] µx is a maximal measure on Y and
represents y. Hence δy = ϕ] µx and
fϕ (y) = f (x) = µx (f )
≤ µx (fϕ ◦ ϕ) = (ϕ] µx )(fϕ ) = δy (fϕ ).
Proof. We start the proof by noticing that Lemma 10.65 yields that fϕ is strongly
affine function for each f ∈ Ac (X). Since ext Y is a closed set, Corollary 5.32 gives
that fϕ is continuous on X if and only if fϕ |ext X is continuous on ext X. Hence (i)
⇐⇒ (iii) by Lemma 10.63.
Further, for any f ∈ Ac (X) and a ∈ R we have
Indeed, if fϕ (y) > a for some y ∈ ext Y , then max f (ϕ−1 (y)) > a. Since ϕ−1 (y) is
a closed face in X (see Proposition 2.72(b)), there exists
such that f (x) = max f (ϕ−1 (y)) (cf. Proposition 2.64(b)). Hence y ∈ ϕ({x ∈
ext X : f (x) > a}). Since the reverse inclusion “⊃” is obvious, (10.13) follows.
Thus (10.13) yields that fϕ |ext Y ∈ C(ext Y ) provided ϕ is open on ext X and
f ∈ Ac (X). Conversely, if each fϕ |ext Y is continuous on ext Y , then ϕ is open on
ext X due to Lemma 10.64. Hence (ii) ⇐⇒ (iii), which concludes the proof.
We apply Proposition 7.11(b) along with Theorem 7.21 to the function cVn+1 ∧ fn to
find a function fn+1 ∈ −S with fn+1 (xn+1 ) > 0 and fn+1 < fn .
This completes the construction. Let now f := infn∈N fn . Since the sequence
{fn } is decreasing, f is an upper semicontinuous S-convex function. Moreover, by a
compactness argument,
∞
\
{x ∈ K : f (x) ≥ 0} = {x ∈ K : fn (x) ≥ 0} 6= ∅.
n=0
By the Minimum principle 7.15(f), f attains its maximum at some point y ∈ ChS (K).
Hence fn (y) ≥ 0, n ≥ 0, and thus
∞
\
y ∈ V0 ∩ Vn ∩ ChS (K).
n=1
Proof. Given an open set U in a compact metric space Y , we find a countable open
cover V = {Vk : k ∈ N} of U such that
• diam V < ε, k ∈ N,
k
• limk→∞ diam Vk = 0,
• dist(Vk , Y \ U ) > 0.
To achieve this, we find compact sets Fn , n ∈ N, such that U = ∞
S
n=1 Fn . For each
n ∈ N we find a finite open cover of Fn consisting of sets of diameter less than 21n ε
that have positive distance from Y \ U . By enumerating all these families into a single
sequence we get the required family V.
We use Theorem 5.1.9 in [169] to find a family {gn : n ∈ N} of positive continuous
functions on U such that
P∞
n=1 gn = cU ,
•
Theorem 10.70. Let P be a Polish space. Then there exists a metrizable simplex X
such that P is homeomorphic to ext X.
Proof. Let P be a Polish space and Y be its metric compactification. We fix a com-
patible metric ρ on TY . By [169], Theorem 4.3.24, P is a Gδ subset of Y , and hence
we may write P = ∞ n=0 Gn , where Y = G0 ⊃ G1 ⊃ · · · are open subsets of Y . For
each n ≥ 0 we use Lemma 10.69 to find continuous functions {hnk }∞ k=1 on Y and a
sequence of strictly positive numbers {βnk }∞
k=1 converging to 0 such that
P∞ n
• h
k=1 k = cGn ,
• diam spt hkn ≤ 2−n βkn , k ∈ N.
We set
pnk := hnk cY \Gn+1 , k ∈ N, n ≥ 0.
Then
∞ X
X ∞
hnk = cY \P .
n=0 k=1
10.7 Several topological results on Choquet boundaries 361
We take into consideration only nonzero functions pnk and inductively find points
xnk , ykn ∈ spt hnk ∩ P such that the family {xnk , ykn : k ∈ N, n ≥ 0} consists of pair-
wise distinct elements. We remark that this is possible because spt hnk ∩ P is infinite,
whenever pnk 6= 0. We define a mapping γ : M(Y ) → M(Y ) as
∞ X
∞ Z
X 1
γ(µ) := µ − pnk dµ (εxn + εykn ). (10.14)
Y 2 k
n=0 k=1
and
Next we verify that γ(µi ) → γ(ν). To this end, we fix f ∈ C(Y ). For each
λ ∈ M(Y \ P ) we denote
Z
1
αkn (λ) := pnk (f − (f (xnk ) + f (ykn )) dλ, k ∈ N, n ≥ 0.
Y 2
Then
∞ ∞ Z
1 XX
pnk dλ (f (xnk ) + f (ykn ))
γ(λ)(f ) = λ(f ) −
2 Y
n=0 k=1
∞ X ∞ Z ∞ ∞ Z
X 1 XX
= pnk dλ − ( pnk dλ)(f (xnk ) + f (ykn ))
Y 2 Y
n=0 k=1 n=0 k=1
X∞ X ∞
= αkn (λ).
n=0 k=1
Let ε > 0 be given. We find δ > 0 such that |f (x) − f (y)| < ε provided ρ(x, y) <
δ. Next we find N ∈ N such that 2−N < 21 δ and K ∈ N satisfying βkn < δ,
n = 1, . . . , N , k ≥ K.
Let A ⊂ N × N consists of all pairs (n, k), where n > N or n = 1, . . . , N and
k > K. Then for any λ ∈ M(Y \ P ), the condition on diameters of the supports hnk
yields
Z N X
X K X
f dγ(λ) − αkn (λ) ≤ |αkn (λ)|
Y n=0 k=1 (n,k)∈A
Z X
≤ε pkn d|λ| ≤ εkλk.
Y (n,k)∈A
For fixed n ≥ 0, k ∈ N,
Z
1
lim αkn (µi ) = lim pnk (f − (f (xnk ) + f (ykn ))) dµi
i→∞ i→∞ Y 2
Z
1
= pnk (f − (f (xnk ) + f (ykn ))) dν n
Y 2
Z
1
= pnk (f − (f (xnk ) + f (ykn ))) dν = αkn (ν).
Y 2
Thus αk (µi ) → αk (ν) and since the series ∞
n n
P P∞ n
n=0 k=1 |αk (λ)| converges uniformly
on BM(Y \P ) , an elementary limit procedure gives
Z ∞ X
X ∞ ∞ X
X ∞ Z
lim f dγ(µi ) = lim αkn (µi ) = αkn (ν) = f dγ(ν).
i→∞ Y i→∞ Y
n=0 k=1 n=0 k=1
We proceed with the proof of the theorem by denoting π : M(Y ) → M(Y )/M the
canonical quotient mapping and by setting X := π(M1 (Y )). Let ε : Y → M1 (Y )
denote the mapping assigning to each x ∈ Y the Dirac measure εx , and let ϕ := π ◦ ε.
We notice the following fact.
Claim 2. If x ∈ P , y ∈ Y and π(εx ) = π(εy ), then x = y.
Proof of Claim 2. Assume that y ∈ Y \ {x} satisfies π(εy ) = π(εx ). Then εy − εx ∈
M , that is, there exists λ ∈ M(Y \ P ) such that
∞ ∞ Z
1 XX 1
εy − εx = γ(λ) = λ − pnk dλ (εxnk + εykn ).
2 Y 2
n=0 k=1
1
But the right-hand side has mass at most 2 at each point of P , a contradiction.
in other words,
∞ X
X ∞
pnk (x)(εxnk − εykn ) ∈ M. (10.15)
n=0 k=1
364 10 Deeper results on function spaces and compact convex sets
Since this measure does not charge Y \ P , (10.15) holds only in case
∞ X
X ∞
pnk (x)(εxnk − εykn ) = 0.
n=0 k=1
As all points xnk , ykn , where pnk is nonzero, are different, this is impossible. Hence
x ∈ P and Claim 3 is proved.
Hence X is a simplex by Proposition 6.9. To finish the proof we verify the follow-
ing fact.
To this end, choose x ∈ ext X. If x ∈ BF for some finite F ⊂ N, then there exists
µ ∈ Mx (X) with µ(AF ) ≥ 21 . Since µ = εx , x ∈ AF . Hence x ∈ AF ∪ (X \ BF )
for every F ⊂ N finite.
Conversely, let x ∈
/ ext X. We choose a maximal measure µ ∈ Mx (X). By the
property of the family {Kn : n ∈ N}, there exists a set N ⊂ N such that
[
ext X ⊂ Kn ⊂ X \ {x}.
n∈N
S
Since µ( n∈N Kn ) = 1 (see Theorem 3.81), there exists a finite set F ⊂ N so that
µ(AF ) ≥ 21 . Since x ∈
/ AF , we get x ∈ BF \ AF and thus x ∈
/ AF ∪ (X \ BF ). This
concludes the proof.
is lower bounded. But f ◦ ϕ(en ) = f (xn ) → −∞, a contradiction (here {en } is the
usual basis of `1 ).
Lemma 10.73. Let f be a fragmented convex function on a compact convex set X and
µ ∈ M+ (X) be a measure on X with kµk > 0. Then for every ε > 0 there exists a
measure ν ∈ M+ (X) such that ν ≤ µ, kνk > 0 and ν(f ) ≥ kνk(f (r( kνk
ν
)) − ε).
Proof. Let µ ∈ M+ (X) and ε > 0 be given. Set Y := co spt µ. By the assump-
tion, there exists an open nonempty set U ⊂ Y so that diam f (U ) ≤ ε. Without
loss of generality we may assume that inf f (U ) = 0 and sup f (U ) ≤ ε. By com-
bining the Krein–Milman and Milman theorems 2.22 and 2.43, we get distinct points
∈ spt µ ∩ ext Y and coefficients λ1 , . . . , λn ∈ (0, 1) so that ni=1 λi = 1
P
y1 , . . . , ynP
and y := ni=1 λi yi ∈ U . We find a closed convex neighborhood V ⊂ U of y and
disjoint open neighborhoods Ui of yi , i = 1, . . . , n, such that
n
X
λi z i ∈ V whenever zi ∈ Ui . (10.16)
i=1
µ|Ui
µi := , i = 1, . . . , n.
µ(Ui )
Let ν̃ := ni=1 λi µi and z be the barycenter of ν̃. From (10.16) we easily obtain that
P
z ∈ V . Thus
ν̃(f ) ≥ 0 ≥ sup f (U ) − ε ≥ f (z) − ε.
We look for ν in the form ν = cν̃ with c > 0. To satisfy the requirement ν ≤ µ, we
choose c ≤ min{µ(U1 ), . . . , µ(Un )}. Then the measure ν = cν̃ has all the required
properties.
Proof. Let P be the positive cone of `1 and S be its base, that is, P = {λ ∈ `1 :
λ ≥ 0} and S = {λ ∈ P : kλk = 1} considered with the norm topology (see Nota-
tion 10.33). Without loss of generality we may assume that f ≥ 0 (see Lemma 10.72).
Let λ0 ∈ S be given. We set
αn := λ0n f (xn ), n ∈ N.
Then {αn } ∈ P .
We define g : P → R as
(
0, λ = 0,
g(λ) :=
kλkf ( ∞ λn
P
n=1 kλk xn ), λ 6= 0.
Since
∞
X λn
λ 7→ xn
kλk
n=1
(Here σλ0 is the sequence {σn λ0n }.) Since σ 7→ σλ0 is continuous on {0, 1}N , ϕ is a
Baire-one function on {0, 1}N .
Set
X∞
N
A := σ ∈ {0, 1} : ϕ(σ) ≤ σn αn .
n=1
P∞
Since the mapping σ 7→ n=1 σn αn is continuous on {0, 1}N and ϕ is Baire-one, A
is a Gδ set (see Theorem A.124). If σ ∈ {0, 1}N is finitely supported, say by a finite
set J ⊂ N, then
X ∞
X 0 X X
ϕ(σ) = g λ0n en ≤ λn g(en ) = λ0n f (xn ) = σn αn .
n∈J n∈J n∈J n=1
λ0 = σλ0 + (N \ σ)λ0 ,
368 10 Deeper results on function spaces and compact convex sets
and thus
ϕ(1) = g(λ0 ) ≤ g(σλ0 ) + g((N \ σ)λ0 )
= ϕ(σ) + ϕ(N \ σ)
∞
X ∞
X ∞
X
≤ σn αn + (1 − σn )αn = αn .
n=1 n=1 n=1
In other words,
∞
X ∞
X
0
λ0n xn λ0n f (xn ).
g(λ ) = f ≤
n=1 n=1
Hence Z XZ
f dµ = f d(µα − µα+1 )
X α<γ X
X
≥ kµα − µα+1 k(f (xα ) − ε)
α<γ
!
X
= kµα − µα+1 kf (xα ) −ε
α<γ
and X
r(µ) = kµα − µα+1 kxα .
α<γ
Proof. Since any Baire-one function is fragmented by Theorems A.124 and A.121,
the preceding Theorem 10.75 asserts that any convex Baire-one function is Ac (X)-
convex. Thus B1 -maximum principle is a particular case of Theorem 3.86.
Proof. Since f is lower bounded, we can find lower bounded sequences {un } and
{−ln } of upper semicontinuous functions such that un < f < ln , un % f and
ln & f (see Theorem A.124).
Let n ∈ N be fixed and x be an arbitrary point of X. Let µ ∈ Mx (Ac (X)) satisfy
µ(ln ) = (ln )∗ (x) (see Lemma 3.21). Then
Hence un < (ln )∗ . We use Corollary 3.49 along with Corollary 4.8 and find a convex
continuous function fn with un < fn < ln .
Obviously, the sequence {fn } is lower bounded and converges to the function f .
370 10 Deeper results on function spaces and compact convex sets
x+y
Φ 1 (x, y) := , (x, y) ∈ X × X,
2 2
is open.
Lemma 10.79. For a compact convex set X the following assertions are equivalent:
(i) f ∗ is continuous for every f ∈ Kc (X),
(ii) ext Xf ∗ is continuous for every f ∈ Kc (X).
Lemma 10.80. Let X be a compact convex set such that f ∗ is continuous for every
f ∈ Kc (X). Then
(a) ext X is a closed set, and
(b) ext Xf ∗ is continuous on X for every f ∈ C(ext X).
Lemma 10.81. Let r : Mmax (X) ∩ M1 (X) → X be an open mapping. Then ext X
is a closed set.
10.9 Function spaces with continuous envelopes 371
Proof. Since the set D := {εx : x ∈ X} is a closed subset of M1 (X), the set
Mmax (X) ∩ M1 (X) \ D is an open set in Mmax (X) ∩ M1 (X). Since the mapping
r is open, the set
is also open.
Theorem 10.84. Let K be a compact space. Then the space M1 (K) is a stable convex
set.
Consequently, any Bauer simplex is a stable convex set.
Proof. Our aim is to prove that the mapping ϕ : M1 (K) × M1 (K) → M1 (K)
defined as
µ+ν
ϕ(µ, ν) := , (µ, ν) ∈ M1 (K) × M1 (K),
2
372 10 Deeper results on function spaces and compact convex sets
is open. Let
n dµ dν o
M := (µ, ν) ∈ M1 (K) × M1 (K) : , are continuous .
d(µ + ν) d(µ + ν)
If we show that M ∩ ϕ−1 (λ) is dense in ϕ−1 (λ) for any λ ∈ M1 (K), we can test by
Lemma 10.82 the openness of ϕ just on the set M.
To check this property, let λ ∈ M1 (K) be given. Let µ, ν ∈ M1 (K) be such that
λ = µ+ν dµ
2 . Let f = d(µ+ν) . Then
1
0≤f ≤1 λ-almost everywhere and λ(f ) = .
2
1
0 ≤ fn ≤ 1, λ(fn ) = and fn → f λ-almost everywhere.
2
Set
µn := fn 2λ and νn := (1 − fn )2λ.
Then (µn , νn ) ∈ M, λ = µn +ν 2
n
and µn → µ, νn → ν. Thus the assumption of
Lemma 10.82 is satisfied for ϕ.
Let (µ, ν) ∈ M be given and let {λi } be a net in M1 (K) converging to λ := µ+ν
2 .
µi +νi
It is enough to find measures (µi , νi ) ∈ M so that λi = 2 and µi → µ, νi → ν.
Since (µ, ν) ∈ M, there exist continuous positive functions f, g on K such that
µ = f λ, ν = gλ and f + g = 2.
and ( µi
µi + νi − kµi k , kµi k ≥ 1,
νi := νi
kνi k , kµi k < 1.
Theorem 10.86. Let X be a compact convex set. Then the following conditions are
equivalent:
(i) X is stable,
(ii) Φλ : (x, y) 7→ λx + (1 − λ)y is an open mapping from X × X onto X for any
λ ∈ [0, 1],
(iii) Φ : (x, y, λ) 7→ λx + (1 − λy) is an open mapping from X × X × [0, 1] onto
X,
(iv) Int C is convex for every convex set C ⊂ X,
(v) co G is open for any open set G ⊂ X,
(vi) r : M1 (X) → X is open,
(vii) r : Mmax (X) ∩ M1 (X) → X is open,
(viii) f ∗ is continuous for any f ∈ C(X),
(ix) f ∗ is continuous for any f ∈ Kc (X),
(x) ext Xf ∗ is continuous for any f ∈ Kc (X).
374 10 Deeper results on function spaces and compact convex sets
Proof. We show (i) =⇒ (ii) =⇒ (iii) =⇒ (iv) =⇒ (v) =⇒ (i), (ii) =⇒ (viii)
=⇒ (ix) =⇒ (vii) =⇒ (i), (viii) ⇐⇒ (vi) and (ix) ⇐⇒ (x).
(i) =⇒ (ii): If Φ 1 is open, Φλ is open for any dyadic rational number λ ∈
2
[0, 1]. Let λ ∈ [0, 1] and y, z ∈ X be given. Set x := λy + (1 − λ)z. Given open
neighborhoods U and V of y and z, respectively, we find a dyadic rational number
α ∈ [0, 1] and a point z 0 ∈ V with x = αy + (1 − α)z 0 . Then
co(U ∪ V ) ⊃ αU + (1 − α)V,
co G ⊂ Int(co G).
Thus co G is open.
(v) =⇒ (i): Let x = y+z 2 , x, y, z ∈ X, and U , V be convex neighborhoods of
y, z, respectively. By (v), co(U ∪ V ) is a neighborhood of x, and thus Lemma 10.85
gives (i).
(ii) =⇒ (viii): By induction it follows that for any n-tuple (λ1 , . . . , λn ) of num-
bers in [0, 1] with λ1 + · · · + λn = 1, the mapping
is open.
Indeed, assuming the validity of the assertion for n − 1, let (λ1 , . . . , λn ) with λ1 +
Pn−1
· · · + λn = 1 be given. If λ := i=1 λi and Φλ : X × X → X be given as
Φλ (x, y) = λx + (1 − λ)y, then
Let f ∈ C(X) be given. Pick x ∈ X and ε > 0. By Lemma 3.21 and Proposition
∗ ε
Pn measure µ ∈ Mx (X) such that µ(f ) > f (x) − 2 . We
4.3, there exists a molecular
writePµ in the form µ = i=1 λi εxi , where x1 , . . . , xn ∈ X, λi are positive numbers
and ni=1 λi = 1. For each i = 1, . . . , n we find an open neighborhood Ui of xi such
that |f (y) − f (xi )| < 2ε for every y ∈ Ui . Due to the previous considerations,
V := Φ(λ1 ,...,λn ) (U1 × · · · × Un )
containing x. We pick y ∈ V and find points yi ∈ Ui , i = 1, . . . , n,
is an open set P
such that y = ni=1 λi yi . Then
n
X n
X
f ∗ (y) ≥ λi f ∗ (yi ) ≥ λi f ∗ (yi )
i=1 i=1
Xn n
X
= λi f (xi ) + λi (f (yi ) − f (xi ))
i=1 i=1
n
εX
≥ µ(f ) − λi
2
i=1
≥ f ∗ (x) − ε.
Thus f ∗ is lower semicontinuous on X. Since f ∗ is always upper semicontinuous, the
proof of the implication is finished.
(vi) ⇐⇒ (viii): We use Lemma 10.63. Let F be a continuous affine function on
M1 (X). Then there exists a function f ∈ C(X) such that µ(f ) = F (µ), µ ∈
M1 (X). Hence we get
Fr (x) = sup{µ(f ) : r(µ) = x, µ ∈ M1 (X)} = f ∗ (x), x ∈ X.
Now the equivalence (vi) ⇐⇒ (vii) is precisely the equivalence (i) ⇐⇒ (ii) in Lemma
10.63.
The implication (viii) =⇒ (ix) is obvious and (ix) ⇐⇒ (x) holds by Lemma 10.79.
The next step is the proof of (ix) =⇒ (vii). By Lemma 10.80(a), the set ext X is
closed. Thus maximal measures are precisely those measures which are carried by
ext X (see Proposition 3.80), in other words Mmax (X) ∩ M1 (X) = M1 (ext X). We
will use Lemma 10.63 again. If F is a continuous affine function on M1 (ext X), there
exists a continuous function f on ext X such that µ(f ) = F (µ), µ ∈ M1 (ext X).
Lemma 3.22 yields
Fr (x) = sup{µ(f ) : r(µ) = x, µ ∈ M1 (ext X)}
= ext Xf ∗ (x), x ∈ X.
(vii) =⇒ (v): By Lemma 10.81, ext X is closed and hence Mmax (X)∩M1 (X) =
M1 (ext X). By Theorem 10.84, the set M1 (ext X) is stable. Thus, M1 (ext X)
satisfies condition (v). Now, we are ready to verify (v) for X: Indeed, if G ⊂ X is
open, then co(r−1 (G)) is open and (vii) implies that
co G = r(co(r−1 (G)))
is open as well.
[ d
X
K \ r(U ) = x∈K: sup si µ(fi ) ≤ 0
s∈S µ∈Mx (H) i=1
[ d
X
= x∈K: sup µ si fi ≤ 0
s∈S µ∈Mx (H) i=1
d
[ X ∗
= x∈K: si fi (x) ≤ 0 .
s∈S i=1
Thus
d
X ∗
K \ r(U ) = πK (x, s) ∈ K × S : si fi (x) ≤ 0 ,
i=1
is continuous on K × S. Thus
d
X ∗
{(x, s) ∈ K × S : si fi (x) ≤ 0}
i=1
10.10 Exercises
Exercise 10.89. Find a boundary of a compact convex set disjoint from the set of
extreme points.
Hint. Take an uncountable set Γ and let X be the unit ball of `∞ (Γ) endowed with
the w∗ -topology. Then ext X = {x ∈ X : |x(γ)| = 1, γ ∈ Γ}. If we set
B := {x ∈ X : |x(γ)| ∈ {0, 1}, x with countable support},
then B is a boundary disjoint from ext X (use Proposition 4.36).
and
∞
X ∞
X
a0 0 + an xn = b0 0 + bn xn .
n=1 n=1
P∞
Then n=1 (an −bn )xn = 0, and thus the strong linear independency yields an = bn ,
n ∈ N. Consequently, a0 = b0 . From this we get that X is a simplex and ext X =
{0} ∪ {xn : n ∈ N}. Hence X is a Bauer simplex.
Exercise 10.92. Let X be a compact convex subset of a Banach space E. Prove that
there exist compact convex Bauer simplices X1 , X2 in E such that X ⊂ co(X1 ∪X2 ).
Xi := co{xn : n ∈ Ni }, i = 1, 2,
Hint. Let F ⊂ X be a face disjoint from F1 and let x ∈ F \ F2 . Then there exist
α ∈ (0, 1) and xi ∈ Fi , i = 1, 2, such that x = αx1 + (1 − α)x2 . Then x1 , x2 ∈ F , a
contradiction with F ∩ F1 = ∅.
380 10 Deeper results on function spaces and compact convex sets
Indeed, the countable additivity follows from the fact that νn is countably additive.
To show its S∞continuity, let σ k → σ in {0, 1}N and ε > 0 be given. Let i0 ∈ N be such
that |νn |( i=i0 Ai ) < ε. Since σ k → σ on each coordinate, there exists k0 ∈ N such
that σ k |i0 = σ|i0 for each k ≥ k0 . Let N := {1, . . . , i0 }. For k ≥ k0 we have
Hence fn is continuous.
If [
f (σ) := ν Ai , σ ∈ {0, 1}N ,
i∈spt σ
then f is finitely additive and it is a pointwise limit of the sequence {fn }. Hence f
has the Baire property in the restricted sense (see Subsection A.2.D) and thus f is
continuous and countably additive by Theorem 10.26. Hence ν is countably additive.
Exercise 10.96. Prove that a compact convex set X with ext X countable is metriz-
able.
Hint. Since any countable space has a countable network, the assertion follows from
Lemma 10.48 and Theorem 10.51.
Hint. Let H1 be the function space on the compact space K1 from Example 3.82 and
let K2 be the quotient of K1 where all points of [0, 1] × {0} are identified with the
point (0, 0) (see [169], Section 2.4). We write q : K1 → K2 for the quotient mapping
and take
Then ϕ(ext X) = ext Y and ϕ is even injective on ext X. On the other hand, if λ is
any continuous probability measure on φ1 ([0, 1] × {0}), then λ is maximal on X, yet
the measure ϕ] λ equals the Dirac measure at the point φ2 (0, 0), and hence ϕ] λ is not
maximal.
and
ext X = φ1 (K1 \ (([0, 1] ∪ [2, 3]) × {0})), ext Y = φ2 (K2 \ ([0, 1] × {0})),
Hint. Let x ∈ K \ ChH (K). By Theorem 3.24, there is f ∈ C(K) such that f (x) <
f ∗ (x). Then U := {y ∈ K : f ∗ (y) − f (y) > 0} is an open neighborhood of x and
U ∩ ChH (K) = ∅.
Exercise 10.103. Let H be a CE-function space on K. Prove that ChH (K) f ∗ is con-
tinuous for every f ∈ C(ChH (K)).
Hint. For necessity use the proof of Lemma 10.63. For sufficiency, asssume that ϕ is
not open. Then there exist an open set U ⊂ K and x ∈ U such that ϕ(x) is not in the
interior of ϕ(U ). Let f ∈ C(K) be positive such that f (x) > 0 and f = 0 on K \ U .
Then the function fϕ is not continuous at ϕ(x).
384 10 Deeper results on function spaces and compact convex sets
The results of Section 10.3 are based mainly upon papers by J. P. R. Christensen.
Theorem 10.24(a) appears in [114], part (b) is standard and we include it just for a
comparison (see, for example, Proposition 272O in [181]). The first version of Theo-
rem 10.26 can be found in [114]; we follow the improved variant of [116], Chapter 5,
and [456], Theorem 2.3.13.
Theorem 10.30 and Exercise 10.92 are taken from J. P. R. Christensen [115]. Theo-
rems 10.31 and 10.37 on automatic boundedness can be found in J. Spurný [430] and
J. Saint Raymond [407]. Theorem 10.32 can be found in J. P. R. Christensen [116],
Theorem 5.8, and [114], Theorem 4. A result related to Theorem 10.32 is due to
M. Talagrand who proved in [442] that an affine function f : [0, 1]N → R measurable
with respect to the sets with the Baire property is continuous. (We note that [0, 1]N is
affinely homeomorphic to (B`∞ , w∗ ).)
We do not know the answer to the following problem.
Section 10.4 follows the paper [220] by R. Haydon. We refer the reader to S. A.
Argyros, G. Godefroy and H. P. Rosenthal [19] for information on spaces not contain-
ing `1 .
Section 10.5 contains results of several authors. Theorem 10.51 can be found in
[350] with a slightly different proof; for the case of extreme points it was proved
by H. H. Corson [127] and a different proof was given by R. Haydon [219]. The
main Theorem 10.56 is an amalgamation of results from papers by M. Hervé [223],
B. MacGibbon [333], J. E. Jayne [247] (see also [394], Section 5.10), G. Debs [141]
and E. A. Reznichenko [387]. We remark that our proof of the result from [141] is
different since it does not use the α-favorability of the set of extreme points of a
compact convex set.
We refer the reader to a paper by E. A. Reznichenko [387] where the following
result (among others) was proved:
Let X be a compact convex set such that
• X is a simplex, or
• ext X is a Lindelöf space.
Then w(X) = w(ext X) = nw(ext X).
(We recall that w(Y ) and nw(Y ) are the weight and network weight, respectively,
of a topological space Y .) It is an open problem whether the conclusion of this theo-
rem holds without additional assumptions on the set X.
Problem 10.108. Let X be a compact convex set. Is it true that w(X) = w(ext X) =
nw(ext X)?
Problem 10.109. Does there exist a nonmetrizable perfectly normal compact convex
set?
(We recall that a topological space is perfectly normal if it is normal and every open
set is of type Fσ .)
Descriptive structure of extreme and exposed points of convex sets was studied by
many authors, we refer the reader to H. H. Corson [126], G. Choquet, H. H. Corson
and V. Klee [112], J. E. Jayne and C. A. Rogers [248], P. Holický and V. Komı́nek
[233], P. Holický and M. Laczkovich [234] or P. Holický and T. Keleti [232].
The problem of transferring properties of a boundary of a compact convex set to
the whole set is investigated in O. Kalenda and J. Spurný [260].
Theorem 10.57 is from M. Kačena and J. Spurný [254]. It was stated in a more
general version without a proof in S. Teleman [455]. However, Example 10.97 from
C. J. K. Batty [32] shows that the result may fail in general. Theorem 10.59 and The-
orem 10.60 are from D. A. Edwards and G. F. Vincent-Smith [161]; our proofs also
use methods of E. A. Reznichenko [387]. An improvement can be found in D. A. Ed-
wards [158], see also G. F. Vincent-Smith [461], A. de la Pradelle [134] and [135].
Theorem 10.66 and Corollary 10.67 are taken from J. Vesterstrøm [460].
Theorem 10.68 for the set of extreme points of a compact convex set is attributed
in [5] to G. Choquet (unpublished result). The general result for Choquet boundaries
is proved in D. A. Edwards [156] (see also Theorem I.5.13 in [5]). It can be found in
G. Choquet [108], Theorem 27.9, that the set of extreme points of a compact convex
set is even α-favorable.
The result of Subsection 10.7.B can be found in G. Choquet [108], Theorem 29.9,
and in R. Haydon [218]. Our exposition follows the paper [218] (see also [24], Sec-
tion 3.6). P. R. Andeneas showed in [15] that every completely metrizable locally
separable space is homeomorphic to ext X for a suitable simplex X. A related result
is due to A. J. Lazar [294], who proved that any uncountable Polish space is home-
omorphic to the set of extreme points of a closed convex body in `2 . A construction
of simplices with a prescribed set of extreme points can be also found in P. J. Stacey
[436]. A related result is due to D. Bensimon [51] and it reads as follows:
Let M be a complete metric space of zero covering dimension. Then there exists a
standard Choquet simplex X such that M is homeomorphic to ext X.
Theorem 10.71 is taken from M. Talagrand [445]. We mention here another result
of M. Talagrand [448]:
There exists a simplex X such that
• ext X is K-analytic,
• ext X is not contained in the smallest family of subsets of ext X containing com-
pact sets and closed with respect to countable unions and intersections,
• ext X is of type Kσδ in β(ext X),
388 10 Deeper results on function spaces and compact convex sets
• there exist an open set U ⊂ ext X and a point ω ∈ ext X such that ext X =
U ∪ {ω},
• ext X \ ext X is discrete.
Section 10.8 follows the paper J. Saint Raymond [408].
Subsection 10.9.A is taken from papers A. Clausing and S. Papadapoulou [121],
S. Papadopoulou [369] and [368] and R. C. O’Brien [365]. We also refer the reader
to G. Debs [138] and [140], Å. Lima [302], A. Clausing and G. Mägerl [120] and
H. Höllein [237]. The results of Subsection 10.9.B follow the paper J. Vesterstrøm
[460].
Exercise 10.90 is rather standard; we refer the reader to Lemma in J. P. R. Chris-
tensen [115] and H. H. Schaefer [409]. Exercise 10.92 is from [115]. Exercise 10.94
can be found in E. M. Alfsen [4], [2] and A. J. Lazar [292]. Exercise 10.95 is a
scalar version of the Hahn–Vitali–Saks theorem (see, for example, N. Dunford and
J. T. Schwartz [150], pp. 158–159). Examples of Exercises 10.97–10.99 are taken
from [254].
Chapter 11
Continuous and measurable selectors
d
Lemma 11.3. Let ϕ : X → 2R be a lower semicontinuous nonempty valued affine
mapping on a simplex X and let U ⊂ Rd be a neighborhood of 0. Then there exists a
continuous affine mapping h : X → Rd such that h(x) ∈ ϕ(x) + U for every x ∈ X.
Since ϕ is lower semicontinuous and affine, g2 , −g1 are lower semicontinuous con-
cave functions with g1 ≤ g2 . By the Edwards in-between theorem 6.6 there exists a
continuous affine function h : X → R such that g1 ≤ h ≤ g2 . By the choice of V ,
h(x) ∈ ϕ(x) + U for every x ∈ X.
d+1
Assume that the assertion holds for Rd . Let ϕ : X → 2R be a lower semi-
continuous affine mapping and U be a neighborhood of 0. Let p and q be canonical
projections of Rd+1 onto R and Rd , respectively. Let Up and Uq be open symmetric
neighborhoods of 0 in R and Rd , respectively, such that Up × Uq ⊂ U .
As p◦ϕ : X → 2R is lower semicontinuous and affine, we may apply the preceding
argument to get a continuous affine function k : X → R such that
p−1 (k(x) + Up ), x ∈ X,
W ⊂ p−1 (k(x) + Up ).
h : x 7→ (k(x), l(x)), x ∈ X,
By setting
d
n
d
X 1 o
ψ(x) := λ = (λ1 , . . . , λn ) ∈ R : λi yi ∈ ϕ(x) + U , x ∈ X,
2
i=1
Indeed, (11.4) yields that ψ is nonempty valued, and, obviously, ψ is affine. Its
lower semicontinuity follows from the observation that
d
n X 1 o
ψ(x) = λ ∈ Rd : ϕ(x) ∩ λi yi − U 6= ∅ .
2
i=1
and
s ∈ ϕ(y) ∩ (t + W ) ⊂ ϕ(y) ∩ V.
Hence
s ∈ ϕ(y) ∩ (h(y) + U ) ∩ V
and
G1 ∩ G2 ⊂ ψ−1 (V ).
This concludes the proof.
for each n ∈ N.
For the inductive step, suppose that the mappings hk have been found for k =
1, . . . , n. Then we apply Lemma 11.4 to
Theorem 11.7. Let X be a compact convex set. Then the following assertions are
equivalent.
(i) X is a simplex,
(ii) any lower semicontinuous affine mapping ϕ : X → 2E with nonempty closed
convex values in a Fréchet space E admits a continuous affine selection.
394 11 Continuous and measurable selectors
Proof. By Theorem 11.6, (i) =⇒ (ii). Conversely, assuming (ii), let f, −g ∈ C(X)
be convex functions such that f ≤ g. Then ϕ : X → 2R defined as
A routine verification shows that ψ is a lower semicontinuous affine mapping and thus
admits a continuous affine selection g. Obviously, g satisfies the required conditions.
r : M1 (F ) ∩ M(H) → coH F
• (Lf )(K) ⊂ co f (F ).
For the proof of (c) we notice that F = co K is a closed metrizable face such that
δx (K) = 1 for each x ∈ F and the mapping x 7→ δx , x ∈ F , is continuous (see
Lemma 11.10 and Lemma 11.11). If ρ : X → F is a continuous affine retraction, the
operator
Lf (x) := δρ(x) (f ), x ∈ X, f ∈ C(K),
is the required extender. This concludes the proof.
Theorem 11.14. For a metrizable simplex X, the following assertions are equivalent:
(i) ext X is uncountable,
(ii) Ac (X) contains an isometric copy of C({0, 1}N ),
(iii) Ac (X) contains an isometric copy of `1 ,
(iv) (Ac (X))∗ is nonseparable.
Proof. Assume that ext X is uncountable. Since ext X is a Polish space (see Proposi-
tion 3.43), ext X contains a homeomorphic copy C of {0, 1}N . By Theorem 11.13(c),
there exists an extender L : C(C) → Ac (X). Hence L(C(C)) is an isometric copy of
C({0, 1}N ) contained in Ac (X). Thus (i) =⇒ (ii).
Since C({0, 1}N ) contains an isometric copy of any separable Banach space (see
[173], proof of Theorem 5.17), (ii) =⇒ (iii). Obviously, (iii) =⇒ (iv). Finally,
assume that ext X is countable. Then Mbnd (X) is isometric to `1 (ext X) and the
mapping π : Mbnd (X) → (Ac (X))∗ defined as π(µ) := µ|(Ac (X))∗ is easily seen to
be an isometry (use Proposition 6.9). Hence (Ac (X))∗ is separable, which proves (iv)
=⇒ (i).
φ(F ) = H ∩ φ(K).
Lh = (E2 ◦ E1 )(h) ◦ φ, h ∈ A,
f (x) := fb(εx ), x ∈ K,
is continuous (we consider the Pettis integral). We remark that fb is well defined by
the assumption on co f (Y ).
By Corollary 11.13, there exists a continuous affine retraction r : M1 (βX) → F .
Then
F (x) := fb(r(εx )), x ∈ βX,
is a continuous extension of f with the required properties. This concludes the proof.
Proof. Let H be as in the hypothesis. Without loss of generality we may assume that
fn (F ) ⊂ [0, 1] for all n ∈ N. We consider a mapping ϕ : F → [0, 1]N (the space
[0, 1]N is considered with the pointwise topology) defined as
x 7→ {fn (x)}n∈N , x ∈ F.
11.3 The weak Dirichlet problem for Baire functions 399
Lemma 11.21. Let H, F , A and r be as in Lemma 11.20. Then, for every f ∈ Hα,b ,
there exists a function h ∈ Aα,b such that h(x) = f (r(x)) for all x ∈ X.
Proof. Assertion (a) follows from Lemma 11.11 and (b) can be obtained by a straight-
forward use of transfinite induction.
Theorem 11.23. Let X be a simplex and K ⊂ ext X be compact. Then for any
bounded function f on K of Baire class α there exists a function h ∈ Aα (X) such
that h = f on K and h(X) ⊂ co f (K).
By (4.9), for each s ∈ X there exists µ ∈ M1 (K) such that π(µ) = s, and then
µ ∈ ϕn (s). Hence all maps ϕn have nonempty values. It is straightforward to verify
that ϕn is affine, and so also ϕn is affine. (Here ϕn is the mapping which assigns to
each s ∈ X the closure of ϕn (s) in M1 (K).)
11.4 Pointwise approximation of maximal measures 401
We show that ϕn and ϕn are lower semicontinuous. To this end, let n ∈ N and let
V be an open subset of M1 (K). If
1
|µ(hk ) − s(hk )| <
n
for any k = 1, . . . , n. There exists an open neighborhood U of s such that for any
t ∈ U and k = 1, . . . , n we have
1
|µ(hk ) − t(hk )| < .
n
Hence µ ∈ ϕn (t) and we see that the set (ϕn )−1 (V ) is open. Since
the mapping ϕn is also lower semicontinuous. By Theorem 11.9, for any n ∈ N there
exists a continuous affine mapping γn : X → M1 (K) such that γn (s) ∈ ϕn (s) for
any s ∈ X. If s ∈ X, then obviously γn (s)(hk ) → s(hk ) for each k ∈ N. Since the
set {hk : k ∈ N} is dense in Ac (H), it immediately follows that γn (s)(h) → s(h) for
any h ∈ Ac (H). This verifies that π(γn (s)) → s.
Proof. As in Section 4.3, let φ be the evaluation mapping from K into X. We may
apply Theorem 11.26 to obtain a sequence of continuous affine mappings γn : X →
M1 (K) such that π(γn (s)) → s for each s ∈ X. If n ∈ N and f ∈ C(K), set
τn (f )(s) := γn (s)(f ), s ∈ X,
Tn f (x) := τn (f )(φ(x)), x ∈ K.
Since εx is the only H-representing measure for x, it would follow that µ = εx . This
contradiction proves (11.6). Given f ∈ C(K), we see that, for any x ∈ ChH (K),
Now, let x ∈ K be arbitrary. Since in metrizable case maximal measures are car-
ried by ChH (K) (see Theorem 3.79), the Lebesgue dominated convergence theorem
assures that
Z Z
Tn f (x) = Tn f (y)dδx (y) → f (y)dδx (y) = δx (f )
ChH (K) ChH (K)
ϕ−1 (U ) := {x ∈ X : ϕ(x) ∩ U 6= ∅} ∈ S
In the sequel we use the notation from Definition 5.2 for a family S of sets in a set
X.
ϕ : X → 2Y be
Proof. We select a compatible metric on Y . For the proof of (a), let S
S -upper measurable and let U ⊂ Y be an open set. We write U = ∞ n=1 Fn , where
Fn ⊂ Y are closed. Then
∞
[
ϕ−1 (U ) = {x ∈ X : ϕ(x) ∩ Fn 6= ∅}
n=1
[∞
X \ ϕ−1 (Y \ Fn )
=
n=1
is in Σ2 (S ).
To verify (b), let ϕ : X → 2Y be S -lower semimeasurable with compact values
and let U ⊂ Y be an open set. Again, we write U = ∞
S
F
n=1 n , where
n 1o
Fn := y ∈ U : dist(y, Y \ Fn ) ≥ , n ∈ N.
n
If K ⊂ Y is compact, then K ⊂ U if and only if K ⊂ Fn for some n ∈ N. Hence
∞
[ ∞
[
ϕ−1 (U ) = ϕ−1 (Fn ) = (X \ ϕ−1 (Y \ Fn )) ,
n=1 n=1
Let V ⊂ Y be an open set such that ϕ1 (x) \ U ⊂ V and dist(V , ϕ2 (x) \ U ) > 0.
Then
ϕ1 (x) ⊂ V ∪ U and ϕ2 (x) ⊂ Int(Y \ V ) ∪ U.
Proof. We select a countable base B of open sets in Y that is stable with respect to
finite unions. We claim that
[
{x ∈ X : diam ϕ(x) < ε} = ϕ−1 (B). (11.8)
B∈B,diam B<ε
Indeed, if diam ϕ(x) < ε, we pick an open set V ⊃ ϕ(x) with diam V < ε. Since
B is stable with respect to finite unions, by the compactness of ϕ(x) we find B ∈ B
such that ϕ(x) ⊂ B ⊂ V .
Since the reverse inclusion is obvious, (11.8) holds. Using Σ2 (S )-upper semimea-
surability of ϕ we conclude the proof.
ψn (x) ⊂ U ⇐⇒ Y \ U ⊂ Y \ ψn (x)
⇐⇒ α > max fn (ϕ(x))
⇐⇒ ϕ(x) ⊂ fn−1 ((−∞, α)).
Hence ψn−1 (U ) ∈ Σ2 (S ).
It follows from the Hahn–Banach theorem and density of {fn : n ∈ N} in Ac (Y )
that
\∞
ψ(x) = ψn (x), x ∈ X.
n=1
It follows that
ϕ(x) ∩ Tj ⊂ ϕ(x) ∩ spt g ⊂ U (y0 , r). (11.10)
By (11.9) and (11.10), we get that x ∈ Dj as required.
Using Lemma 11.31 we obtain a countable partition {Ej : j ≥ 0} of X consisting
of sets from Σ2 (S ) such that Ej ⊂ Dj , j ≥ 0. We define αn : X → N ∪ {0} as
αn (x) := m, x ∈ Em .
is
T∞ a lower semicontinuous S|ϕ(x) -concave function. Further, since {y} = ϕ(x) ∩
n=1 Tαn (x) , we get
Proof. The first part follows from Lemma 11.29 and Theorem 11.35.
If ϕ is upper or lower semicontinuous, then ϕ is Σ2 (Bas(X))-measurable, where
Bas(X) is the algebra from Definition 5.13. An application of the first part yields a
Σ2 (Bas(X))-measurable selection f satisfying (11.11). But this means that f is in
Baf1 (X, Y ).
Proof. We apply Corollary 11.36 to the separable metrizable space E and the cone
S := W({x∗ + c : x∗ ∈ E ∗ , c ∈ R}). Then ChS|ϕ(x) (ϕ(x)) = ext ϕ(x) (see
Proposition 7.6(c) and Exercise 7.63) and hence the assertion follows.
f (x) := fn (x), x ∈ Xn , n ∈ N,
S := W({x∗ + c : x∗ ∈ E ∗ , c ∈ R})
If X ⊂ E is a compact convex set and x ∈ X, then the set Mx (X) of all probabil-
ity Radon measures on X representing x is a compact convex set (in the w∗ -topology).
Moreover, it is an ordered compact convex set (see Section 7.5) and the cone S|Mx (X)
is dense in the cone −Lc (Mx (X)) (here Lc (Mx (X)) is the cone of all continuous
affine isotone functions on Mx (X); see Definition 7.41). Hence by Theorem 7.54(d),
ChS|Mx (X) (Mx (X)) = Ch−Lc (Mx (X)) (Mx (X)) = ext((Mx (X))max ).
(Here (Mx (X))max is the set of all maximal elements in Mx (X).) Hence µ ∈
ext(Mx (X))max is both a maximal and an extreme point of Mx (X) (see Theo-
rem 7.61).
Proof. We recall that r : M1 (K) → K is the barycentric mapping. Let S be the cone
on M+ (X) as above.
1
We define a multivalued mapping M : X → 2M (K) as
Theorem 11.41. Let X be a metrizable compact convex set. Then there exists a map-
ping m ∈ Baf1 (X, M1 (X)) such that
m(x) ∈ ext((Mx (X))max ), x ∈ X.
Proof. We use Corollary 11.36, where Y = M1 (X), S is the convex cone from
Definition 11.39 and ϕ : X → 2Y assigns to x ∈ X the set Mx (X). The role of
the algebra S is played by the algebra Bas(X) generated by closed sets in X (see
Definition 5.13). If r : Y → X is the barycentric mapping and F ⊂ Y is a closed set,
ϕ−1 (F ) = r(F )
is closed in X. Hence ϕ is a Bas(X)-lower semimeasurable mapping. By Corol-
lary 11.36, there exists a Σ2 (Bas(X))-measurable function m : X → Y such that
m(x) ∈ ChS|ϕ(x) (ϕ(x)) = ext((Mx (X))max ), x ∈ X.
Hence m is the required mapping by Theorem 5.19.
412 11 Continuous and measurable selectors
11.6 Exercises
Exercise 11.42. Let K be the one-point compactification of an uncountable discrete
set A and X := M1 (K). Then ext X is uncountable and Ac (X) does not contain `1 .
Hint. Since ext X = {εx : x ∈ K}, ext X is uncountable. Further, Ac (X) is iso-
metric to C(K) and C(K) is isomorphic to c0 (A) (we recall that c0 (A) denotes the
space of all continuous functions on the discrete space A that vanish at infinity, see
Subsection A.3.B). Thus it suffices to show that c0 (A) does not contain `1 . Let
S : `1 → c0 (A) be an isomorphism and let B ⊂ A be countable such that it contains
supports of all vectors Sen , n ∈ N (here {en : n ∈ N} denotes the usual basis in `1 ).
Then any vector in S(`1 ) is supported by B, and thus S(`1 ) is an isomorphic copy of
`1 in c0 , which is impossible.
P f := f − E(f |F ), f ∈ C(βN).
Exercise 11.44. Let X be a simplex such that ext X is countable. Then X is metriz-
able and every strongly affine function is in A1 (X).
Hint. Since ext X is an uncountable Polish space (see Proposition 3.43 and Theo-
rem A.114), it contains a compact perfect subset F ⊂ ext X (see Corollary 6.5 of
[262]). Using Theorem 22.4 in [262] we find a bounded function f ∈ B bα (F ) that
is not of any lower Baire class. Let h ∈ Aα (X) be an extension provided by Theo-
rem 11.23. Then h is not of any lower affine class.
Hint. If f ∈ B bω0 (K), then there exists a bounded sequence {fn } converging to f such
that fn ∈ B bαn (K), where each αn < ω0 . By Exercise 11.46, T fn ∈ (Ac (H))αn +1,b .
Since T fn → T f , we get T f ∈ (Ac (H))ω0 ,b .
Now the assertion follows by transfinite induction.
Hint. If f ∈ Abar (X), then T f = f . Hence the assertion follows from Exercise 11.48
(we recall that Ac (Ac (X)) = Ac (X), see Proposition 4.2).
