Measure
Measure
Measure
Dietmar A. Salamon
ETH Z urich
9 September 2016
ii
Preface
This book is based on notes for the lecture course Measure and Integration
held at ETH Z urich in the spring semester 2014. Prerequisites are the first
year courses on Analysis and Linear Algebra, including the Riemann inte-
gral [9, 18, 19, 21], as well as some basic knowledge of metric and topological
spaces. The course material is based in large parts on Chapters 1-8 of the
textbook Real and Complex Analysis by Walter Rudin [17]. In addition
to Rudins book the lecture notes by Urs Lang [10, 11], the five volumes on
measure theory by David H. Fremlin [4], the paper by Heinz Konig [8] on
the generalized RadonNikod ym theorem, the lecture notes by C.E. Heil [7]
on absolutely continuous functions, Dan Mas Topology Blog [12] on exotic
examples of topological spaces, and the paper by Gert K. Pedersen [16] on
the Haar measure were very helpful in preparing this manuscript.
This manuscript also contains some material that was not covered in the
lecture course, namely some of the results in Sections 4.5 and 5.2 (concerning
the dual space of Lp () in the non -finite case), Section 5.4 on the Gen-
eralized RadonNikod ym Theorem, Sections 7.6 and 7.7 on Marcinkiewicz
interpolation and the CalderonZygmund inequality, and Chapter 8 on the
Haar measure.
I am grateful to many people who helped to improve this manuscript.
Thanks to the students at ETH who pointed out typos or errors in earlier
drafts. Thanks to Andreas Leiser for his careful proofreading. Thanks to
Theo Buehler for many enlightening discussions and for pointing out the
book by Fremlin, Dan Mas Topology Blog, and the paper by Pedersen.
Thanks to Urs Lang for his insightful comments on the construction of the
Haar measure.
iii
iv
Contents
Introduction 1
3 Borel Measures 81
3.1 Regular Borel Measures . . . . . . . . . . . . . . . . . . . . . 81
3.2 Borel Outer Measures . . . . . . . . . . . . . . . . . . . . . . . 92
3.3 The Riesz Representation Theorem . . . . . . . . . . . . . . . 97
3.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
4 Lp Spaces 113
4.1 Holder and Minkowski . . . . . . . . . . . . . . . . . . . . . . 113
4.2 The Banach Space Lp () . . . . . . . . . . . . . . . . . . . . . 115
4.3 Separability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
v
vi CONTENTS
6 Differentiation 185
6.1 Weakly Integrable Functions . . . . . . . . . . . . . . . . . . . 185
6.2 Maximal Functions . . . . . . . . . . . . . . . . . . . . . . . . 190
6.3 Lebesgue Points . . . . . . . . . . . . . . . . . . . . . . . . . . 196
6.4 Absolutely Continuous Functions . . . . . . . . . . . . . . . . 201
6.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
References 289
Introduction
We learn already in high school that integration plays a central role in math-
ematics and physics. One encounters integrals in the notions of area or
volume, when solving a differential equation, in the fundamental theorem of
calculus, in Stokes theorem, or in classical and quantum mechanics. The
first year analysis course at ETH includes an introduction to the Riemann
integral, which is satisfactory for many applications. However, it has certain
drawbacks, in that some very basic functions are not Riemann integrable,
that the pointwise limit of a sequence of Riemann integrable functions need
not be Riemann integrable, and that the space of Riemann integrable func-
tions is not complete with respect to the L1 -norm. One purpose of this book
is to introduce the Lebesgue integral, which does not suffer from these draw-
backs and agrees with the Riemann integral whenever the latter is defined.
Chapter 1 introduces abstract integration theory for functions on measure
spaces. It includes proofs of the Lebesgue Monotone Convergence Theorem,
the Lemma of Fatou, and the Lebesgue Dominated Convergence Theorem.
In Chapter 2 we move on to outer measures and introduce the Lebesgue
measure on Euclidean space. Borel measures on locally compact Hausdorff
spaces are the subject of Chapter 3. Here the central result is the Riesz
Representation Theorem. In Chapter 4 we encounter Lp spaces and show
that the compactly supported continuous functions form a dense subspace of
Lp for a regular Borel measure on a locally compact Hausdorff space when
p < . Chapter 5 is devoted to the proof of the RadonNikod ym theorem
q
about absolutely continuous measures and to the proof that L is naturally
isomorphic to the dual space of Lp when 1/p + 1/q = 1 and 1 < p < .
Chapter 6 deals with differentiation. Chapter 7 introduces product measures
and contains a proof of Fubinis Theorem, an introduction to the convolu-
tion product on L1 (Rn ), and a proof of the CalderonZygmund inequality.
Chapter 8 constructs Haar measures on locally compact Hausdorff groups.
1
2 CONTENTS
Despite the overlap with the book of Rudin [17] there are some differ-
ences in exposition and content. A small expository difference is that in
Chapter 1 measurable functions are defined in terms of pre-images of (Borel)
measurable sets rather than pre-images of open sets. The Lebesgue measure
in Chapter 2 is introduced in terms of the Lebesgue outer measure instead of
as a corollary of the Riesz Representation Theorem. The notion of a Radon
measure on a locally compact Hausdorff space in Chapter 3 is defined in
terms of inner regularity, rather than outer regularity together with inner
regularity on open sets. This leads to a somewhat different formulation of
the Riesz Representation Theorem (which includes the result as formulated
by Rudin). In Chapters 4 and 5 it is shown that Lq () is isomorphic to
the dual space of Lp () for all measure spaces (not just the -finite ones)
whenever 1 < p < and 1/p + 1/q = 1. It is also shown that L () is
isomorphic to the dual space of L1 () if and only if the measure space is
localizable. Chapter 5 includes a generalized version of the RadonNikod ym
theorem for signed measures, due to Fremlin [4], which does not require that
the underying measure is -finite. In the formulation of Konig [8] it asserts
that a signed measure admits a -density if and only if it is both absolutely
continuous and inner regular with respect to . In addition the present
book includes a self-contained proof of the CalderonZygmund inequality in
Chapter 7 and an existence and uniqueness proof for (left and right) Haar
measures on locally compact Hausdorff groups in Chapter 8.
The book is intended as a companion for a foundational one semester
lecture course on measure and integration and there are many topics that it
does not cover. For example the subject of probability theory is only touched
upon briefly at the end of Chapter 1 and the interested reader is referred to
the book of Malliavin [13] which covers many additional topics including
Fourier analysis, limit theorems in probability theory, Sobolev spaces, and
the stochastic calculus of variations. Many other fields of mathematics re-
quire the basic notions of measure and integration. They include functional
analysis and partial differential equations (see e.g. GilbargTrudinger [5]),
geometric measure theory, geometric group theory, ergodic theory and dy-
namical systems, and differential topology and geometry.
There are many other textbooks on measure theory that cover most or
all of the material in the present book, as well as much more, perhaps from
somewhat different view points. They include the book of Bogachev [2]
which also contains many historical references, the book of Halmos [6], and
the aforementioned books of Fremlin [4], Malliavin [13], and Rudin [17].
Chapter 1
3
4 CHAPTER 1. ABSTRACT MEASURE THEORY
1.1 -Algebras
For any fixed set X denote by 2X the set of all subsets of X and, for any
subset A X, denote by Ac := X \ A its complement.
Definition 1.1 (Measurable Space). Let X be a set. A collection A 2X
of subsets of X is called a -algebra if it satisfies the following axioms.
(a) X A.
(b) If A A then Ac A.
(c) Every countable union of elements of A is again an element of A, i.e. if
Ai A for i = 1, 2, 3, . . . then
S
i=1 i A.
A
A measurable space is a pair (X, A) consisting of a set X and a -algebra
A 2X . The elements of a -algebra A are called measurable sets.
Lemma 1.2. Every -algebra A 2X satisfies the following.
(d) A.
(e) If n N and A1 , . . . , An A then ni=1 Ai A.
S
Exercise 1.7. Let X be any set and let I be any nonempty index set.
Suppose that for every
T i I a -algebra Ai 2X is given. Prove that the
intersection A := iI Ai = {A X | A Ai for all i I} is a -algebra.
Lemma 1.8. Let X be a set and E 2X be any set of subsets of X. Then
there is a unique smallest -algebra A 2X containing E (i.e. A is a -
algebra, E A, and if B is any other -algebra with E B then A B).
Proof. Uniqueness follows directly from the definition. Namely, if A and B
are two smallest -algebras containing E, we have both B A and A B
and hence A = B. To prove existence, denote by S 22 the collection of
X
Conditions (a) and (b) in Definition 1.9 are also properties of every -
algebra. However, condition (c) in Definition 1.9 is not shared by -algebras
because it permits arbitrary unions. On the other hand, complements of
open sets are typically not open. Many of the topologies used in this book
arise from metric spaces and are familiar from first year analysis. Here is a
recollection of the definition.
Definition 1.10 (Metric Space). A metric space is a pair (X, d) con-
sisting of a set X and a function d : X X R satisfying the following
axioms.
(a) d(x, y) 0 for all x, y X, with equality if and only if x = y.
(b) d(x, y) = d(y, x) for all x, y X.
(c) d(x, z) d(x, y) + d(y, z) for all x, y, z X.
A function d : X X R that satisfies these axioms is called a distance
function and the inequality in (c) is called the triangle inequality. A
subset U X of a metric space (X, d) is called open (or d-open) if, for
every x U , there exists a constant > 0 such that the open ball
d(x, y) := kx yk
Example 1.12. The set X = R of real numbers is a metric space with the
standard distance function
d(x, y) := |x y|.
R := [, ] := R {, }.
For a, b R define
whenever the right hand side exists. Find an example where the inequality
is strict.
(iv) If an bn for all n N show that lim inf n an lim inf n bn .
Definition 1.15. Let (X, U) be a topological space and let B 2X be the
smallest -algebra containing U. Then B is called the Borel -algebra of
(X, U) and the elements of B are called Borel (measurable) sets.
Lemma 1.16. Let (X, U) be a topological space. Then the following holds.
(i) Every closed subset F X is a Borel set.
(ii) Every countable union
S
i=1 Fi of closed subsets Fi X is a Borel set.
(These are sometimes called F -sets.)
(iii) Every countable intersection
T
i=1 Ui of open subsets Ui X is a Borel
set. (These are sometimes called G -sets.)
Proof. Part (i) follows from the definition of Borel sets and condition (b) in
Definition 1.1, part (ii) follows from (i) and (c), and part (iii) follows from (ii)
and (b), because the complement of an F -set is a G -set.
Consider for example the Borel -algebra on the real axis R with its stan-
dard topology. In view of Lemma 1.16 it is a legitimate question whether
there is any subset of R at all that is not a Borel set. The answer to this
question is positive, which may not be surprising, however the proof of the
existence of subsets that are not Borel sets is surprisingly nontrivial. It will
only appear much later in this book, after we have introduced the Lebesgue
measure (see Lemma 2.15). For now it is useful to note that, roughly speak-
ing, every set that one can construct in terms of some explicit formula, will
be a Borel set, and one can only prove with the Axiom of Choice that subsets
of R must exist that are not Borel sets.
10 CHAPTER 1. ABSTRACT MEASURE THEORY
f 1 (B) := {x X | f (x) B}
f 1 (Y ) = X, f 1 () = , (1.1)
f 1 (Y \ B) = X \ f 1 (B) (1.2)
B AY = f 1 (B) AX .
V UY = f 1 (V ) AX .
Proof. That (i) implies (ii), (iii), (iv), and (v) follows directly from the def-
initions. We prove that (ii) implies (i). Thus let f : X R be a function
such that f 1 ((a, ]) AX for every a R and define
B := f AX = B R | f 1 (B) AX 2R .
14 CHAPTER 1. ABSTRACT MEASURE THEORY
by axiom (c) in Definition 1.1. Hence it follows from (f) in Lemma 1.2 that
(a, b) = [, b) (a, ] B
for every pair of real numbers a < b. Since every open subset of R is a
countable union of sets of the form (a, b), (a, ], [, b), it follows from
axiom (c) in Definition 1.1 that every open subset of R is an element of B.
Hence it follows from Theorem 1.20 that f is measurable. This shows that (ii)
implies (i). That either of the conditions (iii), (iv), and (v) also implies (i) is
shown by a similar argument which is left as an exercise for the reader. This
proves Theorem 1.21.
Our next goal is to show that sums, products, and limits of measurable
functions are again measurable. The next two results are useful for the proofs
of these fundamental facts.
Theorem 1.22 (Vector Valued Measurable Functions). Let (X, A) be
a measurable space and let f = (f1 , . . . , fn ) : X Rn be a function. Then f
is measurable if and only if fi : X R is measurable for each i.
Proof. For i = 1, . . . , n define the projection i : Rn R by i (x) := xi for
x = (x1 , . . . , xn ) R. Since i is continuous it follows from Theorems 1.19
and 1.20 that if f is measurable so is fi = i f for all i. Conversely, suppose
that fi is measurable for i = 1, . . . , n. Let ai < bi for i = 1, . . . , n and define
Q(a, b) := {x Rn | ai < xi < bi i} = (a1 , b1 ) (an , bn ).