414 11 Continuous and measurable selectors
fni+1 (xi+1 ) − fni (xi+1 ) > c + η − (fni (x) + fni (xi+1 ) − fni (x))
(11.14)
> c + η − c − εi = η − ε i .
≤ k(c1 , . . . , cn )k∞ .
Hence
n
X
ck gk ≤ k(c1 , . . . , cn )k∞ .
k=1
To show the estimate from below, assume that |ci | = ci . Then (11.12), (11.13) and
(11.14) give
n
X
ck (fnk+1 (xi+1 ) − fnk (xi+1 ))
k=1
X
≥ ci (fni+1 (xi+1 ) − fni (xi+1 )) − |ci | (fnk+1 (xi+1 ) − fnk (xi+1 ))
k∈{1,...,n}\{i}
X
≥ ci η − εi − (εk+1 + 2εk )
k∈N\{i}
∞
X
≥ ci η − 3εk
k=1
3
≥ ci η.
4
Hence
n
X 3
ck gk ≥ ηk(c1 , . . . , cn )k∞
4
k=1
in this case.
416 11 Continuous and measurable selectors
If i ∈ {1, . . . , n} is such that −ci = k(c1 , . . . , cn )k∞ , then we apply the inequali-
ties above to (−c1 , . . . , −cn ) and get
n n
X X 3 3
ck gk = (−ck )gk ≥ ηk(−c1 , . . . , −cn )k∞ = ηk(c1 , . . . , cn )k∞ .
4 4
k=1 k=1
Exercise 11.51. Let X be a metrizable compact convex set and f be an affine lower
semicontinuous function on X that is not continuous. Then there exists a sequence
{gn } of functions in Ac (X) that is equivalent to the standard basis of c0 .
Hint. Without loss of generality we may assume that 0 < f ≤ 1 (see Lemma 4.20).
By Proposition 4.12, Proposition A.53 and Lemma A.54 there exists an increasing
sequence {fn } of affine continuous functions converging to f . By Exercise 11.50
we can extract a subsequence {fnk } such that gk = fnk+1 − fnk form a sequence
equivalent to the standard basis of c0 .
Exercise 11.52. Prove that the space c of convergent sequences endowed with the
sup-norm is isomorphic to c0 .
Hint. Assume first that ext X is closed. Then ext X contains infinitely many distinct
points xn , n ∈ N, and x such that xn → x. Let F := {x} ∪ {xn : n ∈ N}. Then C(F )
is isometric to the space c of convergent sequences which is isomorphic to c0 (see
Exercise 11.52). By Theorem 11.13(c) there exists an extender L : C(F ) → Ac (X).
Then T (C(F )) is isomorphic to c0 .
If ext X is not closed, there exists a concave continuous function f such that f∗
is not continuous (see Theorem 6.37). Then f∗ is an affine (see Theorem 6.5) lower
semicontinuous function that is not continuous. Hence Exercise 11.51 finishes the
argument.
Ch. Léger [299] and A. J. Lazar and J. Lindenstrauss [296], where a selection theorem
for L1 -preduals is proved.
Theorems 11.13, 11.15 and 11.16 are from A. J. Lazar [293]. Theorem 11.14
is a particular version of A. J. Lazar and J. Lindenstrauss [296], Theorem 2.3, and
E. H. Lacey and P. D. Morris [291], Theorem 1.1. Theorem 11.17 is an affine version
on simplices of the Michael selection theorem (see E. Michael [341] and [342]). The-
orem 11.18 can be found in J. Dugundji [149] who proved there the following result
by topological methods:
If A is a closed subset of a metrizable space X and f : A → E is a continuous
mapping to a locally convex space E, then there exists a continuous mapping F :
X → E extending f such that F (X) ⊂ co f (A).
R. Arens showed in [17] that X can be replaced by a paracompact space if E is a
Banach space.
Theorem 11.19 is a variant of results from J. Dugundji [149] and K. Borsuk [77].
An application to potential theory is given in J. Lukeš and J. Kolář [276] and
D. Werner [465] (see also P. Harmand, D. Werner and W. Werner [216], Chapter II,
pp. 96–98).
Section 11.3 follows the paper J. Spurný [428] and the results of Section 11.4 can
be found in J. Lukeš, J. Malý, I. Netuka, M. Smrčka and J. Spurný [319].
Section 11.5 follows ideas from an unpublished paper by G. Debs [137] (see also
[139]). Theorem 11.41 is proved in M. Talagrand [443]. The existence of a Borel
measurable selection assigning to a point x of a compact convex set X a maximal
measure µ ∈ Mx (X) was established by M. Rao [384].
If X is a simplex, the mapping x 7→ δx has the property that T f (x) := δx (f ),
x ∈ X, is Borel for every f ∈ C(X) (see Theorem 6.8). We do not know the answer
to the following question.
Problem 11.54. Let X be a compact convex set. Does there exist a mapping x 7→ µx ,
x ∈ X such that µx ∈ Mx (X) is maximal and x 7→ µx (f ) is Borel for each f ∈
C(X)?
The question of the existence of Borel measurable selectors is a widely investi-
gated area; we refer the reader for example to K. Kuratowski and A. Maitra [286],
K. Kuratowski and C. Ryll-Nardzewski [287], Ch. Castaign and M. Valadier [102],
D. H. Fremlin [180], R. W. Hansell [207] and [208], J. Kaniewski and R. Pol [261]
and J. Spurný and M. Zelený [433].
Exercises 11.48 and 11.49 are proved in M. Capon [96]. The paper by M. Kačena
and J. Spurný [253] shows that there exists a metrizable simplex X such that
T (B bα (X)) \ B bα (X) 6= ∅ for every α ∈ [0, ω0 ). If U is a bounded open set in Rd and
H(U ) is the function space from Definition 13.26, Theorem 1.3 of [253] shows that
the shift of classes ceases to exist for H(U ) at the second stage.
Exercise 11.49 prompts a question whether the shift of finite classes is essential.
This problem was answered in J. Spurný [420] by the following result.
418 11 Continuous and measurable selectors
Problem 11.55. Given n ≥ 3, does there exist a metrizable simplex X and a strongly
affine function f ∈ B n (X) that is not of affine class n?
The goal of this chapter is to present methods of how to construct new function spaces
from given ones. More precisely, we will deal with products of function spaces and
inverse limits of function spaces. The key feature of both of these constructions is that
they preserve simpliciality.
We begin with products of function spaces, so after the basic definitions and auxil-
iary results we come to the results describing Choquet boundaries and maximal mea-
sures of products of function spaces (see Theorem 12.10 and 12.13). In order to show
that the product preserves simpliciality, we need to prove in Subsection 12.1.C several
technical results on peaked partitions of unity and approximations in product spaces.
These results are essential for the proof of Theorem 12.21, showing that the product
of simplicial spaces is simplicial. This method also opens a way for a construction of
tensor products of compact convex sets which we indicate in Exercises 12.85–12.87.
Inverse limits of function spaces and compact convex sets are studied in Subsec-
tions 12.2.B and 12.2.C. Again, the principal results are the preservation of sim-
pliciality (see Theorem 12.34), its consequence to compact convex simplices (see
Corollary 12.35) and a description of the Choquet boundary of the inverse limit (see
Theorem 12.36).
The general results are applied in the framework of compact convex sets to get in
Theorem 12.40 and Corollary 12.42 that Ac (X) is an L∞,1+ε -space provided X is
a simplex. Perhaps the most important result is Theorem 12.45 that asserts that any
metrizable simplex is the inverse limit of a suitable sequence of finite-dimensional
simplices. A much simpler method yields that any simplex is the inverse limit of a
system of metrizable simplices (see Theorem 12.47).
i∈J
N
The multiaffine product i∈I Hi is defined as
n o
Hi := f ∈ C(K) : πjy (f ) ∈ Hj for any j ∈ I, y ∈
O Y
Ki .
i∈I i∈I\{j}
S
We recall that {Jγ }γ∈Γ is a partition of a set I if γ∈Γ Jγ = I and Jα ∩ Jβ = ∅
Q α, β ∈ Γ. If {Ki }i∈I are compact spaces
for distinct Q and {Jγ }γ∈Γ is a partition of
Q
I, then i∈I Ki is naturally homeomorphic to γ∈Γ ( i∈Jγ Ki ). This provides an
Q Q Q
identification of C( i∈I Ki ) with C( γ∈Γ ( i∈Jγ Ki )).
Proposition 12.2. Let Hi be a function space on a compact space Ki , i ∈ I. Let
{Jγ }γ∈Γ be a partition of I. Then the following assertions hold:
J N
(a) i∈I Hi ⊂ i∈I Hi ,
12.1 Products of function spaces 421
J J J
(b) i∈I Hi = γ∈Γ ( i∈Jγ Hi ),
N N N
(c) i∈I Hi = γ∈Γ ( i∈Jγ Hi ).
Notation 12.3. If J ⊂ I, we write HJ for the space of functions from H that depends
only on coordinates from J; precisely,
Proof. We start the proof by noticing that assertions (a)–(c) are easy to verify from
the definitions. If µ ∈ M+ (K) and h ∈ HJ , then
Hence a1 ⊗ a2 ∈ Ac (H1 ⊗ H2 ).
Finally, let a1 , a2 be arbitrary. Then
Ac (Hi ).
N
Thus h ∈ i∈I
424 12 Constructions of function spaces
Proof. Let µ ∈ Fx (H) be given. By Lemma 12.6(a), (πi )] µ ∈ Mxi (Hi ) for any
i ∈ I. By Lemma A.100(a),
we get
µ(K \ E) ≤ µ((K1 \ E1 ) × K2 ) + µ(K1 × (K2 \ E2 ))
= ((π1 )] µ)(K1 \ E1 ) + ((π2 )] )(K2 \ E2 ).
By Proposition 12.6(a), (πi )] µ ∈ Mxi (Hi ) and thus ((πi )] µ)(KQ
i \ Ei ) = 0,
Ni = 1, 2.
Hence µ(K \E) = 0 and E is H1 ⊗H2 -extremal. By induction, i∈I Ei is i∈I Hi -
extremal provided I is finite. Q
Let now I be an arbitrary set and E := i∈I Ei . Assume that µ(K \ E) > 0
for some x ∈ E and µ ∈ Mx (H). We find f ∈ C(K) such that f = 0 on E and
µ(f ) > 0. Let ε > 0 be arbitrary.
426 12 Constructions of function spaces
Q
By Proposition 12.4(e), there exist a finite set J ⊂ I and g ∈ C( i∈J Ki ) such
that, for K 0 := i∈I\J Ki , we have
Q
kf − g ⊗ cK 0 kC(K) < ε.
Then (πJ )] µ ∈ MπJ (x) (πJ (H)) and, by the
Q first part of the argument, πJ (x) is
contained in the πJ (H)-extremal set EJ := i∈J QEi . Hence spt(πJ )] µ ⊂ Ej .
Further, |g| < ε on EJ . Indeed, we pick t ∈ i∈I\J Ei . Then for any s ∈ EJ ,
(s, t) ∈ E. Hence
|g(s)| = |(g ⊗ cK 0 )(s, t) − f (s, t)| < ε.
Using this we obtain
Z
|µ(g ⊗ cK 0 )| = |((πJ )] µ)(g)| ≤ |g| d((πJ )] µ) < ε.
EJ
Finally,
0 < |µ(f )| ≤ |µ(f − g ⊗ cK 0 )| + |µ(g ⊗ cK 0 )| < 2ε.
Since ε is arbitrary, we get a contradiction.
N Q
Theorem 12.10. Let H = i∈I Hi . Then ChH (K) = i∈I ChHi (Ki ).
Proof. The assertion follows from Proposition 12.9, because we have the following
chain of equivalences for x = (xi )i∈I ∈ K:
x ∈ ChH (K) ⇐⇒ {x} is H-extremal
⇐⇒ {xi } is Hi -extremal for each i ∈ I
⇐⇒ xi ∈ ChHi (Ki ) for each i ∈ I.
N N
Proposition 12.11. Let J ⊂ I, H = i∈I Hi and G := i∈J Hi .
(a) If µ ∈ M+ (K) is H-maximal, then (πJ )] µ is G-maximal.
(b) If µ, ν ∈ M+ (K) with µ ≺H ν, then (πJ )] µ ≺G (πJ )] ν.
Proof. Lemma 12.7 yields that for a verification of (a) it is enough to show that
FπJ (x) (G) ⊂ πJ (Fx (H)) for any x ∈ K. But this follows from Proposition 12.8,
which gives
Y Y
FπJ (x) (G) = Fxi (Hi ) = πJ Fxi (Hi ) = πJ (Fx (H)).
i∈J i∈I
Since
µ(K \ (E1 × E2 )) ≤ µ((K1 \ E1 ) × K2 ) + µ(K1 × (K2 \ E2 ))
= ((π1 )] µ)(K1 \ E1 ) + ((π2 )] µ)(K2 \ E2 ) = 0,
we get that h is constant on
for µ-almost all (x1 , x2 ) ∈ K1 ×K2 (here we use Proposition 12.8). By Theorem 8.32,
µ is H-maximal. This concludes the proof.
N +
Theorem 12.13. Let H := i∈I Hi and µ ∈ M (K). If (πi )] µ is Hi -maximal,
i ∈ I, then µ is H-maximal.
In particular, if µi ∈ M1 (Ki ) are Hi -maximal measures, i ∈ I, then i∈I µi is
N
H-maximal.
Proof. Let µ ∈ M+ (K) be as in the hypothesis.
Step 1. We first prove the assertion by induction for finite index sets I. Assume
that it holds for all finite sets with n elements and let I = J ∪ {i0 }, where J has n
elements. For each i ∈ J,
Since (πi )] ((πJ )] µ) = (πi )] µ for i ∈ J, the first part of the proof yields that (πJ )] µ
is πJ (H)-maximal. Hence (πJ )] µ = (πJ )] ν.
We get that
µ(E) = ((πJ )] µ)(E) = ((πJ )] ν)(E) = ν(E)
Q
for any set E := i∈I Ei , where Ei ⊂ Ki is Borel, i ∈ I, and Ei 6= Ki only for
indices in a finite set J. Hence, by Proposition A.99, µ = ν. We conclude that µ is
H-maximal.
we obtain functions f 0 ∈ Kusc (H) and g 0 ∈ S lsc (H) with f 0 ≤ g 0 . By the Edwards
in-between theorem 6.6 there exists a function a ∈ Ac (H) such that f 0 ≤ a ≤ g 0 .
Then a satisfies the required properties.
Without loss of generality we may assume that all numbers αi1 ≥ 0 (if αi1 < 0 for
some i = 1, . . . , n, we would consider −gi |ChH (K) ≤ (−αi1 )φ1 ≤ −fi |ChH (K) for
this index i in (12.3)).
We set
n 1 o
A1 := fi : i ∈ {1, . . . , n} with αi1 6= 0 ,
αi1
n 1 o
B 1 := gi : i ∈ {1, . . . , n} with αi1 6= 0 .
αi1
By (12.3) and the Minimum principle 3.16, any function from A1 is on K smaller
than any function from B 1 . We set
(
max{f : f ∈ A1 } if A1 =
6 ∅,
a1 :=
min{β1l : l = 1, . . . , p} if A1 = ∅,
(
min{g : g ∈ B 1 } if B 1 =
6 ∅,
b1 :=
max{β1l : l = 1, . . . , p} if B 1 = ∅.
fi ≤ αi1 ψ1 ≤ gi
again by the Minimum principle 3.16. This finishes the first step of the proof.
430 12 Constructions of function spaces
Step 2. Assume now that the assertion holds for m − 1 and we have given functions
{φ1 , . . . , φm } satisfying the assumptions (a) and (b). We rewrite (a) as
m−1
X m−1
X
fi |ChH (K) − αij φj ≤ αim φm ≤ gi |ChH (K) − αij φj , i = 1, . . . , n.
j=1 j=1
As in the first step we may assume that αim ≥ 0 for every i = 1, . . . , n. Then we
obtain for i ∈ {1, . . . , n} with αim 6= 0,
m−1
X αij m−1
X αij
1 1
fi |ChH (K) − φj ≤ φm ≤ gi |ChH (K) − φj . (12.4)
αim αim αim αim
j=1 j=1
on ChH (K). For αim = 0 we get for the functions {φ1 , . . . , φm−1 }
m−1
X
fi |ChH (K) ≤ αij φj ≤ gi |ChH (K) . (12.7)
j=1
We apply the induction hypothesis to (12.6) and (12.7), where we consider the
families
1 1 1 1
fr − gs ∪ {fi }, gr − fs ∪ {gi },
αrm αsm αrm αsm
where r, s are such that αrm 6= 0 6= αsm and i satisfies αim = 0. It yields the
existence of a family {ψ1 , . . . , ψm−1 } ⊂ Ac (H) such that
We set
m−1
X αij
1
Am := fi − ψj : i ∈ {1, . . . , n} with αim 6= 0 ,
αim αim
j=1
m−1
X αij
1
B m := gi − ψj : i ∈ {1, . . . , n} with αim 6= 0 .
αim αim
j=1
By (12.9), any function from Am is smaller than any function from B m . Again we
define (
max{f : f ∈ Am } if Am =
6 ∅,
am :=
min{βml : l = 1, . . . , p} if Am = ∅,
(
min{g : g ∈ B m } if B m =
6 ∅,
bm :=
max{βml : l = 1, . . . , p} if B m = ∅.
Then −am , bm are continuous H-concave functions such that am ≤ bm and am (xl ) ≤
βml ≤ bm (xl ), l = 1, . . . , p (here we use (b), (12.4) and (12.8)). Using Theo-
rem 12.14, we get a function ψm ∈ Ac (H) such that
am ≤ ψm ≤ bm and ψm (xl ) = βml , l = 1, . . . , p. (12.11)
Then {ψ1 , . . . , ψm } is the required family of functions (for αim 6= 0 we get (c) from
(12.11) and the definition of Am , B m , for αim = 0 we obtain (c) from (12.10)). This
concludes the proof.
≤ fi + ε − αim
Using Lemma 12.16 we find functions ψ1 , . . . , ψm−1 in Ac (H) such that (f)–(i) hold
with φj replaced by ψj (in Lemma 12.16, we consider families
m−1
X
ψm := 1 − ψj .
j=1
we obtain (d). Finally, by rewriting (f∗ ) we get (e). This concludes the proof.
12.1 Products of function spaces 433
Let
Vgj := Ugj ∩ ChH1 (K1 ), j = 1, . . . , p.
We set V 0 := ∅ and inductively define families V j , j = 1, . . . , p, as follows. Let
(
V j−1 ∪ {Vgj } if V j−1 ∪ {Vgj +1 , . . . , Vgp } does not cover ChH1 (K1 ),
V j :=
V j−1 otherwise.
After relabeling we can assume that V p = {Vg1 , . . . , Vgm } for some m ∈ {1, . . . , p}.
The family V p is an open cover of ChH1 (K1 ) such that for any l = 1, . . . , m there
exists [
x l ∈ V gl \ Vgj .
j∈{1,...,m}\{l}
434 12 Constructions of function spaces
We denote
ε
C := {g1 , . . . , gp } − co{g1 , . . . , gm } and D := C + U 0, .
3
For any i = 1, . . . , n, the set pi (C) is a compact subset of H2 . By Theorem A.32
and the compactness of K2 , there exist points {ξi1 , . . . , ξiqi } in K2 and their open
neighborhoods {Wi1 , . . . , Wiqi } such that
qi
[ ε
K2 ⊂ Wir and diam h(Wir ) < , r = 1, . . . , qi , h ∈ pi (C).
3
r=1
For any h ∈ pi (C) we find yh ∈ K2 and r ∈ {1, . . . , qi } such that yh ∈ Wir and
|h(yh )| = khk. Hence
ε
khk − < max |h(ξir )| ≤ khk, h ∈ pi (C).
3 r=1,...,qi
Since pi (D) ⊂ pi (C) + U (0, 3ε ), we obtain
2ε
khk − < max |h(ξir )| ≤ khk, h ∈ pi (D). (12.13)
3 r=1,...,qi
Then
m
(
X 1, j = l,
φj ≥ 0, φj = 1, φj (xl ) = j, l = 1, . . . , m.
j=1
0, j 6= l,
Further, for each x ∈ ChH1 (K1 ), there exists a unique index jk such that φjk (x) 6= 0.
Hence, from (12.12) we get
ε
kf (x) − gjk k∞ < .
3
12.1 Products of function spaces 435
Since {Vg1 , . . . , Vgm } is a cover of ChH1 (K1 ), we see from (12.12) that
ε
f (x) ∈ {g1 , . . . , gm } + U (0, ), x ∈ ChH1 (K1 ).
3
Pm
Further j=1 φj (x)gj ∈ co{g1 , . . . , gm }, and thus
m
X
f (x) − φj (x)gj ∈ D.
j=1
i = 1, . . . , n, r = 1, . . . , qi , x ∈ ChH1 (K1 ).
We apply Proposition 12.17 to the family of functions
{Γir ◦ f : i = 1, . . . , n, r = 1, . . . , qi } ⊂ Ac (H1 )
i = 1, . . . , n, r = 1, . . . , qi , x ∈ ChH1 (K1 ),
(
1, j = l,
ψj (xl ) = j, l = 1, . . . , m. (12.18)
0, j= 6 l,
m
X
= max Γir (f (x)) − ψj (x)Γir (gj )
i=1,...,n
r=1,...,qi j=1
ε
≤ , x ∈ K1 .
3
Hence
m
X
f (x) − ψj (x)gj ∞
< ε, x ∈ K1 . (12.19)
j=1
Proof. It follows from Proposition 12.18 that any function in Ac (H1 ) ⊗ Ac (H2 ) can
be uniformly approximated by functions from Ac (H1 ) Ac (H2 ). Using this fact, by
Lemma 12.5 and Proposition 12.6(c) we obtain
Proof. Let H := H1 ⊗ H2 . We show that Ac (H) has the weak Riesz interpolation
property (see Theorem 6.16). To this end, let a, b, c, d be functions from Ac (H) such
that a ∨ b < c ∧ d. We choose ε > 0 such that c ∧ d − a ∨ b > ε. Since Ac (H) ⊂
Ac (H1 ) ⊗ Ac (H2 ) by Proposition 12.6(c), we may use Proposition 12.18 to find a
partition of unity {ψ1 , . . . , ψm } ⊂ Ac (H1 ), points x1 , . . . , xm in ChH1 (K1 ) and a
family of functions {aj , bj , cj , dj : j = 1, . . . , m} ⊂ Ac (H2 ) such that
(
1, j = l,
ψj (xl ) = j, l = 1, . . . , m, (12.20)
0, j 6= l,
we have
a ∨ b < a0 ∨ b0 < c0 ∧ d0 < c ∧ d. (12.22)
Then also
π2x (a0 ) ∨ π2x (b0 ) < π2x (c0 ) ∧ π2x (d0 ), x ∈ K1 . (12.23)
For any j = 1, . . . , m we obtain from (12.20) and (12.21) that
x x x x
π2 j (a0 ) = aj , π2 j (b0 ) = bj , π2 j (c0 ) = cj , π2 j (d0 ) = dj .
aj ∨ bj < cj ∧ dj , j = 1, . . . , m.
Since Ac (H2 ) has the weak Riesz interpolation property (see Theorem 6.16), there
exists a function hj ∈ Ac (H2 ) with
belongs to Ac (H1 ) ⊗ Ac (H2 ) = Ac (H) (see Proposition 12.19). Since the functions
ψj are positive, (12.24) and (12.22) yield
a ∨ b < h < c ∧ d.
is simplicial, and
O O
Ac (H) = Ac H i ⊗ H i 0 = Ac Hi ⊗ Ac (Hi0 )
i∈J i∈J
O O
= Ac (Hi ) ⊗ Ac (Hi0 ) = Ac (Hi ).
i∈J i∈I
N
NStep 2.c Let now Ic be an arbitrary index set and H := i∈I Hi . We first show that
i∈I A (Hi ) = A (H). Let J ⊂ I be finite and
!
O
h∈ Ac (Hi )
i∈I J
(that is, h depends only on coordinates from J). By the first part of the proof,
O O
πJ (h) ∈ Ac (Hi ) = Ac Hi .
i∈J i∈J
By Proposition 12.4(e), ( i∈I Ac (Hi ))f is dense in i∈I Ac (Hi ). It follows that
N N
!f
O O
c
A (Hi ) = c
A (Hi ) ⊂ Ac (H) = Ac (H).
i∈I i∈I
12.1 Products of function spaces 439
Since the second inclusion follows from Proposition 12.6(c), we get the required
c N c
equality A (H) = i∈I A (Hi ).
Let a, b, c, d ∈ Ac (H). Then a, b, c, d ∈ c
N
i∈I A (Hi ), and therefore Proposi-
tion 12.4(e) yields the existence of a finite set J ⊂ I and functions
!
O
0 0 0 0 c
a ,b ,c ,d ∈ A (Hi )
i∈I J
such that
a ∨ b < a0 ∨ b0 < c0 ∧ d0 < c ∧ d.
Then
πJ (a0 ) ∨ πJ (b0 ) < πJ (c0 ) ∧ πJ (d0 ).
N
Since i∈J Hi is simplicial, there exists a function
O O
h0 ∈ Ac Hi = Ac (Hi )
i∈J i∈J
such that
πJ (a0 ) ∨ πJ (b0 ) < h0 < πJ (c0 ) ∧ πJ (d0 ).
Then the function
h := h0 ⊗ cQi∈I\J Ki
Ac (Hi ) = Ac (H) and satisfies
N
belongs to i∈I
a ∨ b < h < c ∧ d.
Theorem
N 12.22. Let Hi be function spaces on compact spaces Ki , i ∈ I, such that
i∈I Hi is simplicial. Then Hi is simplicial for each i ∈ I.
a := aj ⊗ cK 0 , b := bj ⊗ cK 0 , c := cj ⊗ cK 0 , d := dj ⊗ cK 0
By the simpliciality of Ac (H) we find a function h ∈ Ac (H) with a∨b < h < c∧d.
We select an arbitrary y ∈ K 0 . Proposition 12.6(c) yields πjy (h) ∈ Ac (Hj ). Then
and Ac (Hj ) has the weak Riesz interpolation property. Hence Hj is simplicial by
Theorem 6.16.
φ2 ◦ ϕ = ψ ◦ φ1 . (12.25)
By an analogous straightforward argument we get (b). For the proof of (c), let
µ ≺H1 ν for µ, ν ∈ M+ (K1 ) be given. For any g ∈ Kc (H2 ), we obtain
ψ] ◦ (φ1 )] = (φ2 )] ◦ ϕ] ,
Definition 12.27 (Inverse system of compact convex sets). Let I be an up-directed set
and (Xi , pij )i,j∈I be a system of compact convex sets and mappings such that
• pij : Xj → Xi is a continuous affine surjection, i ≤ j,
• pii is identity, i ∈ I,
• pij ◦ pjk = pik , i ≤ j ≤ k.
Then ((Xi , Ac (Xi )), pij )i,j∈I is an inverse system of function spaces as defined in
Definition 12.26. We call it an inverse family of compact convex sets and the compact
convex set X := lim(Xi , pij )i,j∈I is its inverse limit. Sometimes we write briefly
←−
X = lim Xi .
←
Further, πi : X → Xi is an affine continuous surjection for each i ∈ I.
Proposition 12.28. Let ((Ki , Hi ), pij )i,j∈I be an inverse system of function spaces
and (K, H) := lim((Ki , Hi ), pij )i,j∈I be its inverse limit. Then
←−
(a) H is a well-defined function space,
(b) every πi : (K, H) → (Ki , Hi ) is an admissible mapping.
12.2 Inverse limits of function spaces 443
Proof. To show that H is a function space, we first notice that H contains constant
functions. If x, y ∈ K differs at a coordinate i ∈ I, let h ∈ Hi be such that h(xi ) 6=
h(yi ). Then h ◦ πi ∈ H separates the points x and y. Let f ◦ πi , g ◦ πj ∈ H, where
i, j ∈ I, f ∈ Hi and g ∈ Hj . We select k ∈ I such that i, j ≤ k. Then
f ◦ πi + g ◦ πj = f ◦ pik ◦ πk + g ◦ pjk ◦ πk ∈ H,
Proposition 12.29. If ((Ki , Hi ), pij )i,j∈I is an inverse system of function spaces and
(K, H) is its inverse limit, we consider the family (S(Hi ), ψij )i,j∈I , where ψij :
S(Hj ) → S(Hi ) are the restriction mappings.
(a) The family (S(Hi ), ψij )i,j∈I is an inverse system of compact convex sets.
(b) S(H) is affinely homeomorphic to lim S(Hi ).
←−
Proof. The system (S(Hi ), ψij )i,j∈I is inverse by a straightforward verification. For
the proof of (b), let X := lim S(Hi ) be the limit of the system (S(Hi ), ψij )i,j∈I . Let
←
Φ : H → Ac (S(H)) be the mapping from Definition 4.25, that is,
0 ≤ Φ(h)(ϕ(s)) = sj (h0 ), s ∈ X.
Proposition 12.30. Let (Xi , pij )i,j∈I be an inverse system of compact convex sets
and let X := lim Xi be its limit. Let (X, H) be the inverse limit of the family
←
((Xi , Ac (Xi )), pij )i,j∈I , where every (Xi , Ac (Xi )) is regarded as a function space.
Then
(a) H = Ac (X),
(b) S(H) is affinely homeomorphic to X.
Proof. Given the objects as in the hypothesis, we notice that H = {f ◦ πi : f ∈
Ac (Xi ), i ∈ I} is contained in Ac (X). By Proposition 12.28(a), H contains constants
and separates points of X. Thus Exercise 4.49 yields H = Ac (X) and (a) holds.
By Proposition 4.31(a), X is affinely homeomorphic to
S(Ac (X)) = S(H) = S(H).
This concludes the proof.
Theorem 12.31. Let ((Ki , Hi ), pij )i,j∈I be an inverse system of function spaces and
(K, H) be its inverse limit. Let µ ∈ M1 (lim Ki ) be a measure such that (πi )] µ is
←
Hi -maximal for any i ∈ I. Then µ is H-maximal.
Proof. Let K := lim Ki and let ν ∈ M1 (K) satisfy µ ≺H ν. Proposition 12.28(b)
←
and Proposition 12.24(c) yield (πi )] µ ≺Hi (πi )] ν for each i ∈ I. Since (πi )] µ is
Hi -maximal, (πi )] ν = (πi )] µ. Thus Lemma A.103 yields
ν = lim(πi )] ν = lim(πi )] µ = µ.
← ←
Proposition 12.32. Let ((Ki , Hi ), pij )i,j∈I be an inverse system of function spaces
and (K, H) be its inverse limit.
(a) Let x = (xi )i∈I ∈ K and µ ∈ M1 (K). Then µ ∈ Mx (H) if and only if
((πi )] µ, pij )i,j∈I is an inverse system of measures (see Definition A.101) with
(πi )] µ ∈ Mxi (Hi ) for i ∈ I.
(b) Let x = (xi )i∈I ∈ K and let (µi , pij )i,j∈I be an inverse system of measures with
µi ∈ Mxi (Hi ). Then µ := lim µi is an H-representing measure for x.
←−
Proof. For the proof of (a), we first assume that µ ∈ Mx (H). Propositions 12.28(b)
and 12.24(a) yield that (πi )] µ ∈ Mxi (Hi ) for any i ∈ I. Lemma A.103 implies that
((πi )] µ, (pij )] )i,j∈I is an inverse system.
Conversely, let ((πi )] µ, pij )i,j∈I be an inverse system of measures with (πi )] µ ∈
Mxi (Hi ) for i ∈ I. Let h ∈ Hj for some j ∈ I. Then
µ(h ◦ πj ) = ((πj )] µ)(h) = h(xj ) = (h ◦ πj )(x).
Thus µ ∈ Mx (H).
12.2 Inverse limits of function spaces 445
To check (b), Definition A.101 provides a measure µ ∈ M1 (K) such that (πi )] µ =
µi , i ∈ I. By (a), µ ∈ Mx (H), and we are done.
Corollary 12.35. Let (Xi , pij )i,j∈I be an inverse family of Choquet simplices. Then
its inverse limit is a Choquet simplex.
Proof. Let X := lim Xi be the inverse limit and let (X, H) be the inverse limit of
←
the family ((Xi , Ac (Xi )), pij )i,j∈I whose elements are regarded as function spaces.
By Theorem 12.34, H is a simplicial function space. Proposition 12.30 yields H =
Ac (X), and thus Ac (X) is simplicial as well. But this means exactly that X is a
simplex.
Proof. For the proof of (a), let x ∈ K satisfy xi ∈ ChHi (Ki ) for each i ∈ I. Let
µ ∈ Mx (H) be given. By Proposition 12.32, ((πi )] µ, pij )i,j∈I is an inverse system
of measures with (πi )] µ ∈ Mxi (Hi ) for i ∈ I. Then (πi )] µ = εxi for each i ∈ I and
Lemma A.103 gives
µ = lim(πi )] µ = lim εxi = εx .
← ←
Without loss of generality we may assume that h = g ◦πj for some h ∈ Hj and j ≥ i.
Then
(f ◦ pij )(xj ) = (f ◦ pij ◦ πj )(x) = (f ◦ πi )(x)
< (h ◦ πj )(x) < (f ◦ pij )(xj ) + ε
gives
(f ◦ pij )(xj ) < h(xj ) < (f ◦ pij )(xj ) + ε.
This along with Proposition 12.25 yields
Hence
f ∗ (xi ) < f (xi ) + ε.
Since ε > 0 is arbitrary, f ∗ (xi ) = f (xi ) for any f ∈ Kc (Hi ). By Theorem 3.24,
xi ∈ ChHi (Ki ).
12.2 Inverse limits of function spaces 447
Lemma 12.38. (a) Let {a1 , . . . , an } be an admissible basis of `∞ n . Then there exists a
permutation π : {1, . . . , n} → {1, . . . , n} and signs i ∈ {−1, 1}, i = 1, . . . , n,
such that ai = i eπ(i) , i = 1, . . . , n.
If, moreover, {a1 , . . . , an } is positive, all signs i are positive.
(b) Let m > n, T : `∞ ∞
n → `m be an isometry and let {u1 , . . . , un } be an admissible
∞
basis of `n . Then there exist an admissible basis
{v1 , . . . , vm } ⊂ `∞
m
{ai,j : i = 1, . . . , n, j = n + 1, . . . , m}
such that
T ui = vi + m
P
j=n+1 ai,j vj , i = 1, . . . , n,
•
Pn
i=1 |ai,j | ≤ 1, j = n + 1, . . . , m.
•
Fk := span{T u1 , . . . , T un , vn+1 , . . . , vk }
is isometric to `∞
k .
(c) If T in (b) is positive and {u1 , . . . , un } is an admissible positive basis, all num-
bers ai,j , i = 1, . . . , n, j = n + 1, . . . , m, and the basis {v1 , . . . , vm } can be
chosen to be positive.
(d) Let F ⊂ `∞ ∞
m be a subspace isometric to `n , n < m. Then there exist subspaces
Fn+1 , . . . , Fm−1 of `∞ ∞
m such that F ⊂ Fn+1 ⊂ · · · ⊂ Fm−1 ⊂ `m and Fi is
∞
isometric to `i , i = n + 1, . . . , m − 1.
Proof. For the proof of (a), let ai := T ei , i = 1, . . . , n, for some isometry T : `∞
n →
`∞
n . Let e stand for the vector (1, 1, . . . , 1) ∈ `∞ . Since e ∈ ext B ∞ , T e ∈ B ∞ , and
n `n `n
thus there exist signs i ∈ {−1, 1}, i = 1, . . . , n, such that
T e = (1 , . . . , n ).
Let S : `∞ ∞
n → `n be the isometry that maps ei to i ei , i = 1, . . . , n. Then ST is a
positive isometry of `∞
n onto itself.
448 12 Constructions of function spaces
Since
e = ST (e1 + · · · + en ), kST ei k = 1, and ST ei ≥ 0, i = 1, . . . , n,
it follows that ST ei = eπ(i) , i = 1, . . . , n, for some permutation π : {1, . . . , n} →
{1, . . . , n}. Hence T ei = i eπ(i) as needed.
For the proof of (b), let T : `∞ ∞
n → `m be an isometry and {u1 , . . . , un } be an
∞
admissible basis of `n . We write T ui in coordinates with respect to the canonical
basis {e1 , . . . , em } of `∞
m as
T ui = (ci,1 , . . . , ci,m ), i = 1, . . . , n.
For any choice of signs 1 , . . . , n we have
n
X n
X n
X
1= T i ui = i ci,1 , . . . , i ci,m .
i=1 i=1 i=1
{a1,m , . . . , am,m }, m = p, . . . , p + r − 1,
such that
• ei = ei,p+1 + ai,p ep+1,p+1 , i = 1, . . . , p,
• ei,m = ei,m+1 + ai,m em+1,m+1 , i = 1, . . . , m, m = p + 1, . . . , p + r − 1,
Pm
i=1 ai,m = 1, m = p, . . . , p + r − 1.
•
Proof. For the proof of (a), let T and {f1 , . . . , fn } be as in the hypothesis. We find
points {x1 , . . . , xn } such that fi (xi ) = 1, i = 1, . . . , n. Obviously, T is positive. For
a given (a1 , . . . , an ) ∈ `∞
n , let i0 be an index satisfying |ai0 | = k(a1 , . . . , an )k. Then
n
X n
X
ai fi (x) ≤ |ai0 | fi (x) = k(a1 , . . . , an )k, x ∈ X.
i=1 i=1
onto `∞
n if necessary we may assume that T 1 = (1, . . . , 1) =: e. Then T is positive
and the functions fi := T −1 ei , i = 1, . . . , n, form the required partition of unity.
Indeed,
X n n
X
1 = T −1 (e) = T −1 ei = fi ,
i=1 i=1
SIT −1 (`∞ ∞
p ) ⊂ Fp+1 ⊂ `p+r .
and pi=1 ai ≤ 1.
P
Let E := (JS)−1 (`∞ ∞
p+1 ) and V : E → `p+1 denote the restriction of JS. Since
1 ∈ E and V is a positive isometry, V 1 = (1, . . . , 1) ∈ `∞
p+1 . We set
Then {e1,p+1 , . . . , ep+1,p+1 } is a peaked partition of unity in Ep+r . Indeed, all the
functions are obviously positive and have norm one. Since
p+1
X
V 1 = (1, . . . , 1) = vi ,
i=1
we obtain
p+1
X
1= ei,p+1 .
i=1
Pp
We claim that i=1 ai,p = 1. Indeed, from (12.26) we get
Hence
p
X p
X
1= ei,p+1 + ep+1,p+1 ai,p .
i=1 i=1
12.2 Inverse limits of function spaces 451
Now the rest of the argument follows by an application of the this procedure to
the family {e1,m , . . . , em,m } for all m = p + 1, . . . , p + r − 1. This concludes the
proof.
Proof. Given the objects as in the theorem, we first notice that without loss of gen-
erality we may assume that k = 1. Let f ∈ Ac (X), ε > 0 and E be given. Using
the continuity of f and the compactness of X we can find a partition {U1 , . . . , Um }
of ext X consisting of nonempty sets such that diam f (Uj ) < ε, j = 1, . . . , m. We
select points xj ∈ Uj and set φj := cUj , αj := f (xj ), j = 1, . . . , m. Then
m
X
f (x) − αj φj (x) ≤ ε, x ∈ ext X.
j=1
Definition 12.41 (L∞ -spaces). We recall that a Banach space E is termed an L∞,λ -
space if for any finite-dimensional space A ⊂ E there exists a finite-dimensional
subspace F ⊃ A of E whose Banach–Mazur distance from some `∞ n is smaller or
equal than λ where n = dim F (that is, for every ε > 0 there exists an isomorphism
T : F → `∞ n such that kT kkT
−1 k < λ + ε).
Corollary 12.42. If X is a simplex, then Ac (X) is an L∞,1+ε -space for every ε > 0.
It follows that
1
kS −1 k ≤ ε0
. (12.27)
1− n
Since G has dimension n, Theorem 5.6(ii) in [173] yields the existence of a projec-
tion P : F 0 → G with kP k ≤ n. We set H := ker P and
F := E ⊕ H.
T f := Se + h, f = e + h ∈ E ⊕ H,
satisfies by (12.27)
ε0
kf k − kT f k ≤ kf − T f k = kSe − ek ≤ kek
n
ε0 −1 ε0 1
≤ kS kkSek ≤ ε0
kP (Se + h)k
n n1− n
n
≤ ε0 kT f k.
n − ε0
By our choice of ε0 , kf k − kT f k ≤ εkT f k and thus
kT k ≤ (1 − ε)−1 and kT −1 k ≤ 1 + ε.
Remark 12.43. It is known that a Banach space E is an L∞,1+ε -space for every ε > 0
if and only if E ∗ is isometric to L1 (X, Σ, µ) for a suitable measure space (X, Σ, µ)
(see p. 59 in [251]). Hence Corollary 12.42 provides another proof of Theorem 6.25.
Sn := {e∗ ∈ E ∗ : e∗ ≥ 0, e∗ (1) = 1}
Proof. We start the proof by using Lemma 12.39(b) to find a peaked partition of unity
{f1 , . . . , fn } ⊂ Ac (X) such that E = span{f1 , . . . , fn }. Let ψ : Sn → Rn be
defined as
ψ(e∗ ) := (e∗ (f1 ), . . . , e∗ (fn )), e∗ ∈ Sn .
Then ψ is an affine homeomorphism of Sn to Rn . By Lemma 12.39(a) it is enough to
show that ψ(Sn ) = ϕ(X) (ϕ is defined as in Lemma 12.39(a)). Let φ : X → Sn be
the evaluation mapping defined as φ(x)(f ) := f (x), f ∈ E, x ∈ X.
Obviously, for any e∗ ∈ Sn , ψ(e∗ ) has positive coordinates whose sum equals
1, hence ψ(Sn ) ⊂ ϕ(X). Conversely, if xi is a point of X with fi (xi ) = 1, then
ψ(φ(xi )) = ei . Hence ψ(Sn ) ⊃ ext ϕ(X), and the conclusion follows.
Theorem 12.45. Let X be a metrizable simplex. Then the following assertions hold.
(a) There exists an increasing family {En } of finite-dimensional subspaces of Ac (X)
∞
such that E1 = span{1}, En is isometric to `∞ c
S
n and A (X) = n=1 En .
(b) There exists an inverse sequence (Xn , pnm )n,m∈N such that each Xn is an (n−1)-
simplex and X = lim Xn .
←
Proof. For the proof of (a), let {fk : k ∈ N} be a dense subset of BAc (X) . We induc-
tively use Theorem 12.40 to find an increasing sequence {nk } of natural numbers and
spaces Enk ⊂ Ac (X) such that
• n = 1 and E
1 n1 = span{1},
• Enk is isometric to `∞
nk , k ∈ N,
• Enk ⊂ Enk+1 , k ∈ N,
• dist(fi , Enk ) < 2−k , i = 1, . . . , k, k ∈ N.
By Lemma 12.39(a),(c), for each k ∈ N we can find spaces
Corollary 12.46. For any metrizable simplex X,Sthere exists an increasing sequence
{Fn } of finite-dimensional faces of X such that ∞
n=1 Fn is dense in X.
fi (xi ) = 1, i = 1, . . . , n.
Without loss of generality we may assume that all points xi are extreme. Then the set
Fn := co{x1 , . . . , xn }
is a finite-dimensional face of X.
We claim that {Fn } is the required sequence. Indeed, set F := ∞
S
n=1 Fn and
assume the existence of
∞
[
x∈X\ Fn .
n=1
Using the Hahn–Banach theorem we find a function f ∈ Ac (X) and δ > 0 such that
0 < f < 1 and
k(a1 , . . . , an )k`∞
n
= kgk. (12.28)
Since
1
g ≤ δ + m on F,
4
for any i = 1, . . . , n we get
1
0 ≤ ai = g(xi ) ≤ m + δ.
4
12.3 Several examples 455
1 1
m + δ < f (x) ≤ δ + g(x) ≤ m + δ,
4 2
a contradiction. This finishes the proof.
Theorem 12.47. Every simplex is the inverse limit of an inverse system of metrizable
simplices.
Proof. If X is a simplex, let E denote the family of all norm separable closed sub-
spaces of Ac (X) that contain 1 and possess the weak Riesz interpolation property.
By Theorem 9.12, this family is up-directed, that is, for any E1 , E2 ∈ E there exists
E ∈ E with E1 ∪ E2 ⊂ E. Moreover, {E : E ∈ E} is dense in Ac (X).
S
is an affine homeomorphism.
Proof. We start the proof by noticing that z is an extreme point of co(X ∪ {z}).
Further, any point x in co(X ∪{z}) can be uniquely decomposed as x = αy+(1−α)z
for some y ∈ X and α ∈ [0, 1]. Hence there exists a unique maximal measure
µ ∈ Mx (co(X ∪ {z})), namely µ = αδy + (1 − α)εz , where δy is the unique
maximal measure for y on X. Thus co(X ∪ {z}) is a simplex and ext co(X ∪ {z}) =
ext X ∪ {z}.
Lemma 12.49. Let X be a simplex in a Banach space E. Then there exists a simplex
Y ⊂ E ⊕ `2 such that X is a closed face of Y and ext Y is dense in Y .
456 12 Constructions of function spaces
Proof. Let X be a simplex in a Banach space E and let {en : n ∈ N} be the canonical
basis of `2 . We identify x ∈ E with (x, 0) ∈ E ⊕ `2 and y ∈ `2 with (0, y) ∈ E ⊕ `2 .
Let En := span{ek : k = 1, . . . , n} ⊂ `2 and let πn : E ⊕ `2 → E ⊕ En be the
projection.
Let Y0 := X. We construct inductively simplices Yn ⊂ E ⊕ `2 , points zn ∈ E ⊕ `2
and εn > 0, n ∈ N, such that
(a) Yn = co(Yn−1 ∪ {zn }), n ∈ N,
(b) Yn ⊂ E ⊕ En , n ∈ N,
(c) ext Yn ⊂ ext Ym , n ≤ m,
(d) πn (Ym ) = Yn and kπn (x) − xk ≤ 2−n+1 , x ∈ Ym , n ≤ m,
(e) for each k ∈ N there exists n ≥ k such that dist(y, ext Yn ) < εk , y ∈ Yk , and
εn+1 < 21 εk ,
(f) εn+1 ≤ εn , n ∈ N.
To start the construction, we choose y1 ∈ Y0 and set
ε1 := 2−1 ,
z1 := y1 + 2−1 e1 ,
Y1 := co(Y0 ∪ {z1 }).
Assume now that the construction has been completed up to the n–th stage. We denote
Now we distinguish two cases. If Mn (εn ) 6= ∅, we choose yn+1 ∈ Mn (εn ) and set
εn+1 = εn . Otherwise we find εn+1 ∈ (0, 12 εn ) such that Mn (εn+1 ) 6= ∅ and choose
yn+1 ∈ Mn (εn+1 ). In both cases we set
This finishes the construction. Properties (a)–(d) are obviously satisfied. Given k ∈ N,
we find m ≥ k such that 2−m+3 < εk and notice that by (d) and the construction,
By the compactness of Ym we observe that infi,j≥m kπm (zj ) − πm (zi )k = 0, and thus
there must exist j ∈ N with εj < εk . Hence we find n ≥ k such that the (n + 1)-th
step of the construction follows the second case, which verifies property (e).
Let Y := ∞
S
n=1 Yn . Then Y is convex and compact (it follows from (d)).
12.3 Several examples 457
Further, ext Yn ⊂ ext Y . Indeed, let x ∈ ext Yn equal αy1 + (1 − α)y2 for some
y1 , y2 ∈ Y . Then
V := co{αy + (1 − α)ei : i = 0, . . . , n}
ϕ(x) := π −1 (π(x)) ∩ F, x ∈ X.
458 12 Constructions of function spaces
Then ϕ is an affine mapping with nonempty closed convex values. Since the set
co{x1 , . . . , xn } is an (n − 1)-simplex (see Lemma 6.68 and Exercise 6.85), the map-
ping π : F → co{x1 , . . . , xn } is open by Lemma 12.51. Hence for any open set
U ⊂ X,
ei ◦ g = ei ◦ π ◦ g = ei ◦ π = ei , i = 1, . . . , n.
Lemma 12.53. Let F be a closed proper face of a metrizable simplex X with ext X =
X. Let {e1 , . . . , en } ⊂ Ac (X) be a peaked partition of unity, {f1 , . . . , fn+1 } ⊂
Ac (F ) be a partition of unity, and let {a1 , . . . , an } be a set of positive numbers such
that
ei |F = fi + ai fn+1 , i = 1, . . . , n.