Then n
\
f 1
(Q(a, b)) = fi1 ((ai , bi )) A
i=1
by property (f) in Lemma 1.2. Now every open subset of Rn can be expressed
as a countable union of sets of the form Q(a, b). (Prove this!) Hence it follows
from axiom (c) in Definition 1.1 that f 1 (U ) A for every open set U Rn
and hence f is measurable. This proves Theorem 1.22.
1.2. MEASURABLE FUNCTIONS 15
Step Functions
We close this section with a brief discussion of measurable step functions.
Such functions will play a central role throughout this book. In particular,
they are used in the definition of the Lebesgue integral.
Definition 1.25 (Step Function). Let X be a set. A function s : X R
is called a step function (or simple function) if it takes on only finitely
many values, i.e. the image s(X) is a finite subset of R.
Let s : X R be a step function, write s(X) = {1 , . . . , ` } with i 6= j
for i 6= j, and define Ai := s1 (i ) = {x X | s(x) = i } for i = 1, . . . , `.
Then the sets A1 , . . . , A` form a partition of X, i.e.
`
[
X= Ai , Ai Aj = for i 6= j. (1.6)
i=1
a + b = a + c, a< = b = c,
ab = ac, 0<a< = b = c.
Neither of these assertions extend to the case a = .
Definition 1.27 (Measure). Let (X, A) be a measurable space. A measure
on (X, A) is a function
: A [0, ]
satisfying the following axioms.
(a) is -additive, i.e. if Ai A, i = 1, 2, 3, . . . , is a sequence of pairwise
disjoint measurable sets then
!
[ X
Ai = (Ai ).
i=1 i=1
Proof. We prove (i). Choose A1 A such that (A1 ) < and define Ai :=
for i > 1. Then it follows from -additivity that
X
(A1 ) = (A1 ) + ()
i>1
Here the last equation follows from part (ii). This proves part (iv).
1.3. INTEGRATION OF NONNEGATIVE FUNCTIONS 19
P I [0, ] : i 7 i determines
algebra in Example 1.5. Then any function S a
measure : A [0, ] via (AJ ) := jJ j for J I and AJ = jJ Aj .
20 CHAPTER 1. ABSTRACT MEASURE THEORY
where the supremum is taken over all measurable step function s : X [0, )
that satisfy s(x) f (x) for all x X.
The reader may verify that the right hand side of (1.10) depends only on s
and not on the choice of i and Ai . The same definition can be used if f is
only defined on the measurable set E X. Then AE := {A A | A E} is
a -algebra on E and RE := |AE is a measure. So (E, AE , E ) is a measure
space and the integral E f dE is well defined. It agrees with the integral of
the extended function on X, defined by f (x) := 0 for x X \ E.
Theorem 1.35 (Basic Properties of the Lebesgue Integral).
Let (X, A, ) be a measure space and let f, g : X [0, ] be measurable
functions and let E A. Then the following holds.
R R
(i) If f g on E then E f d E g d.
R R
(ii) E f d = X f E d.
R
(iii) If f (x) = 0 for all x E then E f d = 0.
R
(iv) If (E) = 0 then E f d = 0.
R R
(v) If A A and E A then E f d A f d.
R R
(vi) If c [0, ) then E cf d = c E f d.
1.3. INTEGRATION OF NONNEGATIVE FUNCTIONS 21
m X
X n
s+t= (i + j )Ai Bj
i=1 j=1
and hence
Z X n
m X
(s + t) d = (i + j )(Ai Bj E)
E i=1 j=1
Xm X n n
X m
X
= i (Ai Bj E) + j (Ai Bj E)
i=1 j=1 j=1 i=1
Xm n
X Z Z
= i (Ai E) + j (Bj E) = s d + t d.
i=1 j=1 E E
Define f : X [0, ] by
and hence Z
f d. (1.12)
X
This function is a measure by part (ii) of Lemma 1.36 (which asserts that s is
-additive) and by part (iv) of Theorem 1.35 (which asserts that s () = 0).
Now fix a constant 0 < c < 1 and define
Here the last inequality follows from the definition of in (1.11). Hence
s (En ) for all n N. (1.15)
c
Since s : A [0, ] is a measure, by part (i) of Theorem 1.35, it follows
from equation (1.14) and part (iv) of Theorem 1.28 that
Z
s d = s (X) = lim s (En ) . (1.16)
X n c
Here the first and last equations follow from Theorem 1.37 and the second
equation follows from part (i) of Lemma 1.36. This proves (i) for E = X. To
prove it in general, replace f, g by f E , gE and use part (ii) of Theorem 1.35.
26 CHAPTER 1. ABSTRACT MEASURE THEORY
Here the second equation follows from the definition of gn and the third
equation follows from part (i) of the present theorem (already proved). This
proves (ii) for E = X. To prove it in general replace f, fn by f E , fn E and
use part (ii) of Theorem 1.35.
We prove (iii). Let f : X [0, ] be a measurable function and let
Ek A be a sequence of pairwise disjoint measurable sets. Define
[ n
X
E := Ek , fn := f Ek .
k=1 k=1
Then it follows from part (i) of the present theorem (already proved) and
part (ii) of Theorem 1.35 that
Z Z Xn n Z
X n Z
X
fn d = f Ek d = f Ek d = f d.
X X k=1 k=1 X k=1 Ek
is a measure and Z Z
g df = f g d (1.22)
E E
for every measurable function g : X [0, ] and every E A.
Proof. f is -additive by part (iii) of Theorem 1.38 and f () = 0 by
part (iv) of Theorem 1.35. Hence f is a measure (see Definition 1.27). Now
let g := A be the characteristic function of a measurable set A A. Then
Z Z Z
A df = f (A) = f d = f A d.
X A X
Here the first equation follows from the definition of the integral for measur-
able step functions in Definition 1.34, the second equation follows from the
definition of f , and the last equation follows from part (ii) of Theorem 1.35.
Thus equation (1.22) (with E = X) holds for characteristic functions of
measurable sets. Taking finite sums and using part (vi) of Theorem 1.35 and
part (i) of Theorem 1.38 we find that (1.22) (with E = X) continues to hold
for all measurable step functions g = s : X [0, ). Now approximate an
arbitrary measurable function g : X [0, ] by a sequence of measurable
step functions via Theorem 1.26 and use the Lebesgue Monotone Conver-
gence Theorem 1.37 to deduce that equation (1.22) holds with E = X for
all measurable functions g : X [0, ]. Now replace g by gE and use
part (ii) of Theorem 1.35 to obtain equation (1.22) in general. This proves
Theorem 1.40.
28 CHAPTER 1. ABSTRACT MEASURE THEORY
L1 () := L1 (X, A, ) := {f : X R | f is -integrable} .
= c f d c f d
+
ZE E
= c f d.
E
Here the second equation follows from part (vi) of Theorem 1.35. If c < 0
then (cf )+ = (c)f and (cf ) = (c)f + and hence, again using part (iv)
of Theorem 1.35, we obtain
Z Z Z
cf d = (c)f d (c)f + d
E E Z E Z
= (c) f d (c) f + d
Z E E
= c f d.
E
1.4. INTEGRATION OF REAL VALUED FUNCTIONS 31
h+ + f + g = h + f + + g + .
Hence
Z Z Z
h d = +
h d h d
E ZE ZE Z Z
= +
f d + g d +
f d g d
ZE Z E E E
= f d + g d
E E
Proof. f is measurable by part (ii) of Theorem 1.24 and |f (x)| g(x) for all
x X by (1.30) and (1.31). Hence it follows from part (i) of Theorem 1.35
that Z Z
|f | d g d <
X X
Here penultimate step follows from part (i) of Theorem 1.44. This implies
Z
lim sup |fn f | d 0.
n X
1.5. SETS OF MEASURE ZERO 33
Hence Z
lim |fn f | d = 0.
n X
Since Z Z Z Z
fn d f d |fn f | d |fn f | d
E E E X
by part (iii) of Theorem 1.44 it follows that
Z Z
lim fn d
f d = 0,
n E E
The first observation is that every nonnegative function with finite inte-
gral is almost everywhere finite.
R
Lemma 1.47. Let f : X [0, ] be a measurable function. If X f d <
then f < almost everywhere.
R N := {x
Proof. Define R X | f (x) = } and h := N . Then h f and so
(N ) = X h d X f d < by Theorem 1.35. Hence (N ) = 0.
The second observation is that if two integrable, or nonnegative measur-
able, functions agree almost everywhere, then their integrals agree over every
measurable set.
Lemma 1.48. Assume either that f, g : X [0, ] are measurable functions
that agree almost everywhere or that f, g : X R are -integrable functions
that agree almost everywhere. Then
Z Z
f d = g d for all A A. (1.33)
A A
Here the first equation follows from part (iii) of Theorem 1.38 in the non-
negative case and from part (iv) of Theorem 1.44 in the integrable case.
The second equation follows from part (iv) of Theorem 1.35 in the non-
negative case and from part (vi) of Theorem 1.44 in the integrable case.
The third equation follows from part (ii) of Theorem 1.35 in the nonnega-
tive case and from part (v) of Theorem 1.44 in the integrable case. Since
f A\N = gA\N it follows that the integrals of f and g over A agree. This
proves Lemma 1.48.
The converse of Lemma 1.48 fails for nonnegative measurable functions.
For example, if X is a singleton and (X) = then the integrals of any two
positive functions agree over every measurable set. However, the converse of
Lemma 1.48 does hold for integrable functions. Since the difference of two
integrable functions is again integrable, it suffices to assume g = 0, and in
this case the converse also holds for nonnegative measurable functions. This
is the content of the next lemma.
1.5. SETS OF MEASURE ZERO 35
is a set of measure zero. In the integrable case apply this argument to the
function |f | : X [0, ). This proves Lemma 1.49.
Lemma 1.50. Let f L1 (). Then
Z Z
f d =
|f | d (1.34)
X X
takes on the same value on all the elements in a given equivalence class and
so descends to the quotient space L1 (). By Lemma 1.49 it defines a norm
on L1 () and Theorem 1.53 below shows that L1 () is a Banach space with
this norm (i.e. a complete normed vector space).
Theorem 1.52 (Convergent Series of Integrable Functions).
Let (X, A, ) be a measure space and let fn : X R be a sequence of
-integrable functions such that
X Z
|fn | d < . (1.36)
n=1 X
Proof. Define
X
(x) := |fk (x)|
k=1
for x X. This function is measurable by part (ii) of Theorem 1.24. More-
over, it follows from the Lebesgue Monotone Convergence Theorem 1.37 and
from part (i) of Theorem 1.38 that
Z Z X n n Z
X Z
X
d = lim |fk | d = lim |fk | d = |fk | d < .
X n X k=1 n X X
k=1 k=1
These
R functions
R are measurable by part (i) of Theorem 1.24. Moreover,
X
g d = X
d < by Lemma 1.48. Since |gn (x)| g(x) for all n N
and gn converges pointwise to f it follows from the Lebesgue Dominated
Convergence Theorem 1.45 that f L1 () and, for all A A,
Z Z Z X n X Z
f d = lim gn d = lim fk d = fn d.
A n A n A k=1 A
n=1
Here the second step follows from Lemma 1.48 because gn = nk=1 fk almost
P
everywhere. The last step follows by interchanging sum and integral, using
part (i) of
PTheorem 1.44. This proves (1.38). To prove equation (1.39) note
n
that f k=1 fk = f gn almost everywhere, that f (x) gn (x) converges to
zero for all x X, and that |f gn | |f |+g where |f |+g is integrable. Hence,
by Lemma 1.48 and the Lebesgue Dominated Convergence Theorem 1.45
Z X n Z
lim f fk d = lim |f gn | d = 0,
n X n X
k=1
Define
f := fn1 + g.
Pi1
Then fni = fn1 + j=1 gj converges almost everywhere to f . We prove (1.41).
R
Let > 0. By (1.42) there is an ` N such that X |fnkR f | d < /2 for all
k `. By (1.40) the integer ` can be chosen such that X |fn fm | d < /2
for all n, m n` . Then
Z Z Z
|fn f | d |fn fn` | d + |fn` f | d <
X X X
N A, (N ) = 0, EN = E A.
|A = .
N := {x X | f (x) 6= f (x)} A
A, B A,
(E) := (A) where A E B, (1.44)
(B \ A) = 0.
Corollary 1.56. Let (X, A, ) be a measure space and let (X, A , ) be its
completion. Denote the equivalence class of a -integrable function f L1 ()
under the equivalence relation (1.35) in Definition 1.51 by
n o
[f ] := g L1 () {x X | f (x) 6= g(x)} = 0 .
1.7 Exercises
Exercise 1.57. Let X be an uncountable set and let A 2X be the set of
all subsets A X such either A or Ac is countable. Define
0, if A is countable,
(A) :=
1, if Ac is countable,
Prove that Z 1
lim fn (x) dx = 0,
n 0
Exercise 1.63. Let (X, A, ) be a measure space such that (X) < and
let f : X R be a measurable function. Show that f is integrable if and
only if
X
|({x X | |f (x)| > n})| < .
n=1
1.7. EXERCISES 45
function c f .)
46 CHAPTER 1. ABSTRACT MEASURE THEORY
A := B Y | 1 (B) A 2Y .
(1.47)
for all k and some positive real number then the central limit theorem
of LindebergLevy asserts that the sequence Tn := Sn / n converges in dis-
tribution to a so-called standard normal randomR variable with expectation
x 2
zero and variance one, i.e. limn FTn (x) = 12 et /2 dt for all x R.
The hypotheses listed above are quite restrictive and in modern probabil-
ity theory these theorems are often needed in much greater generality. For
proofs, many examples, and comprehensive expositions of probability theory
see Ash [1], Fremlin [4, Chapter 27], Malliavin [13].
An important class of random variables are those where the distribution
functions FX : R [0, 1] are absolutely continuous (Theorem 6.19). This
means that the pushforward measures X P on the Borel -algebra B 2R
admit densities as in Theorem 1.40 with respect to the Lebesgue measure.
The Lebesgue measure is introduced in Chapter 2 and the existence of a
density is the subject of Chapter 5 on the RadonNikod ym Theorem.
Chapter 2
This chapter introduces the most important example, namely the Lebesgue
n
measure on Euclidean space. Let n N and denote by B 2R the -algebra
of all Borel sets in Rn , i.e. the smallest -algebra on Rn that contains all open
sets in the standard topology (Definition 1.15). Then
B + x := {y + x | y B} B for all B B and all x Rn ,
because the translation Rn Rn : y 7 y + x is a homeomorphism. A
measure : B [0, ] is called translation invariant if it satisfies
(B + x) = (B) for all B B and all x Rn . (2.1)
Theorem 2.1. There exists a unique measure : B [0, ] that is trans-
lation invariant and satisfies the normalization condition ([0, 1]n ) = 1.
Proof. See page 64.
Definition 2.2. Let (Rn , B, ) be the measure space in Theorem 2.1 and
denote by (Rn , A, m) its completion as in Theorem 1.55. Thus
n there exist Borel sets B0 , B1 B
A := A R (2.2)
such that B0 A B1 and (B1 \ B0 ) = 0
and m(A) := (B0 ) for A A, where B0 , B1 B are chosen such that
B0 A B1 and (B1 \ B0 ) = 0. The elements of A are called Lebesgue
measurable subsets of Rn , the function m : A [0, ] is called the
Lebesgue measure, and the triple (Rn , A, m) is called the Lebesgue mea-
sure space. A function f : Rn R is called Lebesgue measurable if it
is measurable with respect to the Lebesgue -algebra A on Rn (and the Borel
-algebra on the target space R).
49
50 CHAPTER 2. THE LEBESGUE MEASURE
Step 2. If A A then Ac A.
Let A A. Since
D Ac = D \ A, D \ Ac = D A,
(D) = (D A) + (D \ A)
= (D A) + (D Ac )
= (D A) + (D Ac B) + ((D Ac ) \ B)
((D A) (D Ac B)) + (D Ac B c )
= (D (A B)) + (D (A B)c )
= (D (A B)) + (D \ (A B)).
Here the inequality follows from axioms (a) and (c) in Definition 2.3. Using
axioms (a) and (c) again we obtain (D) = (D (A B)) + (D \ (A B))
for every subset D X and hence A B A.
Step 4. Let Ai A for i N such that Ai Aj = for i 6= j. Then
[
X
A := Ai A, (A) = (Ai ).
i=1 i=1
For k N define
Bk := A1 A2 Ak .
Then Bk A for all k N by Step 3. Now let D X. Then, for all k 2,
(D Bk ) = (D Bk Ak ) + ((D Bk ) \ Ak )
= (D Ak ) + (D Bk1 )
(D) = (D Bk ) + (D \ Bk )
k
X
= (D Ai ) + (D \ Bk )
i=1
Xk
(D Ai ) + (D \ A).
i=1
Here the last two inequalities follow from axiom (c). Hence
X
(D) = (D Ai ) + (D \ A) = (D A) + (D \ A) (2.5)
i=1
given by (2.4) and let B 2X the Borel -algebra of (X, d). Then the
following are equivalent.
(i) B A().
(ii) If A, B X satisfy d(A, B) := inf aA, bB d(a, b) > 0 then
(A B) = (A) + (B).
Proof. We prove that (i) implies (ii). Thus assume that satisfies (i). Let
A, B X such that := d(A, B) > 0. Define
[
U := x X a A such that d(a, x) < = B (a).
aA
For each k N the set Uk is open and d(x, y) 1/k for all x D A and
all y D \ Uk . Hence
1
d(D A, D \ Uk ) .
k
By (ii) and axiom (b) this implies
and hence
D \ A = (D \ Uk ) (D (Uk \ A))
[
= (D \ Uk ) (D (Ui \ Ui+1 )).
i=k
Thus
[
D \ A = (D \ Uk ) Ei , Ei := (D Ui ) \ Ui+1 . (2.7)
i=k
P
Claim 3. The outer measures of the Ei satisfy i=1 (Ei ) < .
Claim 3 implies Claim 2. It follows from Claim 3 that the sequence
X
k := (Ei )
i=k
converges to zero. Moreover, it follows from equation (2.7) and axiom (c) in
Definition 2.3 that
X
(D \ A) (D \ Uk ) + (Ei ) = (D \ Uk ) + k .
i=k
and !
n
X n
[
(E2i1 ) = E2i1 (D)
i=1 i=1
P
for every n N. Hence i=1 (Ei ) 2(D) < and this shows that
Claim 4 implies Claim 3.
Proof of Claim 4. We show that
1
d(Ei , Ej ) for j i + 2.
(i + 1)(i + 2)
Then x
/ Ui+1 because Ei Ui+1 = . (See equation (2.7).) Hence
1
d(a, x) for all a A.
i+1
This implies
n
Theorem 2.10 (The Lebesgue Outer Measure). Let : 2R [0, ]
be the function defined by (2.10). Then the following holds.
(i) is an outer measure.
(ii) is translation invariant, i.e. for all A Rn and all x Rn
(A + x) = (A).
and fix a constant > 0. Then it follows from Definition 2.9 that, for i N,
there exists a sequence of cuboids Qij Qn , j N, such that
[ X
Ai Qij , Vol(Qij ) < + (Ai ).
j=1 j=1
2i
Hence
[ X X X
A Qij , Vol(Qij ) < + (Ai ) = + (Ai ).
i,jN i,jN i=1
2i i=1
This implies
X
(A) < + (Ai )
i=1
P
for every > 0 and thus (A) i=1 (Ai ). This proves part S
(i).
We prove (ii). If A i=1 Qi with Qi Qn , then A + x
S
i=1 (Qi + x)
for every x Rn and Vol(Qi + x) = Vol(Qi ) by definition of the volume.
Hence part (ii) follows from Definition 2.9.
58 CHAPTER 2. THE LEBESGUE MEASURE
d(A, B)
diam(Qi ) := sup |x y| < .
x,yQi 2
I J = , I := {i N | Qi A 6= }, J := {i N | Qi B 6= }.
Hence
X X
(A) + (B) Vol(Qi ) + Vol(Qi )
iI iJ
X
Vol(Qi )
i=1
< (A B) + .
Thus (A) + (B) < (A B) + for all > 0, so (A) + (B) (A B),
and hence (A) + (B) = (A B), by axioms (a) and (c) in Definition 2.3.
This proves part (iii).
We prove (iv) by an argument due to von Neumann. Fix a closed cuboid
Q = I1 In , Ii = [ai , bi ].
We claim that
Vol(Q) (Q). (2.11)
Equivalently, if Qi Qn , i N, is a sequence of closed cuboids then
[
X
Q Qi = Vol(Q) Vol(Qi ). (2.12)
i=1 i=1
2.2. THE LEBESGUE OUTER MEASURE 59
|I| := b a.
Then
|I| 1 #(I Z) |I| + 1.
Hence
N |I| 1 #(N I Z) N |I| + 1
and thus
1 1 1 1
|I| # I Z |I| +
N N N N
for every integer N N. Take the limit N to obtain
1 1
|I| = lim # I Z .
N N N
Thus
n
Y 1 1
Vol(Q) = lim # Ij Z
N
j=1
N N
(2.13)
1 1 n
= lim n # Q Z .
N N N
This implies
Xk Xk
1 1 n 1 1 n 1 1 n
# Q Z # Ui Z # Ui Z .
Nn N i=1
Nn N i=1
Nn N
60 CHAPTER 2. THE LEBESGUE MEASURE
Since > 0 can be chosen arbitrarily small, this proves (2.12) and (2.11).
Thus we have proved that (Q) Vol(Q) (Q) and so (Q) = Vol(Q).
To prove that (int(Q)) = Vol(Q), fix a constant > 0 and choose a closed
cuboid P Qn such that
P int(Q), Vol(Q) < Vol(P ).
Then
Vol(Q) < Vol(P ) = (P ) (int(Q)).
Thus Vol(Q) < (int(Q)) for all > 0. Hence, by axiom (b),
Vol(Q) (int(Q)) (Q) = Vol(Q),
and hence (int(Q)) = Vol(Q). This proves part (iv) and Theorem 2.10.
n
Definition 2.11. Let : 2R [0, ] be the Lebesgue outer measure. A
subset A Rn is called Lebesgue measurable if A is -measurable, i.e.
(D) = (D A) + (D \ A) for all D Rn .
The set of all Lebesgue measurable subsets of Rn will be denoted by
A := A Rn A is Lebesgue measurable .
The function
m := |A : A [0, ]
is called the Lebesgue measure on Rn . A function f : Rn R is called
Lebesgue measurable if it is measurable with respect to the Lebesgue -
algebra A on Rn (and the Borel -algebra on the target space R).
2.2. THE LEBESGUE OUTER MEASURE 61
Proof. Assertion (i) follows from Theorem 2.4 and part (i) of Theorem 2.10.
Assertion (ii) follows from the definitions and part (ii) of Theorem 2.10.
Assertion (iii) follows from Theorem 2.5 and part (iii) of Theorem 2.10. As-
sertion (iv) follows from (iii) and part (iv) of Theorem 2.10.
The restriction of the measure m in Corollary 2.12 to the Borel -algebra
of Rn satisfies the requirements of Theorem 2.1 (translation invariance and
normalization) and hence settles the existence problem. The uniqueness
proof relies on certain regularity properties of the measure m which are
established in the next theorem along with continuity from below for the
Lebesgue outer measure . Theorem 2.14 shows that m is the completion of
its restriction to the Borel -algebra of Rn and, with that at hand, we can
then prove uniqueness in Theorem 2.1.
n
S of subsets of R such that Ai Ai+1 for all i N then
(iii) If Ai is a sequence
their union A := i=1 Ai has Lebesgue outer measure (A) = limi (Ai ).
This shows that A B for every A A. Thus we have proved that (I)
implies (II) and this completes the proof of Theorem 2.14.
Proof of Theorem 2.1. The existence of a translation invariant normalized
Borel measure on Rn follows from Corollary 2.12. We prove uniqueness.
Thus assume that 0 : B [0, ] is a translation invariant measure such
that 0 ([0, 1]n ) = 1. Define := 0 ([0, 1)n ). Then 0 1. We prove in
five steps that = 1 and 0 = .
Step 1. For x = (x1 , . . . , xn ) and k N0 := N {0} define
R(x, k) := [x1 , x1 + 2k ) [xn , xn + 2k ).
Then 0 (R(x, k)) = 2nk = (R(x, k)).
Fix an integer k N0 . Since R(x, k) = R(0, k) + x for every x Rn it follows
from the translation invariance of 0 that there is a constant ck 0 such that
0 (R(x, k)) = ck for all x Rn .
Since R(x, 0) can be expressed as the disjoint union
[
R(x, 0) = R(x + 2k `, k),
`Zn , 0`j 2k 1
this implies
X
= 0 (R(x, 0)) = 0 (R(x + 2k `, k)) = 2nk ck .
`Zn , 0`j 2k 1
Hence ck = 2nk = (R(x, k)). Here the last equation folows from the fact
that (0, 2k )n R(0, k) [0, 2k ]n and hence (R(x, k)) = (R(0, k)) = 2k
by part (iv) of Corollary 2.12. This proves Step 1.
2.2. THE LEBESGUE OUTER MEASURE 65
Remark 2.16. (i) Using Lemma 2.15 one can construct a continuous func-
tion f : R R and a Lebesgue measurable function g : R R such that
the composition g f is not Lebesgue measurable (see Example 6.24).