Then, for each ε > 0, there exists a peaked partition of unity {g1 , . . . , gn+1 } ⊂
Ac (X) such that
• g |
i F = fi , i = 1, . . . , n + 1,
• kei − (gi + ai gn+1 )k < ε, i = 1, . . . , n.
Proof. Let xi ∈ ext X be such that ei (xi ) = 1, i = 1, . . . , n. We find a point
yn+1 ∈ ext X \ F such that
n
X
ei (yn+1 ) − ei aj xj < ε, i = 1, . . . , n. (12.29)
j=1
Let
H := co(F ∪ {y1 , . . . , yn+1 }).
Then H is a closed face of X and by Lemma 12.52 there exists a continuous affine
retraction ϕ : X → H such that ei ◦ ϕ = ei , i = 1, . . . , n. For i = 1, . . . , n + 1 we
find gei ∈ Ac (H) defined by conditions gei = fi on F and
(
0, i 6= j,
gei (yj ) = j = 1, . . . , n + 1,
1, i = j,
12.3 Several examples 459
gi := gei ◦ ϕ, i = 1, . . . , n.
(Use (12.30) for j = 1, . . . , n and (12.29) for j = n + 1.) This finishes the proof.
{a1,m , . . . , am,m }, m ∈ N,
such that
Pm
(1) i=1 ai,m = 1, m ∈ N,
(5) If Em := span{em m
1,m , . . . , em,m }, then for each n ∈ N there exists m ∈ N such
that
dist(xk , Em ) < 2−n , 1 ≤ k ≤ n.
460 12 Constructions of function spaces
{a1,m , . . . , am,m }, p ≤ m ≤ p + r − 1,
such that
Pm
i=1 ai,m = 1, p ≤ m ≤ p + r − 1,
•
p
• ei,p = gi,p+1 + ai,p gp+1,p+1 , 1 ≤ i ≤ p,
• gi,m = gi,m+1 + ai,m gm+1,m+1 , 1 ≤ i ≤ m, p + 1 ≤ m ≤ p + r − 1.
For every 1 ≤ m ≤ p + r, p + 1 ≤ j ≤ p + r and 1 ≤ i ≤ m we set
(
epi,m , 1 ≤ m ≤ p,
eji,m :=
gi,m , p + 1 ≤ m ≤ p + r.
Then (1), (2) and (4) hold for the constructed functions eji,m and (5) holds for n + 1.
We now transfer the constructed objects to Ac (X2 ). Since
{ep+1 p+1
1,p+1 ◦ ϕ, . . . , ep+1,p+1 ◦ ϕ}
Lemma 12.53 provides for any ε > 0 a peaked partition of unity {h1,p+1 , . . . ,
hp+1,p+1 } ⊂ Ac (X2 ) such that
• hi,p+1 (t) = ep+1
i,p+1 (ϕ(t)), t ∈ H2 , 1 ≤ i ≤ p + 1,
12.3 Several examples 461
p
• kfi,p − (hi,p+1 + ai,p hp+1,p+1 )k < ε, 1 ≤ i ≤ p.
We take ε < 2−2(p+1) and define
p+1
fi,p+1 := hi,p+1 , 1 ≤ i ≤ p + 1,
p+1 p+1 p+1
fi,p := fi,p+1 + ai,p fp+1,p+1 , 1 ≤ i ≤ p,
..
.
p+1 p+1 p+1
fi,l := fi,l+1 + ai,l fl+1,l+1 , 1 ≤ i ≤ l,
..
.
p+1 p+1 p+1
f1,1 := f1,2 + a1,1 f2,2 .
Then
p+1 p
kfi,k − fi,k k < 2−p , 1 ≤ k ≤ p, 1 ≤ i ≤ k.
Similarly we construct
m m m m m
{f1,m , . . . , fm,m }, . . . , {f1,2 , f2,2 }, {f1,1 }, p + 2 ≤ m ≤ p + r.
and by (3)
ei,m (ϕ(t)) = fi,m (t), t ∈ H2 , m ∈ N, 1 ≤ i ≤ m. (12.32)
462 12 Constructions of function spaces
Hence
{e1,m , . . . , em,m } and {f1,m , . . . , fm,m }, m ∈ N,
are peaked partitions of unity, in particular, they are linearly independent.
It follows from (12.31) that the equalities
T ei,m = fi,m , 1 ≤ i ≤ m, m ∈ N,
the mapping
φ−1 ∗
1 ◦ T ◦ φ2
Proof. Given the objects as in the hypothesis, we find a function g ∈ Ac (X) such that
0 ≤ g < 21 and δ := g(x1 ) − g(x2 ) > 0. Then
k := cF ∨ g
Since µ is maximal, Lemma 8.13 and Theorem 8.39 yield µ(F 0 ) = 1. Hence
Z
∗
k (z) = µ(k) = k(t) dµ(t) = µ(g) = g(z).
F0
k ≤ a ≤ 1 ∧ h.
g1 ◦ ϕ ≤ f ≤ g2 ◦ ϕ on X.
Assume first that c∗ϕ(F ) (t) = 0, that is, t ∈ (ϕ(F ))0 . Let x ∈ X be such that ϕ(x) = t.
For any ε > 0 there exists a function g ∈ Ac (Y ) such that g ≥ cϕ(F ) and g(t) < ε.
Thus f := g ◦ ϕ ∈ Ac (X) satisfies f ≥ cF and f (x) < ε. Hence c∗F (x) = 0 and
x ∈ F 0 . Thus t ∈ ϕ(F 0 ).
Conversely, let x ∈ F 0 be given. For any ε > 0 we use Theorem 6.6 to find a
function f ∈ Ac (X) such that f = 1 on F and f (x) < ε. For the function g ∈ Ac (Y )
satisfying f = g ◦ ϕ we get g ≥ cϕ(F ) and g(ϕ(x)) < ε. Thus
c∗ϕ(F ) (ϕ(x)) = 0
Further, F as a closed proper face of X has empty interior (see Lemma 2.90), and
thus
ext X ∩ F 0 = ext F 0
is dense in X. Hence ext ϕ(F 0 ) is dense in Y and Y is the Poulsen simplex. This
finishes the proof.
Theorem 12.59. Let Fi , i = 1, 2, be closed faces of the Poulsen simplex and let Fi0 be
their complementary faces, i = 1, 2. Then F10 is affinely homeomorphic to F20 .
Thus ϕ−1 0 0
2 ◦ ϕ ◦ ϕ1 is an affine homeomorphism of F1 onto F2 .
Theorem 12.60 (Properties of the Poulsen simplex). Let S stand for the Poulsen sim-
plex.
(a) Any metrizable simplex is affinely homeomorphic to a face of S.
(b) Let X be a metrizable simplex with the following properties:
• Any metrizable simplex is affinely homeomorphic to a face of X.
• If F1 , F2 are faces of X with dim F1 = dim F2 < ∞, then there exists an
affine homeomorphism ϕ of X onto itself such that ϕ(F1 ) = F2 .
Then X is affinely homeomorphic to S.
12.3 Several examples 465
(c) If K ⊂ ext S is a compact set, then its interior (relative to ext S) is empty.
(d) Every Polish space is homeomorphic to a closed subset of ext S.
(e) If K1 , K2 are homeomorphic compact subsets of ext S, then there exists an affine
homeomorphism ϕ of S onto itself such that ϕ(K1 ) = K2 .
(f) The set ext S is arcwise connected (that is, for any points x1 , x2 ∈ ext S there
exists a homeomorphic copy of [0, 1] contained in ext S that joins x1 with x2 ).
Proof. To prove (a), let X be a metrizable simplex. By Proposition 2.45 we may as-
sume that X is a norm compact subset of `2 . By Lemma 12.49, there exists a compact
simplex Y ⊂ `2 ⊕ `2 such that X is a closed face of Y . Due to Corollary 12.55, Y is
affinely homeomorphic to S.
To verify (b), let X possess the prescribed property. By the assumption, X contains
a face X0 affinely homeomorphic to the Poulsen simplex S. If F ⊂ X is an arbitrary
finite-dimensional closed face, there exist a closed face of the same dimension F0 ⊂
X0 and an affine homeomorphism ϕ : X → X such that ϕ(F ) = F0 . Then
Let
disjoint union of open sets Lx , x ∈ ChH1 (K1 ), there exists x ∈ ChH1 (K1 ) such that
µ(Lx ) > 0. Then
p] µ({x}) = µ(Lx ) > 0,
a contradiction with the assumption.
To prove (d), let µ ∈ M1 (K1 ) be H1 -maximal and continuous. Let ν ∈ M1 (K2 )
be any measure with µ ≺H2 ν. By (b) and Proposition 12.24(c), µ ≺H1 p] ν. Since
µ is H1 -maximal, µ = p] ν. By the continuity of µ and (c), ν is carried by K1 , and
hence µ = ν.
1
P∞is to prove (e). Let µ ∈ M (K1 ) be a discrete H1 -maximal
The next step P∞ measure,
that is, µ = n=1 an εxn , where the numbers an are strictly positive,P∞ n=1 an = 1
and xn are distinct points in ChH1 (K1 ). We claim that ν := n=1 an µxn is the
unique H2 -maximal measure with µ − ν ∈ H⊥ 2 .
To see this, let ν 0 ∈ M1 (K2 ) be an H2 -maximal measure with µ − ν 0 ∈ H⊥ 2 . Then
p] ν 0 − µ ∈ H ⊥1 and p] ν 0 is H -maximal (see Proposition 12.25). Thus p ν 0 = µ by
1 ]
Proposition 6.9. From this and from the fact that ν 0 does not charge anyPpoint in K1
(by Construction 12.61 and H2 -maximality of ν 0 ), we obtain that ν 0 = ∞ n=1 an λxn
for some probability measures λxn on Lxn . Fix n ∈ N. If g ∈ C([0, 1]) is a continuous
R1
function with 0 g = 0, then
(
g on Lxn ,
h :=
0 otherwise,
is in H2 and ν 0 (h) = µ(h) = 0. Hence λxn is a multiple of µxn . Since they are both
probability measures, λxn = µxn . Thus ν 0 = ν.
It follows that any point x ∈ K1 has a unique H2 -maximal measure H2 -repre-
senting x. Indeed, let δx bePthe unique H1 -maximal measure H1 -representing x. We
decompose δx = (δx )c + P∞ n=1 an εxn , where (δx )c is the continuous part of δx , the
numbers an are positive, ∞ n=1 an ≤ 1 and the points xn are distinct elements of
ChH1 (K1 ) (see Proposition A.78). Then we get that
∞
X
δ := (δx )c + a n µ xn
n=1
is the unique H2 -maximal measure H2 -representing x.
To see this, let µ be an H2 -maximal measure with µ − δ ∈ H⊥
2 . Then p] µ is
H1 -maximal and p] µ − δx ∈ H⊥ 1 . Thus
∞
X
(δx )c + an εxn = p] µ = µ|K1 + p] (µK2 \K1 ).
n=1
By (c), p] (µ|K2 \K1 ) is discrete. Hence µ|K1 = (δx )c . Since
∞
X
an µxn − µ|K2 \K1 ∈ H⊥
2,
n=1
468 12 Constructions of function spaces
it follows that
∞
X
an εxn − µ|K2 \K1 ∈ H⊥
2.
n=1
P∞
Hence µ|K2 \K1 = n=1 an µxn by the preceding argument. Thus µ = δ.
Hence H2 is simplicial, which finishes the proof of (e).
Finally, given x ∈ K1 , we repeat the argument from the proof of (e) to deduce that
the unique H2 -maximal measure H2 -representing x is continuous. This concludes the
proof.
pβα := πβ , β < α.
For any β < α, let iβα : Kβ → Kα be the natural homeomorphic embedding, that is,
Lemma 12.64. Let (K, H) := (Kω1 , Hω1 ). Then the following assertions hold.
(a) The function space Hα is simplicial and Hα = Ac (Hα ), α ∈ [0, ω1 ].
(b) If x ∈ iα (Kα ), α < ω1 , then its unique H-maximal measure is carried by
iα+1 (Kα+1 ).
(c) For any continuous measure µ ∈ M1 (K), there exists ordinal α < ω1 such that
µ(iα (Kα )) = 1.
12.3 Several examples 469
S
(d) ChH (K) = K \ α<ω1 iα (Kα ).
= Ac (Hα )
Theorem 12.65. There exists a nonzero strongly affine function g on a Choquet sim-
plex such that g = 0 on ext X.
Consequently, the class of strongly affine functions is not an affinely perfect class
(see Section 5.5).
Proof. Consider the function space (K, H) from Construction 12.63 and the function
f from Lemma 12.64(e). Then X := S(H) is a simplex (see Theorem 6.54) and
ext X = φ(ChH (K)) (φ is the embedding from Definition 4.25). If I is the mapping
from Corollary 5.41, g := If is a nonzero strongly affine function on X that equals
zero on ext X. This concludes the proof.
Lemma 12.67. If a Banach space E has the Schur property and {xn } is a weakly
Cauchy sequence in E (that is, {x∗ (xn )} is Cauchy for each x∗ ∈ E ∗ ), then it is a
norm Cauchy sequence.
Proof. Let {xn } be a weakly Cauchy sequence in E. Assuming that it is not norm
Cauchy, we can find ε > 0 and an increasing sequence {nk } of natural numbers such
that kxnk+1 − xnk k > ε for k ∈ N. Then yk := xnk+1 − xnk , k ∈ N, form a sequence
converging weakly to 0. By the Schur property, kyk k → 0, which is a contradiction.
This finishes the proof.
Corollary 12.68. Let E be a Banach space with the Schur property and X := BE ∗
be considered with the w∗ -topology. Then Aα (X) = Ac (X) for each α < ω1 .
Proof. It follows from Theorem A.7 (see also Proposition 4.36) that any continuous
affine function on X is of the form x + c, where x ∈ E and c ∈ R. Hence if f is
any function in A1 (X), then there exist sequences {xn } in E and {cn } in R such that
f (x∗ ) = limn→∞ xn (x∗ ) + cn , x∗ ∈ X. Then cn → f (0) and {xn } is a weakly
Cauchy sequence. Hence, by Lemma 12.67, {xn } converges in norm S to an element
x ∈ E. Thus f = x + f (0) and f ∈ Ac (X). It follows that α<ω1 Aα (X) =
Ac (X).
I := {(n, p) ∈ N × N : p ≤ n},
and (
1, i = (n, p),
en,p (i) := (n, p) ∈ I.
0, i ∈ I \ {(n, p)},
Let F := span{en,p : (n, p) ∈ I}. For (n, p) ∈ I, let e∗n,p : F → R be the linear
functional defined by
(
∗ 1, (n0 , p0 ) = (n, p),
en,p (en0 ,p0 ) :=
0, (n0 , p0 ) 6= (n, p).
We define a norm k · k1 on F by
∞
X
kyk1 := sup |e∗n,p (y)|, y ∈ F.
n=1 1≤p≤n
Let Σ stand for the set of all increasing sequences of natural numbers (that is,
σ(n) ≤ σ(m) if n ≤ m). For each σ ∈ Σ we define a linear functional gσ : F → R
by setting (
1, σ(n) ≤ p,
gσ (en,p ) := (n, p) ∈ I,
0, σ(n) > p,
and let
kyk2 := sup{|gσ (y)| : σ ∈ Σ}, y ∈ F.
Finally,
kyk := kyk1 ∨ kyk2 , y ∈ F.
Let E1 , E2 , E be the completions of F with respect to norms k · k1 , k · k2 and k · k,
respectively. The mapping
y 7→ (y, y), y ∈ E,
472 12 Constructions of function spaces
Lemma 12.70. Let fn,p := (0, en,p ), (n, p) ∈ I, be considered as elements of G. For
any x∗ ∈ G∗ , the limit limp→∞ limn→∞ x∗ (fn,p ) exists.
Proof. We first show that the set S := {gσ : σ ∈ Σ} is w∗ -compact in E2∗ . To this
end, let σ j ∈ Σ, j ∈ N, be such that {gσj } w∗ -converges to g ∈ E2∗ . Then g ∈ BE2∗
and g(en,p ) ∈ {0, 1}, (n, p) ∈ I. For n ∈ N, we set
(
min{p : p ≤ n, g(en,p ) = 1}, if such p exists,
σ(n) :=
n+1 otherwise.
w∗
By the Milman theorem 2.43 and the w∗ -compactness of S, ext BE2∗ ⊂ S ∪ −S.
It follows from Theorem 2.31 and the Lebesgue dominated convergence theorem that
limp→∞ limn→∞ x∗ (en,p ) exists for each x∗ ∈ E2∗ . Hence limp→∞ limn→∞ x∗ (fn,p )
exists for each x∗ ∈ G∗ , which concludes the proof.
Proof. It is enough to show that x∗∗ is not w∗ -continuous on BG∗ . As in the proof of
Lemma 12.70 we obtain that
1, lim σ(n) < ∞,
x∗∗ (0, gσ ) = n→∞
0, lim σ(n) = ∞.
n→∞
isw∗ -dense in {(0, gσ ) : σ ∈ Σ} (notice that gσ∧n → gσ in the w∗ -topology for any
σ ∈ Σ), x∗∗ is not w∗ -continuous.
Then
x∗ (fbn,k ) = x∗ (b
gn,k ), x∗ ∈ E2∗ ,
x∗ (fbn,k ) = 0, x∗ ∈ E1∗ ,
k
1 ∗ X 1
|x∗ (b en,p ≤ kx∗ kE1∗ , x∗ ∈ E1∗ .
gn,k )| ≤ x
k k
p=1
Since G∗ = E1∗ × E2∗ with the norm k(x∗1 , x∗2 )k = kx∗1 kE1∗ + kx∗2 kE2∗ , we get
1
kfbn,k − gbn,k k ≤ .
k
474 12 Constructions of function spaces
y ∗∗ (y ∗ ) = x∗∗ (x∗ ), x∗ ∈ π −1 (y ∗ ), y ∗ ∈ BE ∗ .
Lemma 12.75. The function y ∗∗ is a discontinuous strongly affine function of the sec-
ond Baire class on BE ∗ .
Proof. Since x∗∗ : BG∗ → R is a strongly affine Baire-two function and π : BG∗ →
BE ∗ is a perfect mapping, y ∗∗ inherits both these properties by Corollary 5.27 and
Proposition 5.29. If y ∗∗ were continuous, x∗∗ would be continuous on BG∗ , and thus
x∗∗ would be in G. But this contradicts Lemma 12.72.
0
We find an element yj+1 ∈ F and kj+1 > kj such that
0
yj+1 ∈ span{en,p : kj ≤ n < kj+1 }
0
and kyj+1 − ym k < ηj+1 .
12.3 Several examples 475
After finishing this construction, we get the required sequence by normalizing the
vectors yj0 , j ∈ N, and denoting them again as {yj }.
Step 2. Assume that kyj k1 → 0.
Then kyj k2 → 1, without loss of generality kyj k2 > 2−1 for all j ∈ N. For each
j ∈ N we find σ j ∈ Σ such that
σ j (n) ≤ n + 1, n ∈ N, (12.34)
Hence we may achieve by choosing another subsequence (still denoted as {yj }) that
kj
∞ X
X 1
|e∗n,p (yj 0 )| < , j ∈ N, j 0 ≥ j. (12.35)
4
n=kj p=1
1
gσ (yj ) ≥ , j ∈ N.
4
Since gσ ∈ E ∗ , this is a contradiction with the requirement that {yj } weakly converge
to 0.
Step 3. Assume that there exists c > 0 such that kyj k1 > c for infinitely many
indices j. We will prove that {yj } contains a subsequence that is equivalent to `1 -
basis (see Definition 10.38), which will yield a contradiction due to Theorem A.9.
476 12 Constructions of function spaces
Without loss of generality, let kyj k1 > c for all j ∈ N. Let c1 , . . . , cm be real
numbers. Since
m
X
e∗n,p cj yj = e∗n,p ci yi ,
ki ≤ n < ki+1 ,
j=1
we obtain
m
X m
X ∞
X m
X
e∗n,p
cj yj ≥ cj yj 1
= sup cj yj
j=1 j=1 n=1 1≤p≤n j=1
X−1
m kj+1
X
= sup |e∗n,p (cj yj )|
j=1 n=kj 1≤p≤n
m kj+1 −1
X X
= |cj | sup |e∗n,p (yj )|
j=1 n=kj 1≤p≤n
m
X
≥c |cj |.
j=1
This means that {yj } is equivalent to `1 -basis, which concludes the proof.
Theorem 12.77. There exists a metrizable compact convex set X and a strongly affine
function f ∈ B 2 (X) such that
S
• f ∈
/ α<ω1 Aα (X),
• Ac (X) =
S
α<ω1 Aα (X).
Proof. Let E be the Banach space from Construction 12.69 and let X := BE ∗ be
endowed with the w∗ -topology. By Lemma 12.76 and Corollary 12.68, Ac (X) =
∗∗ be the function from Definition 12.74. By Lemma 12.75, y ∗∗
S
α<ω1 Aα (X). Let y
is a strongly affine function contained in B 2 (X) that is not continuous. This finishes
the proof.
S 12.78. The construction leading to Theorem 12.77 shows that neither A2 (X)
Remark
nor α<ω1 Aα (X) is an affinely perfect class of functions. Indeed, the mapping π :
BG∗ → ∗∗ ∗∗ ∗∗
S BE is an affine continuous surjection, x = y ◦ π, x ∈ A2 (BG ) and
∗ ∗
∗∗
y ∈ / α<ω1 Aα (BE ∗ ).
12.4 Exercises 477
12.4 Exercises
N
Exercise 12.79. Let (Ki , Hi ), i ∈ I, be Bauer simplicial spaces. Prove that i∈I Hi
is a Bauer simplicial space.
Hint. Use Theorem 12.10 and Theorem 12.21.
1 1
((x1 , x2 ) + (y1 , y2 )) = ((x1 , y2 ) + (y1 , x2 )) ∈ F,
2 2
Q
and thus (x1 , x2 ) ∈ F . By induction it follows that F = i∈I πi (F ) whenever I is
finite.
O
g ◦ S = f on ext Xi .
i∈I
480 12 Constructions of function spaces
N
Since both functions are continuous and affine, g ◦ S = f on i∈I Xi . Hence
Y
g(T (x)) = f (b(x)) = g(S(b(x))), x ∈ Xi .
i∈I
Q
It follows that T (x) = S(b(x)) for any x ∈ i∈I Xi .
Q
Exercise 12.86.NKeeping the notation from Exercise 12.85, let X := i∈I Xi . Prove
that Ac (X) ⊂ i∈I Ac (Xi ).
Exercise 12.87. We keep the notation from Exercise 12.85. Prove the following as-
sertions:
N
(a) The set i∈I Xi is a simplex if and only if all sets Xi are simplices.
N
(b) Assume that every set Xi is a simplex. Prove that the tensor product i∈I Xi
is a unique compact convex set Y (up to an affine homeomorphism) with the
following properties:
Q
i∈I Xi → Y such that
• There exists a continuous multiaffine mapping m :
Q
Y = co m( i∈I Xi ).
F := Xj × (xi )i∈I\{j}
Now, we use Exercise 6.85 to get that b(F ) is a simplex. Since b is an affine
homeomorphism on F , the set F , and consequently Xj , is a simplex.
N Q
To verify (b), denote Z := i∈I Xi and X := i∈I Xi . Let Y be a compact con-
vex set for which there exists a multiaffine mapping m : X → Y with the properties
We notice that m is injective by the second property of Y . We claim that
listed in (b).Q
ext Y ⊂ m( i∈I ext Xi ).
Indeed, since Y = co Qm(X), ext Y ⊂ m(X) by Theorem 2.43. Now it is easy to
verify that ext Y ⊂ m( i∈I ext Xi ).
Consider the multiaffine mapping b : X → Z from Exercise 12.85. By the as-
sumption there exists a continuous affine mapping S : Y → Z such that S ◦ m = b.
Then Y Y
S(ext Y ) ⊂ S m ext Xi = b ext Xi = ext Z
i∈I i∈I
Hint. If g depends on at most one coordinate, then it clearly satisfies the condition in
(ii).
Conversely, assuming that (ii) holds for g, let x1 , y1 ∈ E1 , x2 , y2 ∈ E2 be fixed.
Using (ii) we get
By symmetry,
g(x1 , y2 ) ∨ g(y1 , x2 ) ≤ g(x1 , x2 ) ∨ g(y1 , y2 ).
482 12 Constructions of function spaces
Hint. Let Ei := ext Xi , i = 1, 2. We claim that, for each f ∈ s(A), the function
g := f |E1 ×E2 satisfies (ii) of Exercise 12.88.
Indeed, for fixed xi , yi ∈ Ei , i = 1, 2, and λ ∈ R we take an affine function
h : X1 × X2 → R with
Then
1 1
(h(x1 , x2 ) + h(y1 , y2 )) = (h(x1 , y2 ) + h(y1 , x2 )),
2 2
and so the claim is proved.
It follows from Exercise 12.88 that g depends on at most one variable. Since f is
affine and continuous, an application of the Krein–Milman theorem 2.22 yields the
assertion.
Q
Exercise 12.91. Let Xi , i ∈ I, be compact spaces and X := i∈I Xi . If f : X → R
is continuous and separately constant, then it is constant.
Exercise 12.92. If X is a compact convex set, we denote by V(X) the family of all
linear subspaces A ⊂ Ac (X) with s(A) = A that contain a nonconstant function and
are maximal with respect to inclusion. If Xi , i ∈ I,Qare compact convex sets, we
consider the spaces Ac (Xi ), i ∈ I, as subspaces of Ac ( i∈I Xi ) as in Exercise 12.90.
If Xi , i ∈ I, are compact convex sets with ext Xi affinely independent, then
Y
V Xi = {Ac (Xi ) : i ∈ I}.
i∈I
Q
Hint. We denote X := i∈I Xi . Given A ∈ V(X), let f ∈ A be a nonconstant
function. Exercise 12.91 provides a coordinateQj ∈ I such that f is nonconstant with
respect to the coordinate j. We denote Y := i∈I\{j} Xi and apply Exercise 12.90
to Xj and Y to get f ∈ Ac (Xj ). It follows also from Exercise 12.90 that any other
function g ∈ A is also contained in Ac (Xj ). Indeed, if g ∈ A \ Ac (Xj ), then g ∈
Ac (Y ) by Exercise 12.90, but
/ Ac (Xj ) ∪ Ac (Y ).
f +g ∈
/ Ac (Xj ) ∪ Ac (Y ),
f +g ∈
Exercise 12.93. Find a metrizable compact convex set X such that ext X is a closed
affinely independent set and X is not a simplex.
Hint. Consider K := [0, 1] ∪ {2} and
Z 1 Z 1
H := f ∈ C(K) : f (2) = 2tf (t) dt = 2(1 − t)f (t) dt .
0 0
Hint. Consider X1 = [0, 1], X2 = [0, 1]3 , Y1 = [0, 1]2 , Y2 = [0, 1]2 .
Subsection 12.1.C follows the paper by A. J. Lazar [293]. The results of Subsec-
tion 12.1.D can be found in E.B. Davies and G. F. Vincent-Smith [132] and A. J. Lazar
[293]. Theorem 12.34 is in F. Jellett [249] and E. B. Davies and G. F. Vincent-Smith
[132], Theorem 12.36 appears in E. B. Davies and G. F. Vincent-Smith [132], F. Jellett
[249] and M. W. Grossman [203].
Subsection 12.2.D follows the techniques of the paper A. J. Lazar and J. Linden-
strauss [296], where they provided an isometric classification of L1 -preduals (see also
J. Lindenstrauss and D. E. Wulbert [308], A. J. Lazar and J. Lindenstrauss [295] and
M. Zippin [479]). Lemma 12.38 is from E. A. Michael and A. Pelczynski [344]),
Corollary 12.46 is mentioned in J. Lindenstrauss, G. Olsen and Y. Sternfeld [307].
12.5 Notes and comments 487
The Poulsen simplex was constructed in E. T. Poulsen [376]. Its properties were
proved in J. Lindenstrauss, G. Olsen and Y. Sternfeld [307] using an idea of W. Lusky
who showed in [326] the uniqueness of the Gurari space (a separable L1 -predual that
contains an isometric copy of every other separable L1 -predual). Besides the proper-
ties of the Poulsen simplex S showed in Subsection 12.3.A, we mention the fact that
ext S is homeomorphic to `2 (see [307], Theorem 3.1). We also refer the reader to the
survey paper by W. Lusky [328] (see also [327]).
The construction of simplices with a dense set of extreme points can be also done in
a nonseparable setting as W. Lusky in [329] showed. He proved that for any simplex
X there exists a simplex S of the same density as X with dense set of extreme points
such that X is affinely homeomorphic to a face of S. Moreover, there is an affine
retraction of S onto X. In [330] he proved that Poulsen-like simplices of higher
densities are not unique, precisely, for every cardinal κ ≥ ℵ0 there exist simplices
S1 , S2 of density κ such that they have dense set of extreme points and they are not
affinely homeomorphic.
The representation of metrizable simplices by means of inverse limits can be re-
worded in terms of representing matrices. Let X := lim Xn , where {Xn } are
←
(n − 1)-simplices. For each n ∈ N, let ϕn : Xn → Xn−1 be the affine sur-
jection pn,n−1 . Inductively we choose vertices e1 , . . . , en of Xn that are mapped
PnXn−1 . The remaining vertex en+1 of Xn is mapped onto a point
onto vertices of
ϕn (en+1 ) = i=1 ai,n ϕn (ei ), where a1,n , . . . , an,n are coefficients of the convex
combination. The matrix
a1,1 a1,2 a1,3 . . .
0 a2,2 a2,3 . . .
A= 0
0 a3,3 . . .
.. .. .. ..
. . . .
The question of “how far” are spaces of affine continuous functions on simplices
from C(K) spaces was studied in Y. Benyamini and J. Lindenstrauss [52], where
they constructed a metrizable simplex X such that Ac (X) is not complemented in
any C(K) space. A result due to W. Lusky [332] asserts that any separable complex
L1 -predual is 1-complemented in a suitable C ∗ -algebra. The question of comple-
mentability of the space H(U ) is studied in J. Spurný [429]; nevertheless, the follow-
ing problem seems to be open.
Problem 12.102. Does there exist a bounded open set U ⊂ Rd such that H(U ) is not
complemented in any C(K) space?
A paper by M. Bačák and J. Spurný [28] shows that the question of comple-
mentability of the space Ac (X) on a simplex X cannot be determined by topological
properties of ext X.
It was proved in M. Zippin [479] that any separable L1 -predual E contains a 1-
complemented isometric copy of c0 . If E has nonseparable dual, then it contains an
isometric copy of C({0, 1}N ) (see A. J. Lazar and J. Lindenstrauss [296]). A paper by
I. Gasparis [192] provides a unified approach to both of these results.
A result of D. A. Edwards [159] says that any metrizable simplex is affinely home-
omorphic to the intersection of a decreasing sequence of Bauer simplices. Unfortu-
nately, the proof has a slight gap which is filled in D. A. Edwards, O. Kalenda and
J. Spurný [160].
Subsection 12.3.B follows the paper by M. Talagrand [446], we only changed the
incorrect Lemma 3 in [446]. (Following the notation of [446], take L := [0, 1] and let
A be the space of affine functions on L. If we construct L0 and A0 as in [446], then
Ac (A0 ) = A0 and any measure carried by L0 \ L is A0 -maximal. Hence the point 21
has many A0 -maximal representing measures and thus A0 does not satisfy the Riesz
interpolation property.) The following problem on strongly affine functions seems to
be open.
Problem 12.103. Let f be a Borel strongly affine function on a compact convex set X
such that f = 0 on ext X. Is f = 0 on X?
We do not know the answer even for functions of the first Borel class (see Sec-
tion 5.3).
Problem 12.104. Let f be a strongly affine function on a compact convex set X such
that f ∈ Bof1 (X, R) and f = 0 on ext X. Is f = 0 on X?
Subsection 12.3.C is taken from M. Talagrand [447].
Exercises 12.80, 12.99 and 12.100 are from M. Kačena [252]. Exercises 12.85,
12.86 and 12.87 can be found in E. Behrends [50] and E. Behrends and G. Wittstock
[49] (see also Z. Semadeni [413]). Exercises 12.88–12.95 originate from the problem
posed in Z. Lipecki [309] and are contained in Z. Lipecki, V. Losert and J. Spurný
[310]. The result of Exercise 12.97 can be found in V. Borovikov [76].
Chapter 13
Function spaces in potential theory and the
Dirichlet problem
One of the most important motivations for the theory of function spaces comes from
potential theory. A typical task of potential theory is to find a harmonic function h
on a bounded open set U ⊂ Rd which continuously extends the given boundary data
f ∈ C(∂U ). This is the Dirichlet problem. Its reformulation in terms of function
spaces is to extend f ∈ C(∂U ) to g ∈ H(U ), where H(U ) is the function space of
all g ∈ C(U ) which are harmonic on U . If U is regular, then there exists a (unique)
continuous extension operator F : C(∂U ) → H(U ). For other than regular sets the
matter is more subtle and leads to interesting developments of potential theory. Here
the methods of Choquet’s integral representation theory find numerous applications.
It is also inspiring to observe the parallel between behavior of classes of harmonic
(superharmonic) functions and classes of affine (concave) functions.
Since basic properties of harmonic functions like minimum principles or conver-
gence properties are also shared by other PDE’s, it is possible to study parallel “po-
tential theories” associated with these equations. It seems quite natural that starting
from 1950’s several unifying axiomatic theories were developed.
The Laplace equation can be investigated by methods which can be generalized to
other elliptic equations and also the studies of the heat equation (and other parabolic
equations) yield foundations for thoughts in spirit of potential theory. We want to
present some results in a generality involving the parabolic potential theory. For this
purpose we give an essential part of our exposition in the general framework of har-
monic spaces, not to double our effort in listing definitions and results. On the other
hand, we believe that without serious difficulties, a reader can follow the “Laplace
equation line” reading the phrase “Y is a Bauer harmonic space” as “Y is a bounded
reference domain in R2 or Rd with d ≥ 3” and interpreting all the notions classically
(in the sense which the adjective “classical” relates to the Laplace equation).
Already the theory of function spaces, one of the major themes of our book, can
serve as some kind of axiomatic approach to potential theory. For application of func-
tion spaces to potential theory, it is specific that not a single function space, but a
whole family of function spaces is considered (namely, the family of spaces H(U )
where U runs through the system of all relatively open subsets of Y ). In correspon-
dence with the tradition of harmonic spaces, as a primitive notion we use the sheaf of
“function spaces” on open sets rather than on their closures.
490 13 Function spaces in potential theory and the Dirichlet problem
Now, we describe briefly the contents of this chapter. In Section 13.1, the Dirich-
let problem, one of the central subject of potential theory, is treated. Solution of the
Dirichlet problem is defined by means of balayage on the complement of the set in
question. This enables us to give a sense to Dirichlet problem for non-open domains.
The characterization of regular boundary points, Theorem 13.5, relates boundary be-
havior of the solutions with thinness. Balayage on the essential base of the comple-
ment gives in Subsection 13.1.A rise to another type of solution, which may be, for
instance, for parabolic potential theory different. In Section 13.2, basic geometric,
potential theoretic as well as topological properties of the set of regular points are
summarized with a special emphasis on classical and parabolic potential theories. We
show the characterization of regular points in terms of barriers (Proposition 13.16),
present examples of regular and irregular behavior for the Laplace equation (Exam-
ples 13.18) and for the heat equation (Examples 13.25). We discuss Borel type of the
set of regular points (Theorem 13.19 and Proposition 13.21) and use Evans’ potential
in Lemma 13.22 to prepare a tool for solution methods (Theorem 13.23).
In Section 13.3, function spaces and function cones related to harmonic and super-
harmonic functions are studied. In order to cover various important situations at one
stroke, relatively compact open sets usually considered are replaced by much more
general relatively compact coanalytic sets. An interplay between functional analysis
and potential theory is illustrated by the characterization of the Choquet boundary
and maximal representing measures of function spaces and function cones studied.
We treat the classical Laplace case and the version for Bauer harmonic spaces sepa-
rately. The “Laplace line” culminates in Theorems 13.33–13.35, whereas the abstract
counterpart, applicable to parabolic potential theory, is Theorem 13.41. In Theorem
13.50 we show that H(U )-concavity forces continuity on U , which illustrates that
there is a surprising difference between superharmonicity and H(U )-concavity. The-
orem 13.55 shows that separation properties in convex analysis cannot be extrapolated
to potential theory.
In Section 13.4, various solution methods for the Dirichlet problem are described.
The Perron–Wiener–Brelot method defined by means hyperharmonic upper functions
is recalled in Subsection 13.4.A, in particular, we present Brelot’s resolutivity theorem
13.61. Cornea’s method based on the notion of controlled convergence (not discussed
in a book form before) is treated in Subsection 13.4.B. In Theorem 13.80 we show
that C-resolutivity of continuous functions implies C-resolutivity of resolutive (in
PWB sense) functions. In the context of classical potential theory it makes it possible
to produce the Perron–Wiener–Brelot solution of the generalized Dirichlet problem
using harmonic upper functions only. The method brings a new perspective for under-
standing resolutive functions. Various versions of the Wiener type solution are studied
in Subsection 13.4.C. The classical Wiener result is described in Theorem 13.85 in
combination with Proposition 13.86. The Wiener method has its limitations, Bauer’s
example (Example 13.87) shows the failure of regular exhaustions in parabolic po-
13.1 Balayage and the Dirichlet problem 491
tential theory. Fortunately, some variants of the Wiener method are available also for
the general case. Existence of fine regular exhaustion is shown in Corollary 13.96. In
Subsection 13.4.E, a partial differential equations approach to the Dirichlet problem,
based on weak solutions in Sobolev spaces, is explained and its relation to Perron–
Wiener–Brelot solution is shown. Theorem 13.99 shows an existence and uniqueness
of the solution of the Dirichlet problem with continuous boundary data treated in the
sense of traces.
Special different solution methods known for construction of a solution of the
Dirichlet problem lead to the following important question: What is a reasonable
generalization of the classical Dirichlet problem and to what extent it is uniquely de-
termined? This uniqueness question, not previously treated in standard textbooks
devoted to potential theory, has to do with an operator extension approach to the
Dirichlet problem using the notion of a Keldysh operator. The clue to the solution
of the problem when there is exactly one Keldysh operator is shown to be intimately
connected with the Choquet theory. This approach opens a way to understanding why
uniqueness holds for Laplace equation case (or for other elliptic situations) and does
not hold for the parabolic potential theory. Our presentation in Section 13.5 starts with
the fundamental Keldysh lemma on exposing function (Proposition 13.103). Then,
abstract tools in lattice setting are developed in Subsection 13.5.A. The uniqueness
result for the Laplace equation is Theorem 13.115 in Subsection 13.5.B. Theorem
13.128, which clarifies the situation in harmonic spaces, is presented in Subsection
13.5.C.
Throughout this chapter, Y will be the “entire space” on which our potential-theo-
retic thoughts will be developed. This may be the Euclidean space Rd , d ≥ 3, for the
Laplace equation, or a bounded open subset of R2 for the two-dimensional Laplace
equation (actually, what we need is the existence of the Green function on Y ). For
application to the heat equation, Y will be the space Rd+1 . Finally, Y may be a general
Bauer harmonic space. We always alert (mostly at the beginning of a subsection) if
the corresponding exposition concerns only a specific potential theory (for example,
the Laplace equation).
Our exposition follows terminology, notation and material presented in Appendix,
Sections A.6–A.8.
x 7→ µU
x, x ∈ U,
where µU
x is defined by
µU
x (f ) = Hf (x), f ∈ C(∂U ).
The balayage theory offers a possibility to generalize the Dirichlet problem to other
c
than regular sets. Namely, by Corollary A.193, µVx = εVx if V is regular and x ∈ V .
V c
However, the kernel x 7→ εx is well defined not only for irregular open set, but even
for sets not being open at all.
Definition 13.1 (Solution of the Dirichlet problem). Given a relatively compact set U
in Y , we abbreviate
Uc
µUx := εx , x ∈ U.
We omit the superscript U if the set U is fixed.
Suppose that U ⊂ Y is a relatively compact nonempty set. From Theorem A.198
we know that ∂U 6= ∅. Let f : ∂U → R be a function. If the function
Z
x 7→ f dµx , x ∈ U,
∂U
is well defined and finite, it is called the solution of the Dirichlet problem on U with
boundary data f and denoted by Hf . If the set U on which we solve the Dirichlet
problem is not clear from the context, we use the notation H U f . Notice that the
measure µx is a probability measure carried by ∂U according to Theorem A.198.
Since the solution is defined by means of a positive kernel, the operator H is positive.
Note that for p ∈ P we have
c c
b U = RU
Hp = R on U.
p p
The first equality is just the result of our definitions whereas the second one follows
from Theorem A.195.
If the set U is open, then the solution Hf of the Dirichlet problem is harmonic
whenever it is well defined. We will prove it later (Theorem 13.66; see also H. Bauer
[40], Korollar 4.2.1). For the time being, we only note that Hf is harmonic whenever
f ∈ C(∂U ); namely, this follows directly from the density of P − P in C(∂U ).
The next proposition shows that Hf coincides with the classical solution of the
Dirichlet problem whenever the latter exists, even if U is irregular.
13.1 Balayage and the Dirichlet problem 493
Proposition 13.2. Let U ⊂ Y be relatively compact and open and h ∈ H(U ). Then
h = Hh on U .
Here, the first equality follows form Theorem A.195, for the second equality we used
(A.6) in Definition A.190, and the fact that p is continuous and thus
c
lim RpU (x) = lim p(x) = p(z).
x→z, x∈U
/ x→z, x∈U
/
Remark 13.6. It follows from Theorem 13.5 and Remark A.204 that the concept of a
regular point is local. This means that, given z ∈ ∂U and U 0 ⊂ Y relatively compact
such that U ∩ V = U 0 ∩ V for some neighborhood V of z, then z is regular for U if
and only if z is regular for U 0 .
Remark 13.7. For examples and further properties of regularity we refer to Section
13.2.
can be violated. Therefore certain doubts may occur whether H is the only reasonable
choice of a solution operator. The generalized Dirichlet problem asks for a positive
linear solution operator A such that Af is the classical solution whenever the classical
solution is available. This leaves at least theoretically some freedom how to choose
Af if f does not admit a classical solution.
The function space H(U ) yields a relevant model for the study of the Dirichlet
problem. If H(U ) is simplicial (compare with Chapter 6), then the kernel δx given by
the maximal representing measures is another candidate for a canonical solution op-
erator. In order to show that H(U ) is simplicial, we will use another representation of
this solution operator, namely that of Definition 13.8. The promised role of the forth-
coming operator D in connection with the simpliciality of H(U ) and its representing
measures will be clarified in Section 13.3.
Definition 13.8 (Essential solution, essential regularity). For any f ∈ C(∂U ) we de-
fine Df as Z
c
Df (x) = f dεxβ(U ) , x ∈ U.
∂U
13.1 Balayage and the Dirichlet problem 495
We call Df the essential solution of the Dirichlet problem on U with boundary data
f . We call a point z ∈ ∂U essentially regular if
The set of all essentially regular boundary points for U is denoted by ∂ess U . Since,
for every p ∈ P, Dp ≤ Hp ≤ p, we have ∂ess U ⊂ ∂ reg U .
β(U c )
then f (z) = εz (f ).
and thus (13.2) holds and z is essentially regular. Conversely, if z is essentially regu-
lar, then from (13.2) and Lemma 13.9 we infer that
bβ(U c ) (z) = B U c (z),
p(z) = R p ∈ P,
p p
Remark 13.11. As in Remark 13.6, we obtain from Proposition 13.10 and Remark
A.222 that the concept of an essentially regular point is local.
496 13 Function spaces in potential theory and the Dirichlet problem
whereas
lim inf s(x) > 0 for each y ∈ ∂U \ {z}. (13.3)
x→y, x∈U
lim s(xn ) = 0
n→∞
Remark 13.15. The converse to Lemma 13.14 is also true, see Exercise 13.132.
Proof. Let s be a barrier for z. Then s is a barrier for each sequence {xn } of points
of U converging to z. By Lemma 13.14, each such a sequence {xn } is regular. Hence
z is regular.
Conversely, let z ∈ ∂U be a regular point. We pick a potential p ∈ P such that
p < p on V for each regular set V . (Recall that pV denotes the Poisson modification.
V
For the existence of such a potential, see Proposition A.206.) Let {Vn } be a sequence
of regular sets such that z ∈
/ Vn and
∞
[
U \ {z} ⊂ Vn .
n=1
needed to verify that the regularity axiom of Bauer’s axiomatic theory holds for the
Laplace equation.
The famous Wiener criterion gives a necessary and sufficient condition for the reg-
ularity of a point z ∈ ∂U , namely, if d ≥ 3, z is regular if and only if
∞
X
αn cap(An \ U ) = ∞.
n=1
Here α > 1,
An := {x ∈ Rd : αn ≤ |x − z|2−d ≤ αn+1 }, n ∈ N,
and cap denotes the Newtonian capacity (see [21], Theorem 7.7.2). A similar condi-
tion holds also for d = 2, see [21], Section 7.7.
It follows that the problem of regularity reduces to the task how to measure the
capacity of a set. Since the notion of capacity is not too easy to handle (and we do not
plan to present it here), let us list also some geometric criteria giving conditions for
regularity. These are less sharp but more transparent.
Examples 13.18. (a) If z ∈ ∂U and there is a closed ball B ⊂ U c such that z ∈ ∂B,
then z ∈ ∂ reg U ([21], Theorem 6.6.12).
(b) If d = 2, z ∈ ∂U and z is an endpoint of a line-segment contained in U c , then
z ∈ ∂ reg U ([21], Theorem 6.6.15(i)).
(c) Cone criterion. If d ≥ 3, z ∈ ∂U , C is a closed cone with vertex z, P is a (d − 1)-
dimensional hyperplane containing the axis of C and C ∩ P ∩ B(z, r) ⊂ U c for some
r > 0, then z ∈ ∂ reg U ([21], Theorem 6.6.15(ii)).
(d) Punctured ball. This is the simplest example of an irregular set. If U := U (0, 1) \
{0},
f = 0 on S(0, 1) and f (0) = 1,
then there exists no h ∈ H(U ) such that h|∂U = f . Indeed, according to the minimum
principle (Theorem A.160) every such function h would satisfy
h ≤ εN , ε>0
(N is the fundamental harmonic function, see Definition A.137), and passing to infi-
mum we would obtain that h ≤ 0 in U .
(e) Lebesgue spine. For the so-called Lebesgue spine
q
− 1
n o
L := {0} ∪ (x1 , x2 , x3 ) ∈ R3 : x1 > 0, x22 + x23 ≤ e x1 ,
the origin is not a regular set of U := U (0, 1) \ L. Notice that both U and U c are
connected and that ∂ reg U is not closed. Details can be found in [21], pp. 186–187.
13.2 Boundary behavior of solutions 499
Theorem 13.19. The set of all regular points for U is a Gδ set and the set of all
irregular points is polar.
Proof. The set of all regular points is ∂U ∩ b(U c ) and thus it is polar by the Kellogg
property (Theorem A.215) and Gδ by Corollary A.207. See also [21], Corollary 6.8.4
and Theorem 6.6.8.
Lemma 13.20. There exists a sequence {Bj }∞ j=0 of pairwise disjoint closed balls in
R such that all balls Bj are included in U (0, 31 ),
d
∞
[ ∞
[
Bj \ Bj = {0} (13.4)
j=0 j=0
Proof. The property (13.4) is satisfied by a collection of pairwise disjoint balls if the
sequence of centers converges to the origin. Now, we may assume that Y = U (0, 23 ).
We denote
V := U (0, 13 ),
p(x) := 4
9 − |x|2 , x ∈ Y,
q := p ∧ 31 .
Then p, q ∈ P, p = q on V c and p(0) = 49 = q(0) + 91 . Consider a sequence {zk } of
points of V such that zk → 0, and a sequence {ck } of strictly positive real numbers
such that
∞
X
1
v(0) < 9, where v(y) := ck N (y − zk ), y ∈ V.
k=1
We set Bj := B(zj , rj ) where rj > 0 are chosen small enough to guarantee that
B(zj , rj ) ⊂ V , B(zj , rj ) are mutually disjoint, and v ≥ p on B(zj , rj ) (this is
possible to achieve as v(y) ≥ ck N (y −zk )). Let U be defined by (13.5). The function
c
u := v + q is hyperharmonic on Y and majorizes p on Y \ U , so that RpU ≤ u on U .