(ii) Let E R be the set constructed in the proof of Lemma 2.15. Then
the set E R R2 is not Lebesgue measurable. This follows from a similar
argument as in Lemma 2.15 using the sets ((E [n, n]) + q) [0, 1]. On the
other hand, the set E {0} R2 is Lebesgue measurable and has Lebesgue
measure zero. However, it is not a Borel set, because its pre-image in R
under the continuous map R R2 : x 7 (x, 0) is the original set E and
hence is not a Borel set.
Since ([0, 1)n ) has nonempty interior it follows that () > 0 and since
([0, 1)n ) is contained in the compact set ([0, 1]n ) it follows that () < .
Now define the map : B [0, ] by
((B))
(B) := for B B.
()
Then is a normalized translation invariant Borel measure. The -addi-
tivity follows directly from the -additivity of , the formula () = 0 is
obvious from the definition, that compact sets have finite measure follows
from the fact that (K) is compact if and only if K Rn is compact, the
translation invariance follows immediately from the translation invariance
of and the fact that (B + x) = (B) + (x) for all B B and all x Rn ,
and the normalization condition ([0, 1)n ) = 1 follows directly from the
definition of . Hence = by Theorem 2.1. This proves Step 1.
2.3. THE TRANSFORMATION FORMULA 69
Proof of Theorem 2.17. The proof has seven steps. The first four steps es-
tablish equation (2.14) for the characteristic functions of open sets, compact
sets, Borel sets, and Lebesgue measurable sets with compact closure in U .
Step 1. If W Rn is an open set with compact closure W U then
Z
m((W )) = |det(d)| dm.
W
Fix a constant > 0. Then there exists a constant > 0 that satisfies the
following two conditions.
(a) If a Rn , 0 < r < , R Rn satisfy Br (a) R B r (a) W then
m(R)
m((R)) |det(d(a))| m(R) < .
2 m(W )
Here the second equation follows from Step 1. This proves Step 2.
Step 3. If B B has compact closure B U then (B) B and
Z
m((B)) = |det(d)| dm.
B
That (B) is a Borel set follows from the fact that it is the pre-image of
the Borel set B under the continuous map 1 : V U (Theorem 1.20).
Abbreviate b := m((B)). Assume first that b < and fix a constant > 0.
Then it follows from Theorem 2.13 that there exists an open set W 0 Rn
with compact closure W 0 V such that (B) W 0 and m(W 0 ) < b + and
a compact set K 0 B such that (K 0 ) > b . Define K := 1 (K 0 ) and
W := 1 (W 0 ). Then K is compact, W is open, W U is compact, and
K B W, b < m((K)) m((W )) < b + .
Hence it follows from Step 1 and Step 2 that
Z Z Z
b< |det(d)| dm |det(d)| dm |det(d)| dm < b + .
K B W
R
Thus b < B
|det(d)| dm < b + for every > 0 and so
Z
|det(d)| dm = b = m((B)).
B
X` Z
= i m(Ai ) = s dm.
i=1 V
Here the second equation follows from Step 4. This proves Step 5.
Step 6. We prove (i).
By Theorem 1.26 there is a sequence of Lebesgue measurable step functions
si : V [0, ) such that 0 s1 s2 and f (x) = limi si (x) for
every x V . Choose an exhausing
S sequence of compact sets Ki V such
that Ki Ki+1 for all i and i Ki = V and replace si by si Ki . Then part (i)
follows from Step 5 and the Lebesgue Monotone Convergence Theorem 1.37.
Step 7. We prove (ii).
For E = U part (ii) follows from part (i) and the fact that (f ) = f .
If F AV then f := F |V is Lebesgue measurable, hence f = 1 (F ) |U
2.4. LEBESGUE EQUALS RIEMANN 75
These are finite sums and satisfy supk S(f, k) inf k S(f, k). The function
f : Rn R is called Riemann integrable if supk S(f, k) = inf k S(f, k).
The Riemann integral of a Riemann integrable function f : Rn R is the
real number
Z
R(f ) := f (x) dx := sup S(f, k) = inf S(f, k) = lim S(f, k). (2.29)
Rn kN kN k
76 CHAPTER 2. THE LEBESGUE MEASURE
Remark 2.21. The Riemann integral can also be defined by allowing for
arbitrary partitions of Rn into cuboids (see [19, Definition 2.3]) or in terms of
convergence of the so-called Riemann sums (see [21, Definition 7.1.2]). That
all three definitions agree is proved in [19, Satz 2.8] and [21, Theorem 7.1.8]).
Definition 2.22. A bounded set A Rn us called Jordan measurable if
its characteristic function A : Rn R is Riemann integrable. The Jordan
measure of a Jordan measurable set A Rn is the real number
J (A) := R(A )
Z
= A (x) dx (2.30)
Rn
n o
nk k n
= lim 2 # ` 2 Z R(`, k) A 6= .
k
Then
f (x) f (x) f (x)
for every x Rn . Moreover, |f k | and |f k | are bounded above by the Lebesgue
integrable function cA , where c := supxRn |f (x)| and A := [N, N ]n with
N N chosen such that supp(f ) [N, N ]n . Hence it follows from the
Lebesgue Dominated Convergence Theorem 1.45 that f and f are Lebesgue
integrable and
Z Z
f dm = lim f dm = lim S(f, k) = R(f )
Rn k Rn k k
Z Z
= lim S(f, k) = lim f k dm = f dm.
k k Rn Rn
provided that the limit exists. Here Br Rn denotes the ball of radius r
centered at the origin. There are many
R examples where the limit (2.32) exists
even though the Lebesgue integral Rn |f | dm is infinite and so the Lebesgue
78 CHAPTER 2. THE LEBESGUE MEASURE
2.5 Exercises
Exercise 2.25. Show that the Cantor set in R is a Jordan null set. Show
that Q [0, 1] is a Lebesgue null set but not a Jordan null set. Show that
A Rn is a Lebesgue null set if and only if (A) = 0. Find an open set
U R whose boundary has positive Lebesgue measure.
Exercise 2.26. Prove that every subset of a proper linear subspace of Rn
is Lebesgue measurable and has Lebesgue measure zero. Find a Jordan
measurable subset of Rn that is not a Borel set. Find a bounded Lebesgue
measurable subset of Rn with positive Lebesgue measure that is neither a
Borel set nor Jordan measurable.
Exercise 2.27. Let A, B Rn be Lebesgue null sets. Prove that their sum
A + B := {x + y | x A, y B} is a Lebesgue null set.
Exercise 2.28. Let (X, A, ) be a measure space and define the function
: 2X [0, ] by
(B) := inf (A) A A, B A . (2.33)
(i) Prove that is an outer measure and that A A().
(ii) Assume (X) < . Prove that the measure space (X, A(), |A() ) is
the completion of (X, A, ). Hint: Show that for every subset B X there
exists a set A A such that B A and (B) = (A).
(iii) Let X be a set and A ( X be a nonempty subset. Define
A := {, A, Ac , X}, () := (A) := 0, (Ac ) := (X) := .
Prove that (X, A, ) is a measure space. Given B X, prove that (B) = 0
whenever B A and (B) = whenever B 6 A. Prove that A() = 2X
and that the completion of (X, A, ) is the measure space (X, A , ) with
A = {B X | B A or Ac B} and = |A . (Thus the hypothesis
(X) < cannot be removed in part (ii).)
2.5. EXERCISES 79
Borel Measures
81
82 CHAPTER 3. BOREL MEASURES
for every Borel set B B, and is called regular if it is both outer and inner
regular. A Radon measure is an inner regular Borel measure.
Example 3.4. Let (X, U) be any locally compact Hausdorff space and fix
a point x0 X. Then the Dirac measure = x0 at x0 is a regular Borel
measure (Example 1.31).
The next example occupies three pages and illustrates the subtlety of
the subject (see also Exercise 18 in Rudin [17, page 59]). It constructs a
compact Hausdorff space (X, U) and a Borel measure on X that is not a
Radon measure. More precisely, there is a point X such that the open
set U := X \ {} is not -compact and satisfies (U ) = 1 and (K) = 0
for every compact subset K U . This example is a kind of refinement of
Example 3.5. It is due to Dieudonne.
3.1. REGULAR BOREL MEASURES 83
(iv) We prove that (X, U) is compact. Let {Ui }iI be an open cover of X.
We prove by induction that there exist finite sequences x1 , . . . , x` X and
i1 , . . . , i` I such that xk Uik \ Uik1 and Sxk Ui1 Uik1 for k 2,
and X = `j=1 Uij . Define x1 := and choose i1 I such that Ui1 .
S
If Ui1 = X the assertion holds with ` = 1. Now suppose, by induction,
that x1 , . . . , xk and i1 , . . . , ik have been constructed such that xj Uij for
j = 1, . . . , k and Sxk Ui1 Uik1 . If Ui1 Uik = X we are done with
` = k. Otherwise Ck := X \ Ui1 Uik is a nonempty compact set and we
define xk+1 := max(Ck ) by part (iii). Then xk+1 Ck and Ck Sxk+1 = .
Hence Sxk+1 Ui1 Uik . Choose ik+1 I such that xk+1 Uik+1 .
This completes the induction argument. The induction must stop because
xk+1 xk for all k and every strictly decreasing sequence in X is finite by
the well ordering axiom (d). This shows that (X, U) is compact.
(v) Let Ki X, i N, be a sequence of uncountable compact sets. We
prove that the compact set \
K := Ki
iN
K \ {} =
6 . (3.3)
The next lemma and theorem are included here in preparation for the
Riesz Representation Theorem 3.15. They explain the relation between the
various regularity properties of Borel measures.
Lemma 3.7. Let : B [0, ] be an outer regular Borel measure that is
inner regular on open sets, i.e.
(U ) = sup (K) K U and K is compact (3.4)
for every open set U X. Then the following holds.
(i) Every Borel set B X with (B) < satisfies (3.2).
(ii) If X is -compact then is regular.
Proof. We prove (i). Fix a Borel set B X with (B) < and a constant
> 0. Since is outer regular, there exists an open set U X such that
B U, (U ) < (B) + .
2
Thus U \ B is a Borel set and (U \ B) = (U ) (B) < /2. Use the outer
regularity of again to obtain an open set V X such that
U \ B V, (V ) < .
2
Now it follows from (3.4) that there exists a compact set K X such that
K U, (K) > (U ) .
2
Define C := K \ V . Since X is a Hausdorff space, K is closed, hence C is a
closed subset of K, and hence C is compact (see Lemma A.2). Moreover,
C U \ V B, B \ C (B \ K) V (U \ K) V,
and hence (B \ C) (U \ K) + (V ) < . This proves (i).
We prove (ii). Choose a sequenceSof compact sets Ki X such that
Ki Ki+1 for all i N and X = i=1 Ki . Fix a Borel set B B. If
(B) < then B satisfies (3.2) by (i). Hence assume (B) = . Then
it follows from part (iv) of Theorem 1.28 that limi (B Ki ) = . For
each integer n N choose in N such that
(B Kin ) > n.
Since (B Kin ) (Kin ) < it follows from (i) that (3.2) holds with B
replaced by B Kin . Hence there exists a compact set Cn B Kin such
that (Cn ) > n. This proves (ii) and Lemma 3.7.
3.1. REGULAR BOREL MEASURES 87
(K) 1 (K), 1 (U ) (U )
f |K 1, supp(f ) U, 0 f 1.
and likewise Z Z
1 (K) f d1 = f d (U ).
X X
for
R every compactly
R supported
R continuous function f : X [0, ). Hence
X
f d = X
f d 0 = X
f d 1 for every compactly supported continuous
function f : X [0, ) and hence also for every compactly supported
continuous function f : X R. This proves Step 4 and Theorem 3.8.
3.1. REGULAR BOREL MEASURES 91
Example 3.9. Let (X, U) be the compact Hausdorff space in Example 3.6
and let : B [0, ] be Dieudonnes measure.
(i) Take 1 := and define the function 0 : B [0, ] by (3.5). Then
0 (X) = 1, 0 ({}) = 0, 0 (X \ {}) = 0,
and so 0 is not a measure. Hence the assumptions on 1 cannot be removed
in part (i) of Theorem 3.8.
(ii) Take 1 := to be the Dirac measure at the point X. This is a
regular Borel measure and so the measure 0 in (3.5) agrees with 1 . It is an
easy exercise to show that the integral of a continuous function f : X R
with respect to the Dieudonne measure is given by
Z Z Z
f d = f () = f d0 = f d1 .
X X X
Here the subscript e stands for endlich and indicates that the elements
of Ae have finite measure. We prove in seven steps that A is a -algebra
containing B, that := |A : A [0, ] is a measure, and that (X, A, ) is
a complete measure space. That is outer regular and is inner regular on
open sets follows immediately from conditions (c) and (d) in Definition 3.11.
3.2. BOREL OUTER MEASURES 93
and hence
X
(Ei ) = (E). (3.13)
i=1
Now it follows from (3.12) and (3.13) that
n
X
X
(E) (K1 Kn ) (Ei ) = (E) (Ei )
i=1 i=n+1
Ki , Ui , Ui \ Ki Ae
for i = 0, 1. Define
K := K0 \ U1 E0 \ E1 . (3.15)
Then K is a compact set and
Here the last inequality follows from the definition of K in (3.15) and the
inequalities in (3.14). Since > 0 was chosen arbitrarily it follows that
E0 E1 = (E0 \ E1 ) E1 Ae , E0 E1 = E0 \ (E0 \ E1 ) Ae .
Step 3. A is a -algebra.
First, X A because K Ae for every compact set K X.
Second, assume A A and let K X be a compact set. Then by
definition A K Ae . Moreover K Ae and hence, by Step 2,
Ac K = K \ (A K) Ae .
Ai K Ae
Bi := Ai K Ae
Ei := Bi \ (B1 Bi1 ) Ae
Proof. The proof has nine steps. Step 1 defines a function : 2X [0, ],
Step 2 shows that it is an outer measure, and Steps 3, 4, and 5 show that it
satisfies the axioms of Definition 3.11. Step 6 defines 1 and Step 7 shows
that 1 = . Step 8 defines 0 and Step 9 proves uniqueness.
Step 1. Define the function U : U [0, ] by
U (U ) := sup (f ) f Cc (X), 0 f 1, supp(f ) U (3.17)
0 f 1, K := supp(f ) U.
0 f 1, supp(f ) U, f |K 1.
Hence
a (f ) U (U ).
This shows that a U (U ) for every open set U X containing K. Take
the infimum over all open sets containing K and use the definition of in
equation (3.18) to obtain a (K).
3.3. THE RIESZ REPRESENTATION THEOREM 101
U := {x X | f (x) > } .
(K) U (U ).
(K) (f ).
Hence (K0 ) + (K1 ) < (K0 + K1 ) + for every > 0 and therefore
(K0 ) + (K1 ) (K0 + K1 ). This proves Step 5.
102 CHAPTER 3. BOREL MEASURES
such that
yi yi1 < , i = 1, . . . , n.
For i = 1, . . . , n define
Ei := x K yi1 < f (x) yi .
Then Ei is the intersection of the open set f 1 ((yi1 , )) with the closed set
f 1 ((, yi ]) and hence is a Borel set. Moreover Ei Ej = for i 6= j and
n
[
K= Ei .
i=1
3.3. THE RIESZ REPRESENTATION THEOREM 103
for all i. (For each i, choose first an open set that satisfies the first two
conditions in (3.24) and then intersect it with the open set f 1 ((, yi +)).)
By Theorem A.4 there exist functions 1 , . . . , n Cc (X) such that
n
X n
X
i 0, supp(i ) Ui , i 1, i |K 1. (3.25)
i=1 i=1
This shows that is inner regular on open sets and hence it follows from (i)
that is regular. This proves Theorem 3.18.
Example 3.9 shows that the assumption that every open set is -compact
cannot be removed in part (ii) of Theorem 3.18 even if X is compact. Note
also that Theorem 3.18 provides another proof of regularity for the Lebesgue
measure, which was established in Theorem 2.13.
Corollary 3.19. Let X be a locally compact Hausdorff space such that every
open subset of X is -compact. Then for every positive linear functional
: Cc (X) R there exists a unique Borel measure such that = .
Proof. This follows from Theorem 3.15 and part (ii) of Theorem 3.18.
106 CHAPTER 3. BOREL MEASURES
Remark 3.20. Let X be a compact Hausdorff space and let C(X) = Cc (X)
be the space of continuous real valued functions on X. From a functional
analytic viewpoint it is interesting to understand the dual space of C(X),
i.e. the space of all bounded linear functionals on C(X) (Definition 4.23).
Exercise 5.35 below shows that every bounded linear functional on C(X) is
the difference of two positive linear functionals. If every open subset of X
is -compact it then follows from Corollary 3.19 that every bounded linear
functional on C(X) can be represented uniquely by a signed Borel measure.
(See Definition 5.10 in Section 5.3 below.)
Proof. We prove (i). Let V be a countable basis of the topology and let
U X be an open set. Denote by V(U ) the collection of all sets V V such
that V U and V is compact. Let x U . By Lemma A.3 there is an open
set W X with compact closure such that x W W U . Since V is a
basis of the topology, there is an element V V such that x V W . Hence
V is a closed subset of the compact set W and so is compact by Lemma A.2.
Thus V V(U ) and x V . This shows that
[
U= V.
V V(U )
3.4 Exercises
Exercise 3.23. This exercise shows that the measures 0 , 1 in Theorem 3.15
need not agree. Let (X, d) be the metric space given by X := R2 and
0, if x1 = x2 ,
d((x1 , y1 ), (x2 , y2 )) := |y1 y2 | +
1, if x1 6= x2 .
Let B 2X be the Borel -algebra of (X, d).
(i) Show that (X, d) is locally compact.
(ii) Show that for every compactly supported continuous function f : X R
there exists a finite set Sf R such that supp(f ) Sf R.
(iii) Define the positive linear functional : Cc (X) R by
XZ
(f ) := f (x, y) dy.
xSf
(Here the integrals on the right are understood as the Riemann integrals or,
equivalently by Theorem 2.24, as the Lebesgue integrals.) Let : B [0, ]
be a Borel measure such that
Z
f d = (f ) for all f Cc (X).
X
Exercise 3.25. Let (X, UX ) and (Y, UY ) be locally compact Hausdorff spaces
and denote their Borel -algebras by BX 2X and BY 2Y . Let : X Y
be a continuous map and let X : BX [0, ] be a measure.
(i) Prove that BY BX (See Exercise 1.69).
(ii) If X is inner regular show that X |BY is inner regular.
(iii) Find an example where X is outer regular and X |BY is not outer
regular. Hint: Consider the inclusion of N into its one-point compactification
and use Exercise 3.24. (In this example X is a Borel measure, however, X
is not a Borel measure.)
Exercise 3.26. Let (X, d) be a metric space. Prove that (X, d) is perfectly
normal, i.e. if F0 , F1 X are disjoint closed subsets then there is a continuous
function f : X [0, 1] such that F0 = f 1 (0) and F1 = f 1 (1). Compare
this with Urysohns Lemma A.1. Hint: An explicit formula for f is given by
d(x, F0 )
f (x) := ,
d(x, F0 ) + d(x, F1 )
where
d(x, F ) := inf d(x, y)
yF
for x X and F X.
Exercise 3.27. Recall that the Sorgenfrey line is the topological space
(R, U), where U 2R is the smallest topology that contains all half open
intervals [a, b) with a < b. Prove that the Borel -algebra of (R, U) agrees
with the Borel -algebra of the standard topology on R.
Exercise 3.28. Recall from Example 3.22 that the Double Arrow Space is
X := [0, 1] {0, 1}
with the topology induced by the lexicographic ordering. Prove that B X
is a Borel set for this topology if and only if there is a Borel set E [0, 1]
and two countable sets F, G X such that
B = ((E {0, 1}) F ) \ G. (3.28)
Hint 1: Show that the projection f : X [0, 1] onto the first factor is
continuous with respect to the standard topology on the unit interval.
Hint 2: Denote by B 2X the set of all sets of the form (3.28) with E [0, 1]
a Borel set and F, G X countable. Prove that B is a -algebra.
110 CHAPTER 3. BOREL MEASURES
Let
Ba 2X
be the smallest -algebra that contains Ka . It is contained in the Borel -
algebra B 2X and is called the Baire -algebra of (X, U). The elements of
Ba are called Baire sets. A function f : X R is called Baire measurable
if f 1 (U ) Ba for every open set U R. A Baire measure is a measure
: Ba [0, ] such that (K) < for all K Ka .
(i) Let f : X R be a continuous function with compact support. Prove
that f 1 (c) Ka for every nonzero real number c.
(ii) Prove that Ba is the smallest -algebra such that every continuous func-
tion f : X R with compact support is Ba -measurable.
(iii) If every open subset of X is -compact prove that Ba = B. Hint:
Show first that every compact set belongs to Ka and then that every open
set belongs to Ba .
Exercise 3.30. (i) Let X be an uncountable set and let U := 2X be the
discrete topology. Prove that B X is a Baire set if and only if B is
countable or has a countable complement. Define : Ba [0, 1] by
0, if B is countable,
(B) :=
1, if B c is countable.
R
Show that X f d = 0 for every f Cc (X). Thus positive linear functionals
: Cc (X) R need not be uniquely represented by Baire measures.
(ii) Let X be the compact Hausdorff space of Example 3.6. Prove that the
Baire sets in X are the countable subsets of X \ {} and their complements.
(iii) Let X be the StoneCech compactification of N in Example 4.60 below.
Prove that the Baire sets in X are the subsets of N and their complements.
(iv) Let X = R2 be the locally compact Hausdorff space in Exercise 3.23
(with a nonstandard topology). Show that B X is a Baire set if and only if
the set Bx := {y R | (x, y) B} is a Borel set in R for every x R and one
of the sets S0 := {x R | Bx 6= } and S1 := {x R | Bx 6= R} is countable.
3.4. EXERCISES 111
Exercise 3.31. Let (X, U) be a locally compact Hausdorff space and let
Ba B 2X
be the Baire and Borel -algebras. Let F (X) denote the real vector space of
all functions f : X R. For F F (X) consider the following conditions.
(a) Cc (X) F.
(b) If fi F is a sequence converging pointwise to f F (X) then f F.
Let Fa F (X) be the intersection of all subsets F F (X) that satisfy
conditions (a) and (b). Prove the following.
(i) Fa satisfies (a) and (b).
(ii) Every element of Fa is Baire measurable. Hint: The set of Baire mea-
surable functions on X satisfies (a) and (b).
(iii) If f Fa and g Cc (X) then f + g Fa . Hint: Let g Cc (X). Then
the set Fa g satisfy (a) and (b) and hence contains Fa .
(iv) If f, g Fa then f + g Fa . Hint: Let g Fa . Then the set Fa g
satisfy (a) and (b) and hence contains Fa .
(v) If f Fa and c R then cf Fa . Hint: Fix a real number c 6= 0.
Then the set c1 Fa satisfy (a) and (b) and hence contains Fa .
(vi) If f Fa and g Cc (X) then f g Fa . Hint: Fix a real number c such
that c + g(x) > 0 for all x R. Then the set (c + g)1 Fa satisfy (a) and (b)
and hence contains Fa . Now use (iv) and (v).
(vii) If A X such that A Fa and f Fa then f A Fa . Hint: The
set (1 + A )1 Fa satisfy (a) and (b) and hence contains Fa .
(viii) The set
A := A X | A Fa or X\A Fa
is a -algebra. Hint: If A , B FA then AB = A + B A B Fa . If
X\A , X\B FA then X\(AB) = X\A X\B Fa . If A , X\B FA then
X\(AB) = (X\A)(X\B) = X\B A X\B Fa . Thus
A, B A = A B A.
(ix) A = Ba . Hint: Let K Ka . Use Urysohns Lemma A.1 to construct a
sequence gi Cc (X) that converges pointwise to K .
(x) For every f Fa there exists a sequence
S of compact sets Ki Ka such
that Ki Ki+1 for all i and supp(f ) iN Ki . Hint: The set of functions
f : X R with this property satisfies conditions (a) and (b).
112 CHAPTER 3. BOREL MEASURES
Exercise 3.32. Show that, for every locally compact Hausdorff space X and
any two Borel measures 0 , 1 as in Theorem 3.8, there is a Baire set N X
such that 0 (N ) = 0 and 0 (B) = 1 (B) for every Baire set B X \ N .
Hint 1: Show first that
0 (B) = sup 0 (K) K Ka , K B , (3.29)
where Ka is as in Exercise 3.29. To see this, prove that the right hand side
of equation (3.29) defines a Borel measure on X that is inner regular on
open sets and satisfies 0 and = 0 .
Hint 2: Suppose there exists a Baire set N X such that 0 (N ) < 1 (N ).
Show that 1 (N ) = and that N can be chosen such that 0 (N ) = 0. Next
show that X\N Fa , where Fa is as in Exercise 3.31, and deduce that X \N
is contained in a countable union of compact sets.
Example 3.33. Let X be the StoneCech compactification of N discussed
in Example 4.60 below and denote by Ba B 2X the Baire and Borel -
algebras. Thus B X is a Baire set if and only if either B N or X \B N.
(See part (iii) of Exercise 3.30.) For a Borel set B X define
X1 B U X,
0 (B) := , 1 (B) := inf 0 (U )
.
n U is open
nB
Lp Spaces
4.1 H
older and Minkowski
Assume throughout that (X, A, ) is a measure space and that p, q are real
numbers such that
1 1
+ = 1, 1 < p < , 1 < q < . (4.1)
p q
Then any two nonnegative real numbers a and b satisfy Youngs inequality
1 1
ab ap + bq (4.2)
p q
and equality holds in (4.2) if and only if ap = bq . (Exercise: Prove this by
examining the critical points of the function (0, ) R : x 7 p1 xp xb.)