We have
Z
RbpU c (0) ≤ lim inf u(y) ≤ lim inf − u(x) dx ≤ u(0) < p(0),
y→0 r→0+ B(0,r)
Proposition 13.21. There exists a bounded open set U ⊂ Rd (d ≥ 2) for which the
set ∂ reg U of regular points is not an Fσ set.
Proof. Let {Bj }∞ d
j=0 be a sequence of pairwise disjoint closed balls in R such that all
1
balls Bj are included in U (0, 3 ),
∞
[ ∞
[
Bj \ Bj = {0}
j=0 j=0
Set
∞
[ ∞
[
F := Bj ∪ Bj∗ ∪ {0},
j=1 j=0
Kk := (Kk−1 \ Ck ) ∪ Φk (F ),
Hk := Φk (B0∗ ),
zk := Φk ({0}),
V k := (V k−1 \ {Ck }) ∪ {Φk (B) : B ∈ U}
we finish the construction.
We observe that {Kk } forms a decreasing family of compact sets and its intersec-
tion denoted by K is then a compact set. Further, given a number r > 0, each family
V k contains only a finite number of balls of radius exceeding r. Thus, each ball which
once occurs in V k will be in one of the next steps replaced by an isometric copy of
F . However, the balls Hk do not fall into the families V k and remain in Fk for ever.
We set U := U (0, 31 ) \ K. The points zk are all irregular for U , whereas all boundary
points of the balls Hk are regular for U by the cone criterion (Example 13.18(c)). We
observe that both regular points and irregular points are dense in ∂K and thus ∂reg U
cannot be Fσ .
13.2 Boundary behavior of solutions 501
Now, we will study “negligibility” of irregular points. To this end, we need some
preliminaries concerning polar sets.
Lemma 13.22 (Evans). Let K ⊂ Rd be a compact polar set. Then there exists a
Radon measure µ on K such that N µ = ∞ on K.
Proof. Using a simple scaling argument we may assume that K is contained in an
open ball V of diameter less than 1. This assumption is made in order to avoid prob-
lems with negative values of the logarithmic kernel in the two-dimensional case: now
we can be sure that there exists α > 0 such that N (x − y) ≥ α for x, y ∈ V .
Since K is polar, by Proposition A.214 there exists a Radon measure ν carried by
V such that N ν = ∞ on K. With each z ∈ V \ K we associate the ball Uz :=
U z, 12 dist(z, K) . Choose zk ∈ V \ K such that {V ∩ Uzk : k ∈ N} is a locally
finite covering of V \ K, and yk ∈ K such that
S |zk − yk | = dist(zk , K). Find pairwise
disjoint Borel sets Ek ⊂ V ∩ Uzk such that ∞ k=1 Ek = V \ K.
If z ∈ Ek and x ∈ K, then
|zk − yk | ≤ |zk − x| ≤ |zk − z| + |z − x| ≤ 21 |zk − yk | + |z − x|,
and thus
|zk − yk | ≤ 2|z − x|.
We obtain
|x − yk | ≤ |x − z| + |z − zk | + |zk − yk | ≤ |x − z| + 32 |zk − yk | ≤ 4|x − z|,
and thus, for a suitable constant C ≥ 1,
N (x − z) ≤ CN (x − yk ), x ∈ K, z ∈ Ek . (13.6)
Indeed, (13.6) is clear for d ≥ 3, in the two-dimensional case we estimate
log 4
N (x − z) ≤ log 4 + N (x − yk ) ≤ + 1 N (x − yk ).
α
In addition, we can choose the constant C so that
Z
N (x − z) dν(z) ≤ C, x ∈ K. (13.7)
Vc
then u ≥ 0 on U .
In particular, if h is a bounded harmonic function on U for which limx→z h(x) = 0
for each z ∈ ∂ reg U , then h = 0 on U .
Proof. Let ε > 0 and let k be the function from Theorem 13.23. Then
lim (u + εk)(x) ≥ 0
x→z, x∈U
Examples 13.25. (a) Let V ⊂ Rd be a bounded open set, T > 0 and U := V ×(0, T ).
Suppose that ξ ∈ ∂V and τ ∈ (0, T ). We will show that (ξ, τ ) is regular for the heat
equation if and only if ξ is regular for the Laplace equation.
Let Y := U (0, R) be a ball in Rd containing V . We consider the harmonic structure
of the Laplace equation on Y and the harmonic structure of the heat equation on Y ×R.
Then p(x) = R2 − |x|2 is a potential on Y (see Example A.180). Set q := R bV c .
p p
Denote also ψ(t) := 1 + |t − τ |2 − 1, t ∈ R.
Suppose first that ξ is regular for the Laplace equation. Then q is continuous at ξ,
q ≤ p and q(ξ) = p(ξ). Set
We see that ∆s = −d in U and ∂s ∂t ≥ −1. Thus s|U is a barrier at (ξ, τ ) and (ξ, τ )
is therefore regular for U . (If V is regular for the Laplace equation, then s is S(U )-
exposing at (ξ, τ ).)
If ξ is irregular for the Laplace equation, then q(ξ) < p(ξ) (otherwise p(x)−q(x)+
∗ ∗
|x−ξ|2 would be a barrier for V ). Consider the operator T : H + (Y ) → H + (Y ×R)
defined as T u(x, t) = u(x)et . Then T p is a heat potential on Y ×R (for the reasoning
∗
compare with Subsection A.8.E and Example A.180). Let u ∈ H + (Y ), u ≥ p
on Y \ V . Then T u ≥ T p on (Y × R) \ U . It follows that R bU c ≤ T q. Since
Tp
T q(ξ, τ ) < T p(ξ, τ ), we see that (ξ, τ ) is not regular for the heat equation.
(b) Let U be as above. Then each point (ξ, 0) ∈ V × {0} is a regular boundary point
for the heat equation on U . The function
|x − ξ|2
u : (x, t) 7→ + t, (x, t) ∈ U ,
2d
is H(U )-exposing.
(c) On the other hand, it is not difficult to show that no point of V × {T } is regular.
Indeed, from the Harnack type inequality (Proposition A.151) we easily infer that
there cannot exist a barrier at such a boundary point.
(d) Let V the unit ball in Rd . There exists a strictly positive continuous function η
which attains a strict maximum at 0 such that the point (0, η(0)) is not heat-regular
504 13 Function spaces in potential theory and the Dirichlet problem
and [
S0 (U ) = {S(V )|U : V open, V ⊃ U }.
Proof. The inequality “⊃” being obvious, let h ∈ H0 (U ). We find an open set W ⊃
U and a harmonic function h̃ on W such that h = h̃ on U . We use Tietze’s theorem
to construct a continuous extension f ∈ C(Y ) of h, this means that h = f |U . We find
positive functions u, w ∈ C(Y ) such that
Let
V := {x ∈ W : u(x) + |h̃(x) − f (x)| < w(x)}.
13.3 Function spaces and cones in potential theory 505
Set (
h̃ on V ∩ W,
g :=
h on V \ W.
Since g = h̃ on V , the function g is obviously harmonic on V . Now we want to prove
continuity on V . Since both functions h̃ and h are continuous and coincide on U , we
have h̃ = h on W ∩ U . Thus g = h on U = (U ∩ W ) ∪ (U \ W ).
The equality g = h̃ on V ∩ W shows the continuity of g at all points of V ∩ W and
the continuity of h shows the continuity of g at all points of V \ W . Let z ∈ V ∩ ∂W .
Then z ∈ U ,
lim g(x) = lim h(x) = h(z) = g(z),
x→z, x∈U x→z, x∈U
and
|g(z) − lim g(x)| = |f (z) − lim h̃(x)| ≤ lim sup |h̃(x) − f (x)|
x→z, x∈W x→z, x∈W ∩V x→z, x∈W ∩V
≤ lim w(x) = 0.
x→z
Remark 13.29. In literature, if U is finely open, S(U ) is defined as the function space
of all functions continuous on U which are finely superharmonic on U . These two
definitions are equivalent, see Exercise 13.155. Similar remark applies to the function
space H(U ).
Proposition 13.30. Suppose that U is open. Then Ac (S(U )) = Ac (H(U )) = H(U ).
Proof. Obviously, H(U ) ⊂ Ac (H(U )) ⊂ Ac (S(U )). If h ∈ Ac (S(U )), then h ∈
C(U ). Let V ⊂ U be a regular open set with V ⊂ U and µVx be the harmonic
measure, x ∈ V . Then µVx ∈ Mx (S(U )) and thus H V h = h on V . It follows that h
is harmonic on V , and V being arbitrary, h is harmonic on U , so that h ∈ H(U ).
Proof. Denote u = R bU c .
p
Step 1. First, we suppose that U is open. By Theorem A.179, there exists a measure
µ ∈ M(Y ) such that u is the Green potential Gµ. By Theorems A.181 and A.196,
the measure µ is carried by U c . We use the Lusin theorem to find a sequence {Kn }
of compact subsets of U c such that
∞
[
µ Y \ Kn = 0
n=1
pi,2 = 23 Gνki ,m2 , i = 1, 2. Since νk1 ,m1 has support in L1 , by the comparison princi-
ple (Proposition A.182) we have p2,2 ≥ p1,1 on Y . At the n-th step we find kn > kn−1
and mn > mn−1 such that
n
n+1 Gνkn ,mn ≥ p1,n−1 ∨ · · · ∨ pn−1,n−1 (13.9)
on Ln−1 and observe by Proposition A.182 that (13.9) holds in fact on Y . We set
n
pi,n = n+1 Gνki ,mn , i = 1, . . . , n. The final sequence will be {pn } with pn := pn,n ,
n ∈ N. This sequence is obviously increasing. Since for each i ∈ N we have
bpLki ,
u ≥ sup pn ≥ sup pi,n = R
n∈N n∈N
Theorem 13.33. The convex cone S(U ) is simplicial. For any z ∈ U , the measure
c
εzU is the unique maximal S(U )-representing measure for z. In particular,
ChS(U ) (U ) = ∂ reg U. (13.10)
c
By the density of P − P in C(U ) (Proposition A.169) we obtain that µ = εUz . Thus,
U c
S(U ) is simplicial and εz is the unique representing measure for z.
Now, if z ∈ ∂ reg U and µ ∈ Mx (S(U )) is a maximal representing measure, then
c
µ = εU z = εz , thus z ∈ ChS(U ) (U ). Hence ∂ reg U ⊂ ChS(U ) (U ), which together
with Lemma 13.32 proves (13.10).
Theorem 13.35. The function space H(U ) is simplicial. For any z ∈ U , the measure
c
εzU is the unique maximal H(U )-representing measure for z. In particular,
ChH(U ) (U ) = ∂ reg U.
Proof. Once we know that H(U ) = Ac (S(U )) (Theorem 13.34), this follows from
analogous assertions for the cone S(U ) (Theorem 13.33) by Theorem 7.38. (Alterna-
tively, we can proceed as in the proof of Theorem 13.33.)
Remark 13.36. By Theorems 13.35 and 13.33, the maximal representing measures
for H(U ) and S(U ) coincide. This is no longer true for representing measures that
are not maximal. For x ∈ U , the inclusion Mx (S(U )) ⊂ Mx (H(U )) is obvious. We
are going to show that
Mx (H(U )) \ Mx (S(U )) 6= ∅,
if U is a ball and x is its center. We will consider the case d ≥ 3 only; for d = 2, we
can produce a similar example using the logarithmic kernel. Let U := U (0, 1) and
y = ( 12 , 0, . . . , 0). Then
Z Z
h(0) = − h dx ≥ 2−d − h dy = 2−d h(y)
U U (y, 21 )
for every h ∈ H+ (U ) (this is a weak version of the Harnack inequality, for a version
with the sharp constant see [21], Theorem 1.4.1). Define
Then s ∈ S(U ), s(y) = 22d and s(0) = 2d−2 . According to Key lemma 3.21, there is
a measure µ ∈ M0 (H(U )) such that
We note that
s∗ (0) ≥ 2−d s(y) = 2d > s(0),
from which it follows that
µ(s) = s∗ (0) > s(0).
Hence µ ∈
/ M0 (S(U )).
g := sup{t ∈ −S(U ) : t ≤ f }
is harmonic on U . Now
β(K)
If q ∈ P and q = BqK = Rq , then q is harmonic on K c ⊃ U (Theorem A.196)
and thus
c c
BpU ≤ sup{q ∈ P : q ≤ BpU , q|U ∈ H0 (U )}.
It remains to prove that the supremum can be achieved by an increasing sequence. For
this we need that the family
c
Q = {q ∈ P : q ≤ BpU , q|U ∈ H0 (U )}
is up-directed and that there is a countable subset of Q with the same supremum. The
latter fact follows from Lemma A.55, we will prove the former one. Let s, t ∈ Q, we
want to show that there is q ∈ Q such that s ∨ t ≤ q. There exist open sets V, W ⊃ U
such that s is harmonic on V and t is harmonic on W . Then g := s ∨ t is subharmonic
on V ∩ W and the rest follows from Lemma A.171.
Theorem 13.41. The function space H(U ) and the convex cone S(U ) are simplicial
β(U c )
and H(U ) = S(U ) ∩ −S(U ). For any z ∈ U , the measure εz is the unique
H(U )-maximal H(U )-representing measure and the unique S(U )-maximal S(U )-
representing measure for z. In particular,
Proof. The proof is almost the same as that of Theorems 13.33, 13.34 and and 13.35.
Given z ∈ U , let µ ∈ Mz (S(U )) be a maximal measure. By Theorem 7.26, µ is
carried by the Choquet boundary ChS(U ) (U ). Let p ∈ P. Then by Lemma 13.39
c
there exists a sequence {pn } of potentials from P such that pn % BpU and each
pn |U belongs to H(U ). Using Lemma 13.40 we see that pn → p on U ∩ β(U c ) ⊃
ChS(U ) (U ). We obtain
c c
µ(p) = µ(sup pn ) = sup µ(pn ) = sup pn (z) = BpU (z) = εzβ(U ) (p).
n∈N n∈N n∈N
β(U c )
By the density of P − P in C(U ) (Proposition A.169) we obtain that µ = εz .
β(U c )
Thus, S(U ) is simplicial and εz is the unique S(U )-maximal representing mea-
sure for z. It follows that
c)
ChS(U ) (U ) = {z ∈ U : εβ(U
z = εz } = ∂ess U.
As in the proof of Theorems 13.34 and 13.35 we show that H(U ) = S(U ) ∩ −S(U )
and that the maximal representing measures and Choquet boundaries (when we com-
pare H(U ) with S(U )) are the same.
512 13 Function spaces in potential theory and the Dirichlet problem
By Lemma 13.9,
c
f (z) = εzβ(U ) (f ), z ∈ U.
Then we obtain assertion (a) from Corollary 6.12 and (b) follows from Bauer’s theo-
rem 3.27 (we use the fact that H(U ) = S(U ) ∩ −S(U )).
Remark 13.43. Recall that in potential theory of the Laplace equation, the Choquet
boundary coincides with ∂ reg U = U ∩ b(U c ). (Even for an arbitrary set A we have
β(A) = b(A).) This is no longer true in potential theory of the heat equation as we
will see from Example 13.45. But even if ChH(U ) (U ) = ∂ reg U , the harmonic measure
µU U
(x,t) and the maximal measure δ(x,t) may differ (and the example is even simpler),
this we will see in the following Example 13.44.
Example 13.44. Let
U := (0, 1) × (0, 1) ∪ (1, 2) .
For each (x, t) ∈ U ,
c
µU U
(x,t) = ε(x,t) ,
and
U β(U c )
δ(x,t) = ε(x,t) .
In this example, since horizontal segments are totally thin (see Example A.218),
and
c c
µU U V U
(x,t) = ε(x,t) 6= ε(x,t) = δ(x,t) , t > 1.
To demonstrate the difference, consider f ∈ C(∂U ) such that f > 0 on (0, 1) × {1}
and f = 0 elsewhere. By Examples 13.25(e), Hf (x, t) > 0 for t > 1 close to 1.
Using the Harnack-type inequality (Proposition A.151; together with a scaling and
translating argument), we easily infer that Hf > 0 on (0, 1) × (1, 2). Thus
µU
(x,t) ((0, 1) × {1}) > 0, (x, t) ∈ (0, 1) × (1, 2),
Example 13.45. Let Y = R1+1 be considered with the structure of the heat equation,
n 1 o
U := (0, 1) × (0, 1) \ 1 − : n ∈ N and V := (0, 1) × (0, 1).
n
Then
ChH(U ) (U ) = ∂V \ (0, 1) × {1} while ∂ reg U = ∂V.
Indeed, let [α] denote the integer part of a real number α. Given ξ ∈ (0, 1), set
|x − ξ|2 1
u(x, t) := +t−1+ 1 1, (x, t) ∈ U.
2 1−t − 2
Then u is a barrier at (ξ, 1), and thus (ξ, 1) is regular. On the other hand, using
Example A.218 we infer that β(U c ) ⊂ b(b(U c )) ⊂ b(V c ) = V c \ (0, 1) × {1}).
Examples 13.46. (a) Let U be an open unit ball in Rd and f a bounded discontinuous
Borel function on ∂U . Define
c
h : x 7→ εU
x (f ), x ∈ U.
Then h is H(U )-affine, but not continuous at the points of ∂U where f is discontinu-
ous.
(b) If a set U is not regular, H(U )-affine functions may be discontinuous on U even
for continuous boundary data. Consider, for example, U as the set obtained by remov-
ing the Lebesgue spine (see Example 13.18(e)) with cusp at 0 from the open unit ball
in Rd , d ≥ 3. Define
Z
c
h(x) := |y| dεU
x (y), x ∈ U.
∂U
As before, h is H(U )-affine. Since all points in ∂U \ {0} are regular, there exists a
sequence {xn } in U tending to 0 such that limn→∞ h(xn ) = 0. Since the probability
c
measure εU 0 does not charge 0, it is clear that h(0) > 0. We see that already the
restriction of h to U ∪ {0} is not continuous.
514 13 Function spaces in potential theory and the Dirichlet problem
Proposition 13.48. Let c > 0 and x, y ∈ U such that h(x) ≤ ch(y) for every h ∈
H + (U ). Then s(x) ≤ cs(y) for every H(U )-concave function s ≥ 0.
Moreover, if U is connected, then, for every compact subset K of U and every
x0 ∈ U , there exists c > 0 such that s ≤ cs(x0 ) on K for every H(U )-concave
function s ≥ 0.
In particular, every lower bounded H(U )-concave function which is finite at some
point of U is locally bounded on U .
c
Proof. Recall that the space H(U ) is simplicial and that, for every z ∈ U , εU
z is a
unique measure from Mz (H(U )) carried by ChH(U ) (U ) (see Theorem 13.35). Fix
f ∈ C(∂U ), 0 ≤ f ≤ 1, η > 0, and a compact set L ⊂ ChH(U ) (U ) such that
c Uc
εU
x + εy (U \ L) < η.
Example 13.49. Let U be the open unit ball in Rd , d ≥ 3. There exists a positive
function s ∈ S(U ) which is not H(U )-concave. Choose y ∈ U \ {0} and find c > 0
such that
h(0) ≤ ch(y), h ∈ H + (U ).
Set t(x) := |x|2−d − 1 for x ∈ Rd . Define
s : x 7→ (c + 1)t(y) ∧ t(x), x ∈ Rd .
Then s(0) = (c + 1)t(y) = (c + 1)s(y). By Proposition 13.48 we know that s is not
H(U )-concave.
13.3 Function spaces and cones in potential theory 515
Theorem 13.50. Assume that U is connected. Then every locally lower bounded set
F of H(U )-concave functions, which is upper bounded at some point of U , is locally
bounded and locally equicontinuous on U .
Proof. Assume that the set F is upper bounded at x0 ∈ U . Choose a compact set K
in U containing x0 . It is easy to find an open connected set V containing K such that
V ⊂ U . By adding a suitable constant, we may assume that s|V ≥ 0 for every s ∈ F.
Then, by Proposition 13.48, F is bounded on K, say by a constant a. Fix η > 0. It
follows from D. H. Armitage and S. J. Gardiner [21], Lemma 1.5.6, that there exists
δ > 0 such that
c c
εVx ≤ (1 + η)εVy , x, y ∈ K, |x − y| < δ.
Corollary 13.51. Let U be bounded open connected set and s be a locally lower
bounded H(U )-concave function, which is finite at some point of U . Then s is con-
tinuous on U .
Lemma 13.52. Let h ∈ H + (U ) and h(0) = 1. Then h(x) ≥ ϕ(|x|) for every x ∈ U .
Moreover, if y ∈ U , y 6= 0, and h(y) = ϕ(|y|), then h = P−y/|y| .
516 13 Function spaces in potential theory and the Dirichlet problem
|x − z| ≤ |x| + 1 = x + x/|x|
1 − |x|2
Pz (x) ≥ P−x/|x| (x) = = ϕ(|x|),
(1 + |x|)d
and therefore h(x) ≥ ϕ(|x|). Obviously, h(0) = ϕ(0). Also, for x 6= 0, Pz (x) =
ϕ(|x|) holds if and only if z = −x/|x|.
Suppose now that y ∈ U , y 6= 0, and h(y) = ϕ(|y|). Then
Z
Pz (y) − ϕ(|y|) dµ(z) = h(y) − ϕ(|y|) = 0.
∂U
hz,r : x 7→ Pz (rx), x ∈ U,
and we define
U := {hz,r : z ∈ ∂U, r ∈ (0, 1)}.
Lemma 13.53. Let u := inf U. Then u is an H(U )-concave function and u(x) =
ϕ(|x|) for every x ∈ U .
1 − r2 |x|2
hx,r (x) = Pz (rx) ≥ P−x/|x| (rx) =
(1 + r|x|)d
for every z ∈ ∂U and rR ∈ (0, 1). Hence u(x) = ϕ(|x|). It follows that u ∈ C(U )
and, obviously, u(x) ≥ u dν whenever ν ∈ Mx H(U ) .
For y ∈ U , y 6= 0, put
1 − ϕ(|y|) |x − y|
vy : x 7→ ϕ(|y|) + , x ∈ U.
|y|
13.4 Dirichlet problem: solution methods 517
Lemma 13.54. Let y ∈ U \ {0}. The function vy is continuous and convex on U (in
particular, vy is an H(U )-convex function), u ≤ vy on U , vy (0) = 1 and u(y) =
vy (y).
Proof. Define
1
w : x 7→ 1 − 1 − ϕ(|y|) |x|, x ∈ U.
|y|
Theorem 13.55. There exists a family U ⊂ H(U ) such that u := inf U is continuous
on U and there exists a continuous convex function v on U such that u ≤ v and the
inequalities u ≤ h ≤ v hold for no function h ∈ H(U ).
Proof. The theorem is a consequence of previous Lemmas 13.52, 13.53 and 13.54.
Corollary 13.56. There exist an H(U )-concave function u and an H(U )-convex func-
tion v on U such that u ≤ v on U but u and v cannot be separated by a function from
H(U ).
Functions from the family U (f ) are called upper functions to f while functions from
−U (−f ) are labelled as lower functions to f .
We define the upper solution Hf and the lower solution Hf by
Proposition
P∞ 13.59. Let {fk } be a sequence of positive resolutive functions and f :=
f
k=1 k . Then
∞
X X ∞
Hf = Hf = Hfk = Hfk .
k=1 k=1
13.4 Dirichlet problem: solution methods 519
P∞
Proof. If we choose uk ∈ U (fk ), k ∈ N, then k=1 uk ∈ U (f ), so that
∞
X
Hf ≤ Hfk .
k=1
Since
∞
X
Hgj = Hg
j=1
by the Lebesgue monotone convergence theorem, this verifies the claim for lower
semicontinuous functions.
520 13 Function spaces in potential theory and the Dirichlet problem
and passing to the infimum over all lower semicontinuous majorants we obtain
Z ∗
Hf (x) ≤ f dµx
∂U
Set
∞
X
µ := α n µ xn .
n=1
Then f is µ-integrable. First, if f is lower semicontinuous, then as above, Hf =
limj→∞ Hfj where fj ∈ C(∂U ), j ∈ N, and fj % f . Since the limit Hf is finite
on a dense set, by Doob’s convergence axiom, Hf is harmonic on U . Now, let f be
an arbitrary µ-integrable function. By Lemma A.90, there exist decreasing sequences
{gj } and {kj } of µ-integrable lower semicontinuous functions on ∂U such that −kj ≤
f ≤ gj on ∂U , j ∈ N, and
We use the above reasoning to verify that for each j ∈ N, the functions gj and kj are
resolutive and Hgj and Hkj are harmonic functions. Since
are harmonic. We observe that µ(u) = µ(v), and thus u = v on a dense set. From the
continuity of both u and v and (13.14) we deduce that v = Hf = Hf = u. Thus f
is resolutive and the common value of Hf and Hf is Hf .
Corollary 13.62. If R(U ) is equipped with the pointwise order, then it is a vector
lattice and lattice supremum in R(U ) coincides with the pointwise supremum.
Proposition 13.64. A set M ⊂ ∂U is negligible if and only if there exists a dense set
S ⊂ U such that µx (M ) = 0 for each x ∈ S.
Proof. We find a sequence {xn } of points of U and a sequence {αn } of strictly posi-
tive numbers such that {xn : n ∈ N} is dense in U and
∞
X
αn µxn (|f |) < ∞.
n=1
Set
∞
X
µ := α n µ xn .
n=1
522 13 Function spaces in potential theory and the Dirichlet problem
∞
X ∞
X
f (y) = fj+ (y) − fj− (y) ∈ R
j=1 j=1
holds for each y ∈ ∂U \ N , where the exceptional set N is µxn -null for each n ∈ N.
By Proposition 13.64, N is negligible. Using the Lebesgue monotone convergence
theorem we obtain
∞
X ∞
X ∞ ∞
X X
fj+ − H fj− = Hfj+ − Hfj− .
Hf = H
j=1 j=1 j=1 j=1
Remark 13.67. It may look that the harmonicity of Hf or even Hf follows easily
from considering U (f ) as a saturated family. However, this approach works only
if it is clear that U (f ) contains a superharmonic function. The situation is more
transparent for the Laplace equation. Then we may assume that U is connected. If
Hf is finite at one point, then there exists an upper function finite at this point and
this upper function is already superharmonic.
In this sense the c-solution generalizes the classical solution of the Dirichlet prob-
lem. Under the name “quasisolutions”, a general treatment of solutions of Dirichlet
problem based on the limiting process as in (13.15) is developed in [320].
524 13 Function spaces in potential theory and the Dirichlet problem
and thus X X
u(xj ) ≤ ui (xj ) + ui (xj ) < ∞.
i≤j i>j
It follows that
n n
1 X 1X
lim inf(h(x) + ε(k(x) + q(x))) ≥ lim inf kj (x) ≥ lim inf kj (x).
x→z n x→z n x→z
j=1 j=1
We conclude that
The proof for lim sup is analogous and the assertion is proved.
Then N is a Borel set and for its characteristic function cN we obviously have δk ∈
U (cN ) whenever δ > 0. This shows that N is a negligible set. Fix z ∈ ∂U \ N . Then
we have for every ε > 0
f (z) ≤ lim inf(Hf (x) + εk(x)) ≤ lim sup Hf (x) + ε lim inf k(x).
x→z x→z x→z
526 13 Function spaces in potential theory and the Dirichlet problem
if the sum f1 (z) + f2 (z) is well defined. If f1 (z) = −∞ and f2 (z) = ∞, then
and
lim inf(h(x) + εk(x)) ≥ lim inf(h1 (x) + εk1 (x)) + ∞ = ∞.
x→z x→z
The case f1 (z) = ∞ and f2 (z) = −∞ is analogous. We have obtained (13.16) for
each z ∈ ∂U , and thus h converges to f controlled by k. Since the last part is obvious,
the proof is complete.
We relabel the sequence {hni } as {hi }. There exist functions kn ∈ H + (U ) such that
hn converges to fn controlled by kn . Choose numbers an > 0 such that
∞
X
an kn (yj ) < ∞, j ∈ N,
n=1
and put k := ∞
P P∞
n=1 an kn , g := n=1 n(hn+1 − hn ). By Doob’s convergence axiom,
k, g ∈ H + (U ). Clearly,
∞
X
h = lim hn = (hn+1 − hn ).
n→∞
n=0
= hm − εk,
which yields
We conclude that
s̄(z) = ∞, z ∈ M.
Now, we will study the two most important concrete harmonic spaces from the
point of view of C-resolutivity.
Theorem 13.81. Let Y be a Bauer harmonic space associated with the Laplace equa-
tion and U be a relatively compact open subset of Y . Then U is C-resolutive.
13.4 Dirichlet problem: solution methods 529
Example 13.82. Consider the heat equation in Rd+1 . Denote W = U (0, 1) × (−1, 2)
and set U = W \ {(0, 0)}. Let f ∈ C(∂U ) be given by
(
1, (x, t) = (0, 0),
f (x, t) :=
0 otherwise.
By the Harnack type inequality (Proposition A.151), there exists c > 0 such that
Then by (13.17),
ũ(0, 1) ≥ h(0, 1) ≥ ch(0, 0) ≥ c.
On the other hand, each strictly positive multiple of s|U is an upper function to f and
thus Hf = 0. This shows that Hf cannot be obtained as infimum of caloric upper
functions.
V ∈ V, V ⊃ K =⇒ |H V g(x) − λ| < ε.
In particular, the limit always exists and depends neither on g, nor on the exhaustion.
Proof. Let p ∈ P and x ∈ U . By Theorems A.199 and A.195,
c
Hp(x) = RpU (x) = inf{RpG (x) : G open , G ⊃ U c }.
Given ε > 0, we find a closed (thus compact) set K ⊂ U such that
c
RpK (x) < Hp(x) + ε.
For any V ∈ V with K ⊂ V we then have
c c c
bpU (x) ≤ R
Hp(x) = R bpV (x) ≤ RpK (x) < Hp(x) + ε.
Theorem 13.91 (The Lusin–Menshov property). The fine topology τ has the follow-
ing Lusin–Menshov property: Let F be a finely closed subset of Y and K a Kσ subset
of Y disjoint from F . Then there exists a positive finely continuous and upper semi-
continuous function ϕ on Y such that ϕ = 0 on F and ϕ > 0 on K.
Proof. Let {Kn } be an increasing sequence of compact subsets of Y , ∞
S
n=1 Kn = K
and p a potential as in Proposition A.206. Using Theorems A.195 and A.205 we
obtain that b(F ) ⊂ F and RpF = R bpF < p on Y \ F . According to [66], Lemma
VI.2.5, there exists a sequence {Un } of open sets such that F ⊂ Un ⊂ Y and
ϕ = 0 on Y \ U and ϕ > 0 on K.
K ⊂ V ⊂ V ⊂ {x ∈ Y : g(x) ≤ ϕ(x)} ⊂ U.
Proposition 13.93. The fine interior of any compact subset of Y is a finely regular set.
Proof. Let K ⊂ Y be a compact set and U its fine interior. Then U c is the fine closure
of K c and thus
U c = K c ∪ b(K c ) ⊂ b(U c ) ⊂ U c .
This concludes the proof.
Remark 13.94. The reasoning of the previous proposition shows that the assertion
continues to hold even for finely closed sets (instead of compact ones).
The following proposition shows that the fine solution of the Dirichlet problem is
more efficient from the point of view of insertion of regular sets (recall that we ap-
preciated the possibility to construct “regular exhaustions” for the Wiener method of
Subsection 13.4.C). By Example 13.87, the possibility of ordinary regular exhaustion
fails in the case of the heat equation.
Proposition 13.95. Let U be a finely open set and K ⊂ U compact. Then there exists
a finely regular set V such that K ⊂ V ⊂ V ⊂ U .
Proof. The assertion is an immediate consequence of the Lusin–Menshov property in
Corollary 13.92: There exists a finely open set G such that K ⊂ G ⊂ G ⊂ U . Since
G is compact, the fine interior V of G is a finely regular set and
K ⊂ G ⊂ V ⊂ V ⊂ G ⊂ U.
534 13 Function spaces in potential theory and the Dirichlet problem
Corollary 13.96. Let U ⊂ Y be finely open and relatively compact. Then there exists
a finely regular exhaustion V of U and
|f (y) − f (x)|2
ZZ
dS(x) dS(y) < ∞. (13.18)
∂D×∂D |y − x|2
(see A. Kufner, S. Fučı́k and O. John [284], Theorems 6.8.13, 6.9.2). This condition
is not implied by continuity. Indeed, it is easily seen that if fk is the real part of the
function x 7→ (x1 + ix2 )k , then the integral as in (13.18) tends to infinity as k → ∞,
although the norms of fk in the space C(∂D) remain bounded. Now it is easy to
consider a countable number of pairwise disjoint arcs Ln on ∂D and a continuous
function f ∈ C(∂D) which oscillates as n1 fkn on Ln (with kn large enough) such that
the integral (13.18) diverges.
To overcome this difficulty, a general continuous function f ∈ C(∂U ) is approx-
imated by smooth functions (for example, by mollification). More precisely, we find
a sequence {fn } of functions from C 1 (Rd ) such that fn converge uniformly on Rd
to a continuous extension of f . Each fn |U is in W 1,2 (U ) and thus we can find
hn ∈ W 1,2 (U ) as the unique harmonic function such that hn − fn |U ∈ W01,2 (U ).
13.4 Dirichlet problem: solution methods 535
Now we observe that the operator sending boundary data from W 1,2 (U ) to the vari-
ational solution from H (U ) described in Proposition A.143 is positive (this follows
easily with the aid of testing by the negative part of the solution, cf. D. Gilbarg and
N. S. Trudinger [193], Theorem 8.1), linear and preserves constants. Thus hn con-
verge uniformly to some function h ∈ H (U ) and this h can be regarded as the
solution of the given Dirichlet problem.
We obtained the PDE solution (the abbreviation points to the theory of partial dif-
ferential equations) of the Dirichlet problem, namely “a harmonic function h which
is a limit of harmonic functions which are weak solutions of approximating Dirich-
let problems”. This description of our main achievement is somewhat cumbersome.
Therefore we want to call attention to one aspect of this approach to the Dirichlet
problem which is not so widely known. Let us emphasize that in what follows, we
will not construct a “new solution” but we will characterize the same PDE solution in
terms which allow a more elegant description. In view of Tietze’s theorem, we may
assume that our boundary data is a continuous function on U . The main new ingredi-
ent is the use of the space W01,2 (U ) + C0 (U ) where C0 (U ) is the space of continuous
functions on U vanishing on ∂U .
Lemma 13.97. Let U ⊂ Rd be a regular bounded open set and f ∈ C 1 (Rd ). If h is
the classical solution of the Dirichlet problem on U with boundary data f and u is its
solution in the sense of Proposition A.143, then h = u.
Proof. We choose ε > 0. Let {vn } be a sequence of C ∞c -functions such that vn →
u−f in W01,2 (U ) and denote wn := f +vn −ε. Then the set Vn := {x ∈ U : wn (x) >
h(x)} is open and Vn ⊂ U . Therefore h ∈ W 1,2 (Vn ) and we can use the energy-
minimizing property of h to show that
Z Z
|∇h|2 dx ≤ |∇wn |2 dx,
Vn Vn
Z 1/2 Z 1/2 Z 1/2
|∇h − ∇wn |2 dx ≤ |∇h|2 dx + |∇wn |2 dx
Vn Vn Vn
Z 1/2
≤2 |∇wn |2 dx .
Vn
It follows that Z Z
∇(wn − h)+ dx ≤ 4 |∇wn |2 dx.
U U
This means that (wn − h)+ ∈ Wc1,2 (U ).
Letting n → ∞ we obtain that
Z Z
∇(u − h − ε)+ dx ≤ 4 |∇u|2 dx.
U U
Here we use the notation H V f for the solution of the Dirichlet problem on V with
boundary data f . For the definition of un and gn it is the classical solution, but
by Lemma 13.97 also simultaneously the solution in the sense of (A.2) in Subsec-
tion A.6.A. Now, by the energy-minimizing property, we see that {un } is bounded
in W01,2 (U ). It follows that (passing, if necessary, to a subsequence) that there exists
a weak limit u∞ of un in W01,2 (U ), which is simultaneously a strong limit in L2 (U )
(by the Rellich–Kondrachov compact embedding theorem, see, for example, [334],
Theorem 1.61) and a pointwise limit almost everywhere (at least for a suitable subse-
quence). It is easy to see that u∞ is weakly harmonic on U , thus u∞ = 0 a.e. The
functions gn converge uniformly to 0 by the minimum principle, as
sup |g|(∂Vn ) → 0.
Since obviously
h = gn + un ,
passing to the limit we conclude that h = 0.
Theorem 13.99. Let U ⊂ Rd be a bounded open set and f ∈ C(U ). Then there exists
a unique h ∈ H (U ) such that h − f ∈ W01,2 (U ) + C 0 (U ).
Proof. The uniqueness follows from Lemma 13.98. For the existence, we consider a
sequence {fn } of smooth functions on Rd such that |fn − f | < 2−n−2 on ∂U . We
write h0 = f0 = 0. For each n ∈ N we find the unique harmonic function hn on U
such that hn − fn ∈ W01,2 (U ) and decompose
hn − fn − (hn−1 − fn−1 ) = un + gn ,
where gn ∈ C ∞
c (U ) and kun kW 1,2 < 2
−n . For n ≥ 2 we have |f − f
n n−1 | < 2
−n on
0
∂U and thus also |hn −hn−1 | < 2−n P on U . Therefore we can achieve that |gn | < 2n+1
on U . We observe that the series ∞ n=1 gn is uniformly convergent to a function
P∞
g ∈ C 0 (U ), n=1 un converges in W0 (U ) to a function u ∈ W01,2 (U ), and therefore
1,2
h − f = u + g ∈ W01,2 (U ) + C 0 (U ).
13.5 Generalized Dirichlet problem and uniqueness questions 537
Remark 13.100. From now, we can define the PDE solution as the solution obtained
in Theorem 13.99. It is an obvious feature of the space C 0 (U ) that the solution cannot
depend on values of f off ∂U .
Corollary 13.101. Let f ∈ C(∂U ). Suppose that the Dirichlet problem on U with
boundary data f admits a classical solution h. Then h is the PDE solution of this
Dirichlet problem.
G := {k − h : h ∈ F}.
Theorem 13.107. (a) The operator H is a Riesz homomorphism from the vector lat-
tice R(U ) to H + −H + . The more this is a Riesz homomorphism from C(∂U ) to
H + −H + . (Recall that R(U ) is a vector lattice in view of Corollary 13.62.)
(b) If a Keldysh operator T : C(∂U ) → H + −H + is a Riesz homomorphism, then
T = H.
Proof. (a) Suppose that f, g ∈ R(U ) and h = Hf ∨ Hg. Choose u ∈ U (f ) and
v ∈ U (g). Then h + (u − Hf ) + (v − Hg) ≥ max{u, v} and hence this is an
upper function to f ∨ g. Passing to the infimum over u ∈ U (f ) and v ∈ U (g) we
obtain that H(f ∨ g) ≤ h. Conversely, the function H(f ∨ g) is a common harmonic
majorant of Hf and Hg and thus h ≤ H(f ∨ g).
(b) We will use (a) and the following uniqueness argument: Since T = H on
H(U )|∂U , then also T = H on W(H(U ))|∂U and by the Stone–Weierstrass theorem
13.5 Generalized Dirichlet problem and uniqueness questions 539
D∗ f = Hf ∗ , D∗ f = Hf∗ .
Theorem 13.111. (a) If S is a Keldysh operator on C(∂U ), then there exists an ex-
tended Keldysh operator T : `∞ (∂U ) → H (U ) such that S = T |C(∂U ) .
(b) Let f ∈ `∞ (∂U ), h ∈ H (U ) and D∗ f ≤ h ≤ D∗ f . Then there exists an
extended Keldysh operator T on `∞ (∂U ) such that T f = h.
D∗ f ≤ T f ≤ D∗ f, f ∈ `∞ (∂U ).
To this end, fix z ∈ ∂ reg U . Applying Theorem 13.35 and the Keldysh lemma (Proposi-
tion 13.103) we get a function h ∈ H(U ) such that h(z) = 0 and h > 0 elsewhere on
U . Let ε > 0 and let a neighborhood V of z be chosen in such a way that f ≤ ε+f (z)
on ∂U ∩ V . There exists β > 0 such that f ≤ ε + f (z) + βh on ∂U \ V . If
then f ≤ g on ∂U and
Ag = ε + f (z) + βh on U.
13.5 Generalized Dirichlet problem and uniqueness questions 541
as needed.
Corollary 13.116 (Keldysh theorem). There exists exactly one Keldysh operator on
U.
Proof. Every Keldysh operator A satisfies (a’) from Theorem 13.115.
bb(U c ) = R
R bU c . (13.19)
p p
∗
To this end, let u ∈ H + (Y ) satisfy u ≥ p on b(U c ). By Theorem 13.5, u ≥ p on
∂U ∩ b(U c ) = ∂ reg U , so that by the assumption u ≥ p holds µx -almost everywhere
for each x ∈ U . Thus
u ≥ Hu ≥ Hp = R bpU c on U.
bpb(U c ) = Rpb(U c ) ≥ R
R bpU c on U.
bb(U c ) ≤ R
R bU c
p p
and thus z ∈ b(b(U c )). The converse inclusion is always true, so that b(U c ) =
b(b(U c )). By Proposition A.223, b(U c ) = β(U c ). Using Proposition 13.10 and
Theorem 13.5 we infer that ∂ reg U = ∂ess U .
is negligible.
(ii) =⇒ (iii): This is obvious, as ∂ess U ⊂ ∂ reg U .
(iii) =⇒ (ii): Suppose that ∂ irr U is negligible. Then by Proposition 13.127,
∂ reg U = ∂ess U , but using (iii) again we obtain that ∂U \ ∂ess U is negligible.
(ii) =⇒ (iv): We will show that all functions from C(∂U ) are Keldysh functions.
We choose f ∈ C(U ). By Theorem 3.24,
and this set is negligible by (ii). From Theorem 13.114 we obtain that f is a Keldysh
function.
(iv) =⇒ (i) Let A be a Keldysh operator. If (iv) holds, Lemma 13.126 yields that
Dw ≤ Aw ≤ Hw = Dw for each w ∈ W(H(U )). By the Stone–Weierstrass
theorem (Theorem A.30), A = H on C(∂U ).
Recall that for the purpose of the Wiener solution, there does not exist a regular
exhaustion in general (Example 13.87). Although not every set in parabolic potential
theory is a Keldysh set, we will see that we can take profit from the fact that, from
some point of view, there is “enough” of Keldysh sets.
Proposition 13.129. Let K be a compact subset of U . Then there exists a Keldysh set
V such that K ⊂ V ⊂ V ⊂ U .
{gp,y : p ∈ P 0 , y ∈ S},
bpUsc (y),
gp,y (s) := R s ∈ (0, 1).
13.5 Generalized Dirichlet problem and uniqueness questions 545
The functions gp,y are increasing, and thus they are continuous except for a countable
subset of (0, 1).
Fix a level b ∈ (0, 1) such that all gp,y are continuous at b. If 0 < a < b, then
whereas the converse inequality is always true. Set V := Ub and denote by H V and
DV the corresponding solution operators. If p ∈ P 0 , then H V p = DV p on V ∩ S
and since both functions are harmonic on V , H V p = DV p on V . From the density of
P 0 − P 0 in C(∂V ) we obtain that H V = DV . By Theorem 13.128, V is a Keldysh
set. Since K ⊂ V ⊂ V ⊂ U , the proof is complete.
Example 13.131. There exist a bounded open set U ⊂ R1+1 , a Keldysh operator A
on U and a function p ∈ P such that the inequality Ap ≤ p fails.
Since the above considerations show that H(U )|U 0 ⊂ H (U 0 ), we have c < ∞ in
view of Proposition A.151. By Proposition A.217, there exists a positive supercaloric
function u on R1+1 such that u(0, 0) = ∞ and u(0, 1) = 1. Appealing to Proposi-
tion A.172, there exists a continuous potential p on R1+1 such that 0 ≤ p ≤ u and
p(0, 0) > 2c. Let
p∗ := inf{h ∈ H(U ) : h ≥ p on ∂U }.
Then for each h ∈ H(U ) with h ≥ p on ∂U we have h ≥ 2 on (− 12 , 12 ) × {1} by
the definition of c, as for such a h we have h(0, 0) ≥ p(0, 0) > 2c. Hence p∗ ≥ 2
on (− 21 , 12 ) × {1}. Let f be a continuous function on ∂U such that 0 ≤ f ≤ p∗
and f (0, 1) ≥ 2. By Theorem 13.111 there exists a Keldysh operator A such that
546 13 Function spaces in potential theory and the Dirichlet problem
The barrier (
x2
2 + t − 1, t > 1,
s(x, t) =
1, t<1
shows that {(0, tk )} is a regular sequence for U (see Lemma 13.14), thus, passing to
the limit we obtain
This is a contradiction.
13.6 Exercises
Throughout the exercises, Y will be a Bauer harmonic space.
Exercise 13.132. Let U ⊂ Y be a relatively compact open set. Suppose that {xn } is
a sequence of points of U converging to z ∈ ∂U . If {xn } is regular, then there exists
a barrier for {xn }.
Hint. Follow the lines of the proof of Proposition 13.16.
bpU c (z) − R
= λk R bpU c (z).
k
13.6 Exercises 547
is trivial, we obtain
bU c (zn ) = R
lim R bU c (z), k ∈ N.
n→∞ pk pk
It follows that
c c
εU U
z = lim εzn in M(U ).
n→∞
Exercise 13.134. Let U ⊂ Rd be a bounded open set. Then ∂ reg U , with respect to the
Laplace equation, is dense in ∂U .
Exercise 13.137. Let U ⊂ Y be a relatively compact open set. For any x ∈ U , the
harmonic measure µU x is carried by the set of all nonsemiregular points. The set of all
semiregular points is negligible and open in ∂U .
Hint. Use Exercise 13.136.
Exercise 13.138. A relatively compact open set U is semiregular if and only if ∂ irr U
is negligible and open in ∂U .
Hint. Use Exercises 13.136 and 13.135.
Exercise 13.140. Consider the Bauer harmonic space associated with the heat equa-
tion on Rd+1 . If U ⊂ Rd+1 is a cylinder of type V × (0, T ), where V ⊂ Rd is regular
(a ball, for example), then U is semiregular. The Dirichlet problem for a cylinder is
13.6 Exercises 549
the most typical for the heat equation. Thus, semiregular sets appear already in po-
tential theory of the Laplace equation, but for the heat equation they are of particular
interest.
Hint. Consult [21], Corollary 7.2.4, Theorem 7.3.9 and Theorem 7.5.1.
Exercise 13.142. (a) For the Laplace equation: Let U ⊂ Rd be bounded and open.
Suppose that f ∈ C(∂U ) and h ∈ H (U ) is bounded. Then h = Hf if and only if for
each z ∈ ∂ reg U we have
lim h(x) = f (z).
x→z, x∈U
(b) Prove that an analogy to assertion (a) fails for the heat equation.
(c) In general Bauer harmonic spaces: Let U ⊂ Y be relatively compact and open.
Suppose that f ∈ C(∂U ) and h ∈ H (U ) is bounded. Then h = Hf if and only if for
each z ∈ ∂U and regular sequence {xn } converging to z we have
We will show that L ≥ f (z). Passing to a subsequence we may assume that there
exists a limit limn→∞ µxn in M1 (∂U ). If {xn } is regular, then
is nonempty (and open). The set I of all indices k ∈ N with fk ∈ F is infinite. For
each k ∈ I we have
lim inf(uk (xn ) − Hfk (xn )) ≥ fk (z) − lim Hfk (xn ) > τ.
n→∞ n→∞
then u ≥ 0 on U .
Completeness axiom: U+ (U ) is closed under countable sums.
By analogy with the classical theory of harmonic functions we define for any ex-
tended real-valued function f on ∂U the upper class U (f ) as
U (f ) := u ∈ U(U ) : u is lower bounded on U and
Define Hf and Hf by
Hf := inf U (f ) and Hf := −H(−f ).
By Minimum principle, Hf ≤ Hf on U .
A function f is resolutive if Hf = Hf and if this common value, which we denote
simply by Hf , is finite on U . The function Hf is called the Perron–Wiener–Brelot
solution, briefly PWB solution of the Dirichlet problem for f on U . We denote by
R(U ) the set of all real-valued resolutive functions on U .
As the first task of this series of exercises, prove the following assertions.
13.6 Exercises 551
Exercise 13.144. The following models satisfy the axioms introduced in Exercise
13.143.
(a) A sample example consists of a relatively compact open subset U of a Bauer
∗
harmonic space X, on which we consider as U(U ) the convex cone H (U ) of all
hyperharmonic functions on U . See also Exercise 13.149.