113
114 CHAPTER 4. LP SPACES
Proof. Define
Z 1/p Z 1/q
p q
A := f d , B := g d .
X X
This proves the Holder inequality. To prove the Minkowski inequality, define
Z 1/p Z 1/p Z 1/p
p p p
a := f d , b := g d , c := (f + g) d .
X X X
It follows from the Minkowski inequality (4.4) that the sum of two Lp -
functions is again an Lp -function and hence Lp () is a real vector space.
Moreover, the function
Lp () [0, ) : f 7 kf kp
satisfies the triangle inequality
kf + gkp kf kp + kgkp
for all f, g Lp () by (4.4) and
kf kp = || kf kp
for all R and f Lp () by definition. However, in general kkp is not
a norm on Lp () because kf kp = 0 if and only if f = 0 almost everywhere
by Lemma 1.49. We can turn the space Lp () into a normed vector space
by identifying two functions f, g Lp () whenever they agree almost every-
where. Thus we introduce the equivalence relation
f g f =g -almost everywhere. (4.7)
Denote the equivalence class of a function f Lp () under this equivalence
relation by [f ] and the quotient space by
Lp () := Lp ()/ . (4.8)
This is again a real vector space. (For p = 1 see Definition 1.51.) The
Lp -norm in (4.5) depends only on the equivalence class of f and so the map
Lp () [0, ) : [f ] 7 kf kp
is well defined. It is a norm on Lp () by Lemma 1.49. Thus we have defined
the normed vector space Lp () for 1 p < . It is sometimes convenient to
abuse notation and write f Lp () instead of [f ] Lp (), always bearing
in mind that then f denotes an equivalence class of p-integrable functions. If
(X, A , ) denotes the completion of (X, A, ) it follows as in Corollary 1.56
that Lp () is naturally isomorphic to Lp ( ).
Remark 4.4. Assume 1 < p < and let f, g Lp () such that
kf + gkp = kf kp + kgkp , kf kp 6= 0.
Then it follows from part (ii) of Exercise 4.2 that there exists a real number
0 such that g = f almost everywhere.
4.2. THE BANACH SPACE LP () 117
for all m, n N. To see this, use Lemma 4.8 to find null sets En , Em,n A
such that supX\En |fn | = kfn k and supX\Em,n |fm fn | = kfm fn k for
all m, n N. Then the union E of the sets En and Em,n is measurable and
satisfies (4.12). Since [fn ] is a Cauchy sequence in L () we have
Since |fm (x) fn (x)| n for all m n and all x X \ E it follows that
(fn (x))nN is Cauchy sequence in R and hence converges for every x X \ E.
Define f : X R by f (x) := limn fn (x) for x X \ E and by f (x) := 0
for x E. Then
4.3 Separability
Definition 4.11. Let X be a topological space. A subset S X is called
dense (in X) if its closure is equal to X or, equivalently, U S 6= for every
nonempty open set U X. A subset S X of a metric space is dense if and
only if every element of X is the limit of a sequence in S. The topological
space X is called separable if it admits a countable dense subset.
Every second countable topological space is separable and first countable
(see Lemma 3.21). The Sorgenfrey line is separable and first countable but is
not second countable (see Example 3.22). A metric space is separable if and
only if it is second countable. (If S is a countable dense subset then the balls
with rational radii centered at the points of S form a basis of the topology.)
The Euclidean space X = Rn with its standard topology is separable (Qn is
a countable dense subset) and hence is second countable. The next lemma
gives a criterion for a linear subspace to be dense in Lp ().
Lemma 4.12. Let (X, A, ) be a measure space and let 1 p < . Let X
be a linear subspace of Lp () such that [A ] X for every measurable set
A A with (A) < . Then X is dense in Lp ().
Proof. Let Y denote the closure of X in Lp (). Then Y is a closed linear
subspace of Lp (). We prove in three steps that Y = Lp ().
Step 1. If s Lp () is a measurable step function then [s] Y .
Write s = R`i=1 i Ai where Ri R \ {0} and Ai = s1 (i ) A. Then
P
|i |p (Ai ) = X |i Ai |p d X |s|p d < and hence (Ai ) < for all i.
This implies [Ai ] Y for all i. Since Y is a linear subspace of Lp () it
follows that [s] Y . This proves Step 1.
4.3. SEPARABILITY 121
Lp (1 ) Lp (0 ) : [f ]1 7 [f ]0 (4.14)
Thus the map (4.14) is injective and has a closed image. To prove (4.15),
define E := {x X | |f (x)| > } for > 0. Then 1 (E ) < and hence
R 1 and 0 agree
R onp all Borel subsets of E by Lemma 3.7. This implies
p
E
|f | d0 = E |f | d1 , and (4.15) follows by taking the limit 0.
We prove that the map (4.14) is surjective. Denote its image by X . This
is a closed linear subspace of Lp (0 ), by what we have just proved. Let B B
such that 0 (B) < . By (3.5) there is a sequence of compact sets Ki B
i
such thatS Ki Ki+1 and 1 (Ki ) = 0 (Ki ) > 0 (B) 2 for all i. Define
A := iN Ki B. Then 1 (A) = 0 (A) = limi 0 (Ki ) = 0 (B). This
implies A Lp (1 ) and [B ]0 = [A ]0 X . By Lemma 4.12, it follows
that X = Lp (0 ) and this proves Lemma 4.17.
Proof of Theorem 4.13. Let V U be a countable basis for the topology. As-
sume without of generality that V is compact for all V V. (If W U is any
countable basis for the topology then the set V := V W | V is compact
is also a countable basis for the topology by Lemma A.3.) Choose a bijection
N V : i 7 Vi and let I := {I N P| #I < } be the set of finite subsets
of N. Then the map I N : I 7S iI 2 i1
is a bijection, so the set I is
countable. For I I define VI := iI Vi . Define the set V Lp () by
( `
)
X
V := s = j VIj ` N and j Q, Ij I for j = 1, . . . , ` .
j=1
n o
C := [f ] Lp (1 ) | > 0 g Cc (X) such that kf gkp < .
Remark 4.18. The reader may wonder whether Theorem 4.15 continues to
hold for all Borel measures : B [0, ] that are inner regular on open sets.
To answer this question one can try to proceed as follows. Let 0 , 1 be the
Borel measures on X in Theorem 3.15 that satisfy 0 = 1 = . Then
0 is a Radon measure, 1 is outer regular and is inner regular on open sets,
and 0 (B) (B) 1 (B) for all B B. Thus Lp (1 ) Lp () Lp (0 )
and one can consider the maps
Lp (1 ) Lp () Lp (0 ).
Now define
x0
y := .
kx0 k2
Fix an element x H and define := (x). Then (xx0 ) = (x) = 0.
Hence it follows from (4.26) that
0 = hx0 , x x0 i = hx0 , xi kx0 k2 .
This implies
hx0 , xi
hy, xi = = = (x).
kx0 k2
Thus y satisfies (4.24).
We prove (4.25). Assume y H satisfies (4.24). If y = 0 then = 0 and
so kyk = 0 = kk. Hence assume y 6= 0. Then
kyk2 (y) |(x)|
kyk = = sup = kk .
kyk kyk 06=xH kxk
The next theorem is the first step towards understanding the dual space
of Lp () and is a fairly easy consequence of the Holder inequality. It asserts
that for 1/p + 1/q = 1 every element of Lq () determines a bounded lin-
ear functional on Lp () and that the resulting map Lq () Lp () is an
isometric embedding (for p = 1 under the semi-finite hypothesis). The key
question is then whether every bounded linear functional on Lp () is of that
form. That this is indeed the case for 1 < p < (and for p = 1 under the
localizable hypothesis) is the content of Theorem 4.35 below. This is a much
deeper theorem whose proof for p 6= 2 requires the RadonNikod ym theorem
and will be carried out in Chapter 5.
Theorem 4.33. Let (X, A, ) be a measure space and fix constants
1 1
1 p , 1 q , + = 1. (4.30)
p q
Then the following holds.
(i) Let g Lq (). Then the formula
Z
g ([f ] ) := f g d for f Lp () (4.31)
X
Lq () Lp () : [g] 7 g . (4.33)
khkq < it follows that kgkq = khkq c and this proves part (ii).
We prove (iii). Assume (X, A, ) is semi-finite and 1 < q < . Suppose,
by contradiction, that kgkq > c. We will prove that there exists a measurable
function h : X [0, ) such that
Moreover E :=
S
i=1 Ei = {x A | f (x) > 0} is not a null set by (4.36).
Hence one of the sets Ei has positive measure. Thus (X, A, ) is semi-finite.
This proves Step 4 and Theorem 4.33.
The next theorem asserts that, for 1 < p < , every bounded linear
functional on Lp () has the form (4.31) for some g Lq (). For p 6= 2
this is a much deeper result than Corollary 4.28. The proof requires the
RadonNikod ym Theorem and will be deferred to the next chapter.
that
P
Choose N N such n=N |yn |
=: < 1 and define x = (xn )nN c by
xn := 0 for n < N and xn := 1 for n N . Then y (x) < 1 = (x)
and hence y 6= . This shows that does not belong to the image of the
isometric inclusion `1 , (` ) .
Exercise 4.37. Let 0 : c R be the functional in Example 4.36 and
denote its kernel by c0 := ker 0 . Thus c0 is the set of all sequences of real
numbers that converge to zero, i.e.
c0 = x = (xn )nN ` lim xn = 0 .
n
We close this section with two results that will be needed in the proof
of Theorem 4.35. The first asserts that every bounded linear functional on
Lp () can be written as the difference of two positive bounded linear function-
als (Theorem 4.39). The second asserts that every positive bounded linear
functional on Lp () is supported on a -finite subset of X (Theorem 4.40).
When : Lp () R is a bounded linear functional it will be convenient to
abuse notation and write (f ) := ([f ] ) for f Lp ().
Definition 4.38. Let (X, A, ) be a measure space and let 1 p < . A
bounded linear functional : Lp () R is called positive if
f 0 = (f ) 0
= + , k+ k + k k = kk. (4.39)
Take the supremum over all E A with E A and (E) < to obtain
X
+ (A) + (Ai ).
i=1
To prove the converse inequality, assume first that + (Ai ) = for some i;
since Ai A this implies + (A) = = i=1 + (Ai ). Hence it suffices to
P
assume + (Ai ) < for all i. Fix a constant > 0 and choose a sequence of
measurable sets Ei A such that Ei Ai and (Ei ) > + (Ai ) 2i for
all i. Since E1 En A it follows from the definition of + that
n
X n
X
+
(A) (E1 En ) = (Ei ) > + (Ai ) .
i=1 i=1
P
Take the limit n to obtain + (A) i=1 + (Ai ) for all > 0, so
X
+
(A) + (Ai )
i=1
In particular, Lp () L1 (+ ) L1 ( ).
Assume first that f = s : X [0, ) is a measurable step function in
Lp (). Then there exist real numbers i > 0 and measurable sets Ai A
4.5. THE DUAL SPACE OF LP () 139
`
X
i i = .
i=1
2
Then
Z Z `
X
+
i + (Ai ) + (Ai )
s d + s d =
X X i=1
X`
i (Ei+ ) (Ei ) + 2i
i=1
`
!
X
= i Ei+ Ei +
i=1
`
X
c
i Ei
+
+
Ei
i=1 p
`
!1/p
X
= c ip (Ei+ \ Ei ) + (Ei \ Ei+ ) +
i=1
`
!1/p
X
c ip (Ai ) +
i=1
= c kskp + .
Now let > 0 and choose E A such that E A and (E ) > + (A) .
Since (A\E ) (A) this implies
`
X `
X Z Z
+ +
s d .
(s) = i (Ai ) = i (Ai ) (Ai ) = s d
i=1 i=1 X X
This proves (4.43) for p-integrable step functions. Now let f Lp () and
assume f 0. By Theorem 1.26 there is a sequence of measurable step
functions 0 s1 s2 that converges pointwise to f . Then (f sn )p
converges pointwise to zero and is bounded above by f p L1 (). Hence
limn kf sn kp = 0 by the Lebesgue Dominated Convergence Theorem
and hence limn (sn ) = (f ). Moreover, X f d c kf kp < by
R
Theorem 4.40. Let (X, A, ) be a measure space, let 1 < p < , and let
: Lp () R be a positive bounded linear functional. Define
(A) := sup {(E ) | E A, E A, (E) < } (4.44)
for A A. Then the map : A [0, ] is a measure, Lp () L1 (), and
Z
(f ) = f d for all f Lp (). (4.45)
X
Moreover, there are measurable sets N A and Xn A for n N such that
[
X \N = Xn , (N ) = 0, (Xn ) < , Xn Xn+1 (4.46)
n=1
for all n N.
Proof. That is a measure satisfying Lp () L1 () and (4.45) follows from
Theorem 4.39 and the fact that + = and = 0 because is positive.
Now define c := kk. We prove in three steps that there exist measurable
sets N A and Xn A for n N satisfying (4.46).