(b) U is a finely open subset of a Bauer harmonic space and U(U ) the family of all
finely hyperharmonic functions, or, alternatively, the family all lower semicontinuous
finely hyperharmonic functions. For definition, see Exercise 13.152.
(c) X is a metrizable compact convex set in a locally convex space, U = X \ ext X
and U(U ) is the family of all lower semicontinuous concave functions on U , see Ex-
ercises 13.150, 13.151.
Exercise 13.145. Under the setting of Exercise 13.143, the set U is said to be resolu-
tive (with respect to U(U )) if any function from C(∂U ) is resolutive.
If U is resolutive and x ∈ U , then the Riesz representation theorem A.72 provides
a unique Radon measure µx on ∂U such that
Z
Hf (x) = f dµx for any f ∈ C(∂U ).
∂U
Exercise 13.146. Under the setting of Exercise 13.143, let U be a coanalytic resolutive
set (Exercise 13.145). Then the harmonic measure µx is carried by U \ U for any
x ∈ U.
552 13 Function spaces in potential theory and the Dirichlet problem
Hint. It is enough to show that µx (K) = 0 for any compact set K ⊂ U . Then the
characteristic function cK is upper semicontinuous and cK = 0 on U \ U . In light of
the preceding Exercise 13.145 we conclude that
Exercise 13.147. Under the setting of Exercise 13.143, let U be a coanalytic resolutive
set (Exercise 13.145). Let f be an arbitrary extended real-valued function on ∂U .
Then Z ∗
Hf (x) = f dµx for any x ∈ U.
∂U
u
b(z) := lim inf u(y), z ∈ ∂U.
y→z,y∈U
≤ inf Hs : s ∈ S .
Exercise 13.148. Under the setting of Exercise 13.143, let U be a coanalytic resolutive
set (Exercise 13.145). Let N ⊂ ∂U be a negligible set (this means, analogously as in
Definition 13.63, that µx (N ) = 0 for each x ∈ U ). Let u ∈ U(U ) be lower bounded.
Suppose that
lim inf u(x) ≥ 0, z ∈ U \ (U ∪ N ).
x→z, x∈U
Then u ≥ 0 on U .
Exercise 13.150. Let X be a compact convex set in a locally convex space and f be
a lower semicontinuous concave function on X \ ext X. If lim infy→x f (y) ≥ 0 for
every x ∈ ext X, then f ≥ 0 on X.
Hint. If
f (x), x ∈ X \ ext X,
fb(x) :=
lim inf f (y), x ∈ ext X,
y→x
then fb is lower semicontinuous on X. Assume that m := min fb(X) < 0 and set
M := {x ∈ X : fb(x) = m}. The set M is nonempty and closed. It is also extremal.
Indeed, if λx + (1 − λ)y ∈ M , where x, y ∈ X and λ ∈ (0, 1), then fb = m on the
open segment (x, y). Hence x, y ∈ X \ ext X by the assumption. We see that fb is
concave on the closed segment [x, y], and therefore x, y ∈ M .
Since M is extremal and M ∩ ext X = ∅, we arrive at a contradiction to Proposi-
tion 2.20.
Exercise 13.152. Let Y be a Bauer harmonic space and U ⊂ Y be a finely open set.
We say that a function u : U → (−∞, ∞] is finely hyperharmonic on U is for each
x ∈ U and each relatively compact finely open set V with x ∈ V ⊂ V ⊂ U we have
Z
c
u dεVx ≤ u(x)
V
(in particular, we require the integral on the left-hand side to be well defined). Let
∗ ∗
H (U ) be the set of all finely hyperharmonic functions on U and H lsc (U ) be the set
of all lower semicontinuous finely hyperharmonic functions on U . If U is relatively
∗ ∗
compact and coanalytic, then both H (U ) and H lsc (U ) satisfy the axioms of Exercise
13.143. Moreover, in both cases the set U is resolutive (in the sense of Exercise
13.145) and the PWB solution of the Dirichlet problem with continuous boundary
data f is just H U f .
Hint. Consult [320], Section 13.A.
e ) = H(U
S(U e ) + P| (13.20)
U
T0 f = f,
c
Tn f (x) = εVx (Tn−1 f ), x ∈ U, n ∈ N.
Hint. Use the density of P − P in C(U ). If p ∈ C(U ), then the sequence {Tn p}
decreases to the greatest harmonic minorant of p in U , which is Hp by Proposition
13.3.
Exercise 13.157. If ∂ irr U is negligible, then for any x ∈ ∂ reg U there exists a function
h ∈ H+ (U ) such that h(x) = 0 and h > 0 on U \ {x}.
Hint. This follows from Proposition 13.127, Theorem 13.41 and Proposition 13.103.
in the work of C. F. Gauss (1840). Méthode du balayage was used by H. Poincaré for a
construction of the solution of the Dirichlet problem in 1890; see [375].
Ch. de la Vallée Poussin introduced the harmonic measure in [288], [289] and showed
c
that the balayage εUx (for a bounded open set U ) coincides with the harmonic measure
U
µx ; see also O. Frostman [188].
Since about 1940, balayages studied originally for measures by means of Green
potentials, was systematically investigated by M. Brelot as a kind of a dual operation
on superharmonic functions. The notions of the réduite RsA and balayage R bA were
s
introduced by M. Brelot [84]. This opened the way to a study of the Dirichlet problem
for general sets, not restricting attention only to open domains.
Balayage of measures as well as of hyperharmonic functions turns out to be one of
the most efficient tools not only in classical theory, but also in abstract potential theory,
such as the theory of harmonic spaces, of balayage spaces, H-cones etc. Readers
are referred, for instance, to H. Bauer [40], [43], J. Bliedtner and W. Hansen [66],
M. Brelot [87], N. Boboc, Gh. Bucur and A. Cornea [71], C. Constantinescu and
A. Cornea [123], J. L. Doob [145], I. Netuka [362].
The notion of a regular boundary point was introduced by H. Lebesgue [298]. A
c
characterization of regular points by means of the equality εU x = εx (Theorem 13.5)
goes back to O. Frostman [189].
Essential balayage was invented by J. Bliedtner and W. Hansen in connection with
their studies of simplicial cones in potential theory [62], [66]. Essential solution is
only our language used to make a parallel between two extreme cases of solution of
the generalized Dirichlet problem and to describe corresponding regular points. The
term “weak Dirichlet problem” is used in J. Bliedtner and W. Hansen [65] in this
connection.
The idea of a barrier goes back to H. Poincaré [375]. Explicitly this notion appears
in H. Lebesgue [298] and the relation between a barrier and regularity in Proposi-
tion 13.16 is due to G. Bouligand [78].
Example 13.18(d) of an isolated irregular boundary point is given in S. Zaremba
[478]; the striking Example 13.18(e) of the Lebesgue spine appears in H. Lebesgue
[297]. Wiener’s criterion mentioned at the beginning of Subsection 13.2.A is due to
N. Wiener [469].
O. D. Kellog [267] and F. Vasilesco [457] proved that the set of irregular points is of
capacity zero (which here means polar) for the plane case. For the Laplace equation
case (see Theorem 13.19) for R3 , the result is due to G. C. Evans [170]. For a detailed
account of the early history of regular points, see F. Vasilesco [458].
Examples of bounded open sets U for which ∂ reg U is not an Fσ set were pro-
posed by S. J. Gardiner, A. Cornea and W. Hansen in 2004 (private communication).
The construction in Proposition 13.21 seems to be new. The result of Lemma 13.22
is due to G. C. Evans [171]. A similar result to that of Theorem 13.23 appears in
M. G. Arsove [22]. A relation between regularity for the Laplace equation and the
13.7 Notes and comments 557
and show that the Choquet boundary for H∗ (U ) is the set of all non-semiregular
boundary points of U . The cone H∗ (U ) is simplicial, but not geometrically simplicial
in general (these notions thus can really differ for cones of discontinuous functions).
Note that the cone H∗ (U ) is not convex.
Results of Subsection 13.3.C are taken from W. Hansen and I. Netuka [214], where
the topic is studied in more detail and in the framework of harmonic spaces. The
relation between the notion of H(U )-concavity and the notion of superharmonicity
(see Example 13.49) has been clarified in I. Netuka and J. Veselý [363].
In [214], W. Hansen and I. Netuka derived the following theorem for strong Brelot
spaces.
Theorem. Let U be a relatively compact open connected set and s a lower bounded
function on U that is not identically ∞ . Then the following assertions are equivalent:
(i) for every open set V with V ⊂ U , the restriction s|V belongs to the uniform
closure of W(H(V )),
558 13 Function spaces in potential theory and the Dirichlet problem
Theorem 13.54 and its Corollary 13.55 are taken from I. Netuka [361]. It is shown
there that the separation is impossible in general even if we suppose that a strict in-
equality between H(U )-convex and H(U )-concave function is assumed. It should be
mentioned that the inequality h(x) ≥ ϕ(|x|) from Lemma 13.52 is well known; see
M. Brelot [87], p. 203.
The material of Section 13.4 summarizes several methods of solution of the Dirich-
let problem. This problem has a long and interesting history. Various methods for its
solution were discovered before 1920 but none of them applied to the case of general
domains (although nowadays we know that, for example, the balayage method is well
adapted to work for general domains). The first method requiring no limitations on
the shape of the boundary is due to O. Perron [372] and R. Remak [386].
In [468], N. Wiener defined a completely different type of solution and showed in
[470] that his solution coincides with that of Perron, thus establishing the resolutivity
of continuous functions (Theorem 13.60). The question of whether such a solution
of the generalized Dirichlet problem is the only bounded harmonic function having
preassigned values at each regular point was settled later (cf. Exercise 13.142).
A general treatment of the Perron method in the case of classical potential the-
ory is due to M. Brelot [83] who considers arbitrary boundary data and lower semi-
continuous functions. His main result, the so-called Brelot resolutivity theorem (see
Theorem 13.61), shows that resolutivity is equivalent to integrability with respect to
harmonic measures. For the case of open sets in axiomatic theories of harmonic func-
tions, Theorem 13.61 is due to H. Bauer [39].
It could be of interest to the reader to mention how M. Brelot was actually led to
the question of resolutivity. In [470], N. Wiener notes that in the simplest cases of
discontinuous boundary values, the Perron generalization of the Dirichlet problem is
still satisfactory. He continues: However, we do not need to go much further to find
an instance where the Perron generalization breaks down. For example, let G be the
interior of a circle with boundary R. Let us establish a system of polar coördinates
with the centre of G in pole. Let f (P ) be 1 for those points P of R with a θ-coördinate
rational in terms of 2π, and let f (P ) be 0 elsewhere on R. Clearly any upper function
of f is at least 1 and no lower function is greater than 0. Hence the Perron method
yields us here no unique generalization of the Dirichlet problem. M. Brelot told of
this example in 1973: All mathematicians (including himself) had shared N. Wiener’s
belief that the Perron approach had been of no good for general discontinuous bound-
ary data. This fact was frequently reproduced among specialists in potential theory
(needless to say that N. Wiener was an authority in mathematical circles). It was one
day in 1937, when M. Brelot was slowly walking to the Sorbonne to lecture on the
Dirichlet problem and he also intended to mention Wiener’s example. Thinking on
13.7 Notes and comments 559
upper functions (which are, of course (!), at least 1), he suddenly stopped his flow of
ideas by saying to himself: Am I really able to establish this? It did not take long
time to realize that the upper solution is in fact 0, and this moment opened the way to
Brelot’s resolutivity investigations.
The exposition of Subsection 13.4.B is based on papers [124] and [125] by
A. Cornea.
In Subsection 13.4.C we present the Wiener method and consider its generaliza-
tions. It was observed by H. Bauer ([40], p. 147; cf. Example 13.87) that in the case
of potential theory of the heat equation, one cannot, in general, insert a regular set in
between a compact set and an open set. Example 13.122 is taken from J. Lukeš [314]
and Example 13.45 from the paper by J. Köhn and M. Sieveking [275].
In [190], B. Fuglede developed the so-called fine potential theory based on the no-
tion of finely harmonic functions on finely open sets. In general harmonic spaces, first
J. Bliedtner and W. Hansen [62], [64] considered cones of continuous finely superhar-
monic functions, and then J. Lukeš and J. Malý [318] proposed the foundations of
“full” fine potential theory. This has been systematically studied by J. Lukeš, J. Malý
and L. Zajı́ček in [320]. Notice that the minimum principle for Fuglede’s finely hy-
perharmonic functions is based on the so-called quasi-Lindelöf property of the fine
topology, while in the “full” setting it is based on the Lusin–Menshov property 13.91
(cf. J. Lukeš [316] and J. Lukeš and J. Malý [318]). A part of this research concerning
the fine Wiener method is presented in Subsection 13.4.D. The reader can consult a
survey paper [191] by B. Fuglede.
Solution methods of classical potential theory rely on special properties of the La-
place equation. Namely, the Poisson integral is mostly used for the solution of the
Dirichlet problem on a ball. Alternatively, the method of integral equations uses
knowledge of the fundamental solution. The theory of partial differential equations,
however, provides methods which are more flexible without extra difficulty and cover
a wide class of elliptic equations.
For example, for an equation of the form
− div A(x, ∇u) = 0, (13.21)
where A : Ω×Rd → Rd is a continuous function satisfying some structure conditions,
the existence of solutions can be proved by the following methods:
• Riesz representation theorem in Hilbert spaces if A is the identity (so that we are
again restricted to the Laplace equation);
• Lax–Milgram theorem if A is linear in the second variable;
• direct methods of the calculus of variations if (13.21) is the Euler–Lagrange equa-
tion of a functional of type
Z
Ju = F (x, ∇u) dx,
U
which is the case if A(x, ξ) = ∇ξ F (x, ξ);
560 13 Function spaces in potential theory and the Dirichlet problem
• the theory of monotone operators in the general (possibly nonlinear and non-
variational) case. (Note that Banach’s contraction principle is powerful enough
provided the structure is simple.)
All these methods lead to weak solutions of the equation (13.21). In Subsec-
tion 13.4.E, we illustrate this alternative PDE approach in the simple case of the
Laplace equation. For an introduction to more general models and more detailed
treatment of the ideas, we refer to D. Gilbarg and N. S. Trudinger [193], J. Heinonen,
T. Kilpeläinen and O. Martio [221], J. Malý and W. P. Ziemer [334]. In our presenta-
tion we show how the PDE approach to continuous Dirichlet data can be viewed; we
follow J. Malý and W. P. Ziemer [334], Section 2.3.9.
A. F. Monna emphasized in 1938 and 1939 the fact that the methods of Perron
and Wiener are only special constructions and investigated the uniqueness question
of the solution of the Dirichlet problem from the functional analysis point of view
(see [349] where relevant references and interesting comments on the subject may be
found). He asked whether an operator of the Dirichlet problem (submitted to certain
natural conditions) was uniquely determined. A similar question was investigated by
M. V. Keldysh, who proved in [263] that there is exactly one positive linear opera-
tor sending continuous functions on ∂U into harmonic functions on U such that its
value is the solution of the classical Dirichlet problem, provided it exists (see Corol-
lary 13.116).
An operator satisfying conditions (a’) and (b) from Theorem 13.115 is sometimes
called a K-operator.
In our presentation, the Keldysh lemma 13.103 is a statement from Choquet the-
ory. For the classical potential theory, in view of Theorem 13.35, it provides a “high
quality” barrier for any regular boundary point. For a characterization of all exposed
sets, see J. Lukeš, T. Mocek and I. Netuka [321]. A hard analysis proof was given
by M. V. Keldysh in [264]. This proof, based on the Wiener criterion of regular-
ity, is very involved. A proof in the same spirit for the plane case can be found in
P. C. Curtis, Jr. [129]. In fact, knowing the Keldysh lemma, the proof of the Keldysh
theorem 13.116 is very easy. This was the approach of M. V. Keldysh in [263]. How-
ever, M. Brelot observed in [86] that a weakened form of the Keldysh lemma (which
can be proved in a simpler way) is sufficient for a proof of the uniqueness of a Keldysh
operator. He also realized that linearity plays no role; see Theorem 13.115. A modi-
fication of the definition of the Keldysh operator, the so-called Ninomiya operator, is
investigated from the point of view of uniqueness by I. Netuka in [360].
The lattice approach in Subsection 13.5.A follows I. Netuka [359]. The envelopes
∗
D f and D∗ f in Definition 13.109 were considered by M. Brelot in [85]; see also
I. Netuka [357], [359], H. and U. Schirmeier [410]. Extended Keldysh operators
(Definition 13.108) are studied by I. Netuka in [357] and [359]. An elementary ap-
proach to the Keldysh theorem in classical potential theory is available in I. Netuka
[358].
13.7 Notes and comments 561
In [314], J. Lukeš noticed that the Keldysh theorem fails for potential theory for
the heat equation; see Example 13.122 and J. Lukeš and I. Netuka [324]. The key
result saying that ∂ irr U is negligible if and only if the complement of ChH(U ) U is
negligible (Theorems 13.41 and 13.128) is a consequence of deep results of J. Bliedt-
ner and W. Hansen [62]. Results in the spirit of Theorem 13.128 are established in
H. and U. Schirmeier [410], J. Lukeš [314] and I. Netuka [357], [359]. The fact that a
Keldysh set can be inserted between a compact and an open set (Proposition 13.129)
was proved by J. Lukeš and I. Netuka [323]. Keldysh type operators satisfying addi-
tional conditions of Remark 13.130 are studied by W. Hansen in [213]. The construc-
tion from Example 13.131 seems to be new.
The so-called principal solution of the Dirichlet problem considered in J. Lukeš
[313] is a further example of a Keldysh operator. In this connection, we formulate the
following problem.
Problem 13.159. Determine the extreme points of the convex set K of all Keldysh
operators on U . (The operators H and D of Section 13.1 are extreme points of K.
What else can be said?)
Concerning the exercises, the relation between regularity and barriers was general-
ized to filters by N. Boboc, C. Constantinescu and C. Cornea [72]; Exercise 13.132
concerning regular sequences is then a special case.
The notion of a semiregular set goes back to H. Bauer [39]. Exercise 13.136 sum-
marizes results of I. Netuka [355], J. Lukeš [315], J. Lukeš and J. Malý [317] and
J. Bliedtner and W. Hansen [63]. Accumulation points of harmonic and balayaged
measures were investigated in detail in [317]. Exercise 13.133 follows Köhn and
Sieveking [275]. The behavior of the heat equation described in Exercise 13.140 mo-
tivated J. Köhn [273] to introduce an axiomatic system of potential theory built up on
the notion of semiregular set.
Exercise 13.142(c) is from J. Malý’s thesis (Prague, 1980).
Exercises 13.143–13.149 generalize classical results of M. Brelot and H. Bauer (cf.
above notes to Subsection 13.4.A) to a weak axiomatic system proposed by J. Malý in
his thesis (Prague 1980); see also J. Lukeš, J. Malý and L. Zajı́ček [320] (in particular
11.C.1).
Exercises 13.152–13.154 offer a short excursion into fine potential theory, follow-
ing [320], parts of Sections 12 and 13. Exercise 13.155 clarifies the relations between
the approximation approach and the fine approach to the introduction of function
spaces or cones. This is based on ideas from Section VII.9 of [66].
The method of Exercise 13.156 goes back to H. A. Schwarz (1870), H. Poincaré
and H. Lebesgue; cf. generalization and survey by J. Veselý [459]. Exercise 13.158 is
due to I. Netuka [357].
At the end of these notes, we add a short historical remark concerning “abstract”
harmonic spaces (cf. Section A.8). It seemed quite natural to develop an axiomatic
562 13 Function spaces in potential theory and the Dirichlet problem
system of “harmonic functions” which would unify the classical theory of harmonic
functions, the theory of caloric functions, or, more generally, solutions of some partial
differential equations, and extend classical potential theory to these axiomatics. These
theories were constructed around the 1950’s by G. Tautz, J. L. Doob, M. Brelot and
H. Bauer. In the sixties N. Boboc, C. Constantinescu and A. Cornea, in a series of
papers, built up a further more general theory, in fact, a localized version of Bauer’s
axiomatic system, culminating by the appearance of the monograph [123]. For the
purpose of our monograph, we decided to confine ourselves to a simplified version of
Bauer’s axiomatic. Among many articles on axiomatic theories, we refer the reader
to a survey paper [43] by H. Bauer.
Chapter 14
Applications
In classical analysis, there is a series of results where more complicated objects are
expressed as a mixture (or an average) of simpler ones. This chapter shows that such
simple objects are typically extreme points of an appropriate compact convex set and
the mixture is an integral representation in the spirit of the Choquet theory. Usually,
the extreme points are parametrized in a suitable way and representing measures are
carried by the set of parameters rather than by the set of extreme points itself.
The simplest example of such a representation is an expression of double stochastic
matrices by means of permutation matrices (Theorem 14.16). In this finite-dimen-
sional situation, Minkowski’s theorem 2.11 provides a sufficient tool. Using the Inte-
gral representation theorem 2.31, we present several results on representation of con-
vex functions and concave functions on a one-dimensional interval (Theorem 14.4,
Corollary 14.9, Corollary 14.12).
Positive harmonic functions on a ball in Rd admit a representation by means of
Poisson kernels (Theorem 14.18). This is the classical Riesz–Herglotz theorem. We
note that this is a very special case of the Martin type representation extensively stud-
ied in potential theory. (Space limitations do not allow us to include this elegant
part of mathematical analysis here.) In view of the relation between holomorphic
functions and harmonic functions on a disc, Theorem 14.18 also offers a represen-
tation of holomorphic functions with positive real part. Nevertheless, we included
this representation independently in Section 14.6. We believe that a reader could find
it interesting to see how methods of complex analysis lead to the direct identifica-
tion of extreme points (Theorem 14.36) without recourse to the Riesz–Herglotz the-
orem 14.18. In Section 14.5, elementary function theory combined with the Integral
representation theorem 2.31 makes it possible to obtain a representation of typically
real holomorphic functions. Extreme completely monotonic functions are identified
in Section 14.7 and the classical Bernstein theorem on representability of completely
monotonic functions as the Laplace transform of measures is proved (Theorem 14.43).
For the special case of discrete abelian groups, we present a proof of the Bochner–
Weil theorem saying that positively definite functions admit a representation by means
of characters (Theorem 14.53).
It seems to be of interest to show that an application of the Krein–Milman theo-
rem turns out to be an efficient tool in establishing results having nothing to do with
integral representation. This is illustrated in the proofs of the Lyapunov theorem on
the range of nonatomic vector measures with values in Rd (Theorem 14.58) and the
Stone–Weierstrass theorem (Theorem 14.61 and Theorem 14.62).
564 14 Applications
All the situations described above follow the same scenario: look at a suitable
convex set, prove its compactness, identify the extreme points, show that they form a
closed set. Then the Integral representation theorem does the job. A different situation
appears in Section 14.11 where a representation of invariant measures is investigated
(Theorem 14.70). The extreme points, that is, ergodic measures, do not necessarily
form a closed set and the full strength of the Choquet representation theorem 3.45 is
used.
g 7→ g(x), g ∈ R[0,1] ,
ϕ1 := c{1} , ψ0 := c{0} ,
(x − y)+
ϕy : x 7→ , x ∈ [0, 1], y ∈ [0, 1),
1−y
and
(y − x)+
ψy : x 7→ , x ∈ [0, 1], y ∈ (0, 1].
y
14.1 Representation of convex functions 565
Proposition 14.3. If E0 := {ϕy : y ∈ [0, 1]} and E1 := {ψy : y ∈ [0, 1]}, then
ext K = E0 ∪ E1 .
Proof. It is easy to verify that, for each y ∈ [0, 1], the functions ϕy and ψy are extreme
points of K. Suppose that k ∈ K \ (E0 ∪ E1 ). If we show that k = f + g for two
positive convex functions that are not a real multiples of k, then k is not an extreme
point of K. Indeed, we denote α := f (0)+f (1). Then α ∈ (0, 1), 1−α = g(0)+g(1)
and we can write k as a nontrivial convex combination
f g
h=α + (1 − α) .
α 1−α
Now, we will distinguish several cases. If k is discontinuous but not of the form ϕ1
or ψ0 , then such a decomposition is easy. If k is affine and yet not in E0 ∪ E1 , then
c := inf k > 0 and thus we can decompose as k = k(0)(1 − x) + k(1)x. From
now we assume that k is continuous and not affine. Suppose that we can find a point
a ∈ (0, 1) such that k is affine neither on [0, a] nor on [a, 1]. Let h be an affine
function supporting k at a. If h is increasing, we can write k = f + g, where f , g
are nonnegative convex, f = k on [0, a] and f = h on [a, 1]. If h is decreasing, we
interchange the role of intervals [0, a] and [a, 1]. Anyway, f is not a real multiple of
k. Now, if no such a point a as above exists, we use another rule for the choice of the
distinguished point a, namely, now a will be the supremum of all points t from (0, 1)
such that k is affine on [0, t]. Then 0 < a < 1 and k is affine both on [0, a] and [a, 1]
(but not on [0, 1]). We can use the same trick with the supporting function as above,
in addition, we will assume that h will touch k only at a. Taking into account that
k∈ / E0 ∪ E1 we observe again that the function f as above is not a real multiple of
k.
If we decompose
σ = aε0,1 + bε1,0 + µ ⊗ ε0 + ν ⊗ ε1 ,
we can summarize our achievement as follows.
566 14 Applications
Theorem 14.4. Let k be a positive convex function on [0, 1]. Then there exist a ≥ 0,
b ≥ 0 and Radon measures µ and ν on [0, 1] such that
x−ξ ξ−x
Z Z
k(x) = ac{0} (x) + bc{1} (x) + dµ(ξ) + dν(ξ)
[0,x) 1 − ξ (x,1] ξ
Proposition 14.5. The set K is not a Choquet simplex. Moreover, there is a unique
maximal measure representing an element k ∈ K if and only if either k is affine on
(0, 1) or inf k = 0.
Proof. The function k can be uniquely written as a convex combination of the func-
tions ϕ1 , ψ0 and a continuous function from K. Hence we may assume that k is
continuous.
If k is (continuous) affine, the only possibility of representation is a unique combi-
nation of ϕ0 and ψ1 .
If inf k = 0, then there exist points a, b ∈ [0, 1] such that a ≤ b, k > 0 on
[0, a) ∪ (b, 1] and k = 0 on [a, b]. Then we decompose
kc(a,1] kc[0,a)
k = k(1) + k(0) ,
k(1) k(0)
which is a convex combination of monotone functions from K and this decomposition
is unique.
Hence we may assume that min k = 0 and k is monotone, say, increasing. If µ, ν
are the measures from Theorem 14.4 for the function k, then µ = 0 and thus
ξ−x
Z
k(x) = dν(ξ).
(x,1] ξ
Let η be a C 2 “test function” with support in (0, 1]. Using integration by parts and
Fubini’s theorem we obtain
Z 1 Z 1 Z ξ 0
η(ξ) η (x)
dν(ξ) = dx dν(ξ)
0 ξ 0 0 ξ
Z 1 Z ξ
ξ − x 00
= η (x) dx dν(ξ)
0 0 ξ
Z 1 Z
ξ − x 00
= η (x) dν(ξ) dx
0 (x,1] ξ
Z 1
= k(x)η 00 (x) dx.
0
14.2 Representation of concave functions 567
This shows that the expression on the left-hand side does not depend on the measure
ν, and using density of C 2 functions in C c ((0, 1]) we deduce that ν is unique.
Now, we need to show that if k is not affine and inf k > 0, then the representation is
not unique. Namely, for the representation we write first k = h + (k − h), where h is
a supporting affine function such that 0 ≤ h ≤ k on [0, 1] and h is a positive multiple
of ϕ0 or ψ1 . Then the representing measure for k − h does not charge ϕ0 and ψ1 (this
would contradict the fact that h supports k). Since there exist at least two different
supporting affine functions h1 , h2 which are distinct combinations of ϕ0 and ψ1 , the
representation of k cannot be unique.
kf kp ≤ 1.
Define (
(1 − x)y, 0 ≤ y ≤ x ≤ 1,
G(x, y) :=
x(1 − y), 0 ≤ x ≤ y ≤ 1.
Proof. First we observe that C((0, 1)) is metrizable (see Examples 1.44 in [403]).
Therefore it is enough to consider a sequence {gk } and show that a subsequence is
convergent in Sp . It is easy to see that the original sequence is bounded and thus
the functions gk are equilipschitz on each compact subinterval of (0, 1). We exhaust
(0, 1) by a sequence of compact intervals, use the Arzèla-Ascoli theorem A.32 and the
diagonal method, and find a subsequence which converges uniformly to a function g
in C((0, 1)) on every compact subinterval of (0, 1). The function g is concave and we
can use the Lebesgue dominated convergence theorem (if p < ∞) or a trivial estimate
(if p = ∞) to conclude that kgkp ≤ 1.
Theorem 14.8. Extreme points of S1 are the zero function and the functions
nx 1 − xo
gy : x 7→ 2 min , , x ∈ (0, 1),
y 1−y
where y runs through [0, 1]. (Of course, g0 : x 7→ 2(1 − x), g1 : x 7→ 2x.)
568 14 Applications
Proof. Let y ∈ [0, 1]. To show that gy is an extreme point of S1 , suppose that gy + f ,
gy − f are in S1 for some f ∈ C((0, 1)). Then f is affine on (0, y] and [y, 1), as
−gy + (gy − f ) and −gy + (gy + f ) are concave thereon. Trivially f is continuous
and f (0+ ) = f (1− ) = 0. It follows that f is a multiple of gy , and thus also gy + f
and gy − f are multiples of gy . The restriction on norm shows that f = 0.
Conversely, let g be a nonzero extreme point of S1 . We want to prove that g = gy
for some y ∈ [0, 1]. Find the greatest interval (0, a) on which g coincides with some
multiple of x 7→ x. If none such interval exists, set a = 0. Similarly, find the greatest
interval (b, 1) on which g coincides with some multiple of x 7→ 1 − x. If none such
interval exists, set b = 1. If a = b, then it easily follows that g = ga (= gb ). Suppose
that a < b. Fix c with a < c < b. Find an affine function h such that h ≥ g and
h(c) = g(c). Set
(
g(x) − h(1)x, x ≤ c,
f0 (x) =
h(x) − h(1)x, x > c,
and (
g(x) − h(0)(1 − x), x ≥ c,
f1 (x) =
h(x) − h(0)(1 − x), x < c.
Denote Z 1
λ0 = f0 (x) dx,
0
Z 1
λ1 = f1 (x) dx.
0
and
f0 f1
g = λ0 + λ1 ,
λ0 λ1
that is, g is a convex combination of functions from S1 . Since g is an extreme point,
it follows that
f0 f1
g= = ,
λ0 λ1
in particular
xh(1)
g(x) = , x ∈ (0, c),
λ1
which contradicts the maximality of a.
14.2 Representation of concave functions 569
Corollary 14.9. Let f be a positive concave function on (0, 1). Then there exists a
unique Radon measure µ on (0, 1) such that
Z 1
f (x) = G(x, y) dµ(y) + (1 − x)f (0+ ) + xf (1− ), x ∈ (0, 1),
0
and
Z 1 Z 1
y(1 − y) dµ(y) + f (0+ ) + f (1− ) = 2 f (x) dx.
0 0
Proof. We can assume that f is not the zero function and normalize f by kf k1 = 1.
We rewrite the extreme points
nx 1 − xo
gy (x) = 2 min , (14.3)
y 1−y
2G(·, y)
gy = , y ∈ (0, 1).
y(1 − y)
The Krein–Milman theorem with transfer 2.33 yields a probability Radon measure P
on [0, 1] such that
Z Z 1
2G(x, y)
f (x) = gy (x) dP (y) = 2(1 − x) P ({0}) + 2x P ({1}) + dP (y).
[0,1] 0 y(1 − y)
(Due to the normalization, the representing measure on ext S1 does not charge the
zero function.) By the Lebesgue dominated convergence theorem and (14.3)
Z Z
lim gy (x) dP (y) = lim gy (x) dP (y) = 0.
x→0+ (0,1) x→1− (0,1)
It follows that
2P ({0}) = f (0+ ), 2P ({1}) = f (1− ).
For the uniqueness, we pick a C 2 “test function” η with support in (0, 1). Using
integration by parts, for each y ∈ (0, 1) we have
Z 1 Z y Z 1
00 00
G(x, y)η (x) dx = x(1 − y)η (x) dx + (1 − x)yη 00 (x) dx
0 0 y
Z y
= y(1 − y)η 0 (y) − (1 − y)η 0 (x) dx
0
Z 1
− y(1 − y)η 0 (y) + yη 0 (x) dx
y
(This calculation reflects that G is the Green function.) Hence using Fubini’s theorem
we obtain Z 1 Z 1 Z 1
η(y) dµ(y) = − G(x, y)η 00 (x) dx dµ(y)
0 0 0
Z 1 Z 1
=− G(x, y)η 00 (x) dµ(y) dx
0 0
Z 1
=− f (x)η 00 (x) dx.
0
This shows that the expression on the left-hand side does not depend on µ. Since
any continuous function on (0, 1) with compact support can be uniformly approxi-
mated by C 2 functions on a suitable interval [a, b] ⊂ (0, 1), the measure µ is uniquely
determined.
R1
Remark 14.10. The Green potential u(x) = 0 G(x, y) dµ(y) solves in the sense of
distributions the Dirichlet problem −u00 = µ, u(0) = u(1) = 0. Hence u0+ (0) −
u0− (1) = µ((0, 1)) and if µ is a finite measure, then u is not only concave but also
Lipschitz. Therefore to obtain the representation of general concave functions we
need to consider for µ also infinite measures (but the measure P from (14.4) is a
probability measure).
Theorem 14.11. Extreme points of S∞ are the zero function and the functions
x
a,
0 < x < a,
ga,b (x) := 1−x
, b < x < 1,
1−b
1, a ≤ x ≤ b,
where
0 ≤ a ≤ b ≤ 1.
14.2 Representation of concave functions 571
Proof. Suppose that one of the function ga,b is expressed as a nontrivial convex com-
bination of some functions f0 and f1 from S∞ . Then obviously f0 = f1 = 1
on [a, b]. Further, if a > 0, then f0 (0+ ) = f1 (0+ ) = 0, and if b < 1, then
f0 (1− ) = f1 (1− ) = 0. Also, the functions f0 and f1 must be affine on (0, a] and
on [b, 1). Then there is no other choice than f0 ≡ f1 ≡ ga,b .
Conversely, let g be a nonzero extreme point of S∞ . Then obviously sup g = 1. Let
[a, b] be the greatest (possibly degenerate) interval such that g = 1 on [a, b] ∩ (0, 1),
b = 1 if g(1− ) = 1 and a = 0 if g(0+ ) = 1. Suppose that a > 0. We must show that
g = xa on (0, a]. It is not difficult to deduce from the extremality of g that g(0+ ) = 0.
Consider the function
(
g(x) − xa , 0 < x < a,
h(x) =
0, a ≤ x < 1,
Ra
and set α := 0 h.
If g−0 (a) > 0, then there exists ε > 0 such that both g + εh, g − εh belong to S .
∞
Let us consider the most difficult case g− 0 (a) = 0. By the maximality of the interval
[a, b], g cannot be affine on any left neighborhood of a. Then by Theorem 14.8, h/α
is not an extreme point of
n Z a o
S1 ((0, a)) := f concave on (0, a) : f ≥ 0, f ≤1 .
0
Corollary 14.12. Let f be a positive concave function on (0, 1). Then there exists a
Radon measure µ on
T := {(s, t) ∈ [0, 1] × [0, 1] : s ≤ t}
such that
Z
x
f (x) = µ([0, x] × [x, 1]) + s dµ(s, t)
T ∩((x,0]×[0,1])
Z
1−x
+ 1−t dµ(s, t), x ∈ (0, 1).
T ∩([0,1]×[0,x))
572 14 Applications
Theorem 14.13. Let 1 < p < ∞. Then extreme points of Sp are the zero function and
all functions f ∈ K with kf kp = 1.
Proof. Obviously, any nonzero extreme point has Lp -norm one. Conversely, let us
assume that f is a function of norm one, which is expressed as a nontrivial convex
combination of some functions f0 and f1 from Sp . Then due to the strict convexity of
the Lp -norm, the norm of f is strictly smaller than one, which is a contradiction.
Remark 14.14. The set S1 is simplex. To show that none of Sp with 1 < p < ∞ is
simplicial, consider the equality
1
2 αp x + 12 αp (1 − x) = 12 αp ,
R1 R1
where αp is chosen so that 0 (αp x)p dx = 0 (αp (1−x))p dx = 1. Then the constant
function 21 αp is represented as a convex combination of the functions αp x and 12 αp (1−
x) as well as a convex combination of the functions 1 and 0. A similar argument works
for S∞ with α∞ = 1.
Theorem 14.16. The equality ext D = P holds and every doubly stochastic matrix is
a convex combination of permutation matrices.
Proof. Suppose first that A = (ai,j )di,j=1 is a permutation matrix. Then all entries are
0 or 1 and it easily follows that A ∈ ext D.
Conversely, consider A ∈ ext D and let the set S of all positions [i, j] such that
0 < ai,j < 1 be nonempty. We will show that there exist p, m ∈ N and sequences
{ik }∞ ∞
k=1 , {jk }k=1 of indices from {1, 2, . . . , d} such that m > 1 and the following
properties hold:
(a) [ik , jk ], [ik , jk+1 ] ∈ S for each k ∈ N,
14.4 The Riesz–Herglotz theorem 573
The matrix E has the property that sum of its elements over each row and each column
is zero. If we take ε > 0 so small that ε ≤ ai,j ≤ 1 − ε for each [i, j] ∈ S, then both
the matrices A + εE and A − εE are doubly stochastic and thus A is not an extreme
point of D, which is a contradiction.
It remains to construct the sequences {ik }∞ ∞
k=1 , {jk }k=1 . We start with an arbitrary
[i1 , j1 ] ∈ S. There must be at least one column index j2 6= j1 such that [i1 , j2 ] ∈ S,
otherwise the sum over the i1 -th row would be ai1 ,j1 < 1. Similarly we find i2 6= i1
such that [i2 , j2 ] ∈ S.
We proceed inductively and always find jk+1 6= jk such that [ik , jk+1 ] ∈ S and
ik+1 6= ik such that [ik+1 , jk+1 ] ∈ S. We go on with the construction till we find
p ∈ N when it happens that jp or ip is chosen for the second time.
Then we continue with the construction more carefully. If this happens first with
the column index, jp = jp−m , then we set ip = ip−m 6= ip−1 . If this happens first
with the row index, ip = ip−m , then we set jp+1 = jp+1−m 6= jp . At all the next steps
we follow the rule of period m. It is clear that the sequences constructed according to
this algorithm satisfy the properties (a), (b) and (c).
Once we know the equality ext D = P, Theorem 2.11 concludes the proof.
Remark 14.17. It is easy to show that the set D is a Choquet simplex in the case d = 2
whereas it is not if d ≥ 3. Namely, the arithmetic mean of all permutation matrices
with positive determinant is the matrix with all entries n1 , as well as the arithmetic
mean of all permutation matrices with negative determinant.
1 − |x|2
P : (x, z) 7→ , z ∈ ∂U, x ∈ U,
|x − z|d
and
P z : y 7→ P (y, z), y ∈ U.
Differentiation under the integral sign shows that the Poisson integral
Z
P µ(x) := Px dµ, x ∈ U,
∂U
Proof. For r ∈ (0, 1) and x ∈ U (0, 1/r) let hr (x) := h(rx). Then the function hr is
harmonic on U (0, 1/r), hence hr = P hr on U by Remark A.133. If µr := hr σ, then
Z
µr (∂U ) = hr dσ = hr (0) = 1
∂U
by the mean value theorem (Theorem A.173). By the compactness of M1 (∂U ), there
exists a measure µ ∈ M1 (∂U ) and a sequence {rn } so that rn % 1 and µrn → µ.
Then, for any x ∈ U , we have
Z
h(x) = lim h(rn x) = lim Px hrn dσ = lim (P µrn )(x)
n→∞ n→∞ ∂U n→∞
Z Z
= lim Px dµrn = Px dµ = P µ(x).
n→∞ ∂U ∂U
ψ : z 7→ P z , z ∈ ∂U,
{P z : z ∈ ∂U } ⊂ ext H(U
e ),
Theorem 14.21. Let f ∈ T (U ). Then there exist real numbers a, b and a Radon
probability measure µ on [−1, 1] such that
Z
f (z) = a + b ht (z) dµ(t), z ∈ U.
[−1,1]
T 0 (U ) := g ∈ T (U ) : g(0) = 0, g 0 (0) = 1 .
(1 − z 2 )g(z)
f (z) := , z ∈ U \ {0} ,
z
and f (0) = 1, we have f ∈ A(U ) and Re f ≥ 0. In fact, Re f > 0 on U by the strong
minimum principle for harmonic functions (see Proposition A.134). Define
1+z
ϕ(z) := , z ∈ U.
1−z
Then ϕ ∈ A(U ), ϕ(0) = 1, ϕ is injective, ϕ(U ) = (0, ∞) × R and
w−1
ϕ−1 (w) = , w ∈ ϕ(U );
w+1
cf. R. B. Burckel [95], Exercise 2.5, p. 43. The function F := ϕ−1 ◦ f has the
following properties:
F ∈ A(U ), F (0) = 0, F (U ) ⊂ U.
The left-hand side of the inequality is minorized by |1 − z 2 ||g(z)| − |z| and the right-
hand side is majorized by |z|(|1 − z 2 ||g(z)| + |z|). This yields
and
|z|
|g(z)| ≤ , z ∈ U.
(1 − |z|)2
The result now follows by the maximum modulus principle; see R. B. Burckel [95],
Exercise 5.7 and Corollary 5.9, p. 127 and 128.
Now we are going to characterize the set of extreme points of T 0 (U ). To this end,
fix β > 2 and g ∈ T 0 (U ) and notice that the functions
(
(β − z −1 − z)g(z), z ∈ U \ {0} ,
f1 (z) :=
−1, z = 0,
578 14 Applications
and (
(β + z −1 + z)g(z), z ∈ U \ {0} ,
f2 (z) :=
1, z = 0,
belong to T (U ). We will show this for f1 , the proof for f2 follows the same lines.
We can suppose that g is holomorphic on a neighborhood of U , otherwise we would
consider the function g(rz)r−1 and let r → 1+ . If z is real, −1 ≤ z ≤ 1, then f1 (z)
is real, so that Im f1 (z) = 0. For z = eiθ , θ real,
Since Im g(z) > 0 if Im z > 0, we have Im f1 (eiθ ) ≥ 0 for θ ∈ [−π, π], thus the
harmonic function Im f1 is positive on U + by Proposition A.135. In fact, Im f1 is
strictly positive there since g(z) 6= 0 if Im z > 0. Similarly, Im f1 < 0 on U − . This
shows that f1 ∈ T (U ).
Proof. Suppose that g ∈ ext T 0 (U ) and t = − 41 g 00 (0). We will show that t ∈ [−1, 1]
and g = ht . Fix α > 2|t| + 2 and define
1 (1 − (z −1 + z + 2t)g(z)), z ∈ U \ {0} ,
f (z) := α
0, z = 0.
1 1 1
(g + f )(z) = + (α − 2t − − z)g(z), z ∈ U \ {0} ,
α α z
hence the preceding result (with β = α − 2t > 2) shows that the function g + f is
typically real. Similarly, g − f is typically real since α + 2t > 2. Also,
(g + f )(0) = (g − f )(0) = 0
and
(g + f )0 (0) = 1, (g − f )0 (0) = 1.
Proof. Suppose that the sequence {gn } of functions from T 0 (U ) converges locally
uniformly on U to a function g. Obviously, g ∈ A(U ), g(0) = 0 and g 0 (0) = 1. Since
Im z · Im gn (z) ≥ 0 for z ∈ U by Lemma 14.22, we have Im gn ≥ 0, and therefore
Im g ≥ 0 on U + . Actually, since g 0 (0) = 1, Im g > 0 there by the strong minimum
principle (Proposition A.134). A similar reasoning shows that Im g < 0 on U − . We
conclude that g ∈ T 0 (U ) and T 0 (U ) is closed. By Lemma 14.23, T 0 (U ) is bounded.
We see that T 0 (U ) is compact.
Theorem 14.26. The set {ht : t ∈ [−1, 1]} is compact and is equal to ext T 0 (U ).
Hence
Fs (ht ) = 24t2 − 48st − 6 > 24s2 − 48s2 − 6 = Fs (hs ),
Proof of Theorem 14.21. We may (and do) suppose that f ∈ T0 (U ). Using the
Krein–Milman theorem with transfer 2.33 (see also Remark 2.32(c)), we obtain the
required representation
Z
f (z) = ht (z) dµ(z).
[−1,1]
580 14 Applications
Theorem 14.28. Let f ∈ P(U ). Then there exists a Radon probability measure µ on
T such that Z
1 + ηz
f (z) = dµ(η), z ∈ U.
T 1 − ηz
We will prove this representation theorem using the Krein–Milman theorem with
transfer 2.33. At first, we establish several auxiliary results.
1+r
max{|f (z)| : |z| ≤ r} ≤ .
1−r
Proof. Note that the inverse of the function ϕ from (14.6) is given by
w−1
ϕ−1 (w) = , w ∈ R.
w+1
Let f ∈ P(U ). Then f (0) = 1 and Re f > 0 on U in view of the strong minimum
principle for harmonic functions (see Proposition A.134). Hence f (U ) ⊂ R. Define
F := ϕ−1 ◦ f . Then F ∈ A(U ), F(0) = 0 and F (U ) ⊂ U . By Schwarz’s lemma (see
R. B. Burckel [95], Theorem 6.1),
|F (z)| ≤ |z|, z ∈ U.
14.6 Holomorphic functions with positive real part 581
It follows that
|f (z)| − 1 f (z) − 1
≤ ≤ |z|, z ∈ U.
|f (z)| + 1 f (z) + 1
If |z| ≤ r, then
1+r
|f (z)| ≤ .
1−r
We will consider the space A(U ) endowed with the topology of locally uniform
convergence. It is known that a subset of A(U ) is compact if and only if it is closed
and bounded (see W. Rudin [403], Sec. 1.45).
M := {hη : η ∈ T },
Proof. The set P(U ) is obviously convex and closed. It is bounded by Lemma 14.29,
hence compact. Define
Φ : η 7→ hη , η ∈ T. (14.8)
Then Φ : T → A(U ) is an injective mapping and Φ(T ) = M . If 0 < r < 1, η, ξ ∈ T
and |z| ≤ r, then
2|z||η − ξ| 2r
|hη (z) − hξ (z)| = ≤ |η − ξ|.
|(1 − ηz)(1 − ξz)| (1 − r)2
∞
X
f (z) := cn z n , z ∈ U, (14.9)
n=0
define
∞
X
Af := αn cn .
n=0
and define
∞
X 1
a : z 7→ αn z n , z ∈ U.
r
n=0
Fix ρ ∈ (r, 1) and define
ψ(t) := ρeit , t ∈ [0, 2π].
Then for f of the form (14.9) we have
Z ∞
1 dz X
a 1/z f (z) = αn cn = Af.
2πi ψ z
n=0
Theorem 14.33. Let A be a continuous linear functional on A(U ). Then there exists
a unique sequence {αn }∞
n=0 ∈ S such that
∞
X
Af := αn cn , (14.10)
n=0
Remark 14.34. For a characterization of the dual space of the space of holomorphic
functions on a general open set in C, see D. H. Luecking and L. A. Rubel [312].
and
Z
1 dw
p(z/w)q(w)
2πi ψ w
Z ∞ ∞ dw
1 X
−n
X
= 1+2 n
pn z w 1+2 qm ρ2m w−m = 1.
2πi ψ w
n=1 m=1
We arrive at Z ∞
1 dw X
p(z/w) Re q(w) =1+2 pn qn z n . (14.11)
2πi ψ w
n=1
The integral on the left-hand side is equal to
Z 2π
1 dt
p zρ−1 e−it Re q ρeit iρeit it
2πi 0 ρe
584 14 Applications
Proof. We first prove that ext P(U ) ⊂ M . Recall that M is closed and co M is
compact, hence it is sufficient to show that co M = P(U ). Indeed, by the Milman
theorem 2.43, we then have ext P(U ) ⊂ M = M . Obviously, co M ⊂ P(U ). We
will show that the assumption P(U ) \Pco M 6= ∅ leads to a contradiction.
∞ n
Let g ∈ P(U ) \ co M , g(z) := n=0 dn z for z ∈ U . By the Hahn–Banach
theorem, there are a continuous linear functional B on A(U ) and β ∈ R such that
Put
α0 := Re β0 − β and αn := βn for n ≥ 1.