Step 1. For every > 0 there exists a measurable set A A and a measur-
able function f : X [0, ) such that
f |X\A = 0, inf f > 0, kf kp = 1, (f ) > c . (4.47)
A
4.6 Exercises
Many of the exercises in this section are taken from Rudin [17, pages 7175].
Exercise 4.41. Let (X, A, ) be a measure space and let
f = (f1 , . . . , fn ) : X Rn
R
be a measurable function such that X |fi | d < for i = 1, . . . , n. Define
Z Z Z
f d := f1 d, . . . , fn d Rn .
X X X
Hint: Prove the inequality first for vector valued integrable step functions
s : X Rn . Show that for all R> 0 there is a vector valued
R integrable step
function s : X Rn such that k X (f s) dk < and X kf sk d < .
Exercise 4.42. Let (X, A, ) be a measure space such that (X) = 1. Let
f L1 () and let : R R be convex. Prove Jensens inequality
Z Z
f d ( f ) d. (4.50)
X X
Exercise 4.46. For each of the following three conditions find an example
of measure space (X, A, ) that satisfies it for all p, q [1, ].
(a) If p < q then Lp () ( Lq ().
(b) If p < q then Lq () ( Lp ().
(c) If p 6= q then Lp () 6 Lq () and Lq () 6 Lq ().
4.6. EXERCISES 145
Exercise 4.47. Let (X, U) be a locally compact Hausdorff space and define
f is continuous and
C0 (X) := f : X R > 0 K X such that
K is compact and supX\K |f | <
Prove that X is a Banach space with respect to the sup-norm. Prove that
Cc (X) is dense in C0 (X).
Exercise 4.48. Let (X, A, ) be a measure space such that (X) = 1 and
let f, g : X [0, ] be measurable functions such that f g 1. Prove that
kf k1 kgk1 1.
Exercise 4.49. Let (X, A, ) be a measure space such that (X) = 1 and
let f : X [0, ] be a measurable function. Prove that
q Z p
2
1 + kf k1 1 + f 2 d 1 + kf k1 . (4.55)
X
and Z
f p d (E)1p for 0 < p < 1. (4.57)
E
for all > 0. (On page 47 this is called convergence in probability.) Assume
(X) < and prove the following.
(i) If fn converges to f almost everywhere then fn converges to f in measure.
Hint: See page 47.
(ii) If fn converges to f in measure then a subsequence of fn converges to f
almost everywhere.
(iii) If 1 p and fn , f Lp () satisfy limn kfn f kp = 0 then fn
converges to f in measure.
148 CHAPTER 4. LP SPACES
Then (Ef ) = 0.
(vii) Define g : X R by
0, if x E0 ,
g(x) := (4.62)
a, if x X s for all s > a and x
/ X r for all r < a.
Then g is well defined and measurable and g = f on Af \ Ef for all f F .
Example 4.60. This example is closely related to Exercise 3.24, however,
it requires a considerable knowledge of Functional Analysis and the details
go much beyond the scope of the present manuscript. It introduces the
StoneCech compactification X of the natural numbers. This is a com-
pact Hausdorff space containing N and satisfying the universality property
that every continuous map from N to another compact Hausdorff space Y
extends uniquely to a continuous map from X to Y . The space C(X) of
continuous functions on X can be naturally identified with the space ` .
Hence the space of positive bounded linear functionals on ` is isomorphic
to the space of Radon measures on X by Theorem 3.15. Thus the Stone
The weak- topology U 2X is the smallest topology such that the map
f : X R, f () := (),
is continuous for each ` . The topological space (X, U) is a separable
compact Hausdorff space, called the StoneCech compactification of N.
It is not second countable and one can show that the complement of a point
in X that is not equal to one of the n is not -compact. The only continuous
functions on X are those of the form f , so the map ` C(X) : 7 f is
a Banach space isometry. (Verify that kf k := supX |f ()| = kk for all
` .) Thus the dual space of ` can be understood in terms of the Borel
measures on X.
By Theorem 3.18 every Radon measure on X is regular. However, the
Borel -algebra B 2X does carry -finite measures : B [0, ] that are
inner regular but not outer regular (and must necessarily satisfy (X) = ).
Here is an example pointed out to me by Theo Buehler. Define
X 1
(B) :=
nN
n
n B
for every Borel set B X. This measure is -finite and inner regular but
is not outer regular. (The set U := {n | n N} is open, its complement
K := X \ U is compact and has measure zero, and every open set containing
K misses only a finite subset of U and hence has infinite measure.) Now let
X0 X be the union of all open sets in X with finite measure. Then X0
is not -compact and the restriction of to the Borel -algebra of X0 is a
Radon measure but is not outer regular.
Chapter 5
The RadonNikod
ym Theorem
151
152 THEOREM
CHAPTER 5. THE RADONNIKODYM
An := {E A | E An } , n := |An .
En := {x An | f (x) 6= g(x)}
Define : L2 ( + ) R by
Z
(f ) := f d.
X
for f L2 ( + ). (Here we abuse notation and use the same letter f for a
function in L2 ( + ) and its equivalence class in L2 ( + ).) Then
Z Z
|(f )| |f | d |f | d( + ) c kf kL2 (+)
X X
= f d( + ) f d (5.5)
ZX X
= f d
X
for all f L2 ( + ).
5.1. ABSOLUTELY CONTINUOUS MEASURES 155
By Theorem 1.40 this implies that equation (5.5) continues to hold for every
measurable function f : X [0, ), whether or not it belongs to L2 ( + ).
Now define the measurable function h : X [0, ) by
g(x)
h(x) := for x X.
1 g(x)
By equation (5.4) with f = A and equation (5.5) with f = A h it satisfies
Z Z
(A) = A d = A g d( + )
ZX X Z
= A h(1 g) d( + ) = A h d
ZX X
= h d
A
Example 5.5. Let X be a one element set and let A := 2X . Define the
measure : 2X [0, ] by () := 0 and (X) := .
(i) Choose () := 0 and (X) := 1. Then R but there does not exist a
(measurable) function f : X [0, ] such that X f d = (X). Thus the
hypothesis that (X, A, ) is -finite cannot be removed in Theorem 5.4.
R
(ii) Choose := . Then (A) = A f d for every nonzero function
f : X [0, ). Thus the hypothesis that (X, A, ) is -finite cannot be
removed in Step 1 in the proof of Theorem 5.4.
Then and and are not -finite. There R does not exist any
measurable function f : X [0, ] such that (X) = R X f d. Nor is there
any measurable function h : X R such that (A) = A h d for all A A.
(The only possible such function would be h := H which is not measurable.)
= a + s = 0a + 0s , a , 0a , s , 0s .
s (A) = 0, (X \ A) = 0, 0s (A0 ) = 0, (X \ A0 ) = 0.
a (E A A0 ) = (E A A0 ) = 0a (E A A0 ).
Moreover (E \ (A A0 )) = 0, hence a (E \ (A A0 )) = 0 = 0a (E \ (A A0 ))
and hence
s (E \ (A A0 )) = (E \ (A A0 )) = 0s (E \ (A A0 )).
158 THEOREM
CHAPTER 5. THE RADONNIKODYM
This implies
a (E) = a (E A A0 ) = 0a (E A A0 ) = 0a (E)
s (E) = s (E \ (A A0 )) = 0s (E \ (A A0 )) = 0s (E).
This proves uniqueness.
We prove existence. The measure
:= + : A [0, ]
is -finite. Hence it follows from the RadonNikod
ym Theorem 5.4 that there
exist measurable functions f, g : X [0, ) such that
Z Z
(E) = f d, (E) = g d for all E A. (5.7)
E E
Define
A := x X g(x) > 0 (5.8)
and
a (E) := (E A), s (E) := (E Ac ) for E A. (5.9)
Then it follows directly from (5.9) that the maps a , s : A [0, ] are
measures and satisfy a + s = . Moreover, it follows from (5.9) that
s (A) = (A Ac ) = () = 0
and from (5.8) that g|Ac = 0, so by (5.7)
Z
c
(A ) = g d = 0.
Ac
This shows that s . It remains to prove that a is absolutely continuous
with respect to . To see this, let E A such that (E) = 0. Then by (5.7)
Z Z
E g d = g d = (E) = 0.
X E
Hence it follows from Lemma 1.49 that E g vanishes -almost everywhere.
Thus EA g = A E g vanishes -almost everywhere. Since g(x) > 0 for all
x E A, this implies
(E A) = 0.
Hence Z
a (E) = (E A) = f d = 0.
EA
This shows that a and completes the proof of Theorem 5.3.
5.2. THE DUAL SPACE OF LP () REVISITED 159
L () L1 () : g 7 g (5.13)
is bijective.
5.2. THE DUAL SPACE OF LP () REVISITED 161
Proof. The proof that (i) implies (ii) is outlined in Exercise 4.59.
We prove that (ii) implies (iii). Since (X, A, ) is semi-finite, the linear
map (5.13) is injective by Theorem 4.33. We must prove that it is surjective.
Assume firstthat : L1 () R is a positive bounded linear functional.
Define E := E A (E) < and, for E E, define
AE := {A A | A E} , E := |AE . (5.14)
E = E : L1 (E ) R.
satisfies the hypotheses of condition (F) on page 148. Thus it follows from (ii)
that there exists a measurable function g : X R such that, for all E E,
the restriction g|E agrees with gE almost everywhere on E.
We prove that g 0 almost everywhere. Suppose otherwise that the
set A := {x X | g(x) < 0} has positive measure. Since (X, A, ) is semi-
finite there exists a set E E such that E A and (E) > 0. Since
g(x) < 0 gE (x) for all x E it follows that g|E does not agree with gE
almost everywhere, a contradiction. This contradiction shows that g 0
almost everywhere.
162 THEOREM
CHAPTER 5. THE RADONNIKODYM
Hence
(f ) = lim (Ei f ) = lim Ei (f |Ei )
i i
Z Z Z
= lim f gEi d = lim f g d = f g d.
i Ei i Ei X
Here the last step follows from the Lebesgue Monotone Convergence Theo-
rem 1.37. This shows that (f ) = g (f ) for every nonnegative integrable
function f : X [0, ). It follows that
(f ) = (f + ) (f ) = g (f + ) g (f ) = g (f )
for all f L1 (). Thus = g as claimed.
This shows that every positive bounded linear functional on L1 () be-
longs to the image of the map (5.13). Since every bounded linear functional
on L1 () is the difference of two positive bounded linear functionals by The-
orem 4.39, it follows that the map (5.13) is surjective. Thus we have proved
that (ii) implies (iii).
5.2. THE DUAL SPACE OF LP () REVISITED 163
We prove that (iii) implies (i). Assume that the map (5.13) is bijective.
Then (X, A, ) is semi-finite by part (iv) of Theorem 4.33. Now let E A
be any collection of measurable sets. Assume without loss of generality that
E1 , . . . , E` E = E1 E` E.
(Otherwise, replace E by the collection E 0 of all finite unions of elements
of E; then every measurable envelope of E 0 is also an envelope of E.) For
E E define AE and E by (5.14) and define the bounded linear functional
E : L1 (E ) R by
Z
E (f ) := f dE for f L1 (E ). (5.16)
E
Here the second equation follows from (5.18), the third follows from (5.16),
and the last follows from the fact that E A E \ G for all E E. Since
g > 0 on A it follows from Lemma 1.49 that (A) = 0, a contradiction. This
contradiction shows that our assumption that (H \ G) > 0 must have been
wrong and so (H \ G) = 0 as claimed. Thus we have proved that every
collection of measurable sets E A has a measurable envelope, and this
completes the proof of Theorem 5.8.
5.2. THE DUAL SPACE OF LP () REVISITED 165
Theorem 4.40 also asserts that there exists a measurable set N A such
that (N ) = 0 and the restriction of to X \ N is -finite. Define
X0 := X \ N, A0 := {A A | A X0 } , 0 := |A0
as in (5.14), let 0 : Lp (0 ) Lp () be the extension operator as in (5.15),
and define 0 := 0 : Lp (0 ) R. Then 0 is a positive bounded linear
functional on Lp (0 ) and
Z Z
(f ) = f d = f d = 0 (f |X0 ) for all f Lp ().
X X\N
Since (X0 , A0 , 0 ) is -finite it follows from Theorem 5.7 that there exists a
function g0 Lq (0 ) such that g0 0 and
Z
0 (f0 ) = f0 g0 d0 for all f0 Lp (0 ).
X0
Take the supremum over all pairs of measurable sets E, F satisfying (5.22)
to obtain
X
||(A) ||(Ai ) (5.23)
i=1
To prove the converse inequality, fix a constant > 0. Then there are
sequences of measurable sets Ei , Fi A such that
Ei Fi = , Ei Fi = Ai , (Ei ) (Fi ) > ||(Ai ) i
2
S S
for all i N. The sets E := i=1 Ei and F := i=1 Fi satisfy (5.22) and so
X X
||(A) (E) (F ) = (Ei ) (Fi ) > ||(Ai ) .
i=1 i=1
= a + s , a , s . (5.27)
Hence
Z Z
(h 1) d = (P ) (P ) = 0, (h + 1) d = (N ) + (N ) = 0.