Then {αn }∞
n=0 ∈ S and the equality
∞
X
Af := αn cn , f ∈ A(U ),
n=0
for f of the form (14.9) defines a continuous linear functional on A(U ). We have
hence
∞
X
Re Ahη = α0 + 2 Re αn η n > 0.
n=1
Since the function η 7→ Re Ahη is strictly positive on T , the strong minimum principle
for harmonic functions of Proposition A.134 yields
∞
X
α0 + 2 Re αn z n > 0 for any z ∈ U.
n=1
Proof of Theorem 14.28. Let f ∈ P(U ). Using the Krein–Milman theorem with
transfer 2.33 (see Remark 2.32(c)), we obtain the desired representation
Z
f (z) = hη (z) dµ(η).
T
586 14 Applications
1 − |z|2
Z 1 + ηz Z
h(z) = Re dµ(η) = 2
dµ(η).
T 1 − ηz T |η − z|
Let E := R[0,∞) be equipped with the product topology. Then E is a locally convex
space (with the pointwise convergence topology) and ∆α : E → E is a linear operator.
A real-valued function f on [0, ∞) is said to be completely monotonic if f ≥ 0 on
[0, ∞) and
(−1)n ∆αn . . . ∆α1 f ≥ 0 on [0, ∞) (14.15)
holds for all αj > 0 and n ∈ N. Let CM denote the set of all completely monotonic
functions on [0, ∞).
Proof. The monotonicity of f follows from (14.15) for n = 1 and convexity from
(14.15) for n = 2.
Proof. Denote Λ := {fλ : λ ∈ [0, ∞]}. It is easy to check that the sum and a positive
multiple of completely monotonic functions is again in CM. Hence, K is convex.
It follows from Lemma 14.40 that 0 ≤ f ≤ 1 for any f ∈ K. Therefore,
K ⊂ [0, 1][0,∞) ⊂ E.
Since [0, 1][0,∞) is compact and K is obviously a closed subset of [0, 1][0,∞) , K is
compact.
Assume that ϕ ∈ ext K. When ϕ(z) = 1 for some z ∈ [0, ∞), taking into account
that ϕ is decreasing and convex, we see that ϕ = f0 .
So suppose that ϕ(t) < 1 for all t ∈ (0, ∞). Fix s > 0 such that ϕ(s) > 0 and
define functions
ϕ(t) − ϕ(t + s) ϕ(t + s)
hs : t 7→ and gs : t 7→ , t > 0.
1 − ϕ(s) ϕ(s)
Hence,
ϕ(t)ϕ(s) = ϕ(s + t) (14.17)
588 14 Applications
for any pair s, t > 0 with ϕ(s) > 0. By a symmetry reason, (14.17) is also true if
ϕ(t) > 0. Finally, if ϕ(s) = ϕ(t) = 0, then ϕ(s + t) = 0 as well by monotonicity.
Completely monotonic functions are convex, and therefore continuous on (0, ∞).
As a continuous solution of (14.17), we get
Tb (fλ ) = fλb
is an extreme point of K. Since this holds for every b > 0, we see that all functions
from Λ are extreme.
As the set Λ is obviously closed in K, the set ext K is closed.
Theorem 14.43 (Bernstein). Let f be a real-valued function on [0, ∞). Then the
following assertions are equivalent:
(i) f ∈ K,
(ii) there exist a unique Radon measure µ ∈ M+ ([0, ∞)) and ω > 0 such that
kµk + ω = 1 and
Z
f (x) = e−xt dµ(t) + ωc{0} (x), x ∈ [0, ∞), (14.18)
[0,∞)
Proof. (i) =⇒ (ii): We use the Krein–Milman theorem with transfer 2.33 to obtain
the sought representation
Z
f (x) = fλ (x) dµ(λ) + ωc{0} (x), x ∈ [0, ∞).
[0,∞]
the linear span of the functions fλ , λ ∈ [0, ∞). Then A is an algebra and by the
Stone–Weierstrass theorem A.29, A is dense in the space of all continuous functions
on [0, ∞] (here we understand [0, ∞] as the one-point compactification of [0, ∞)). It
follows that the representation (14.18) of a completely monotonic function is unique.
(ii) =⇒ (iii): This follows from the fact that the measure µ is finite and therefore
we can differentiate the integral with respect to t under the integral sign.
(iii) =⇒ (i): Let CCM be the set of all C ∞ functions on (0, ∞) satisfying
(14.19). From the mean value theorem we easily infer that each mapping f 7→ −∆α f
maps CCM into itself. By iteration we see that (−1)n ∆αn . . . ∆α1 f ∈ CCM for any
f ∈ CCM. Hence the functions satisfying (iii) are completely monotonic.
The aim of this section is to show that every positive definite function admits an
integral representation in terms of very special positive definite functions, the so-
called characters defined below.
Denote by Md (G) the set of all finite linear combinations of Dirac measures on G.
So to say that µ belongs to Md (G) means that there are complex numbers λ1 , . . . , λn
and elements x1 , . . . , xn ∈ G such that
n
X
µ= λj εxj . (14.21)
j=1
e ∈ Md (G) by
With µ of the form (14.21), we associate the measure µ
n
X
µ
e := λj ε−xj .
j=1
e ? f )(0) ≥ 0
(µ ? µ
whenever µ ∈ Md (G).
Fix a positive definite function f on G, c ∈ C, x ∈ G and put µ := ε0 + cεx . Then
The choice c = 0 gives f (0) ≥ 0, the choice c = 1 and c = i yields 2f (0) + f (x) +
f (−x) ≥ 0 and 2f (0) + i(f (x) − f (−x)) ≥ 0, respectively. Thus both numbers
f (x) + f (−x) and i(f (x) − f (−x)) are real, which shows that f (−x) = f (x). If
f (x) 6= 0, the choice c = −|f (x)|/f (x) then gives |f (x)| ≤ f (0). In particular,
f (0) = 0 implies that f vanishes identically.
Let P denote the set of all positive definite functions f on G such that f (0) = 1
and Gb be the set of all characters of G. Obviously, P is a convex set in the vector
space of complex functions on G and G b ⊂ P.
Notice first that for a positive definite function f and for ν ∈ Md (G), the function
g := ν ? νe ? f is positive definite. Indeed, for an arbitrary µ ∈ Md (G) we have
µ?µ
e?g =µ?µ
e ? ν ? νe ? f = (µ ? ν) ? (µ]
? ν) ? f,
e ? g)(0) ≥ 0.
so that (µ ? µ
Suppose that g(0) = 0. Since |g| ≤ g(0), the function λfy + λf−y is a real multiple
of f . Similarly for the case h(0) = 0. If g(0) 6= 0 and h(0) 6= 0, we have
g(0) g h(0) h
f= 2
+ 2
1 + |λ| g(0) 1 + |λ| h(0)
g h
and g(0) ∈ P, h(0) ∈ P . We know that g(0) ≥ 0, h(0) ≥ 0 and, obviously,
Thus f is a convex combination of functions from P , and therefore both g and h are
real multiples of f because f ∈ ext P . The same is true for g − h.
We conclude that there exists a number a(λ) ∈ R such that
we get f ∈ G.
b
Definition 14.48 (The space S). Consider P as a subset of the topological vector space
S of complex functions on G endowed with the topology of pointwise convergence.
Obviously, for every x ∈ G, the mapping g 7→ g(x), g ∈ S, is a continuous linear
functional on S. Both P and Gb are closed subsets of {h ∈ S : |h| ≤ 1}, hence P and
Gb are compact.
b : γ 7→ γ(x),
x γ ∈ G.
b
Then
A := {b
x : x ∈ G}
is a dense subset of the space C(G).
b
592 14 Applications
Lemma 14.51. For every f ∈ P there exists a unique Radon measure ν ∈ M1 (G)
b
ν
such that f = f .
for each x ∈ G. The density result from Lemma 14.49 shows that ν is uniquely
determined.
Example 14.54. Consider the additive group Z of integers, so that complex functions
on G = Z are simply complex sequences indexed by Z. Fix a character γ on Z and
1
denote z := γ(1) . Since
n
γ(n) = γ(1) and γ(−n) = γ(n)
14.9 Range of vector measures 593
Our result in the situation of the group G = Z reads as follows: A sequence {ak }∞
k=−∞
is positive definite if and only if there exists a Radon measure µ on T such that ak =
b(k) for every k ∈ Z. Thus positive definite functions on Z are exactly sequences of
µ
Fourier coefficients.
P := {g ∈ L∞ (ν) : 0 ≤ g ≤ 1} .
ext P = {cM : M ∈ Σ} .
a≤g ≤1−a on Q.
for every positive f ∈L1 (ν), P is a w∗ -closed subset of the closed unit ball in L∞ (ν),
where L∞ (ν) is considered as the dual of L1 (ν). Consequently, P is w∗ -compact by
the Banach–Alaoglu theorem.
594 14 Applications
a≤f ≤1−a on Q.
Then Z n
X Z
fj h dν = bk fj dν = 0, j = 1, . . . , d,
X k=1 Qk
hence h ∈ N . Set
ah
h0 := , g1 := f + h0 , g2 := f − h0 .
khk∞
µ : M 7→ (µ1 (M ), . . . , µd (M )) ∈ Rd , M ∈ Σ.
P := {g ∈ L∞ (ν) : 0 ≤ g ≤ 1} ,
T (P ) ⊂ T (N + ext P ) = T (ext P ) ⊂ T (P ).
Proof. Obviously, there exists f ∈ A such that f (s) = 0 and f (t) 6= 0. Then the
function
ff
h := ,
1 + kf k2
belongs to A, 0 ≤ h < 1 and h(s) 6= h(t). Now, the function
1
g := h + 1 − sup h(K)
2
has the required properties.
Lemma 14.60. Let C(K) be a real or complex space, A ⊂ C(K) be an algebra and
n o
X := µ ∈ A⊥ : kµk ≤ 1 . (14.22)
Let µ ∈ ext X and g ∈ A be a real-valued function such that 0 < g < 1. Then g is
constant on spt µ.
κ := gµ, ν := (1 − g)µ.
Since 0 < g < 1, g and 1 − g are in A and A is an algebra, the measures κ and ν are
nontrivial and belong to X. Notice that
κ ν
µ = kκk + kνk
kκk kνk
and Z Z
kκk + kνk = g d|µ| + (1 − g) d|µ| = |µ|(K) = 1.
K K
Using the Choquet theory we also present a proof of the lattice version of the Stone–
Weierstrass theorem.
Theorem 14.62 (Stone–Weierstrass). Let K be a compact space and L a linear sub-
space of the real space C(K). If L is a lattice containing the constant functions and
separating points of K, then L is dense in C(K).
Proof. Obviously, L is a function space. We will show that ChL (K) = K. To this
end, pick x ∈ K, ν ∈ Mx (L) and h ∈ L. Since the function
ϕ : t 7→ |h(t) − h(x)|, t ∈ K,
belongs to L, we get
Z Z
0 = ϕ(x) = ϕ dν = |h(t) − h(x)| dν(t).
K K
Hence
spt ν ⊂ {t ∈ K : h(t) = h(x)} .
Since L separates points of K,
\
spt ν ⊂ {t ∈ K : h(t) = h(x)} = {x} .
h∈L
Hence ν = εx , so that x ∈ ChL (K). It follows that Ac (L) = C(K) (cf. Bauer’s
characterization of the Choquet boundary 3.24 and Corollary 3.23(a)).
Select now f ∈ C(K) = Ac (L) ⊂ Kc (L) and ε > 0. By Proposition 3.55, there
exists h ∈ −W(L) = L such that f − ε ≤ h ≤ f . This implies f ∈ L and finishes
the proof.
In other words, for the image measure S] µ we have S] µ = µ. It is easy to see that
T -invariant measures form a convex cone.
If µ is a T -invariant measure, a Borel measurable function f on K is said to be
µ-invariant if
µ(B4S −1 B) = 0, S ∈T.
The following result will be useful for a description of extreme T -invariant mea-
sures.
For the converse, fix S ∈ T and suppose that S] ν = ν. For a given α ∈ R, let
Then f − α > 0 on A1 , so
Z
ν(A1 ) − αµ(A1 ) = (f − α)dµ ≥ 0
A1
Note that
ν(A1 ) = ν(S −1 A) − ν(S −1 A ∩ A) = ν(A) − ν(S −1 A ∩ A) = ν(A2 )
and, similarly, µ(A1 ) = µ(A2 ). Now
ν(A1 ) ≥ αµ(A1 ) = αµ(A2 ) ≥ ν(A2 ) = ν(A1 ),
which yields ν(A1 ) − αµ(A1 ) = 0, hence 0 = µ(A1 ) = µ(A2 ). We conclude that
the sets A and S −1 A = {y ∈ K : f ◦ S (y) ≤ α} differ only by a set of µ-measure
zero.
Now we will use the following equality for real-valued functions ϕ and ψ on K:
[
{x ∈ K : ϕ(x) > ψ(x)} = {x ∈ K : ϕ(x) > r > ψ(x)}
r∈Q
[
= {x ∈ K : ψ(x) ≤ r} \ {x ∈ K : ϕ(x) ≤ r} .
r∈Q
Let us remark that X may be empty in general (see the simple Example 14.67
below). The Markov–Kakutani theorem A.3 shows that X is nonempty provided that
T is a commuting family of continuous maps. The theorem is applied as follows:
the set C is M1 (K) with the w∗ -topology, the mapping TS : C → C is for S ∈ T
defined by TS µ = S] µ, µ ∈ M1 (K), and T = {TS : S ∈ T }.
Example 14.67. Set K := {x, y} (equipped with the discrete topology) and T :=
{S1 , S2 } where
Proof. (i) =⇒ (ii): Assume that the measure µ is not ergodic. Fix a µ-invariant set
A such that 0 < µ(A) < 1. Define
−1 −1
µ1 := µ(A) µ|A and µ2 := µ(K \ A) µ|K\A .
(S −1 B) ∩ (A4S −1 A) = (S −1 B ∩ A)4(S −1 B ∩ S −1 A)
= µ|A (S −1 B).
(ii) =⇒ (iii): Assume that the measure µ is ergodic and f is a positive Borel
µ-invariant function. Then f −1 (U ) is µ-invariant for any U ⊂ R Borel, and thus
µ(f −1 (U )) ∈ {0, 1}. If {Un : n ∈ N} is a countable base of open sets in R, the set
[
L = K \ {f −1 (Un ) : µ(f −1 (Un )) = 0}
Now, the Choquet theorem 3.45 shows that every T -invariant measure has a repre-
sentation as an “integral average” of ergodic measures.
Theorem 14.70. Let K be a metrizable compact convex set and X be as in Defini-
tion 14.66. For every µ ∈ X there exists a unique Radon measure m ∈ M1 (X) such
that m is carried by the set of ergodic measures and
Z
µ(f ) = ν(f ) dm(ν), f ∈ C(K). (14.24)
X
As already mentioned, the set of ergodic measures need not be closed, as the fol-
lowing example reproduced from R. R. Phelps [374] shows.
Example 14.71. Let I := [0, 1], let J be the unit circle represented as R(mod 1).
Consider K := I × J with the product topology and T consisting of the single
mapping
S : (x, y) 7→ (x, x + y), (x, y) ∈ K.
For n ∈ N denote by µn the Radon measure on K which gives mass 1/n to the n
points (1/n, k/n), 0 ≤ k ≤ n − 1. Then µn is an ergodic measure. Indeed, let
B ⊂ K be a Borel set and
Kn := (1/n, k/n) : 0 ≤ k ≤ n − 1 .
K := {0, 1}G , the following dichotomy result: the set of G-invariant measures on
K is either a Bauer simplex or the Poulsen simplex. Actually, the former case occurs
if and only if G has the so-called property T of Kazhdan. For further results and
literature on simplices of invariant measures we refer to T. Downarowicz [148].
14.12 Exercises
Exercise 14.74. Let µ ∈ M1 ([0, 1]) and
x−ξ ξ−x
Z Z
1 1
k : x 7→ dµ(ξ) + dµ(ξ), x ∈ [0, 1].
2 [0,x) 1 − ξ 2 (x,1] ξ
2n!
|f (n) (z)| ≤ , z ∈ U.
(1 − |z|)n+1
Hint. The function f 7→ |f (n) (z)| is convex and continuous on A(U ). By Bauer’s
concave minimum principle (see Corollary 2.24) and Theorem 14.36,
2n!
max |f (n) (z)| : f ∈ P(U ) = max |g (n) (z)| : g ∈ M ≤
.
(1 − |z|)n+1
Exercise 14.76. Let W be an infinite subset of U , g ∈ A(U ), g (n) (0) 6= 0 for any
n = 0, 1, . . . . For w ∈ W define
gw : z 7→ g(wz), z ∈ U.
Hint. Assume that there exists a nontrivial continuous linear functional A on A(U )
annihilating the linear hull of {gw : w ∈ W }. It follows from Theorem 14.33 that
there exists {αn }∞n=0 ∈ S such that (14.10) holds for f of the form (14.9). For
w ∈ W we have
∞
X g (n) (0) n n
gw (z) = w z , z ∈ U,
n!
n=0
hence
∞
X g (n) (0) n
Agw = αn w = 0. (14.25)
n!
n=0
604 14 Applications
p
n
Since lim supn→∞ |αn | < 1, the function
∞
X g (n) (0)
z 7→ wn z n
n!
n=0
Exercise 14.78. Let K be a compact space and L a subspace of the real space C(K)
containing the constant functions. If L is a lattice, then L is an algebra.
First hint. Fix f ∈ L, f 6= 0, and prove that f 2 ∈ L. For this, put a := kf k and
denote by A the set of all functions ϕ ∈ C([−a, a]) such that ϕ ◦ f ∈ L. Then A is a
closed linear subspace of C([−a, a]), A is a lattice containing the constant functions
and separating points of [−a, a] (consider the identity function). By Theorem 14.62,
A = C([−a, a]). In particular, the function ϕ : t 7→ t2 , t ∈ [−a, a], belongs to A, thus
f 2 ∈ L.
Second hint. Fix f ∈ L, f 6= 0, and prove again that f 2 ∈ L. Put a := kf k. Given
ε > 0, there are affine functions ϕ1 , . . . , ϕn on [−a, a] such that
t2 − (ϕ1 ∨ · · · ∨ ϕn )(t) ≤ ε,
whenever t ∈ [−a, a] (a polygonal line inscribed to the graph of the function t 7→ t2
on [−a, a]). Since ϕj ◦ f ∈ L, j = 1, . . . , n, and L is obviously a lattice, f 2 is a
uniform limit of functions of L, thus f 2 ∈ L.
Exercise 14.79. Let f be a real-valued function on [0, ∞). Prove that f is completely
monotonic if and only if
n
X n
(−1)k f (x + kδ) ≥ 0
k
k=1
Then there exists a uniquely determined Radon measure µ on (0, 1) such that
Z
f (x) = f (1) + (x − 1)f (0) + G(x, y) dµ(y), x ∈ [0, 1].
[0,1]
In fact, µ turns out to be the second distributive derivative of f . The same result
(formulated in terms of one-dimensional potential theory) can be found in J. L. Doob
[145], p. 260. A representation theorem similar to that of Theorem 14.4 is proved in
W. Blaschke und G. Pick [60] using tools of classical analysis.
The proof of main Theorem 14.16 on a representation of doubly stochastic matri-
ces was inspired by a hint to Exercise 5.15 (p. 60) in C. Berg, J. P. R. Christensen
and P. Ressel [54] where matrices are considered as chessboards. Another proof
can be based on M. Hall Jr. [205] (Theorem 5.1.9), and a similar proof is given in
K. Jacobs [242]. A still different proof is presented in A. W. Roberts and D. E. Var-
berg [388] (Sec. 63, Theorem A). The result on the representation of doubly stochas-
tic matrices goes back to G. Birkhoff [57]. The history of representation of doubly
stochastic matrices is described in the paper by S. King and R. Shiflett [269] where
related results on infinite doubly stochastic matrices and operators are reviewed. In
particular, doubly stochastic measures are considered. Let I = [0, 1] and let λ stand
606 14 Applications
For various generalizations of the classical Bochner theorem (where the group is R),
consult G. Choquet [108], volume II, or R. R. Phelps [374], 2nd edition, pp. 103–105,
or G. B. Folland [177], pp. 76–121, where further references can be found. Note that
14.13 Notes and comments 607
the general case usually requires an extension of the representation theorems introduc-
ing, for instance, the notion of universal cap of a cone. However, in [110], G. Choquet
gives a proof of a general Bochner–Weil theorem on abelian locally compact groups
based on a reduction to the discrete case presented here.
Theorem 14.58 was proved by A. A. Lyapunov in [300]. The idea to use the Krein–
Milman theorem for the proof of the Lyapunov convexity theorem 14.58 goes back to
J. Lindenstrauss [306]. Lemma 14.56 is taken from R. B. Holmes [238]. Let us remark
that the assertion of the Lyapunov convexity theorem holds only in finite-dimensional
spaces.
The proof of the Stone–Weierstrass theorem 14.61 follows L. de Branges [133].
For a generalization, see W. Rudin [403], Chap. 5. The proof of Theorem 14.62 was
inspired by a paper of J. Bliedtner [61].
The exposition of Section 14.11 is based on Phelps’ book [374] where a more
general situation is considered; see also the survey paper [179] by V. P. Fonf, J. Lin-
denstrauss and R. R. Phelps, or G. Choquet [108] (volumes II and III). A discussion
on various definitions of ergodic measures is contained there and uniqueness of the
representation (14.24) is established. Also, further references related to the subject
are to be found there.
The hints of Exercise 14.78 are taken from a short paper of C. Dellacherie [142];
cf. also a paper by G. Nöbeling and H. Bauer [364]. Exercise 14.80 is taken from a
paper by D. Milman [345]. In general, an operator T ∈ ext BL(X,Y ) need not be an
isometry. For a partial converse and further discussions, see N. M. Roy [401]. For a
finite-dimensional Banach space X, every extreme point of BL(X) is an isometry if
and only if X is a Hilbert space; see M. A. Navarro [354].
Chapter A
Appendix
This chapter presents a survey of notions and facts needed throughout the book. As
was already mentioned in the introduction, our key references are W. Rudin [403]
and M. Fabian, P. Habala, P. Hájek, V. Montesinos Santalucı́a, J. Pelant and V. Zi-
zler [173] for functional analysis, L. Asimow and A. J. Ellis [24] for ordered vector
spaces, D. H. Fremlin [182], [181] and [183] for measure theory, R. Engelking [169]
and K. Kuratowski [285] for topology, A. S. Kechris [262] and C. A. Rogers and J. E.
Jayne [394] for descriptive set theory, D. H. Armitage and S. J. Gardiner [21] for clas-
sical potential theory and J. Bliedtner and W. Hansen [66] for abstract potential theory.
Section A.5 also uses the paper G. Koumoullis [280].
Theorem A.2 (Mazur). Let A be a convex subset of a locally convex space. Then A is
closed if and only if it is weakly closed.
Theorem A.3. Let X be a nonempty compact convex set in a locally convex space Y
and T be a commuting family of continuous affine mappings carrying X into itself.
Then there exists a point x ∈ X such that T x = x for every T ∈ T .
Proof. The assertions (a) and (b) can be found as Theorem 4.9 of [403], and (c) fol-
lows from (a) and (b). The verification of the fact that T is a (w∗ –w∗ )-homeomor-
phism is straightforward.
• if x, y ∈ E, x ≤ y and y ≤ x, then x = y,
• if x, y, z ∈ E, x ≤ y and y ≤ z, then x ≤ z,
A.1 Functional analysis 611
Proposition A.11. For an ordered vector space E the following properties are equiv-
alent:
(i) (Riesz decomposition property): If b1 , b2 , a ∈ E + and 0 ≤ a ≤ b1 + b2 , then
there exist a1 , a2 ∈ E + such that a = a1 + a2 and a1 ≤ b1 , a2 ≤ b2 .
(ii) (Riesz refinement property): If a1 , a2 , b1 , b2 ∈ E + and a1 + a2 = b1 + b2 , then
there exist cij ∈ E + , i, j = 1, 2, such that
Proof. We first show by induction Pn+1the equivalences (i) ⇐⇒ (i’), (ii) ⇐⇒ (ii’) and
(iii) ⇐⇒ (iii’). Let 0 ≤ a ≤ i=1 bi for b1 , . P
. . , bn+1 positive. Then there exist
c, an+1 ∈ E with a = c+an+1 such that 0 ≤ c ≤ ni=1 bi and 0 ≤ an+1 ≤ Pbnn+1 . The
inductive assumption yields the existence of a1 , . . . , an positive with a = i=1 ai and
ai ≤ bi , i = 1, . . . , n. Hence (i) =⇒ (i’).
To show (ii) =⇒ (ii’), we first prove the assertion inductively for the case of a pair
a1 , a2 . Let a1 , a2 and b1 , . . . , bm+1 be such that a1 , a2 ≤ bj , 1 ≤ j ≤ m + 1. Using
the inductive assumption we find c0 with a1 , a2 ≤ c0 ≤ bj , 1 ≤ j ≤ m, and then we
select c such that a1 , a2 ≤ c ≤ c0 , bm+1 .
We proceed analogously in the general case a1 , . . . , an .
To verify (iii) =⇒ (iii’)P we use a similar reasoning as in the previous
Pn case.P First we
handle the case a1 + a2 = j=1 bj and then we pass to the case i=1 ai = m
m
j=1 bj .
Now we prove (i) =⇒ (ii) =⇒ (iii) =⇒ (i).
(i) =⇒ (ii): Given a1 +a2 = b1 +b2 with positive terms, we have 0 ≤ a1 ≤ b1 +b2 .
By (i) there exist c11 , c12 ∈ E + such that c11 ≤ b1 , c12 ≤ b2 and a1 = c11 + c12 . Now
it suffices to put c21 := b1 − c11 and c22 := b2 − c12 .
612 A Appendix
and each member of this equality is positive. By (ii), there exist cij ∈ E + , i, j = 1, 2,
such that
b1 − a1 = c11 + c12 , b2 − a2 = c21 + c22 ,
b1 − a2 = c11 + c21 , b2 − a1 = c12 + c22 .
We set c := a1 + c12 . Then
0, a − b2 ≤ a, b1 .
0, a − b2 ≤ c ≤ a, b1 .
Definition A.12 (Supremum and infimum). If (E, ≤) is a partially ordered set and
M a nonempty subset of E, then z is called an upper bound of M if m ≤ z for all
m ∈ M , and z is called a least upper bound of M , or supremum of M , if z is an upper
bound of M and z ≤ y for any upper bound y of M . Any subset of E has at most one
supremum. W
The supremum of M will be denoted by M . The supremum of a two-point set
{x, y} will be denoted by x ∨ y. TheVnotion of greatest lower bound or infimum is
defined in a similar way with notation M and x ∧ y.
Definition A.13 (Vector lattices). A partially ordered set (E, ≤) is called a lattice if
every two-point subset of E has a supremum and an infimum. The ordered vector
space E is called a vector lattice if it is a lattice.
Given x ∈ E, we write x+ := x ∨ 0, x− := (−x) ∨ 0 = −(x ∧ 0) and |x| :=
x ∨ (−x).
Lemma A.14. Let E be an ordered vector space and assume that a, b ∈ E are such
that a ∧ b exists. Then
(a) for any c ∈ E, (a + c) ∧ (b + c) exists and equals (a ∧ b) + c,
(b) if λ ≥ 0, then λa ∧ λb exists and equals λ(a ∧ b),
(c) (−a) ∨ (−b) exists and equals −(a ∧ b).
A.1 Functional analysis 613
Proof. The proof follows by a straightforward verification; see also Proposition 2.5.1
in [24].
Proposition A.15. Let E be an ordered vector space. Then the following assertions
are equivalent:
(i) E is a vector lattice,
(ii) a ∧ b exists for each a, b ∈ E + ,
(iii) a ∨ 0 exists for all a ∈ E.
Proof. Obviously, (i) =⇒ (ii) and (i) =⇒ (iii). If a, b ∈ E are arbitrary, we can
find c ∈ E such that a + c, b + c ∈ E + (we recall that E = E + − E + ). Then
a ∧ b = (a + c) ∧ (b + c) − c and a ∨ b = −((−a) ∧ (−b)). Hence (ii) =⇒ (i).
Now assume (iii) and choose a, b ∈ E. By (iii), (a − b) ∨ 0 exists. Then (a − b) ∨
0 + b is the supremum a ∨ b. So (iii) =⇒ (i).
Proof. We follow the proof of Proposition 2.5.3 in [24]. For the proof of (a),
a − (a ∨ b) = a + (−a ∧ −b) = 0 ∧ (a − b)
= 0 ∧ (a − b) + b − b = (b ∧ a) − b.
Hence (b) follows with b = 0 in (a) and (c) is obvious from the definition of a+ , a− .
Since
|a| + a = (a ∨ −a) + a = (2a) ∨ 0,
we get
|a| = 2a+ − (a+ − a− ) = a+ + a− .
Hence |a| ≥ 0 and thus
Proposition A.17. If an ordered vector space E is a lattice, then it has the Riesz
decomposition property.
614 A Appendix
Definition A.18 (Ordered locally convex spaces and their duals). A locally convex
space E is an ordered locally convex space if E is ordered by a proper closed convex
cone E + . If E is a normed linear space or even a Banach space, we call it an ordered
normed linear space or an ordered Banach space.
A functional x∗ ∈ E ∗ is positive if x∗ (x) ≥ 0 for all x ∈ E + . If (E ∗ )+ denotes the
proper convex cone of all positive elements x∗ ∈ E ∗ , then (E ∗ , w∗ ) is also an ordered
locally convex space provided E ∗ = (E ∗ )+ − (E ∗ )+ . If E is an ordered normed
linear space, we note that (E ∗ )+ is w∗ -closed and thus norm closed. Hence E ∗ is an
ordered Banach space.
Examples A.19. (a) A function space H with the pointwise ordering serves as an
example of an ordered normed linear space. (In the case of the pointwise ordering
considered on families of functions we sometimes also use the term natural ordering.)
(b) The dual space H∗ of a function space H with the ordering defined as ϕ ≤ ψ if
ψ(h) − ϕ(h) ≥ 0 for all h ∈ H+ is an ordered Banach space.
Indeed, if ϕ ∈ H∗ is given, we extend it to a measure µ ∈ (C(K))∗ . If µ = µ+ −µ−
is the decomposition of µ into its positive and negative part, we set ϕ+ := µ+ |H and
ϕ− := µ− |H . Then ϕ+ , ϕ− are positive functionals on H and ϕ = ϕ+ − ϕ− .
(c) The spaces Ac (X) (X a compact convex set), C(K) (K a compact space), C c (X),
C 0 (X) (X a σ-compact locally compact space) and their duals are ordered Banach
spaces.
x∗ (x) < 0.
Lemma A.21. Let E be an ordered vector space. Then every additive and positively
homogeneous functional ϕ on E + can be uniquely extended to a linear functional
on E.
Proposition A.22. Let E be an ordered Banach space. Then there exists α > 0
such that for every x ∈ E we can find x1 , x2 ≥ 0 satisfying x = x1 − x2 and
kx1 k + kx2 k ≤ αkxk.
Proof. The proof is a variant of the proof of the open mapping theorem; see Theo-
rem 2.1.2 of [24].
Remark A.23. In all cases of ordered normed linear spaces that we encounter, the
existence of α > 0 from Proposition A.22 is easily shown.
Theorem A.24. Let E be an ordered Banach space with the Riesz decomposition
property such that E ∗ = (E ∗ )+ − (E ∗ )+ . Then E ∗ is a lattice.
Then w∗ (x1 + x2 ) ≤ w∗ (x1 ) + w∗ (x2 ) for all x1 , x2 ∈ E + and w∗ (λx) = λw∗ (x)
for all x ∈ E + and λ ≥ 0. It follows from the Riesz refinement property that w∗ (x1 +
x2 ) = w∗ (x1 ) + w∗ (x2 ), x1 , x2 ∈ E + .
Using Lemma A.21 we extend w∗ to the whole space E. Then w∗ is positive on
E and we claim that w∗ is continuous.
+
Indeed, let α > 0 be the constant from Proposition A.22. Given x ∈ E, let x =
x1 − x2 , where x1 , x2 ∈ E + and kx1 k + kx2 k ≤ αkxk. Then
If z ∗ ≤ x∗ , y ∗ , x ≥ 0 and x = x1 + x2 , x1 , x2 ≥ 0, then
Hence z ∗ ≤ w∗ and w∗ = x∗ ∧ y ∗ .
A.2 Topology
We consider all topological spaces to be Hausdorff (the only exception are topologies
on extreme points considered in Chapter 9). We recall that Tychonoff spaces are
just completely regular spaces (see [169], Section 1.5). A topological space X is
metrizable if there exists a metric on X that induces the same family of open sets. A
topological space is Polish if it is separable and completely metrizable. It is Lindelöf
if any open cover has a countable subcover. We recall that a topological space X is σ-
compact if it can be written as a countable union of compact sets; it is locally compact
if any point of X has a neighborhood whose closure is compact. A topological space
616 A Appendix
X is normal if for a given pair of disjoint closed sets F1 , F2 in X there exist disjoint
open sets U1 , U2 in X such that Fi ⊂ Ui , i = 1, 2. We note that any σ-compact
locally compact or metrizable space is normal (see [169], Sections 3.3, 3.8 and 4.1).
A set A in a topological space X is perfect if A is closed and has no isolated points.
The set Y
(xi )i∈I ∈ Ki : pij (xi ) = xj , j ≤ i, j, i ∈ I
i∈I
is the inverse limit of the system (Ki , pij )i,j∈I and is denoted as lim(Ki , pij )i,j∈I .
←
Sometimes we write only lim Ki .
←
By [169], Theorem 3.2.13, the inverse limit is compact and nonempty provided the
spaces Ki are nonempty. Also, the projections
πi : lim(Ki , pij )i,j∈I → Ki
←
is an isometric isomorphism (here C(K1 , C(K2 )) denotes the space of all continuous
mappings on K1 with values in C(K2 )).
Proof. If
F := {F ⊂ X : F closed, ϕ(F ) = Y }
is ordered by inclusion (that is, F1 ≤ F2 if F1 ⊃ F2 ), then it is easy to see by Zorn’s
lemma that there exists a maximal element X 0 ∈ F.
Definition A.39 (Up and down-directed families of sets and functions). A family F
of extended real-valued functions on a set X is down-directed if for every f1 , f2 ∈ F
there exists f3 ∈ F with f3 ≤ f1 ∧ f2 . A family F of sets is down-directed if
{cF : F ∈ F} is down-directed. Analogously we define up-directed families of
functions and sets.
{ϕ−1 (y) ∩ F : F ∈ F}
is a down-directed family T
of nonempty compact sets, and thus their intersection con-
tains a point x. Then x ∈ F ∈F F and ϕ(x) = y.
(b) If X is normal and A ⊂ X, then A is a zero set if and only if A is a closed Gδ set.
(c) If X is metrizable, any closed set is a zero set.
Proof. For the proof of (a) see [169], p. 42, (b) can be found as Corollary 1.5.12 in
[169] and (c) follows from Corollary 4.1.12 in [169].
Definition A.44 (Baire and Borel sets). If X is a topological space, the σ-algebra
generated by all open sets is the σ-algebra of Borel sets. The σ-algebra of Baire sets
is generated by all cozero sets. Hence, in the case of a metrizable space X, these
families coincide (see Proposition A.43(c)).
Definition A.45 (Baire and Borel functions). If X, Y are topological spaces, then
f : X → Y is a Borel measurable mapping (or Borel mapping) if f −1 (U ) is Borel
for each U ⊂ Y open. If Y = R we speak about Borel functions.
We say that f is a Baire measurable mapping if f −1 (U ) is a Baire set for every
U ⊂ Y open. In the case when Y = R, it follows from Corollary 5.22 that every
Baire measurable function is of some Baire class as defined in Definition 5.20 and
thus we usually speak about Baire functions.
Proposition A.46. Let K be a compact space and let F be a vector space of bounded
functions on K satisfying
• F contains all bounded upper semicontinuous functions (see Definition A.49),
Proof. Set
B := {B ⊂ K Borel : cB ∈ F}.
It follows from the assumptions that B is a Dynkin system (see Definition A.67). Since
B contains all open sets, B contains all Borel sets by Proposition A.68.
If f is an arbitrary bounded Borel function, we may assume that 0 ≤ f ≤ 1. For
n ∈ N, let An,k := {x ∈ K : k−1 k n
2n < f (x) ≤ 2n }, k = 1, . . . , 2 , and
n
X k−1
fn := cAn,k .
2n
k=1
Proposition A.47. Let K be a compact space and let F be a vector space of bounded
functions on K satisfying
• F contains all continuous functions,
B := {B ⊂ K Baire : cB ∈ F}.
Then B is a Dynkin system and it is easy to see that B contains all cozero sets. Indeed,
if U ⊂ K is a cozero set, using Tietze’s theorem we construct fn ∈ C(K), n ∈ N,
such that 0 ≤ fn ≤ 1 and fn % cU . Hence cU ∈ F and U ∈ B.
Again by Proposition A.68, B contains all Baire sets and we may finish the proof
as in Proposition A.46.
Proposition A.48. Let A be a Baire set in a normal space X. Then there exist contin-
uous functions fn , n ∈ N, on X with values in [0, 1] such that A is contained in the
σ-algebra generated by the family {fn−1 ((0, 1]) : n ∈ N}.
Proof. Let A denote the family of all Baire sets satisfying the condition. Then A
contains all cozero sets by definition and is obviously closed with respect to taking
complements and countable unions. Thus A contains all Baire sets.
Proof. Proof of (i) ⇐⇒ (iii) is elementary and (i) ⇐⇒ (ii) follows from [169], Exer-
cise 1.7.15.
It is not difficult to realize that for any completely regular space X we have fb =
sup{g ∈ C(X) : g ≤ f }.
Proof. The implication (i) =⇒ (ii) is straightforward. For the proof of the converse
implication, we assume that f is positive. For any q ∈ Q∩(0, ∞), set Uq := {x ∈ X :
f (x) > q}. Using Urysohn’s lemma A.25 we find a sequence of positive continuous
functions {fn,q } such that fn,q % qcUq . Then
Proof. Without loss of generality, we may assume that {hn } is a strictly decreasing
sequence.
We fix n ∈ N. For each x in X, there exist a function gx ∈ F and an open neigh-
borhood U (x) of x such that f ≤ gx and gx < hn on U (x). The Lindelöf property
ensures the existence of a countable family {U (xm )} chosen from {U (x)}x∈X which
covers X. We relabel the corresponding functions {gxm } as {gn,m }∞ m=1 . Then we
have inf{gn,m : m ∈ N} < hn .
We enumerate the functions {gn,m : n, m ∈ N} as a single sequence {gn }. Then
f = inf{gn : n ∈ N}. Since F is down-directed, we may inductively find a sequence
{fn } such that g1 = f1 and fn+1 ≤ gn+1 ∧ fn . Then {fn } has the desired properties.
Lemma A.55. Let X be a topological space with a countable base of open sets and
F be a family of lower semicontinuous functions on X. Then there exists a countable
family F 0 ⊂ F such that sup F = sup F 0 .
Proof. For a number q ∈ Q, we use the existence of a countable base to find a count-
able family F q ⊂ F such that
[ [
{x ∈ X : f (x) > q} = {x ∈ X : f (x) > q}.
f ∈F f ∈F q
Definition A.59 (Meager sets and sets with the Baire property). A set A ⊂ X is
meager if A can be covered by countably many nowhere dense sets. It is residual
(sometimes comeager), if X \ A is meager.
The set A is said to have the Baire property, if A is contained in the σ-algebra
generated by Borel sets and meager sets. It has the Baire property in the restricted
sense if A ∩ F has the Baire property in F for every F ⊂ X.
A function f : X → R has the Baire property (or the Baire property in the re-
stricted sense) if f −1 (U ) has the Baire property (or the Baire property in the restricted
sense) for any open set U ⊂ R.
Proposition A.60. Any Borel set has the Baire property in the restricted sense.
If (P) is a property of a measure space (X, Σ, µ), where no confusion can arise we
sometimes say that the measure µ has property (P).
If (X, Σ, µ) is a measure space and Y is a topological space, a mapping f : X → Y
is µ-measurable (briefly measurable) if f −1 (U ) ∈ Σ for each U ⊂ Y open.
If Σ is a σ-algebra of sets in X and µ : Σ → R is σ-additive, it is called a signed
measure (analogously we define a complex measure). Any signed measure can be
decomposed as µ = µ+ − µ− , where µ+ , µ− is the positive and negative part of µ,
respectively. Moreover, the measures µ+ , µ− are finite (see [181], Corollary 231F).
We write |µ| for the variation of µ, that is,
X∞
|µ|(A) := sup{ |µ(Ai )| : {Ai } is a partition of A}, A ∈ Σ.
n=1
Then |µ|(A) = µ+ (A) + µ− (A) for any A ∈ Σ (see Section 231 of [181]). (Similarly
we define the variation of a complex measure µ.)
S a measure space, µ is σ-finite if there exist sets Xn ∈ Σ, n ∈ N,
If (X, Σ, µ) is
such that X = n∈N Xn and µ(Xn ) < ∞.
Definition A.62 (Inner and outer measure). If (X, Σ, µ) is a measure space, it induces
an inner measure µ∗ on the family 2X of all sets in X by the formula
Theorem
P∞ A.66. Let (X, Σ, µ) be a measure space and An ∈ Σ, n ∈ N, satisfy
n=1 µ(A n ) < ∞. Then the set {n ∈ N : x ∈ An } is finite for µ-almost all
x ∈ X.
Definition A.69 (Upper and lower integral). If (X, Σ, µ) is a measure space and f :
X → [−∞, ∞] is a function, the upper integral of f is defined as
Z ∗ Z
f dµ := inf g dµ : g is µ-integrable, f ≤ g µ-almost everywhere .
X X
The lower integral is defined analogously. We refer the reader to Proposition 133J of
[182] for more information on this subject.
Remark A.74 (Inner and outer regularity of Radon measures). Let X be a locally
compact σ-compact space and µ be a measure defined on Borel sets in X that is
finite on compact sets. It is not difficult to realize that µ is inner regular with respect
to compact sets if and only if it is outer regular with respect to open sets (that is,
µ(A) = inf{µ(G) : G ⊃ A open} for A ⊂ X Borel).
Indeed, let Xn ⊂ X, n ∈ N, be compact sets with X = ∞
S
n=1 Xn . If µ is inner
regular with respect to compact sets, A ⊂ X is Borel and ε > 0, let Kn ⊂ Xn \ A be
628 A Appendix
compact sets such that µ((Xn \ A) \ Kn ) < 2−n ε, n ∈ N. Let Un ⊂ X be open such
that µ(Un \ Xn ) < 2−n ε, n ∈ N. Then
∞
[
U := Un ∩ (X \ Kn )
n=1
Theorem A.75 (Riesz representation theorem for (C 0 (X))∗ ). Let X be a locally com-
pact σ-compact space and let C 0 (X) denote the Banach space of continuous functions
vanishing at infinity. Then for anyR T ∈ (C 0 (X))∗ there exists a unique signed Radon
measure µ on X such that T f = X f dµ, f ∈ C 0 (X).
Proof. The assertion is essentially Proposition 437I of [183] (or, by Remark A.74, we
can use Theorem 6.19 from [402]). We only explain how to get the complete measure
representing T .
Given T ∈ (C 0 (X))∗ , we use the aforementioned Proposition 437I of [183] to find
a signed measure ν defined on the σ-algebra B of all Borel sets in X such that
• if ν + and ν − are the positive and negative parts of ν, respectively, then they are
inner regular with respect to compact sets,
R
• T f = X f dν for any f ∈ C 0 (X).
Let (X, Σ+ , µ+ ) and (X, Σ− , µ− ) be the unique Radon measures extending ν + and
ν − , respectively (see Proposition A.73). Let Σ := Σ+R ∩ Σ− and µ := µ+ − µ− on Σ.
Then (X, Σ, µ) is a Radon measure space and T f = X f dµ for any f ∈ C 0 (X).
a finite Radon measure µ is discrete if and only if there exists a sequence of positive
numbers {λj } and points xj ∈ X such that
∞
X ∞
X
λj < ∞ and µ= λ j εxj .
j=1 j=1
Pk
A Radon measure µ is molecular if µ = j=1 λj εxj , where x1 , . . . , xk ∈ X,
Pk
λ1 , . . . , λk are positive and j=1 λj = 1.
A Radon measure µ is called continuous if µ({x}) = 0 for each x ∈ X.
µ = µd + µc ,
Lemma A.81. Let µ and ν be Radon measures on a locally compact σ-compact space
X. Then spt(µ + ν) = spt µ ∪ spt ν.
a way to understand M(X) as a Banach space with the norm kµk = |µ|(X) for any
µ ∈ M(X) (see Proposition 437I of [183]). Then µ ∈ M(X) is a limit of a sequence
{µn } in M(X), if and only if |µ − µn |(A) → 0 for every Borel set A in X.
Further, if we define an order “≤” on M(X) by
then M(X) becomes an ordered RBanach space. Moreover, the mapping T : M(X)
→ (C 0 (X))∗ given by T µ(f ) = X f dµ, f ∈ C 0 (X), is an isometric isomorphism
preserving order. (We recall that (C 0 (X))∗ is an ordered Banach space as defined
in Subsection A.1.C that is even a lattice.) Hence it follows from Theorem A.24
that M(X) is a lattice. Further, if µ ∈ M(X), then the positive part µ+ of µ is
precisely the supremum µ ∨ 0 in the ordered Banach space M(X). Hence the cone
of all positive finite Radon measures coincides with the cone of all positive elements
in (C 0 (X))∗ . There is a strong relation between the order and the norm on M(X),
namely kµk = k|µ|k, µ ∈ M(X), and kµ + νk = kµk + kνk for µ, ν ∈ M+ (X).
Remark A.83. The space C(K) of continuous functions on a compact space K, the
space C 0 (X) of continuous functions vanishing at infinity on a locally compact space
X and M(X) are examples of Banach lattices (see Section 5 in [251]).
Theorem A.84 (Lebesgue monotone convergence theorem for nets). Let µ be a Radon
measure on a compact space K and F a nonempty up-directed set of lower semicon-
tinuous functions on K. Then
Z Z
(sup F) dµ = sup f dµ : f ∈ F .
K K
Theorem A.85. Let X be a locally compact σ-compact space and M(X) be identified
with (C 0 (X))∗ as in Definition A.82. Consider M(X) to be endowed with the w∗ -
topology.
(a) BM(X) and M1 (X) are compact convex sets.
(b) If f is lower semicontinuous function on X, then µ 7→ µ(f ), µ ∈ M1 (X), is a
lower semicontinuous function on M1 (X).
(c) If X is metrizable, BM(X) and M1 (X) are metrizable as well.
Proof. The dual unit ball BM(X) is compact by Theorem A.5 and M1 (X) is compact
because
M1 (X) = {µ ∈ BM(X) : µ ≥ 0, µ(1) = 1}.
Further, (b) follows from Theorem A.84 and (c) from Theorem A.5 and Proposi-
tion A.33.
A.3 Measure theory 631
Lemma A.90. Let X be a locally compact σ-compact space and µ be a Radon mea-
sure on X. If f : X → [−∞, ∞] is a function, then
Z ∗
f dµ = inf{µ(g) : g ≥ f, g lower semicontinuous}.
X
R∗
Proof. If X f dµ = ∞, it is enough to take g = ∞. We assume that h ≥ f µ-
almost everywhere and it is µ-integrable. Let ε > 0. Using Theorem A.76, we find
an increasing S∞sequence {Kn } of compact sets in X such that h|Kn is continuous
and µ(X \ n=1 Kn ) = 0. For n ∈ N, we find open sets Gn ⊃ Kn such that
< 2−n ε. Let hn be a continuous function on X such that |hn | ≤ |h|
R
Gn \Kn |h| dµ
on X, hn = h on Kn and hn = 0 on X \ Gn .
Then g1 := supn∈N hn is a lower semicontinuous function on X such that g1 ≥ h
µ-almost everywhere. Let N be a set with µ(N ) = 0 such that h ≥ f and T∞g1 ≥ h on
X \ N . Let {Un } be a decreasing sequence
P∞ of open sets such that N ⊂ n=1 Un and
−n
µ(Un ) < ε2 , n ∈ N. Then g2 := n=1 cUn is lower semicontinuous on X and
g := g1 + g2 is a lower semicontinuous function satisfying g ≥ f . Since
Z Z Xn Z
(h1 ∨ · · · ∨ hn ) dµ ≤ h dµ + |h| dµ, n ∈ N,
X Kn i=1 Gi \Ki
we get
Z Z
µ(g) = µ(g2 ) + lim (h1 ∨ · · · ∨ hn ) dµ ≤ h dµ + 2ε.
n→∞ X X
This finishes the proof.
Proof. For the surjectivity of ϕ] , see Theorem 418L in [183]. The verification of the
w∗ -continuity is straightforward.
defined as ϕ((xi )i∈I ) = ((xi )i∈Ij )j∈J identifies i∈I µi with j∈J µj .
N N
Theorem A.97 (Fubini’s theorem). Let (K1 , Σ1 , µ1 ) and (K2 , Σ2 , µ2 ) be Radon mea-
Nspaces on compact spaces K1 and K2 , respectively. If f : K1 × K2 → R is a
sure
µ1 µ2 -integrable function, then
Z Z Z
f (x, y) d(µ1 ⊗ µ2 )(x, y) = f (x, y) dµ2 (y) dµ1 (x)
K1 ×K2 K1 K2
Z Z
= f (x, y) dµ1 (x) dµ2 (y).
K2 K1
Q
Proof. Let U be the family of all finite unions of sets of the form i∈I Ui , where
Q
Ui ⊂ Ki are open and Ui 6= Ki for finitely many i ∈ I. If Fb ⊂ i∈I Ki is compact
and Vb ⊃ Fb is open, we can find an element U ∈ U with Fb ⊂ U ⊂ Vb . Given A ⊂ K
µ-measurable, by regularity of µ there are F ⊂ A ⊂ V such that F is closed, V open
and µ(V \ F ) < 12 ε. By the above reasoning we choose U ∈ U such that F ⊂ U ⊂ V
and get the sought set.
Q
Proposition
Q A.99. If two Radon measures on i∈I Ki coincide on the cylinder sets
i∈I Ui , where Ui ⊂ Ki are open and Ui 6= Ki for finitely many i ∈ I, then they are
equal.
Q
Lemma A.100. Let K := i∈I Ki be the product of compact spaces Ki , i ∈ I. Then
Y
0 = µ(Gj × Ki ) = ((πj )] µ)(Gj ),
i∈I\{j}
Y
0 = ((πj )] µ)(Gj ) = µ(Gj × Ki ),
i∈I\{j}
Definition A.101 (Inverse limits of Radon measures). Let (Ki , pij )i,j∈I be an inverse
system of compact spaces and (Ki , Σi , µi )i∈I be Radon probability spaces such that
(pij )] µj = µi , i ≤ j. By Proposition 418M of [183], there exists a Radon proba-
bility measure µ on K = lim(Ki , pij )i,j∈I such that (πi )] µ = µi for each i ∈ I.
←
Lemma A.102 below shows that µ is uniquely determined.
We call the family ((Ki , Σi , µi )i∈I , (pij )i,j∈I ) the inverse system of measures and
µ the inverse limit of this system. Sometimes we write (µi , pij )i,j∈I for the inverse
system and lim µi for the inverse limit.
←
Lemma A.102. Let ((Ki , Σi , µi )i∈I , (pij )i,j∈I ) be an inverse system of Radon prob-
ability spaces. Then there is a unique Radon probability measure µ on K := lim Ki
←
such that (πi )] µ = µi for all i ∈ I.
By Proposition A.99, µ = ν.
Lemma A.103. Let (Ki , pij )i,j∈I be an inverse system of compact spaces and let µ be
a Radon probability measure on lim Ki . Then ((πi )] µ, pij )i,j∈I is an inverse system
←
of measures and µ = lim(πi )] µ.
←
636 A Appendix
Proof. To check that ((πi )] µ, pij )i,j∈I is an inverse system is easy and the fact µ =
lim µi follows from Lemma A.102.
←
Lemma A.104. Let (Ki , pij )i,j∈I be an inverse system of compact spaces and let µ be
a continuous Radon probability measure on lim Ki . Then for each δ > 0 there exists
←
i ∈ I such that ((πj )] µ)({x}) < δ for each j ≥ i and x ∈ Kj .
Proof. For each i ∈ I, let
J := {i ∈ I : Li 6= ∅}
is cofinal (that is, for every i ∈ I there exists j ∈ J such that i ≤ j). Obviously, each
set Li is finite and, moreover,
pij (Lj ) ⊂ Li , i ≤ j, i, j ∈ J.
Hence we can find x := (xi )i∈I ∈ K such that xi ∈ Li for each i ∈ J. Then
\
{x} = πi−1 (xi )
i∈J
and
δ ≤ ((πi )] µ)({xi }) = µ(π −1 (xi )), i ∈ J.
Thus µ({x}) ≥ δ, a contradiction.
We can extend the domain of the mapping T to all bounded universally measurable
functions on X by defining
Z
T f (x) := f (t) dTx (t), x ∈ X.
X
We use the same letter T for the mapping assigning a bounded universally measurable
function f ∈ U(X) the function T f .
We can further consider the mapping T to be defined on the space M(X) by the
formula Z
T µ(f ) := Tx (f ) dµ(x), f ∈ C 0 (X).
X
By Theorem A.75, T µ is a well-defined signed Radon measure. Again we use the
letter T for the mapping T : M(X) → M(X).
We remark that T : M(X) → M(X) is positive if and only if T : U(X) →
`∞ (X) is positive and this is the case if and only if Tx ≥ 0 for all x ∈ X.
Theorem A.106 (Disintegration theorem). Let K, L be compact spaces and µ be a
Radon probability measure on K × L. Let π : K × L → K be the projection. Then
there exists a family {Tx }x∈K of Radon probability measures on L such that
Z Z Z
f dµ = f (x, y) dTx (y) d((π)] µ)(x), f ∈ L1 (µ).
K×L K L
Proof. This follows from Proposition 452O in [183] applied to the projection π.
ϕ−1 (U ) := {x ∈ X : ϕ(x) ⊂ U }
is open in X for each U ⊂ Y open.
638 A Appendix
We recall that NN stands for the usual product space endowed with the pointwise
topology and N<N for the set of all finite sequences of natural numbers. If σ :=
(s1 , s2 , . . . ) ∈ NN and n ∈ N, we write σ|n for the finite sequence (s1 , . . . , sn ). The
length of a finite sequence s is denoted as |s|. A similar notation is used for {0, 1}N .
Proof. For the proof of (a) and (b) see Section 2.5 of [394], (c) can be found in
Section 2.6 of [394], (d) follows from the proof of Theorem 2.7.1 in [394] and (e)
in Section 5.1, p. 96 of [394]. Assertion (f) follows from Theorem 5.5.1 and Theo-
rem 5.5.2 of [394], Section 2.9 of [394] yields (g) and (h) follows from Section 2.4 of
[394] and (a).
If X is a compact space, let F be the smallest family containing compact sets and
stable with respect to countable unions and intersections. By (a), every member of F
is K-analytic. Since every Baire subset of X is contained in F, (i) follows.
Inductively we define for α ∈ (0, ω1 ) the sets of additive and multiplicative class α
as
∞
[
Σ0α (X) := An : An ∈ Π0αn (X), αn < α, n ∈ N ,
n=1
\∞
Π0α (X) := An : An ∈ Σ0αn (X), αn < α, n ∈ N .
n=1
Thus Σ01 (X) are Fσ sets and Π01 (X) are Gδ sets in X. Also,
[ [
Σ0α (X) = Π0α (X)
α<ω1 α<ω1
It follows from §12, II and V in [285] that for a resolvable set A ⊂ X there exist
an ordinal κ and an increasing transfinite sequence of open sets
∅ = U0 ⊂ U1 ⊂ U2 ⊂ · · · ⊂ Uα ⊂ · · · ⊂ Uκ = X,
S
such that {Uα : α < λ} S = Uλ for every limit ordinal λ ∈ [0, κ] and there exists
I ⊂ [0, κ) such that A = {Uα+1 \ Uα : α ∈ I}.
It is easy to see that an equivalent description of a resolvable set H is a requirement
that for every nonempty closed set A ⊂ X there exists a relatively open set U ⊂ A
such that either U ⊂ H or U ⊂ A \ H. We refer the reader to §12, VI of [285] for the
basic properties of resolvable sets.
Proof. The proof of (a) can be found in §12, VI of [285], (b) is easy to observe, (c)
readily follows from the definition and (d) is proved in §12, VI in [285].
If A ⊂ X is resolvable, then A \ Int A is resolvable by (a), and thus it is nowhere
dense. Thus if A is a dense resolvable set, then A contains a dense open set and A is
residual.
Assertion (f) is proved in §26, X of [285] and (g) is easy to verify. Indeed, if F ⊂ X
is a closed set, then both A and F \ A cannot be dense in F due to the Baire category
theorem. Hence the conclusion follows.
Then [
B := (A ∩ Uβ )
β∈J
is µ-measurable,
[ [ [ [
B ⊂A∩ Uβ and A∩ Uβ \B ⊂ Uβ \ Uβ .
β<α β<α β<α β∈J
S
Since the latter set is of µ-measure 0, A ∩ β<α Uβ \B is of µ-measure 0, and thus
[
A = B ∪ (A ∩ Uβ ) \ B
β<α
Remark A.125. It is easy to see that sequences {ln }, {un } from Theorem A.124(iii)
can be chosen to be bounded if f is bounded.
Proposition A.126. If X is a topological space, then the family of all Baire-one func-
tions on X is a vector lattice (in the pointwise ordering), algebra (with respect to the
pointwise multiplication), contains constant functions and is closed with respect to
uniform limits. Moreover, f /g is Baire-one provided f , g are Baire-one and g(x) 6= 0
for all x ∈ X.
644 A Appendix
Proof. The stability of Baire-one functions with respect to addition, lattice operations
and pointwise multiplication easily follows from Definition A.123. Obviously, it con-
tains constant functions. If f , g are Baire-one and g(x) 6= 0 for all x ∈ X, let {fn },
{gn } be sequence of continuous functions on X with fn → f and gn → g. Then
fn gn f
→ ,
gn2 + n−1 g
and f /g is Baire-one.
If {fn } is a sequence of Baire-one functions uniformly converging to a function f ,
we extract a subsequence (denoted likewise) such that kfn+1 − fn k < 2−n . For each
n ∈ N, let {fnk }k∈N be sequences of continuous functions on X pointwise converging
to fn+1 − fn . Without loss of generality we may assume that kfnk k ≤ 2−n for each
n, k ∈ N. Then
Xn
sn := fnk , n ∈ N,
k=1
is continuous and
∞
X
lim sn = (fn+1 − fn ) = f − f1 .
n→∞
n=1
Thus f is Baire-one.
Proof. Assertions (i) ⇐⇒ (ii) ⇐⇒ (iii) follow from Theorem A.121 since X is hered-
itarily Baire. The equivalence (i) ⇐⇒ (iv) follows by Proposition A.117(f) and The-
orem A.121. Finally, Theorem A.124 yields (iv) ⇐⇒ (v) since X is normal.
Since finite sets in K are Gδ sets, gn are Baire-one functions by Theorem A.127(iv).
Since {gn } converges to g uniformly, g is Baire-one by Proposition A.126.
∂2 ∂2
∆ := + · · · + .
∂x21 ∂x2d
Definition A.130 (Dirichlet problem). For the theory of the Laplace equation, a prin-
cipal problem is to find a solution in an open set which attains given boundary data.
This is the so-called Dirichlet problem. Let U ⊂ Rd be a bounded open set. Given a
continuous function f on ∂U , a function h ∈ H (U ) is called a classical solution of
the Dirichlet problem (on U with boundary data f ) if h has a continuous extension to
U and this coincides with f on ∂U .
Definition A.131 (Poisson kernel and integral). For every U (x, r), the function
r2 − |y − x|2
Px,r : (y, z) 7→ rd−2 , z ∈ S(x, r), y ∈ U (x, r),
|y − z|d
is called the Poisson kernel of U (x, r). Given a function f on S(x, r), the integral
Z
Px,r f (y) := − Px,r (y, z)f (z) dS(z), y ∈ U (x, r),
S(x,r)
Proposition A.132. Let U (x, r) be an open ball in Rd . Then for each f ∈ C(S(x, r))
there exists a unique solution of the Dirichlet problem on U (x, r) with boundary data
f and this is given by the Poisson integral Px,r f .
1 log 1 ,
x ∈ R2 \ {0} ,
N (x) := σ2 |x|
∞, x = 0.
This is the so-called logarithmic kernel. For d ≥ 3, define the Newtonian kernel by
1
|x|2−d , x ∈ Rd \ {0} ,
N (x) := (d − 2)σd
∞, x = 0.
The kernel is also called the fundamental harmonic function with pole at 0. Given a
Radon measure µ in Rd satisfying
Z
|N (y)| dµ(y) < ∞,
Rd \B(0,1)
A.6 The Laplace equation 647
Definition A.138 (Sobolev spaces). (See D. Gilbarg and N. S. Trudinger [193], Chap-
ter 7.) Let U ⊂ Rd be an open set. We write V ⊂⊂ U if V is a bounded open
set and V ⊂ U . The Sobolev space W 1,2 (U ) consists of all locally integrable func-
tions on U which together with their first distributional derivatives belong to L2 (U ).
1/2
The Sobolev norm of a function u ∈ W 1,2 (U ) is U (u2 + |∇u|2 ) dx
R
. The lo-
1,2 1,2
cal version of this space is denoted by Wloc (U ). More precisely, u ∈ Wloc (U ) if
u ∈ W 1,2 (V ) for each V ⊂⊂ U . We identify two Sobolev functions which differ
on a set of measure zero. A function from a Lebesgue or a Sobolev space which is
considered set-theoretically (without the convention on identification) is called a rep-
resentative. The space Wc1,2 (U ) is defined as the space of all functions u ∈ W 1,2 (U )
which can be represented as compactly supported in U . The closure of Wc1,2 (U ) in
W 1,2 (U ) is denoted by W01,2 (U ); this is the space of functions with “zero on the
boundary”. The set Cc∞ (U ) (the space of infinitely differentiable functions with com-
pact support in U ) is dense in W01,2 (U ). If two functions differ by a function from
W01,2 (U ), it has the meaning that they “coincide on the boundary”. Thus the quotient
space W 1,2 (U )/W01,2 (U ) is a natural choice for the space of boundary data for the
Dirichlet problem.
In view of the following proposition, we will use the notion “weakly harmonic func-
tion” for a harmonic function modulo modifications on sets of measure zero.
For a Sobolev function u, U (z, r0 ) ⊂⊂ U and 0 < r < R < r0 , we then have
Z R Z
AR u(z) − Ar u(z) = − ∇u(z + ty) · y dy dt.
r U (0,1)
Indeed, this is clear for smooth functions and the general case follows by mollification.
Now, if u is weakly harmonic, then we test by
1
v(x) := (|x − z|2 − t2 ) ∈ W01,2 (U (z, t)), x ∈ U (z, t),
2t
and obtain
Z Z
1
− ∇u(z + ty) · y dy = − ∇u(x) · ∇v(x) dx = 0.
U (0,1) tm U (z,t)
It follows that t 7→ At u(z) is constant. If z is a Lebesgue point of u, we infer that the
mean value property holds for z, namely that Ar u(z) = u(z). Now, it is not difficult
to see that the mean value operator is smoothing: if u is locally integrable, then Ar u
is continuous (on its domain {x ∈ U : U (x, r) ⊂⊂ U }), Ar Ar u is C 1 , Ar Ar Ar u is
C 2 and so on. Hence by induction we obtain infinite smoothness of a representative of
a weakly harmonic function.
A.7 The heat equation 649
Definition A.142 (The Dirichlet problem and the Dirichlet integral). From the point
of view of the calculus of variations, a solution of the Dirichlet problem is the func-
tion which minimizes the Dirichlet integral among all functions satisfying the bound-
ary data. Throughout the rest of this subsection we will assume that U is bounded.
To begin with, we interpret the boundary data as the elements of the quotient space
W 1,2 (U )/W01,2 (U ). In other words, we look for a solution u of
Z nZ o
|∇u(x)|2 dx = min |∇v(x)|2 dx : v ∈ W 1,2 (U ), v − u0 ∈ W01,2 (U ) (A.2)
U U
where a ∈ Rd . (We consider the real and imaginary parts of ga to produce real-valued
examples.) The function ga is bounded on Rd × (0, ∞). The canonical example is the
heat kernel
1 |x|2
exp − , t > 0,
u(x, t) := (4πt)d/2 4t (A.4)
t ≤ 0,
0,
which plays the role analogous to that of the Newtonian kernel. This function is
caloric outside the origin (0, 0). Convolutions with the heat kernel are used to produce
“heat potentials”, in order to solve the equation − ∩ h = µ or to solve the Cauchy
initial value problem for ∩ h = 0 on Rd × (0, ∞).
Theorem A.146. The space H + (Rd+1 ) contains positive constants and separates
points of Rd+1 .
Proof. Obviously, constants are caloric. To separate points we will use the functions
ha as in (A.3). Choose points (x, t), (x0 , t0 ) ∈ Rd+1 and a ∈ Rd with |a| = 1. If
ha (x, t) = ha (x0 , t0 ) and h−a (x, t) = h−a (x0 , t0 ), then
a · x + t = a · x0 + t0 ,
−a · x + t = −a · x0 + t0 .
Adding and subtracting these equations we obtain
t = t0 , a · x = a · x0 .
It follows that also x = x0 (otherwise we would specify a as
x0 − x
a=
|x0 − x|
to obtain a contradiction). Hence at least one of the functions ha separates the given
points.
Corollary A.148 (Minimum principle). Let U ⊂ Rd+1 be a bounded open set and
h ∈ H (U ). If
lim inf h(x, y) ≥ 0 for each (ξ, τ ) ∈ ∂U,
(x,t)→(ξ,τ )
then u ≥ 0 on U .
Proof. Assume the contrary. Then the zero extension of h to the compact set U is
lower semicontinuous and hence there exists (x, t) ∈ U such that
h(x, t) = min h(U ) < 0.
From the strong minimum principle (Proposition A.147) we deduce that h is equal to
a negative constant on the connected component of (x, t) in U , which contradicts the
limes inferior assumption.
Remark A.149. In literature, the minimum principles are usually stated for the so-
called parabolic boundary. However, in this presentation we prefer an approach
which is consistent with a general potential theory and not specific to parabolic prob-
lems. This point of view is reflected also in the following treatment of the “classical”
Dirichlet problem.
Theorem A.150. Given (ξ, τ ) ∈ Rd+1 , let U be the ball
U ((ξ, τ ), r) := {(x, t) ∈ Rd+1 : |x − ξ|2 + |t − τ |2 < r2 }.
Then for each f ∈ C(∂U ) there exists a unique function h ∈ C(U ) such that h is
caloric on U and h|∂U = f . (In other words, the “caloric Dirichlet problem” with
the boundary data f is solvable.)
Proof. See [301], Theorem 3.18.
Proposition A.151 (Harnack type inequality). Let R, T > 0, r < R, 4r2 < T . Then
there exists a constant c such that
h(ξ, τ ) ≤ ch(x, t)
for each positive h ∈ H (U (0, R) × (0, T )), (x, t) ∈ U (0, r) × (T − r2 , T ) and
(ξ, τ ) ∈ U (0, 12 r) × (T − 2r2 , T − r2 ].
Proof. See [301], Corollary 7.27.
Proof. We easily deduce from the Harnack inequality A.151 that the limit function h
is locally bounded. From Proposition A.152 we deduce that the sequence is locally
equicontinuous, so that by the Arzelà–Ascoli theorem A.32, the sequence converges
locally uniformly and the limit h is continuous. If V is a ball with V ⊂ U , then h
is caloric on V since, by Proposition A.153, H (V ) is closed under uniform conver-
gence. Hence h is caloric on U .
f ≥ 0 on ∂U =⇒ Hf ≥ 0 on U
holds.
A.8 Axiomatic potential theory 653
Proposition A.158. Each Bauer harmonic space Y in the sense of Definition A.156
satisfies the axioms of balayage spaces of [66].
Proof. According to [40], Korollar 4.2.7, any nonempty intersection of two regular
sets is again a regular set. The Doob’s convergence property is evidently stronger
than “Bauer’s convergence property”. It follows that Y is a normal Bauer space in the
sense of Section VIII.1 in [66]. By [66], Corollary 1.2, Y is a harmonic space in the
sense of [66], Section III.8. However, these harmonic spaces are balayage spaces by
definition.
Examples A.159. (a) The sheaf of harmonic functions (in the classical sense, that is,
C 2 -solutions of the Laplace equation) on an open set Y ⊂ Rd , d ≥ 1, satisfies Bauer’s
axioms (a)–(d), as shown in Section A.6. (For the base axiom, we notice that any open
ball is a regular set.) If the remaining separation axiom (e) is satisfied on Y , we say
that Y is a Greenian set. This is the case, for instance, if d ≥ 3 or Y is bounded. See
also [40], I.§2, II.§5.
(b) In the course of Section A.7 we have shown that all Bauer’s axioms are verified by
the sheaf of caloric functions on Rd+1 , see also [40], I.§2 and [66], Section V.6. We
will continue to call these functions “caloric” when dealing specifically with the heat
equation, but “harmonic” if we think of them in the context of Bauer harmonic spaces.
Terminological derivatives like “supercaloric function” (that is, superharmonic for the
heat equation) will be also used in the sequel.
(c) The backward Kolmogorov operator
∂2 ∂ ∂
2
+x −
∂x ∂y ∂t
generates a Bauer harmonic space on R3 ; for details see the paper by W. Hoh and
N. Jacob [228].
(d) For further examples of Bauer harmonic spaces and references see, for example,
H. Bauer [43], C. Constantinescu and A. Cornea [123] and [66], Chapter VIII.
The following assertions easily follow from the definition of hyperharmonic func-
tions and the Lebesgue monotone convergence theorem A.84.
∗
Theorem A.161. The family H (U ) forms a min-stable convex cone of functions. If
∗ ∗
F ⊂ H (U ) is a nonempty up-directed set, then sup F ∈ H (U ).
is called the Poisson modification of s. By [40], Satz 2.2.2, the Poisson modification
sV is always hyperharmonic. If s is superharmonic on U , then sV is superharmonic
on U and harmonic on V (see [40], Korollar 2.3.2).
∗
Definition A.164 (Saturated families). A nonempty family F ⊂ H (U ) is called sat-
urated if F is min-stable and there exists a family V of open subsets of U which forms
a base of the topology on U and the following condition is satisfied: If V ∈ V, and
u ∈ F, then uV ∈ F.
Proof. Let V be the family of open sets as in the definition of a saturated family.
Given V ∈ V, the family
G := {u|V : u ∈ F, u is harmonic on V }
Proof. Set
A.8.C Potentials
In this subsection, Y will be a Bauer harmonic space in the sense of Definition A.156.
V := {(p − q)|K : p, q ∈ P}
Proof. This is a standard application of the Stone–Weierstrass theorem A.30, see also
[40], Satz 2.7.4, [66], Proposition I.1.2, Corollary III.6.10.
Proof. This follows from [66], Lemma VII.5.1. Note that the function
q := inf{s ∈ S + (Y ) : q ≥ f }
where R − r d
c= .
R − 3r
Proof. If x, x0 ∈ B(z, r), then B(x, R − 3r) ⊂ B(x0 , R − r) ⊂ U , and thus
Z R − r dZ
h(x) = − h(y) dy ≤ − h(y) dy = ch(x0 ).
B(x,R−3r) R − 3r 0
B(x ,R−r)
Proof. If x ∈ K, then x can be joined with z by a chain of balls for which the
inequality (A.5) holds, so that h(y) ≤ cx h(z) for y in a neighborhood of x. Now the
claim follows by the compactness of K.
658 A Appendix
Definition A.177 (Green potentials). In this text, we want to have a unified treatment
of the potential theories under consideration (namely, we mean axiomatic potential
theory covering the Laplace equation and the heat equation). We can skip the one-
dimensional Laplace equation theory because it is not very interesting. However, in
the two-dimensional case we encounter the difficulty that the entire space does not
satisfy the separation axiom A.156(e). Therefore, we are frequently forced to restrict
our attention to Greenian sets.
Recall that an open set Y ⊂ Rd is called Greenian if the cone S + (Y ) separates
points of Y . Each open set Y ⊂ Rd is Greenian if d ≥ 3; we can separate points
by the family {N a : a ∈ Y }, where N a (x) := N (x − a), x ∈ Rd . On a bounded
open set Y ⊂ R2 we can separate points by the family {N a + c : a ∈ Y }, where c is
a suitable constant. The full characterization of Greenian sets is given by Myrberg’s
theorem (see [21], Theorem 5.3.8).
Throughout Chapter 13 we deal with Dirichlet problem on a relatively compact set
U , so that we can always insert a Greenian set Y between U and Rd .
Let Y ⊂ Rd be a Greenian set. For each y ∈ Y we find the greatest harmonic
minorant hy of the function x 7→ N (x − y), x ∈ Y . (By Proposition A.166, this
is possible if, for instance, d ≥ 3 or Y is bounded. The general case follows from
Theorem 5.3.8 of [21]. Note that we have made a different choice from the equivalent
conditions for our definition of Greenian sets.) The function G : Y × Y → [0, ∞]
defined as
G(x, y) := N (x − y) − hy (x), (x, y) ∈ Y × Y,
is called the Green function for Y . Given a Radon measure µ in Y , we define the
potential Gµ by Z
Gµ(x) := G(x, y) dµ(y), x ∈ Y.
Y
Remark A.183. For a more general statement, the so-called Maria–Frostman domi-
nation principle, see [21], Theorem 5.1.11.
Theorem A.184 (Evans–Vasilesco continuity principle). Let Y be a Greenian set and
µ be a measure on Y . If the restriction of Gµ to spt µ is continuous, then Gµ is
continuous on Y .
Proof. See [21], Theorem 4.5.1.
660 A Appendix
Lemma A.187 (Minimum principle for the Cauchy problem). Let T > 0 and u be a
lower bounded lower semicontinuous function on Rd × [0, T ), which is supercaloric
on Rd × (0, T ). If u ≥ 0 on Rd × {0}, then u ≥ 0 on Rd × (0, T ).
Proof. Consider the auxiliary supercaloric function
1
v(x, t) := 2dt + |x|2 + , (x, t) ∈ Rd × (0, T ).
T −t
Given ε > 0 and R > 0, there exists r > R such that
lim inf u(x, t) + εv(x, t) ≥ 0, (y, s) ∈ ∂(B(0, r) × (0, T )).
(x,t)→(y,s),
(x,t)∈B(0,r)×(0,T )
Hence Theorem A.160 implies that u+εv ≥ 0 on B(0, r)×(0, T ) ⊃ B(0, R)×(0, T ).
Letting ε → 0 and R → ∞ we obtain the assertion.
A.8.F Balayage
This subsection is again situated in the full generality of Bauer harmonic spaces (in
the sense of Definition A.156).
∗
Proposition A.189. Let W ⊂ H + (Y ) and u := inf W. Then the lower semicontinu-
ous regularization u
b of u is hyperharmonic (in fact, this is the greatest hyperharmonic
minorant of W).
the reduced function or réduite of u on A. Since, in general, the reduced function RuA
is not lower semicontinuous, we pass to the lower semicontinuous regularization of
the reduced function (compare with Definition A.51) and define the balayage of u on
A as its regularization RbA . The function RbA is hyperharmonic on Y by Proposition
u u
A.189 and obviously satisfies 0 ≤ R bA ≤ RA ≤ u on Y .
u u
It is useful to know that
bA (x) = lim inf RA (x),
R (A.6)
u y→x u
The proof of this fact relies on an abstract Choquet’s lemma; see [66], Proposition
I.1.4.
Note that the operation of balayage is not restricted to Dirac measures; this is only
a matter of simplification motivated by the fact that we do not need to sweep out more
general measures.
∗
Proposition A.192. Let V ⊂ Y be a regular set and u ∈ H + (Y ). Then uV =
c
RuV = RbuV c .
c
Corollary A.193. If V ⊂ Y is a regular set and x ∈ V , then µVx = εVx .
∗
Theorem A.194. Let U ⊂ Y be a relatively compact open set, u ∈ H (Y ) and x ∈ U .
Then Z
c
u dεUx ≤ u(x).
∂U
bA = sup{R
R bK : K compact, K ⊂ A}.
p p
Proof. See [66], Proposition VI.1.9. Note that in the statement of the proposition to
which we refer, the set A is supposed to be Borel. However, this assumption is used
only to conclude that it is analytic.
Proof. Recall that we assume that constants are harmonic. Assume first that ∂U = ∅.
Then U is open and the function u = −1 is hyperharmonic on U . Using the minimum
principle (Theorem A.160), we arrive at a contradiction.
c
Now, in the general case, by [66], Proposition VI.2.1, εUx (Int U ) = 0. By [66],
c Vc
Proposition VI.2.2, there exists an Fσ set V ⊂ U such that x ∈ V and εUx = εx . Let
W be a relatively compact open set containing U . Using Theorem A.197 and Lemma
A.54 we can find an increasing sequence {Kj } of compact subsets of V c such that
c
bpKj
bV c = sup R bpKj ∪W
bV c = sup R
Rp and thus also Rp
j∈N j∈N
for each p from a countable subset P 0 of P with the property that P 0 − P 0 is dense
Vc c c
in C(W ). Then we deduce that εxj → εVx = εU x , where Vj = W \ Kj . By [66],
A.8 Axiomatic potential theory 663
Vc
Proposition III.8.1, the measures εxj are carried by V j ⊂ W and thus passing to the
c Uc
limit we infer that εUx is carried by W . Varying W we conclude that εx is carried
by U .
c
Finally we will prove that εxU is a probability measure. To this end, we use Propo-
sition A.168 to find a potential p0 ∈ P such that p0 > 1 on W and set p = p0 ∧ 1.
∗
If u ∈ H + (Y ) is such that u ≥ p on U c , then the minimum principle (Theorem
c
A.160) applied to u|W shows that u ≥ 1 on W . Hence RpU ≥ 1 on W . The converse
inequality and passing from reduction to balayage are obvious, so that R bpU c = 1 on
W . It follows that
1=R bpU c (x) = εU c Uc
x (p) = εx (1),
c
so that εU
x is a probability measure.
bU c (x) = inf{R
R bG (x) : G open, G ⊃ U c }.
p p
Proof. For (a) and (b) see [66], Proposition VI.4.1, (c) is in [66], Corollary VI.4.2.
Proof. By Proposition A.206, R bpA is lower semicontinuous and thus of the first Baire
class. Hence the assertion follows from Theorem A.124. See also [66], Proposition
VI.4.1.
Proof. For the second equality see [66], Corollary VI.4.17. Thus it follows that RpA
is lower semicontinuous, and hence the first equality holds.
Definition A.210 (Polar sets). In potential theory, polar sets play a role of exceptional
sets. A set A ⊂ Y is called polar if there is a superharmonic function v ≥ 0 on Y
such that A ⊂ {x ∈ Y : v(x) = ∞}. By [66], Proposition VI.5.3, the collection of
all polar sets forms a σ-ideal of subsets of Y .
We recall that a family I of subsets of a set X is a σ-ideal if it stable with respect
to countable unions and contains with its every element all its subsets.
bpA = 0 for each p ∈ P.
Proposition A.211. A set A ⊂ Y is polar if and only if R
Definition A.212 (Totally thin and semipolar sets). We say that A ⊂ Y is totally thin
if b(A) = ∅. Countable unions of totally thin sets are called semipolar sets. The
collection of all semipolar sets is a σ-ideal of subsets of Y containing all polar sets
(this is easily seen from Proposition A.211).
Proof. By [66], Corollary II.4.2, Y endowed with the fine topology is a Baire space.
Now, we refer to [66], Proposition VI.5.9, to find that each semipolar set is finely
meager.
Corollary A.216. Any semipolar set is polar in potential theory of the Laplace equa-
tion.
Proof. If A is totally thin, A = A \ b(A) is polar by Theorem A.215. Since polar sets
form a σ-ideal, semipolar sets are polar as well.
Proof. Let u be the heat kernel (A.4). We find tk & 0 such that u(0, tk ) = 2k . Set
∞
X
1 ∧ 2−k u(x, t + tk ) , (x, t) ∈ Rd+1 .
w(x, t) = (A.7)
k=1
Then obviously w is supercaloric and w(0, 0) = ∞. Consider a point (x, t) 6= (0, 0).
Then limk→∞ u(x, t + tk ) = u(x, t) and thus the sum (A.7) converges. Thus, w is
finite outside the origin.
Definition A.220 (Essential base). The essential base β(A) of a set A ⊂ Y is defined
as the intersection of all finely closed sets E such that A \ E is semipolar. By [66],
Proposition VI.6.1, the set β(A) itself is also finely closed and A \ β(A) is semipolar.
Note that by Theorem A.215 above, the operator β simply coincides with the base
operator b in potential theory of the Laplace equation.
Remark A.222. Similarly as in Remark A.204 we observe that the concept of essen-
tial base is local.
Proof. By the definition of the essential base, W is finely open and W \U is semipolar.
If V ⊂ W is nonempty and finely open, then by Theorem A.213, V ∩ U 6= ∅. Hence
U is finely dense in W .
bpβ(A) .
Proposition A.226. Let A ⊂ Y and p ∈ P. Then BpA = R
[1] E. M. Alfsen, On the geometry of Choquet simplexes, Math. Scand. 15 (1964), 97–110.
[2] , Boundary values for homomorphisms of compact convex sets, Math. Scand.
19 (1966), 113–121.
[3] , Facial structure of compact convex sets, Proc. London Math. Soc. (3) 18
(1968), 385–404.
[4] , On the Dirichlet problem of the Choquet boundary, Acta Math. 120 (1968),
149–159.
[5] , Compact convex sets and boundary integrals, Springer-Verlag, New York,
1971, Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 57.
[6] E. M. Alfsen and T. B. Andersen, Split faces of compact convex sets, Arhus University
(1968/69).
[7] , Split faces of compact convex sets, Proc. London Math. Soc. (3) 21 (1970),
415–442.
[8] E. M. Alfsen and B. Hirsberg, On dominated extensions in linear subspaces of CC (X),
Pacific J. Math. 36 (1971), 567–584.
[9] E. M. Alfsen and F. W. Shultz, State spaces of operator algebras, Mathematics: Theory
& Applications, Birkhäuser Boston Inc., Boston, MA, 2001, Basic theory, orientations,
and C ∗ -products.
[10] , Geometry of state spaces of operator algebras, Mathematics: Theory & Ap-
plications, Birkhäuser Boston Inc., Boston, MA, 2003.
[11] E. M. Alfsen and C. F. Skau, Simplicial decomposition of boundary measures on con-
vex compact sets., Math. Scand. 26 (1970), 62–72.
[12] Ş. Alpay, A localization of the equal support property, J. Pure Appl. Sci. 11 (1978),
233–235 (1980).
[13] F. Altomare and M. Campiti, Korovkin-type approximation theory and its applications,
de Gruyter Studies in Mathematics 17, Walter de Gruyter & Co., Berlin, 1994, Ap-
pendix A by Michael Pannenberg and Appendix B by Ferdinand Beckhoff.
[14] D. Amir, On isomorphisms of continuous function spaces, Israel J. Math. 3 (1965),
205–210.
[15] P. R. Andenaes, Extreme boundaries and continuous affine functions, Math. Scand. 40
(1977), 197–208.
[16] T. B. Andersen and H. R. Atkinson, On Banach algebra-valued facially continuous
functions, J. London Math. Soc. (2) 9 (1974/75), 381–384.
670 Bibliography
[17] R. Arens, Extension of functions on fully normal spaces, Pacific J. Math. 2 (1952),
11–22.
[18] R. F. Arens and J. L. Kelley, Characterization of the space of continuous functions over
a compact Hausdorff space, Trans. Amer. Math. Soc. 62 (1947), 499–508.
[19] S. A. Argyros, G. Godefroy and H. P. Rosenthal, Descriptive set theory and Banach
spaces, Handbook of the geometry of Banach spaces, Vol. 2, North-Holland, Amster-
dam, 2003, pp. 1007–1069.
[20] D. H. Armitage, The Riesz-Herglotz representation for positive harmonic functions
via Choquet’s theorem, Potential theory—ICPT 94 (Kouty, 1994), de Gruyter, Berlin,
1996, pp. 229–232.
[21] D. H. Armitage and S. J. Gardiner, Classical potential theory, Springer Monographs in
Mathematics, Springer-Verlag London Ltd., London, 2001.
[22] M. G. Arsove, The Wiener-Dirichlet problem and the theorem of Evans, Math. Z. 103
(1968), 184–194.
[23] E. Artin, A proof of the Krein–Milman Theorem, Collected papers, Springer-Verlag,
New York, 1982, Edited by Serge Lang and John T. Tate, Reprint of the 1965 original.
[24] L. Asimow and A. J. Ellis, Convexity theory and its applications in functional analysis,
London Mathematical Society Monographs 16, Academic Press Inc. [Harcourt Brace
Jovanovich Publishers], London, 1980.
[25] R. E. Atalla, Markov operators, peak points, and Choquet points, Proc. Amer. Math.
Soc. 41 (1973), 103–109.
[26] I. Babuška and R. Výborný, Reguläre und stabile Randpunkte für das Problem der
Wärmeleitungsgleichung, Ann. Polon. Math. 12 (1962), 91–104.
[27] M. Bačák, Point simpliciality in Choquet representation theory, Illinois J. Math., to
appear.
[28] M. Bačák and J. Spurný, Complementability of spaces of affine continuous functions
on simplices, Bull. Belg. Math. Soc. Simon Stevin 15 (2008), 465–472.
[29] R. Baire, Sur les fonctions de variables rélees, Ann. di Mat. Pura ed Appl. 3 (1899),
1–123.
[30] M. V. Balashov, An analogue of the Kreı̆n-Milman theorem for a strongly convex hull
in a Hilbert space, Mat. Zametki 71 (2002), 37–42.
[31] A. Barvinok, A course in convexity, Graduate Studies in Mathematics 54, American
Mathematical Society, Providence, RI, 2002.
[32] C. J. K. Batty, Maximal measures on tensor products of compact convex sets, Quart. J.
Math. Oxford Ser. (2) 33 (1982), 1–10.
[33] , A characterisation of simplexes by an extension property, Quart. J. Math.
Oxford Ser. (2) 34 (1983), 391–397.
[34] , Some properties of maximal measures on compact convex sets, Math. Proc.
Cambridge Philos. Soc. 94 (1983), 297–305.
Bibliography 671
[35] , Topologies and continuous functions on extreme points and pure states, Math.
Proc. Cambridge Philos. Soc. 98 (1985), 501–511.
[36] H. Bauer, Minimalstellen von Funktionen und Extremalpunkte, Arch. Math. 9 (1958),
389–393.
[37] , Un problème de Dirichlet pour la frontière de Šilov d’un space compact, C.
R. Acad. Sci. Paris 247 (1958), 843–846.
[38] , Šilovscher Rand und Dirichletsches Problem, Ann. Inst. Fourier Grenoble 11
(1961), 89–136, XIV.
[39] , Axiomatische Behandlung des Dirichletschen Problems für elliptische und
parabolische Differentialgleichungen, Math. Ann. 146 (1962), 1–59.
[40] , Harmonische Räume und ihre Potentialtheorie, Ausarbeitung einer im Som-
mersemester 1965 an der Universität Hamburg gehaltenen Vorlesung. Lecture Notes in
Mathematics, No. 22, Springer-Verlag, Berlin, 1966.
[41] , Aspects of modern potential theory, in: Proceedings of the International
Congress of Mathematicians (Vancouver, B.C., 1974), Vol. 1, pp. 41–51, Canad. Math.
Congress, Montreal, Que., 1975.
[42] , Approximation and abstract boundaries, Amer. Math. Monthly 85 (1978),
632–647.
[43] , Harmonic spaces—a survey, Confer. Sem. Mat. Univ. Bari (1984), 34.
[44] , Simplicial function spaces and simplexes, Exposition. Math. 3 (1985), 165–
168.
[45] , Measure and integration theory, de Gruyter Studies in Mathematics 26, Walter
de Gruyter & Co., Berlin, 2001, Translated from the German by Robert B. Burckel.
[46] H. S. Bear, The Silov boundary for a linear space of continuous functions, Amer. Math.
Monthly 68 (1961), 483–485.
[47] R. Becker, Convex cones in analysis, Travaux en Cours [Works in Progress] 67, Her-
mann Éditeurs des Sciences et des Arts, Paris, 2006, With a postface by G. Choquet,
Translation of the 1999 French version.
[48] J. B. Bednar, On the Dirichlet problem for functions on the extreme boundary of a
compact convex set, Math. Scand. 27 (1970), 141–144 (1971).
[49] E. Behrends and G. Wittstock, Tensorprodukte kompakter konvexer Mengen, Invent.
Math. 10 (1970), 251–266.
[50] , Tensorprodukte und Simplexe, Invent. Math. 11 (1970), 188–198.
[51] D. Bensimon, The topological space of all extreme points of a compact convex set,
Rend. Circ. Mat. Palermo (2) 37 (1988), 177–200.
[52] Y. Benyamini and J. Lindenstrauss, A predual of l1 which is not isomorphic to a C(K)
space, in: Proceedings of the International Symposium on Partial Differential Equa-
tions and the Geometry of Normed Linear Spaces (Jerusalem, 1972), 13, pp. 246–254
(1973), 1972.
672 Bibliography
[53] S. K. Berberian, Compact convex sets in inner product spaces, Amer. Math. Monthly 74
(1967), 702–705.
[54] C. Berg, J. P. Reus Christensen and P. Ressel, Harmonic analysis on semigroups, Grad-
uate Texts in Mathematics 100, Springer-Verlag, New York, 1984, Theory of positive
definite and related functions.
[55] S. Bernstein, Sur les fonctions absolument monotones, Acta Math. 52 (1929), 1–66.
[56] C. Bessaga and T. Dobrowolski, Affine and homeomorphic embeddings into l2 , Proc.
Amer. Math. Soc. 125 (1997), 259–268.
[57] G. Birkhoff, Three observations on linear algebra, Univ. Nac. Tucumán. Revista A. 5
(1946), 147–151.
[58] E. Bishop and K. de Leeuw, The representations of linear functionals by measures on
sets of extreme points, Ann. Inst. Fourier. Grenoble 9 (1959), 305–331.
[59] B. Blackadar, Operator algebras, Encyclopaedia of Mathematical Sciences 122,
Springer-Verlag, Berlin, 2006, Theory of C ∗ -algebras and von Neumann algebras, Op-
erator Algebras and Non-commutative Geometry, III.
[60] W. Blaschke and G. Pick, Distanzschätzungen im Funktionenraum II, Math. Ann. 77
(1916), 277–300.
[61] J. Bliedtner, Approximation by harmonic functions, Potential theory—ICPT 94 (Kouty,
1994), de Gruyter, Berlin, 1996, pp. 297–302.
[62] J. Bliedtner and W. Hansen, Simplicial cones in potential theory, Invent. Math. 29
(1975), 83–110.
[63] , Cones of hyperharmonic functions, Math. Z. 151 (1976), 71–87.
[64] , Simplicial cones in potential theory. II. Approximation theorems, Invent.
Math. 46 (1978), 255–275.
[65] , The weak Dirichlet problem, J. Reine Angew. Math. 348 (1984), 34–39.
[66] , Potential theory, Universitext, Springer-Verlag, Berlin, 1986, An analytic and
probabilistic approach to balayage.
[67] N. Boboc and Gh. Bucur, Sur le prolongement des fonctions affines, Math. Scand. 26
(1970), 42–50.
[68] , Cônes convexes de fonctions continues sur un espace compact. Topologies
sur la frontière de Choquet, Rev. Roumaine Math. Pures Appl. 17 (1972), 1307–1316.
(1 plate), Collection of articles dedicated to G. Călugăreanu on his seventieth birthday.
[69] , Simplicial measures on a compact space, Rev. Roumaine Math. Pures Appl.
17 (1972), 1155–1163.
[70] , Conuri convexe de funcţii continue pe spaţii compacte, Editura Academiei
Republicii Socialiste România, Bucharest, 1976, With an English summary.
[71] N. Boboc, Gh. Bucur and A. Cornea, Order and convexity in potential theory: H-
cones, Lecture Notes in Mathematics 853, Springer, Berlin, 1981, In collaboration with
Herbert Höllein.
Bibliography 673
[72] N. Boboc, C. Constantinescu and A. Cornea, On the Dirichlet problem in the axiomatic
theory of harmonic functions, Nagoya Math. J. 23 (1963), 73–96.
[73] N. Boboc and A. Cornea, Convex cones of lower semicontinuous functions on compact
spaces, Rev. Roumaine Math. Pures Appl. 12 (1967), 471–525.
[74] H. Bohman, On approximation of continuous and of analytic functions, Ark. Mat. 2
(1952), 43–56.
[75] F. F. Bonsall, On the representation of points of a convex set, J. London Math. Soc. 38
(1963), 332–334.
[76] V. Borovikov, On the intersection of a sequence of simplexes, Uspehi Matem. Nauk
(N.S.) 7 (1952), 179–180.
[77] K. Borsuk, Über Isomorphie der Funktionalräume, Bull. Acad. Polonaise (1933), 1–10.
[78] G. Bouligand, Sur le problème de Dirichlet, Ann. Soc. Polon. Math. 4 (1926), 59–112.
[79] N. Bourbaki, Integration. I. Chapters 1–6, Elements of Mathematics (Berlin), Springer-
Verlag, Berlin, 2004, Translated from the 1959, 1965 and 1967 French originals by
Sterling K. Berberian.
[80] J. Bourgain, D. H. Fremlin and M. Talagrand, Pointwise compact sets of Baire-
measurable functions, Amer. J. Math. 100 (1978), 845–886.
[81] J. Bourgain and M. Talagrand, Compacité extrémale, Proc. Amer. Math. Soc. 80 (1980),
68–70.
[82] R. D. Bourgin, Geometric aspects of convex sets with the Radon-Nikodým property,
Lecture Notes in Mathematics 993, Springer-Verlag, Berlin, 1983.
[83] M. Brelot, Familles de Perron et problème de Dirichlet, Acta Litt. Sci. Szeged 9 (1939),
133–153.
[84] , Minorantes sous-harmoniques, extrémales et capacités, J. Math. Pures Appl.
(9) 24 (1945), 1–32.
[85] , Remarque sur le prolongement fonctionnel linéaire et le problème de Dirich-
let, Acta Sci. Math. Szeged 12 (1950), 150–152.
[86] , Sur un théorème de prolongement fonctionnel de Keldych concernant le
problème de Dirichlet, J. Analyse Math. 8 (1960/1961), 273–288.
[87] , Éléments de la théorie classique du potentiel, 3e édition. Les cours de Sor-
bonne. 3e cycle, Centre de Documentation Universitaire, Paris, 1965.
[88] , Historical introduction, Potential Theory (C.I.M.E., I Ciclo, Stresa, 1969),
Edizioni Cremonese, Rome, 1970, pp. 1–21.
[89] , Les étapes et les aspects multiples de la théorie du potentiel, Enseignement
Math. (2) 18 (1972), 1–36.
[90] , Le balayage de Poincaré et l’épine de Lebesgue, in: Proceedings of the 110th
national congress of learned societies (Montpellier, 1985), pp. 141–151, Com. Trav.
Hist. Sci., Paris, 1985.
674 Bibliography
[91] E. Briem, Facial topologies for subspaces of C(X), Math. Ann. 208 (1974), 9–13.
[92] , Extreme orthogonal boundary measures for A(K) and decompositions for
compact convex sets, Spaces of analytic functions (Sem. Functional Anal. and Function
Theory, Kristiansand, 1975), Springer, Berlin, 1976, pp. 8–16. Lecture Notes in Math.,
Vol. 512.
[93] , Enlarging a subspace of C(X) without changing the Choquet boundary, Math.
Scand. 44 (1979), 218–224.
[94] , A characterization of simplexes by parallel faces, Bull. London Math. Soc. 12
(1980), 55–59.
[95] R. B. Burckel, An introduction to classical complex analysis. Vol. 1, Pure and Applied
Mathematics 82, Academic Press Inc. [Harcourt Brace Jovanovich Publishers], New
York, 1979.
[96] M. Capon, Sur les fonctions qui vérifient le calcul barycentrique, Proc. London Math.
Soc. (3) 32 (1976), 163–180.
[97] P. Cartier, J. M. G. Fell and P.-A. Meyer, Comparaison des mesures portées par un
ensemble convexe compact, Bull. Soc. Math. France 92 (1964), 435–445.
[98] B. Cascales and G. Godefroy, Angelicity and the boundary problem, Mathematika 45
(1998), 105–112.
[99] B. Cascales, G. Manjabacas and G. Vera, A Krein-Šmulian type result in Banach
spaces, Quart. J. Math. Oxford Ser. (2) 48 (1997), 161–167.
[100] B. Cascales and R. Shvydkoy, On the Krein-Šmulian theorem for weaker topologies,
Illinois J. Math. 47 (2003), 957–976.
[101] B. Cascales and G. Vera, Topologies weaker than the weak topology of a Banach space,
J. Math. Anal. Appl. 182 (1994), 41–68.
[102] C. Castaing and M. Valadier, Convex analysis and measurable multifunctions, Lecture
Notes in Mathematics, Vol. 580, Springer-Verlag, Berlin, 1977.
[103] M. M. Choban, On some problems of descriptive set theory in topological spaces, Us-
pekhi Mat. Nauk 60 (2005), 123–144.
[104] G. Choquet, Theory of capacities, Ann. Inst. Fourier, Grenoble 5 (1953–1954), 131–
295 (1955).
[105] , Le théorème de représentation intégrale dans les ensembles convexes com-
pacts, Ann. Inst. Fourier Grenoble 10 (1960), 333–344.
[106] , Remarque à propos de la démonstration de l’unicité de P.-A. Meyer, Séminaire
Brelot–Choquet–Deny (Théorie du Potentiel) 6 (1961/62), Exposé No. 8.
[107] , Deux exemples classiques de représentation intégrale, Enseignement Math.
(2) 15 (1969), 63–75.
[108] , Lectures on analysis. Vol. I–III: Infinite dimensional measures and problem
solutions, Edited by J. Marsden, T. Lance and S. Gelbart, W. A. Benjamin, Inc., New
York-Amsterdam, 1969.
Bibliography 675
[109] , La naissance de la théorie des capacités: réflexion sur une expérience person-
nelle, C. R. Acad. Sci. Sér. Gén. Vie Sci. 3 (1986), 385–397.
[110] , Une démonstration du théorème de Bochner-Weil par discrétisation du
groupe, Results Math. 9 (1986), 1–9.
[111] , Existence et unicité des représentations intégrales au moyen des points
extrémaux dans les cônes convexes, Séminaire Bourbaki, Vol. 4, Soc. Math. France,
Paris, 1995, pp. Exp. No. 139, 33–47.
[112] G. Choquet, H. Corson and V. Klee, Exposed points of convex sets, Pacific J. Math. 17
(1966), 33–43.
[113] G. Choquet and P.-A. Meyer, Existence et unicité des représentations intégrales dans
les convexes compacts quelconques, Ann. Inst. Fourier (Grenoble) 13 (1963), 139–154.
[114] J. P. R. Christensen, Borel structures and a topological zero-one law, Math. Scand. 29
(1971), 245–255 (1972).
[115] , Compact convex sets and compact Choquet simplexes, Invent. Math. 19
(1973), 1–4.
[116] , Topology and Borel structure, North-Holland Publishing Co., Amsterdam,
1974, Descriptive topology and set theory with applications to functional analysis and
measure theory, North-Holland Mathematics Studies, Vol. 10. (Notas de Matemática,
No. 51).
[117] C. H. Chu, Anti-lattices and prime sets, Math. Scand. 31 (1972), 151–165.
[118] , A note on faces of compact convex sets, J. London Math. Soc. (2) 5 (1972),
556–560.
[119] C. H. Chu and H. B. Cohen, Isomorphisms of spaces of continuous affine functions,
Pacific J. Math. 155 (1992), 71–85.
[120] A. Clausing and G. Mägerl, Generalized Dirichlet problems and continuous selections
of representing measures, Math. Ann. 216 (1975), 71–78.
[121] A. Clausing and S. Papadopoulou, Stable convex sets and extremal operators, Math.
Ann. 231 (1977/78), 193–203.
[122] H. B. Cohen, A bound-two isomorphism between C(X) Banach spaces, Proc. Amer.
Math. Soc. 50 (1975), 215–217.
[123] C. Constantinescu and A. Cornea, Potential theory on harmonic spaces, Springer-
Verlag, New York, 1972, With a preface by H. Bauer, Die Grundlehren der mathe-
matischen Wissenschaften, Band 158.
[124] A. Cornea, Résolution du problème de Dirichlet et comportement des solutions à la
frontière à l’aide des fonctions de contrôle, C. R. Acad. Sci. Paris Sér. I Math. 320
(1995), 159–164.
[125] , Applications of controlled convergence in analysis, Analysis and topology,
World Sci. Publ., River Edge, NJ, 1998, pp. 257–275.
676 Bibliography
[126] H. H. Corson, A compact convex set in E 3 whose exposed points are of the first cate-
gory, Proc. Amer. Math. Soc. 16 (1965), 1015–1021.
[127] , Metrizability of compact convex sets, Trans. Amer. Math. Soc. 151 (1970),
589–596.
[128] A. Curnock, J. Howroyd and Ngai-Ching Wong, The unique decomposition property
and the Banach-Stone theorem, Function spaces (Edwardsville, IL, 2002), Contemp.
Math. 328, Amer. Math. Soc., Providence, RI, 2003, pp. 151–156.
[129] P. C. Curtis, Jr., On a theorem of Keldysh and Wiener, Abstract Spaces and Approxi-
mation (Proc. Conf., Oberwolfach, 1968), Birkhäuser, Basel, 1969, pp. 351–356.
[130] L. Danzer, B. Grünbaum and V. Klee, Helly’s theorem and its relatives, Proc. Sympos.
Pure Math., Vol. VII, Amer. Math. Soc., Providence, R.I., 1963, pp. 101–180.
[131] I. K. Daugavet, A property of completely continuous operators in the space C, Uspehi
Mat. Nauk 18 (1963), 157–158.
[132] E. B. Davies and G. F. Vincent-Smith, Tensor products, infinite products, and projective
limits of Choquet simplexes, Math. Scand. 22 (1968), 145–164 (1969).
[133] L. de Branges, The Stone-Weierstrass theorem, Proc. Amer. Math. Soc. 10 (1959),
822–824.
[134] A. de la Pradelle, À propos du mémoire de G. F. Vincent-Smith sur l’approximation
des fonctions harmoniques, Ann. Inst. Fourier (Grenoble) 19 (1969), 355–370 (1970).
[135] , Approximation des fonctions harmoniques à l’aide d’un théorème de G. F.
Vincent-Smith, Séminaire de Théorie du Potentiel, dirigé par M. Brelot, G. Choquet et
J. Deny (1969/70), Exp. 3, Secrétariat Mathématique, Paris, 1971, p. 16.
[136] M. De Wilde, Pointwise compactness in spaces of functions and R. C. James theorem,
Math. Ann. 208 (1974), 33–47.
[137] G. Debs, Sélections maximales d’une multi–application, Unpublished manuscript.
[138] , Applications affines ouvertes et convexes compacts stables, Bull. Sci. Math.
(2) 102 (1978), 401–414.
[139] , Sélections maximales d’une multi–application, Le séminaire d’Initiation a
l’Analyse 29 (1979).
[140] , Some general methods for constructing stable convex sets, Math. Ann. 241
(1979), 97–105.
[141] , Quelques propriétés des espaces α-favorables et applications aux convexes
compacts, Ann. Inst. Fourier (Grenoble) 30 (1980), vii, 29–43.
[142] C. Dellacherie, Un complément au théorème de Weierstrass-Stone, Séminaire de Prob-
abilités (Univ. Strasbourg, Strasbourg, 1966/67), Vol. I, Springer, Berlin, 1967, pp. 52–
53.
[143] R. Deville and C. Finet, An extension of Simons’ inequality and applications, Rev. Mat.
Complut. 14 (2001), 95–104.
Bibliography 677
[216] P. Harmand, D. Werner and W. Werner, M -ideals in Banach spaces and Banach alge-
bras, Lecture Notes in Mathematics 1547, Springer-Verlag, Berlin, 1993.
[217] F. Hausdorff, Set theory, Second edition. Translated from the German by John R. Au-
mann et al, Chelsea Publishing Co., New York, 1962.
[218] R. Haydon, A new proof that every Polish space is the extreme boundary of a simplex,
Bull. London Math. Soc. 7 (1975), 97–100.
[219] , An extreme point criterion for separability of a dual Banach space, and a new
proof of a theorem of Corson, Quart. J. Math. Oxford (2) 27 (1976), 379–385.
[220] , Some more characterizations of Banach spaces containing l1 , Math. Proc.
Cambridge Philos. Soc. 80 (1976), 269–276.
[221] J. Heinonen, T. Kilpeläinen and O. Martio, Nonlinear potential theory of degenerate
elliptic equations, Oxford Mathematical Monographs, The Clarendon Press Oxford
University Press, New York, 1993, Oxford Science Publications.
[222] L. L. Helms, Introduction to potential theory, Robert E. Krieger Publishing Co., Hunt-
ington, N.Y., 1975, Reprint of the 1969 edition, Pure and Applied Mathematics, Vol.
XXII.
[223] M. Hervé, Sur les représentations intégrales à l’aide des points extrémaux dans un
ensemble compact convexe métrisable, C. R. Acad. Sci. Paris 253 (1961), 366–368.
[224] H. U. Hess, On a theorem of Cambern, Proc. Amer. Math. Soc. 71 (1978), 204–206.
[225] E. Hewitt and L. J. Savage, Symmetric measures on Cartesian products, Trans. Amer.
Math. Soc. 80 (1955), 470–501.
[226] B. Hirsberg, A measure theoretic characterization of parallel and split faces and their
connections with function spaces and algebras, Various Publications Series, No. 16,
Matematisk Institut, Aarhus Universitet, Aarhus, 1970.
[227] J. Hoffmann-Jørgensen, Probability in Banach space, École d’Été de Probabilités de
Saint-Flour, VI-1976, Springer-Verlag, Berlin, 1977, pp. 1–186. Lecture Notes in
Math., Vol. 598.
[228] W. Hoh and N. Jacob, On the potential theory of the Kolmogorov equation, Math.
Nachr. 154 (1991), 51–66.
[229] P. Holický, Čech analytic and almost K-descriptive spaces, Czechoslovak Math. J.
43(118) (1993), 451–466.
[230] , Luzin theorems for scattered-K-analytic spaces and Borel measures on them,
Atti Sem. Mat. Fis. Univ. Modena 44 (1996), 395–413.
[231] , Extensions of Borel measurable maps and ranges of Borel bimeasurable maps,
Bull. Pol. Acad. Sci. Math. 52 (2004), 151–167.
[232] P. Holický and T. Keleti, Borel classes of sets of extreme and exposed points in Rn ,
Proc. Amer. Math. Soc. 133 (2005), 1851–1859 (electronic).
[233] P. Holický and V. Komı́nek, Two remarks on the structure of sets of exposed and ex-
treme points, Extracta Math. 15 (2000), 547–561.
682 Bibliography
[234] P. Holický and M. Laczkovich, Descriptive properties of the set of exposed points of
compact convex sets in R3 , Proc. Amer. Math. Soc. 132 (2004), 3345–3347 (electronic).
[235] P. Holický and J. Pelant, Internal descriptions of absolute Borel classes, Topology Appl.
141 (2004), 87–104.
[236] P. Holický and J. Spurný, Perfect images of absolute Souslin and absolute Borel Ty-
chonoff spaces, Topology Appl. 131 (2003), 281–294.
[237] H. Höllein, A geometrical characterization of Choquet simplexes, Math. Z. 160 (1978),
249–254.
[238] R. B. Holmes, Geometric functional analysis and its applications, Springer-Verlag,
New York, 1975, Graduate Texts in Mathematics, No. 24.
[239] L. Hörmander, Notions of convexity, Progress in Mathematics 127, Birkhäuser Boston
Inc., Boston, MA, 1994.
[240] J. Hotta, A remark on regularly convex sets, Kōdai Math. Sem. Rep. 3 (1951), 37–40,
{Volume numbers not printed on issues until Vol. 7 (1955).}.
[241] A. Hulanicki and R. R. Phelps, Some applications of tensor products of partially-
ordered linear spaces, J. Functional Analysis 2 (1968), 177–201.
[242] K. Jacobs, Extremalpunkte konvexer Mengen, Selecta Mathematica, III, Springer,
Berlin, 1971, pp. 90–118. Heidelberger Taschenbücher, 86.
[243] R. C. James, Weak compactness and reflexivity, Israel J. Math. 2 (1964), 101–119.
[244] , Weakly compact sets, Trans. Amer. Math. Soc. 113 (1964), 129–140.
[245] G. Jameson, Ordered linear spaces, Lecture Notes in Mathematics, Vol. 141, Springer-
Verlag, Berlin, 1970.
[246] J. E. Jayne, Generation of σ-algebras, Baire sets and descriptive Borel sets, Mathe-
matika 24 (1977), 241–256.
[247] , Metrization of compact convex sets, Math. Ann. 234 (1978), 109–115.
[248] J. E. Jayne and C. A. Rogers, The extremal structure of convex sets, J. Functional
Analysis 26 (1977), 251–288.
[249] F. Jellett, Homomorphisms and inverse limits of Choquet simplexes, Math. Z. 103
(1968), 219–226.
[250] , On affine extensions of continuous functions defined on the extreme boundary
of a Choquet simplex, Quart. J. Math. Oxford Ser. (2) 36 (1985), 71–73.
[251] W. B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces,
Handbook of the geometry of Banach spaces, Vol. I, North-Holland, Amsterdam, 2001,
pp. 1–84.
[252] M. Kačena, Products and projective limits of function spaces, Comment. Math. Univ.
Carolin. 49 (2008), 547–578.
[253] M. Kačena and J. Spurný, Affine Baire functions on Choquet simplices, preprint.
Bibliography 683
[254] , Affine images of compact convex sets and maximal measures, Bull. Sci. Math.
133 (2009), 493–500.
[255] V. M. Kadets, R. V. Shvidkoy, G. G. Sirotkin and D. Werner, Banach spaces with the
Daugavet property, Trans. Amer. Math. Soc. 352 (2000), 855–873.
[256] R. V. Kadison, A representation theory for commutative topological algebra, Mem.
Amer. Math. Soc., 1951 (1951), 39.
[257] O. F. K. Kalenda, (I)-envelopes of closed convex sets in Banach spaces, Israel J. Math.
162 (2007), 157–181.
[258] , (I)-envelopes of unit balls and James’ characterization of reflexivity, Studia
Math. 182 (2007), 29–40.
[259] O. F. K. Kalenda and J. Spurný, Extending Baire-one functions on topological spaces,
Topology Appl. 149 (2005), 195–216.
[260] , Boundaries of compact convex sets and fragmentability, J. Funct. Anal. 256
(2009), 865–880.
[261] J. Kaniewski and R. Pol, Borel-measurable selectors for compact-valued mappings in
the non-separable case, Bull. Acad. Polon. Sci. Sér. Sci. Math. Astronom. Phys. 23
(1975), 1043–1050.
[262] A. S. Kechris, Classical descriptive set theory, Graduate Texts in Mathematics 156,
Springer-Verlag, New York, 1995.
[263] M. V. Keldych, On the Dirichlet problem, Dokl. Akad. Nauk SSSR 32 (1941), 308–309
(Russian).
[264] , On the solubility and the stability of Dirichlet’s problem, Uspekhi Matem.
Nauk 8 (1941), 171–231 (Russian).
[265] O.-H. Keller, Die Homoiomorphie der kompakten konvexen Mengen im Hilbertschen
Raum, Math. Ann. 105 (1931), 748–758.
[266] J. L. Kelley, Note on a theorem of Krein and Milman, J. Osaka Inst. Sci. Tech. Part I. 3
(1951), 1–2.
[267] O. D. Kellog, Unicité des fonctions harmoniques, C. R. Acad. Sci. Paris 187 II (1928),
526–527.
[268] S. S. Khurana, Pointwise compactness on extreme points, Proc. Amer. Math. Soc. 83
(1981), 347–348.
[269] S. King and R. Shiflett, Doubly stochastic operators and the history of Birkhoff’s prob-
lem 111, Stochastic processes and functional analysis, Lecture Notes in Pure and Appl.
Math. 238, Dekker, New York, 2004, pp. 411–440.
[270] K. Kivisoo and E. Oja, Extension of Simons’ inequality, Proc. Amer. Math. Soc. 133
(2005), 3485–3496 (electronic).
[271] V. Klee, Some new results on smoothness and rotundity in normed linear spaces., Math.
Ann. 139 (1959), 51–63 (1959).
684 Bibliography
[272] V. L. Klee, Jr., Extremal structure of convex sets. II, Math. Z. 69 (1958), 90–104.
[273] J. Köhn, Harmonische Räume mit einer Basis semiregulärer Mengen, Seminar über
Potentialtheorie, Springer, Berlin, 1968, pp. 1–12.
[274] , Barycenters of unique maximal measures, J. Functional Analysis 6 (1970),
76–82.
[275] J. Köhn and M. Sieveking, Reguläre und extremale Randpunkte in der Potentialtheorie,
Rev. Roumaine Math. Pures Appl. 12 (1967), 1489–1502.
[276] J. Kolář and J. Lukeš, Simultaneous solutions of the weak Dirichlet problem, Potential
Anal. 15 (2001), 17–21, ICPA98 (Hammamet).
[277] P. P. Korovkin, On convergence of linear positive operators in the space of continuous
functions, Doklady Akad. Nauk SSSR (N.S.) 90 (1953), 961–964.
[278] , Linear operators and approximation theory, Translated from the Russian ed.
(1959). Russian Monographs and Texts on Advanced Mathematics and Physics, Vol.
III, Gordon and Breach Publishers, Inc., New York, 1960.
[279] R. A. Kortram, The extreme points of a class of functions with positive real part, Bull.
Belg. Math. Soc. Simon Stevin 4 (1997), 449–459.
[280] G. Koumoullis, A generalization of functions of the first class, Topology Appl. 50
(1993), 217–239.
[281] M. Kraus, A note on the uniform approximation of continuous affine functions, Expo.
Math. 27 (2009), 73–78.
[282] U. Krause, Der Satz von Choquet als ein abstrakter Spektralsatz und vice versa, Math.
Ann. 184 (1970), 275–296.
[283] M. Krein and D. Milman, On extreme points of regular convex sets, Studia Math. 9
(1940), 133–138.
[284] A. Kufner, O. John and S. Fučı́k, Function spaces, Noordhoff International Publish-
ing, Leyden, 1977, Monographs and Textbooks on Mechanics of Solids and Fluids;
Mechanics: Analysis.
[285] C. Kuratowski, Topologie. Vol. I, Monografie Matematyczne, Tom 20, Państwowe
Wydawnictwo Naukowe, Warsaw, 1958, 4ème éd.
[286] K. Kuratowski and A. Maitra, Some theorems on selectors and their applications to
semi-continuous decompositions, Bull. Acad. Polon. Sci. Sér. Sci. Math. Astronom.
Phys. 22 (1974), 877–881.
[287] K. Kuratowski and C. Ryll-Nardzewski, A general theorem on selectors, Bull. Acad.
Polon. Sci. Sér. Sci. Math. Astronom. Phys. 13 (1965), 397–403.
[288] C. De La Vallée Poussin, Sur quelques extensions de la méthode du balayage de
Poincaré et sur le problème de Dirichlet, C. R. Acad. Sci. Paris 192 (1931), 651–653.
[289] , Extension de la méthode du balayage de Poincaré et problème de Dirichlet,
Ann. Inst. H. Poincaré 2 (1932), 169–232.
Bibliography 685
[290] H. E. Lacey, The isometric theory of classical Banach spaces, Springer-Verlag, New
York, 1974, Die Grundlehren der mathematischen Wissenschaften, Band 208.
[291] H. E. Lacey and P. D. Morris, On spaces of type A(K) and their duals, Proc. Amer.
Math. Soc. 23 (1969), 151–157.
[292] A. J. Lazar, Affine products of simplexes, Math. Scand. 22 (1968), 165–175 (1969).
[293] , Spaces of affine continuous functions on simplexes, Trans. Amer. Math. Soc.
134 (1968), 503–525.
[294] , Extreme boundaries of convex bodies in l2 , Israel J. Math. 20 (1975), 369–
374.
[295] A. J. Lazar and J. Lindenstrauss, On Banach spaces whose duals are L1 spaces, Israel
J. Math. 4 (1966), 205–207.
[296] , Banach spaces whose duals are L1 spaces and their representing matrices,
Acta Math. 126 (1971), 165–193.
[297] H. Lebesgue, Sur le cas d’impossibilité du problème de Dirichlet ordinaire, C. R.
Séances Soc. Math. France 17 (1912).
[298] , Conditions de régularité, conditions d’irregularité, conditions d’impossibilité
dans le problème de Dirichlet, C. R. Acad. Sci. Paris 178 (1924), 349–354.
[299] C. Léger, Une démonstration du théorème de A. J. Lazar sur les simplexes compacts,
C. R. Acad. Sci. Paris Sér. A-B 265 (1967), A830–A831.
[300] A. Liapounoff, Sur les fonctions-vecteurs complètement additives, Bull. Acad. Sci.
URSS. Sér. Math. [Izvestia Akad. Nauk SSSR] 4 (1940), 465–478 (Russian).
[301] G. M. Lieberman, Second order parabolic differential equations, World Scientific Pub-
lishing Co. Inc., River Edge, NJ, 1996.
[302] Å. Lima, On continuous convex functions and split faces, Proc. London Math. Soc. (3)
25 (1972), 27–40.
[303] H. Lin, An introduction to the classification of amenable C ∗ -algebras, World Scientific
Publishing Co. Inc., River Edge, NJ, 2001.
[304] J. Lindenstrauss, Extension of compact operators, Mem. Amer. Math. Soc. No. 48
(1964), 112.
[305] , A remark on extreme doubly stochastic measures, Amer. Math. Monthly 72
(1965), 379–382.
[306] , A short proof of Liapounoff’s convexity theorem, J. Math. Mech. 15 (1966),
971–972.
[307] J. Lindenstrauss, G. Olsen and Y. Sternfeld, The Poulsen simplex, Ann. Inst. Fourier
(Grenoble) 28 (1978), vi, 91–114.
[308] J. Lindenstrauss and D. E. Wulbert, On the classification of the Banach spaces whose
duals are L1 spaces, J. Functional Analysis 4 (1969), 332–349.
686 Bibliography
[309] Z. Lipecki, On compactness and extreme points of some sets of quasi-measures and
measures. III, Manuscripta Math. 117 (2005), 463–473.
[310] Z. Lipecki, V. Losert and J. Spurný, Products of simplices, preprint.
[311] L. H. Loomis, Unique direct integral decompositions on convex sets, Amer. J. Math. 84
(1962), 509–526.
[312] D. H. Luecking and L. A. Rubel, Complex analysis, Universitext, Springer-Verlag, New
York, 1984, A functional analysis approach.
[313] J. Lukeš, Principal solution of the Dirichlet problem in potential theory, Comment.
Math. Univ. Carolinae 14 (1973), 773–778.
[314] , Théorème de Keldych dans la théorie axiomatique de Bauer des fonctions
harmoniques, Czechoslovak Math. J. 24(99) (1974), 114–125.
[315] , Semiregular sets in harmonic spaces, Časopis Pěst. Mat. 100 (1975), 195–
197.
[316] , The Lusin-Menchoff property of fine topologies, Comment. Math. Univ. Car-
olinae 18 (1977), 515–530.
[317] J. Lukeš and J. Malý, On the boundary behaviour of the Perron generalized solution,
Math. Ann. 257 (1981), 355–366.
[318] , Fine hyperharmonicity without axiom D, Math. Ann. 261 (1982), 299–306.
[319] J. Lukeš, J. Malý, I. Netuka, M. Smrčka and J. Spurný, On approximation of affine
Baire-one functions, Israel J. Math. 134 (2003), 255–287.
[320] J. Lukeš, J. Malý and L. Zajı́ček, Fine topology methods in real analysis and potential
theory, Lecture Notes in Mathematics 1189, Springer-Verlag, Berlin, 1986.
[321] J. Lukeš, T. Mocek and I. Netuka, Exposed sets in potential theory, Bull. Sci. Math. 130
(2006), 646–659.
[322] J. Lukeš, T. Mocek, M. Smrčka and J. Spurný, Choquet like sets in function spaces,
Bull. Sci. Math. 127 (2003), 397–437.
[323] J. Lukeš and I. Netuka, The Wiener type solution of the Dirichlet problem in potential
theory, Math. Ann. 224 (1976), 173–178.
[324] , What is the right solution of the Dirichlet problem?, Romanian-Finnish Sem-
inar on Complex Analysis (Proc., Bucharest, 1976), Lecture Notes in Math. 743,
Springer, Berlin, 1979, pp. 564–572.
[325] , Extreme harmonic functions on a ball, Expo. Math. 22 (2004), 83–91.
[326] W. Lusky, The Gurarij spaces are unique, Arch. Math. (Basel) 27 (1976), 627–635.
[327] , On separable Lindenstrauss spaces, J. Functional Analysis 26 (1977), 103–
120.
[328] , Separable Lindenstrauss spaces, Functional Analysis: surveys and recent re-
sults (Proc. Conf., Paderhorn, 1976), Notas Mat. 63, North-Holland, Amsterdam, 1977,
pp. 15–28.
Bibliography 687
[348] G. Mokobodzki and D. Sibony, Cônes de fonctions et théorie du potentiel. I. Les noy-
aux associés à un cône de fonctions, Séminaire de Théorie du Potentiel, dirigé par M.
Brelot, G. Choquet et J. Deny: 1966/67, Exp. 8, Secrétariat mathématique, Paris, 1968,
p. 35.
[349] A. F. Monna, Note sur le problème de Dirichlet, Nieuw Arch. Wisk. (3) 19 (1971),
58–64.
[350] W. B. Moors and E. A. Reznichenko, Separable subspaces of affine function spaces on
convex compact sets, Topology Appl. 155 (2008), 1306–1322.
[351] W. B. Moors and J. Spurný, On the topology of pointwise convergence on the bound-
aries of L1 -preduals, Proc. Amer. Math. Soc. 137 (2009), 1421–1429.
[352] I. Namioka and R. R. Phelps, Tensor products of compact convex sets, Pacific J. Math.
31 (1969), 469–480.
[353] I. Namioka and R. Pol, σ-fragmentability and analyticity, Mathematika 43 (1996), 172–
181.
[354] M. A. Navarro, Some characterizations of finite-dimensional Hilbert spaces, J. Math.
Anal. Appl. 223 (1998), 364–365.
[355] I. Netuka, A remark on semiregular sets, Časopis Pěst. Mat. 98 (1973), 419–421
(Czech).
[356] , Thinness and the heat equation, Časopis Pěst. Mat. 99 (1974), 293–299.
[357] , The classical Dirichlet problem and its generalizations, Potential theory,
Copenhagen 1979 (Proc. Colloq., Copenhagen, 1979), Lecture Notes in Math. 787,
Springer, Berlin, 1980, pp. 235–266.
[358] , The Dirichlet problem for harmonic functions, Amer. Math. Monthly 87
(1980), 621–628.
[359] , Extensions of operators and the Dirichlet problem in potential theory, Rend.
Circ. Mat. Palermo (2) Suppl. 10 (1985), 143–163.
[360] , The Ninomiya operators and the generalized Dirichlet problem in potential
theory, Osaka J. Math. 23 (1986), 741–750.
[361] , Separation properties involving harmonic functions, Expo. Math. 18 (2000),
333–337.
[362] , The work of Heinz Bauer in potential theory, Selecta, de Gruyter, Berlin,
2003, pp. 29–41.
[363] I. Netuka and J. Veselý, On harmonic functions (solution of the problem 6393 [1982,
502] proposed by G.A. Edgar), Amer. Math. Monthly 91 (1984), 61–62.
[364] G. Nöbeling and H. Bauer, Allgemeine Approximationskriterien mit Anwendungen,
Jber. Deutsch. Math. Verein. 58 (1955), 54–72.
[365] R. C. O’Brien, On the openness of the barycentre map, Math. Ann. 223 (1976), 207–
212.
Bibliography 689
[403] , Functional analysis, second ed, International Series in Pure and Applied
Mathematics, McGraw-Hill Inc., New York, 1991.
[404] M. Ruiz Galán and S. Simons, A new minimax theorem and a perturbed James’s theo-
rem, Bull. Austral. Math. Soc. 66 (2002), 43–56.
[405] J. J. Saccoman, On the origin of extreme points and the Kreı̆n-Mil0 man theorem, Aus-
tral. Math. Soc. Gaz. 17 (1990), 10–12.
[406] J. Saint-Raymond, Fonctions boréliennes sur un quotient, Bull. Sci. Math. (2) 100
(1976), 141–147.
[407] , Fonctions convexes sur un convexe borné complet, Bull. Sci. Math. (2) 102
(1978), 331–336.
[408] , Fonctions convexes de première classe, Math. Scand. 54 (1984), 121–129.
[409] H. H. Schaefer, Topological vector spaces, The Macmillan Co., New York, 1966.
[410] H. Schirmeier and U. Schirmeier, Einige Bemerkungen über den Satz von Keldych,
Math. Ann. 236 (1978), 245–254.
[411] G. Schober, Univalent functions—selected topics, Lecture Notes in Mathematics, Vol.
478, Springer-Verlag, Berlin, 1975.
[412] Z. Semadeni, Free compact convex sets, Bull. Acad. Polon. Sci. Sér. Sci. Math. As-
tronom. Phys. 13 (1965), 141–146.
[413] , Categorical methods in convexity, Proc. Colloquium on Convexity (Copen-
hagen, 1965), Kobenhavns Univ. Mat. Inst., Copenhagen, 1967, pp. 281–307.
[414] , Banach spaces of continuous functions. Vol. I, PWN—Polish Scientific Pub-
lishers, Warsaw, 1971, Monografie Matematyczne, Tom 55.
[415] D. Sibony, Cônes des fonctions et potentiels, Cours de 3éme Cycle, Fac. des Sci. de
Paris, mimeographed, 1967–68, p. 150.
[416] W. Sierpiński, Les fonctions continues et la propriété de Baire, Fund. Math. 28 (1937),
120–121.
[417] S. Simons, A convergence theorem with boundary, Pacific J. Math. 40 (1972), 703–708.
[418] , An eigenvector proof of Fatou’s lemma for continuous functions, Math. Intel-
ligencer 17 (1995), 67–70.
[419] R. Sine, Geometric theory of a single Markov operator, Pacific J. Math. 27 (1968),
155–166.
[420] J. Spurný, Baire classes of Banach spaces and strongly affine functions, Trans. Amer.
Math. Soc., to appear.
[421] , Borel sets and functions in topological spaces, preprint.
[422] , Boundary problem for L1 –preduals, Illinois J. Math., to appear.
[423] , On the Dirichlet problem of extreme points for non-continuous functions,
Israel J. Math., to appear.
692 Bibliography
[424] , On the Dirichlet problem for functions of the first Baire class, Comment. Math.
Univ. Carolin. 42 (2001), 721–728.
[425] , Representation of abstract affine functions, Real Anal. Exchange 28
(2002/03), 337–354.
[426] , The Dirichlet problem for Baire-one functions, Cent. Eur. J. Math. 2 (2004),
260–271 (electronic).
[427] , Affine Baire-one functions on Choquet simplexes, Bull. Austral. Math. Soc.
71 (2005), 235–258.
[428] , The weak Dirichlet problem for Baire functions, Proc. Amer. Math. Soc. 134
(2006), 3153–3157 (electronic).
[429] , Complementability of spaces of harmonic functions, Potential Anal. 29
(2008), 271–302.
[430] , Automatic boundedness of affine functions, Houston J. Math. 35 (2009), 553–
561.
[431] , The Dirichlet problem for Baire–two functions on simplices, Bull. Austral.
Math. Soc. 79 (2009), 285–297.
[432] J. Spurný and O. F. K. Kalenda, A solution of the abstract Dirichlet problem for Baire-
one functions, J. Funct. Anal. 232 (2006), 259–294.
[433] J. Spurný and M. Zelený, Additive families of low Borel classes and Borel measurable
selectors, Canad. Math. Bull., to appear.
[434] S. M. Srivastava, A course on Borel sets, Graduate Texts in Mathematics 180, Springer-
Verlag, New York, 1998.
[435] P. J. Stacey, Split faces and projective sets in a metrizable compact convex set, Math.
Ann. 219 (1976), 167–170.
[436] , Choquet simplices with prescribed extreme and Šilov boundaries, Quart. J.
Math. Oxford Ser. (2) 30 (1979), 469–482.
[437] S. Sternberg, Lectures on differential geometry, second ed, Chelsea Publishing Co.,
New York, 1983, With an appendix by Sternberg and Victor W. Guillemin.
[438] Y. Sternfeld, Characterization of Bauer simplices and some other classes of Choquet
simplices by their representing matrices, Notes in Banach spaces, Univ. Texas Press,
Austin, Tex., 1980, pp. 306–358.
[439] A. H. Stone, Non-separable Borel sets, Rozprawy Mat. 28 (1962), 41.
[440] E. Størmer, On partially ordered vector spaces and their duals, with applications to
simplexes and C ∗ -algebras, Proc. London Math. Soc. (3) 18 (1968), 245–265.
[441] S. Straszewicz, Über exponierte Punkte abgeschlossener Punktmengen, Fund. Math.
24 (1935), 139–143.
[442] M. Talagrand, Les fonctions affines sur [0, 1]N ayant la proprieté de Baire faible sont
continues, Séminaire Choquet (Initiation a l’analyse) 15 (1975/76).
Bibliography 693
[443] , Sélection mesurable de mesures maximales simpliciales, Bull. Sci. Math. (2)
102 (1978), 49–56.
[444] , Deux généralisations d’un théorème de I. Namioka, Pacific J. Math. 81 (1979),
239–251.
[445] , Sur les convexes compacts dont l’ensemble des points extrémaux est K-
analytique, Bull. Soc. Math. France 107 (1979), 49–53.
[446] , Three convex sets, Proc. Amer. Math. Soc. 89 (1983), 601–607.
[447] , A new type of affine Borel function, Math. Scand. 54 (1984), 183–188.
[448] , Choquet simplexes whose set of extreme points is K-analytic, Ann. Inst.
Fourier (Grenoble) 35 (1985), 195–206.
[449] S. Teleman, An introduction to Choquet theory with the applications to reduction the-
ory, Increst Preprint Series in Mathematics (1980), 1–294.
[450] , Measure–theoretic properties of the Choquet and maximal topologies, Increst
Preprint Series in Mathematics (1982), 1–41.
[451] , On the regularity of the boundary measures, Complex analysis—fifth
Romanian-Finnish seminar, Part 2 (Bucharest, 1981), Lecture Notes in Math. 1014,
Springer, Berlin, 1983, pp. 296–315.
[452] , Topological properties of the boundary measures, Studies in probability and
related topics, Nagard, Rome, 1983, pp. 457–463.
[453] , Measure–theoretic properties of the maximal orthogonal topology, Increst
Preprint Series in Mathematics (1984), 1–35.
[454] , On the boundary barycentric calculus, J. Funct. Anal. 78 (1988), 85–98.
[455] , Sur les mesures maximales, C. R. Acad. Sci. Paris Sér. I Math. 318 (1994),
525–528.
[456] F. Topsøe and J. Hoffmann-Jørgensen, Analytic spaces and their applications, Ana-
lytic sets, Academic Press Inc. [Harcourt Brace Jovanovich Publishers], London, 1980,
pp. 317–401.
[457] F. Vasilesco, Sur les singularités des fonctions harmoniques, J. Math. Pures Appl. 9
(1930), 81–111.
[458] , La notion de point irrégulier dans le problème de Dirichlet, Hermann, Paris,
1938, Actualités Scientifiques et Industrielles, no. 660, 61 pp.
[459] J. Veselý, Sequence solutions of the Dirichlet problem, Časopis Pěst. Mat. 106 (1981),
84–93, 102, With a loose Russian summary.
[460] J. Vesterstrøm, On open maps, compact convex sets, and operator algebras, J. London
Math. Soc. (2) 6 (1973), 289–297.
[461] G. F. Vincent-Smith, Uniform approximation of harmonic functions, Ann. Inst. Fourier
(Grenoble) 19 (1969), 339–353 (1970).
694 Bibliography
[462] H. von Weizsäcker, Der Satz von Choquet-Bishop-de Leeuw für konvexe nicht kom-
pakte Mengen straffer Maße über beliebigen Grundräumen, Math. Z. 142 (1975), 161–
165.
[463] , A note on infinite dimensional convex sets, Math. Scand. 38 (1976), 321–324.
[464] H. von Weizsäcker and G. Winkler, Noncompact extremal integral representations:
some probabilistic aspects, Functional analysis: surveys and recent results, II (Proc.
Second Conf. Functional Anal., Univ. Paderborn, Paderborn, 1979), Notas Mat. 68,
North-Holland, Amsterdam, 1980, pp. 115–148.
[465] D. Werner, Some lifting theorems for bounded linear operators, Functional analysis
(Essen, 1991), Lecture Notes in Pure and Appl. Math. 150, Dekker, New York, 1994,
pp. 279–291.
[466] , Recent progress on the Daugavet property, Irish Math. Soc. Bull. (2001), 77–
97.
[467] D. V. Widder, The Laplace Transform, Princeton Mathematical Series, v. 6, Princeton
University Press, Princeton, N. J., 1941.
[468] N. Wiener, Certain notions in potential theory, J. Math. Mass. 3 (1924), 24–51.
[469] , The Dirichlet problem, J. Math. Mass. 3 (1924), 127–146.
[470] , Note on a paper of O. Perron, J. Math. Mass. 4 (1925), 21–32.
[471] S. Willard, Some examples in the theory of Borel sets, Fund. Math. 71 (1971), 187–
191.
[472] W. Wils, The ideal center of partially ordered vector spaces, Acta Math. 127 (1971),
41–77.
[473] G. Winkler, Choquet order and simplices with applications in probabilistic models,
Lecture Notes in Mathematics 1145, Springer-Verlag, Berlin, 1985.
[474] R. Wittmann, On the existence of Shilov boundaries, Proc. Amer. Math. Soc. 89 (1983),
62–64.
[475] , Shilov points and Shilov boundaries, Math. Ann. 263 (1983), 237–250.
[476] J. D. Maitland Wright, On approximating concave functions by convex functions, Bull.
London Math. Soc. 5 (1973), 221–222.
[477] K. Yosida and M. Fukamiya, On regularly convex sets, Proc. Imp. Acad. Tokyo 17
(1941), 49–52.
[478] S. Zaremba, Sur le principe de Dirichlet, Acta Math. 34 (1911), 293–316.
[479] M. Zippin, On some subspaces of Banach spaces whose duals are L1 spaces, Proc.
Amer. Math. Soc. 23 (1969), 378–385.
List of symbols
Basic notation
cA characteristic function of A; p. vii
A4B symmetric difference of A and B; p. viii
Ac complement of a set A; p. viii
f ∧ g, f ∨ g pointwise infimum and supremum of functions f, g;
p. viii
f + , f − , |f | positive part, negative part, and absolute value of a
function f ; p. viii
f |A restriction of a function f to a set A; p. viii
F b, F + bounded and positive functions from F; p. viii
ω0 , ω1 the first infinite and uncountable ordinal; p. viii
Z,N, Q, R, C set of integers, natural, rational, real and complex
numbers; p. vii
RRe z, Im z real and imaginary part of a complex number; p. vii
−
RA f (y) dy integral mean value; p. viii
S(x,r) f (y) dS(y) surface integral; p. viii
∇ gradient; p. viii
W(H) min-stable cone generated by H; p. 55
N<N finite sequences of natural numbers; p. 638
σ|n restriction of a sequence σ ∈ NN ; p. 638
|s| length of a sequence s ∈ N<N ; p. 638
oscF f (x) oscillation of a function; p. 641
Functional analysis
span A, co A, co A linear span, convex and closed convex hull of A;
p. viii
ker T kernel of an operator T ; p. viii
UE , BE , SE open unit ball, closed unit ball and unit sphere of
E; p. viii
E/F quotient space; p. viii
E⊕F sum of normed linear spaces; p. viii
E∗ dual space of E; p. viii
(x, y) scalar product in a Hilbert space; p. viii
aff A affine hull of A; p. 5
H⊥⊥ double annihilator of H in (M(K))∗ ; p. 154
coσ (A) σ-convex hull of A; p. 78
f f g, f g g infimum and supremum of {f, g} in an ordered
space; p. 200
List of symbols 697
Spaces
c0 space of sequences converging to 0; p. viii
C(X), C b (X) space of continuous and bounded continuous func-
tions on X; p. viii
`p space of p-summable sequences; p. viii
Lp (µ) space of p-integrable functions; p. viii
C n (U ), C n , C ∞ (U ), C ∞ space of n-times continuously differentiable func-
tions on U or infinitely differentiable functions on
U ; p. viii
H(U ) continuous functions on U harmonic on U ; p. 53
`∞
n Rn with the maximum norm; p. 446
H0 (U ), S0 (U ) functions continuous on U that are harmonically or
superharmonically extendable on U ; p. 504
H(U ), S(U ) function spaces on a general set U ; p. 504
T (U ) typically real functions on U ; p. 575
P(U ) holomorphic functions with positive real part;
p. 579
CM completely monotonic function; p. 586
C c (X) continuous functions with compact support; p. 627
C 0 (X) continuous functions vanishing at infinity; p. 627
1,2
W 1,2 (U ), Wloc (U ) Sobolev and local Sobolev space on U ; p. 647
C∞c infinitely differentiable functions with compact
support; p. 647
Function spaces
Kor P Korovkin closure; p. 2
Mx (H) measures H-representing x; p. 54
ChH (K) Choquet boundary for a function space H; p. 54
expH (K) H-exposed points; p. 55
A(H), Ac (H) H-affine and H-affine continuous functions; p. 55
Kc (H), Kusc (H), Klsc (H) continuous, upper semicontinuous and lower semi-
continuous H-convex functions; p. 55
S c (H), S usc (H), S lsc (H) continuous, upper semicontinuous and lower semi-
continuous H-concave functions; p. 55
Qµ,F , Qµ sublinear functional; p. 57
Ff ∗ , f ∗ , Ff , f upper and lower envelopes of f ; p. 57
∗ ∗
≺H , ≺ Choquet ordering given by a function space H;
p. 58
M(H) H-representing measures; p. 66
r : M(H) → K H-barycenter mapping; p. 67
Mmax (H), Mbnd (H) H-maximal and H-boundary measures; p. 73
Bµ Bauer functional; p. 95
l Bishop–de Leeuw preordering; p. 103
S(H) state space of H; p. 120
φ : K → S(H) evaluation mapping; p. 120
Φ : H → Ac (S(H)) embedding of H; p. 120
Hα,b functions of H-affine class α; p. 154
δx unique maximal measure representing x; p. 169
ChH (K)f ∗,α , ChH (K)f envelopes defined by non-continuous functions;
∗,α
p. 182
HT function space associated to a Markov projection
T ; p. 188
µTx representing measure associated to a Markov pro-
jection T ; p. 188
coH A closed H-convex hull of A; p. 247
Fx (H) union of supports of measures representing x;
p. 251
H2⊥ measures in H⊥ of norm less or equal than 2; p. 259
τM topology generated by M-extremal sets; p. 275
List of symbols 699
Function cones
Mx (S) measures S-representing x; p. 217
ChS (K) Choquet boundary for a function cone S; p. 217
expS (K) S-exposed points; p. 217
A(S), Ac (S) S-affine and continuous S-affine functions; p. 217
Kc (S), Kusc (S), Klsc (S) continuous, upper semicontinuous and lower semi-
continuous S-convex functions; p. 217
S c (S), S usc (S), S lsc (S) continuous, upper semicontinuous and lower semi-
continuous S-concave functions; p. 217
≺S , ≺ Choquet ordering given by a function cone S;
p. 217
Mmax (S), Mbnd (S) S-maximal and S-boundary measures; p. 217
Convex sets
ext A extreme points of A; p. 7
Ac (X) affine continuous functions on X; p. 12
r(µ) barycenter of measure µ; p. 12
Mx (Ac (X)), Mx (X) measures representing x; p. 13
r : M1 (X) → X barycenter mapping; p. 14
exp X exposed points of X; p. 20
far X farthest points of X; p. 21
face A face generated by A; p. 27
face x face generated by x; p. 27
700 List of symbols
Potential theory
c
µUx , µx abbreviations for εUx ; p. 492
Hf solution of the Dirichlet problem; p. 492
∂ reg U , ∂ irr U regular and irregular points of U ; p. 493
∂ess U essentially regular boundary points of U ; p. 494
Df essential solution of the Dirichlet problem; p. 494
Hf, Hf upper and lower solution; p. 517
U (f ) upper functions to f ; p. 517
R(U ) resolutive functions; p. 518
H +−H + harmonic lattice; p. 538
∆ Laplace operator; p. 645
H (U ) harmonic functions on U ; p. 645
Px,r Poisson kernel; p. 645
N logarithmic or Newtonian kernel; p. 646
Nµ logarithmic or Newtonian potential of a measure µ;
p. 646
∩ heat operator; p. 649
(Y, H ) harmonic structure; p. 652
∗
H (U ), S (U ) hyperharmonic and superharmonic functions on U ;
p. 652
List of symbols 701
c- B
resolutive function, 523 Baire
solution, 523 function, 147, 154, 398
c0 , 414, 416 mapping, 147
n-simplex, 5, 43, 205 measurable
function, 146, 620
A mapping, 146, 620
absolutely continuous measure, 625 property, 624
abstract property in the restricted sense,
Baire classes of mappings, 138 331, 333, 624
Borel classes, 136
set, 143, 620
Dirichlet problem, 550
space, 358, 623
additive
Baire-one function, 128, 147, 164,
Baire class, 143
369, 643
Borel class, 143
balayage
admissible
essential, 667
basis of `∞
n , 447
of a Dirac measure, 661
positive, 447
of a function, 661
mapping, 440
Banach–Alaoglu, 609
affine
Banach–Dieudonné, 127, 610
combination, 5
function, 10, 108 Banach–Mazur distance, 451
hull, 5 Banach–Stone, 323
mapping, 389 barrier, 496
subspace, 5 barycenter, 12, 17
codimension, 5 mapping, 14
dimension, 5 barycentric formula, 12, 55, 110, 116,
affinely 126, 338
independent, 5 base, 664
perfect class, 150, 185, 307, 470, essential, 666
476 Bauer, 18, 61, 63
algebra, 135, 300, 595, 604, 617 functional, 95
algebraic tensor product, 420 harmonic space, 653
analytic simplex, 185, 203, 209, 210, 358,
set, 638 371, 379, 602
simplicial space, 192 simplicial space, 185, 383
angelic space, 317 Bishop–de Leeuw, 83, 103
annihilator, 609 preordering, 103
antilattice, 200 Bishop–Phelps, 610
Archimedean set, 257, 258, 269 Borel
arcwise connected set, 464 function, 620
Arzelà–Ascoli, 617 mapping, 620
Index 705
resolutive H-
function, 518, 550 convex, 247
set, 551 exposed, 257, 267, 269
resolvable set, 640 extremal, 56, 246
restricted function space, 204 M-extremal, 274
retract, 395 S-extremal, 222
retraction, 395 µ-invariant, 598
Riesz, 176, 177, 611, 659 analytic, 638
decomposition property, 177, 611 Archimedean, 257, 258, 269
decomposition theorem, 659 arcwise connected, 464
interpolation property, 176, 611 Baire, 143, 620
refinement property, 611 Borel, 143, 620
representation theorem, 627, 628 bounded, 608
weak interpolation property, 176, Cantor, 323, 396
206 Choquet, 250, 258, 266, 394
Riesz–Herglotz, 574 clopen, 163
Rosenthal, 100 coanalytic, 638
comeager, 324, 624
S complementary, 244, 247, 320,
saturated 379
family, 655 cozero, 142, 619
set, 115 determined, 638
scattered space, 162, 164 distinguishable, 338
Schur, 610 exposed, 200, 267
property, 470 extremal, 10, 26
semicontinuous finely regular, 532
function, 621 functionally
mapping, 389 closed, 142, 619
semiextremal set, 307 open, 142, 619
semipolar set, 665 Greenian, 654
semiregular hereditary upwards, 233
point, 547 Keldysh, 542
sequence, 547 meager, 97, 154, 624
separation theorem, 639 measure
set convex, 30, 31, 48
Fσ , 619 extremal, 36, 241
Gδ , 619 negligible, 521
K- of abstract
analytic, 638 additive class, 136
countably determined, 364, 638 ambiguous class, 136
c
A (H)-exposed, 266, 267 multiplicative class, 136
Index 713