P N
5.4 RadonNikod
ym Generalized
This section discusses an extension of the RadonNikod ym Theorem 5.18 for
signed measures to all measure spaces. Thus we drop the hypothesis that
is -finite. In this case Examples 5.5 and 5.6 show that absolute conti-
nuity of with respect to is not sufficient for obtaining the conclusion
of the RadonNikod ym Theorem and a stronger condition is needed. In [4,
Theorem 232B] Fremlin introduces the notion truly continuous, which is
equivalent to absolutely continuous whenever is -finite. In [8] Konig re-
formulates Fremlins criterion in terms of inner regularity of with respect
to . We shall discuss both conditions below, show that they are equiva-
lent, and prove the generalized RadonNikod ym Theorem. As a warmup we
rephrase absolute continuity in the familiar - language of analysis.
Standing Assumption. Throughout this section (X, A, ) is a measure
space and : A R is a signed measure.
Lemma 5.21 (Absolute Continuity). The following are equivalent.
(i) is absolutely continuous with respect to .
(ii) For every > 0 there exists a constant > 0 such that
Proof. That (ii) implies (i) is obvious. Conversely, assume (i). Then ||
by Lemma 5.16. Assume by contradiction that (ii) does not hold. Then there
exists a constant > 0 and a sequence of measurable sets Ai A such that
For n N define
[
\
Bn := Ai , B := Bn .
i=n n=1
Then
1
Bn Bn+1 , (Bn ) , ||(Bn ) ||(An ) |(An )|
2n1
for all n N. Hence (B) = 0 and ||(B) = limn ||(Bn ) by part (v)
of Theorem 1.28. This contradicts the fact that || . This contradiction
shows that our assumption that (ii) does not hold must have been wrong.
Thus (i) implies (ii) and this proves Lemma 5.21.
174 THEOREM
CHAPTER 5. THE RADONNIKODYM
(A P ) 0, (A \ P ) 0, ||(P ) = (A P ) (A \ P ) (5.32)
for all A A. Such a measurable set P will be fixed throughout the proof.
We prove that (i) implies (ii). Fix a set A A such that ||(AE) = 0 for
all E E. Then it follows from (5.32) that (A E P ) = (A E \ P ) = 0
for all E E. By (i) this implies (A P ) = (A \ P ) = 0 and hence
||(A) = 0 by (5.32). This shows that (i) implies (ii).
We prove that (ii) implies (i). Fix a set A A such that (A E) = 0
for all E E. Since E P E and E \ P E for all E E this implies
(A E P ) = (A E \ P ) = 0 for all E E. Hence it follows from (5.32)
that ||(A E) = 0 for all E E. By (ii) this implies ||(A) = 0 and hence
(A) = 0 because |(A)| ||(A). This shows that (ii) implies (i).
We prove that (ii) implies (iii). Fix a set A A and define
Define
[
B := A \ F, F := Fi . (5.36)
i=1
Then ||(F ) = limi ||(Fi ) = c by part (iv) of Theorem 1.28 and hence
Definition 5.25. The signed measure is called inner regular with re-
spect to if it satisfies the equivalent conditions of Lemma 5.24.
Theorem 5.26 (Generalized RadonNikod ym Theorem).
Let (X, A, ) be a measure space and let : A R be a signed measure.
Then the following are equivalent.
(i) is truly continuous with respect to .
(ii) is absolutely continuous and inner regular with respect to .
(iii) There exists a function f L1 () such that (5.28) holds.
If these equivalent conditions are satisfied then the function f in (iii) is
uniquely determined by (5.28) up to equality -almost everywhere.
First proof of Theorem 5.26. This proof is due to Konig [8]. It has the advan-
tage that it reduces the proof of the generalized RadonNikod ym Theorem
to the standard RadonNikod ym Theorem 5.18 for -finite measure spaces.
We prove that (i) implies (ii). To see that is absolutely continuous
with respect to , fix a measurable set A A such that (A) = 0 and fix a
constant > 0. Choose > 0 and E A such that (E) < and (5.31)
holds. Then (A E) (A) = 0 < and hence |(A)| < by (5.31). Thus
|(A)| < for all > 0 and hence (A) = 0. This shows that .
We prove that is inner regular with respect to by verifying that
satisfies condition (i) in Lemma 5.24. Fix a set A A such that
E A, (E) < = (A E) = 0.
We must prove that (A) = 0. Let > 0 and choose > 0 and E A
such that (E) < and (5.31) holds. Then ((A \ E) E) = 0 < , hence
|(A \ E)| < by (5.31), and hence
This shows that |(A)| < for all > 0 and so (A) = 0 as claimed. Thus
we have proved that (i) implies (ii).
We prove that (ii) implies (iii). Since is inner regular with respect to
there exists a sequence of measurable sets Ei A such that Ei Ei+1 and
(Ei ) < for all i N and ||(X) = limi ||(Ei ). Define
[
X0 := Ei , A0 := {A A | A X0 } , 0 := |A0 , 0 := |A0 .
i=1
GENERALIZED
5.4. RADONNIKODYM 177
Hence Z Z
(A) = 0 (A X0 ) = f0 d0 = f d
AX0 A
for all A A. This shows that (i) implies (iii). The uniqueness of f up to
equality -almost everywhere follows immediately from Lemma 1.49.
We prove that (iii) implies (i). Choose f L1 () such that (5.28) holds.
Define Z
c := ||(X) = |f | d
X
and
En := x X | 2n |f (x)| 2n ,
[
E := {x X | f (x) 6= 0} = En .
nN
Second proof of Theorem 5.26. This proof is due to Fremlin [4, Chapter 23].
It shows directly that (i) implies (iii) and has the advantage that it only uses
the Hahn Decomposition Theorem 5.19. It thus also provides an alternative
proof of Theorem 5.18 (assuming the Hahn Decomposition Theorem) which
is of interest on its own. By Lemma 5.23 it suffices to consider the case where
: A [0, ) is a finite measure that is truly continuous with respect to .
Consider the set
f is measurable and
F := f : X [0, ) R .
A
f d (A) for all A A
all f F and
f, g F = max{f, g} F . (5.38)
(Let f, g F and A A and define the sets Af := {x A | f (x) > g(x)} and
Ag := R{x A | g(x) f (x)};
R then AfR, Ag A, Af Ag = , and Af Ag = A;
hence A max{f, g} d = Af f d + Ag g d (Af ) + (Ag ) = (A).)
Now define Z
c := sup f d (X)
f F X
(A P E)
Z
0
(A P ) = h d.
(E) A
5.5 Exercises
Exercise 5.27. Let (X, A, ) be a measure space such that (X) < .
Define
(A, B) := (A \ B) + (B \ A) for A, B A. (5.45)
Define an equivalence relation on A by A B iff (A, B) = 0. Prove that
descends to a function : A/ A/ [0, ) (denoted by the same
letter) and that the pair (A/,
R ) is a complete metric space. Prove that the
function A R : A 7 A f d descends to a continuous function on A/
for every f L1 ().
Exercise 5.28 (Rudin [17, page 133]). Let (X, A, ) be a measure space.
A subset F L1 () is called uniformly integrable if, for every > 0,
there is a constant > 0 such that, for all E A and all f F ,
Z
(E) < = f d < .
E
Now use part (i) to find a constant 0 > 0 such that, for all A A,
Z
0
(A) < = sup fn d < 3.
(5.48)
nN A
Exercise 5.29 (Rudin [17, page 134]). Let (X, A, ) be a measure space
such that (X) < and fix a real number p > 1. Let f : X R be
a measurable function and let fn L1 () be a sequence that converges
pointwise to f and satisfies
Z
sup |fn |p d < .
nN X
Prove that Z
1
f L (), lim |f fn | d = 0.
n X
Hint: Use Vitalis Theorem in Exercise 5.28.
Exercise 5.30. Let X := R, denote by B 2X the Borel -algebra, and
let : B [0, ] be the restriction of the Lebesgue measure to B. Let
: B [0, ] be a measure. Prove the following.
(i) If B B and 0 < c < (B) then there exists a Borel set A B such that
(A) = c. Hint: Show that the function f (t) := (B [t, t]) is continuous.
(ii) If there exists a constant 0 < c < such that
(B) = c = (B) = c.
1 , 2 , 1 2 , 2 1 ,
and
6 1 , 6 2 .
5.5. EXERCISES 183
Exercise 5.34. Let (X, A, ) be a measure space. Show that the signed
measures : A R form a Banach space M = M(X, A) with norm
kk := ||(X).
Show that the map
L1 () M : [f ] 7 f
defined by (5.53) is an isometric linear embedding and hence L1 () is a closed
subspace of M.
Exercise 5.35. Let (X, U) be a compact Hausdorff space such that every
open subset of X is -compact and denote by B 2X its Borel -algebra.
Denote by C(X) := Cc (X) the space of continuous real valued functions
on X. This is a Banach space equipped with the supremum norm
kf k := sup|f (x)|.
xX
Let M(X) denote the space of signed Borel measures as in Exercise 5.34.
For M(X) define the linear functional : C(X) R by
Z
(f ) := f d.
X
Prove the following.
(i) k k = kk. Hint: Use the Hahn Decomposition Theorem 5.19 and the
fact that every Borel measure on X is regular by Theorem 3.18.
(ii) Every bounded linear functional on C(X) is the difference of two positive
linear functionals. Hint: For f C(X) with f 0 prove that
+ (f ) := sup (hf ) h C(X), 0 h 1
(5.52)
= sup (g) g C(X), 0 g f .
Here the second supremum is obviously greater than or equal to the first. To
prove the converse inequality show that, for all g C(X) with 0 g f
and all > 0 there is an h C(X) such that 0 h 1 and |(g hf )| < .
Namely, find C(X) such that 0 1, (x) = 0 when f (x) /2 kk
and (x) = 1 when f (x) / kk; then define h := g/f . Once (5.52) is
established show that + extends to a positive linear functional on C(X).
(iii) The map M(X) C(X) : 7 is bijective. Hint: Use the Riesz
Representation Theorem 3.15.
(iv) The hypothesis that every open subset of X is -compact cannot be
removed in part (i). Hint: Consider Example 3.6.
184 THEOREM
CHAPTER 5. THE RADONNIKODYM
Differentiation
185
186 CHAPTER 6. DIFFERENTIATION
and only if (A(t, f )) > for all t > 0. When f () < it is the smallest
number t such that the domain A(t, f ) (on which |f | > t) has measure at
most . This is spelled out in the next lemma.
Lemma 6.1. Let 0 < and 0 t < . Then the following holds.
(i) f () = if and only if f (s) > for all s 0.
(ii) f () = t if and only if f (t) and f (s) > for 0 s < t.
(iii) f () t if and only if f (t) .
Proof. It follows directly from the definition of f in (6.2) that f () =
if and only if (s, f ) > for all s [0, ) and this proves (i).
To prove (ii), fix a constant 0 t < . Assume first that (t, f )
and (s, f ) > for 0 s < t. Since f is nonincreasing this implies
(s, f ) (t, f ) for all s t and hence f () = t by definition.
Conversely, suppose that f () = t. Then it follows from the definition of
f that (s, f ) for s > t and (s, t) > for 0 s < t. We must prove
that (t, f ) . To see this observe that
[
A(t, f ) = A(t + 1/n, f ).
n=1
Proof. For 0 < t, c < it follows from part (iii) of Lemma 6.1 that
t(t, f ) c (t, f ) ct1 f (ct1 ) t ct1 f (ct1 ) c.
This shows that supt>0 t(t, f ) = sup>0 f (). Moreover,
Z Z
t(t, f ) = t(A(t, f )) |f | d |f | d
A(t,f ) X
Example 6.3. This example shows that the weak triangle inequality (6.6) is
sharp. Let (R, A, m) be the Lebesgue measure space and define f, g : R R
by
1 1
f (x) := , g(x) := for 0 < x < 1
x 1x
and f (x) := g(x) := 0 for x 0 and for x 1. Then
For the Lebesgue measure space (Rn , A, m) we write L1, (Rn ) := L1, (m)
and L1, (Rn ) := L1, (m).
Proof. We prove that (i) implies (ii). Fix a constant > 0. Since f is
differentiable at x and f 0 (x) = A, there exists a constant > 0 such that,
for all y R,
f (x) f (y)
0 < |x y| < =
xy A . (6.12)
or, equivalently,
Since ([a, b)) = f (b) f (a) and m([a, b)) = b a it follows that
([a, b))
m([a, b)) A .
6.2. MAXIMAL FUNCTIONS 191
Thus (6.12) holds for x < y < x + and an analogous argument proves
the inequality for x < y < x. Thus (ii) implies (i) and this proves
Theorem 6.6.
The main theorem of this chapter will imply that, when is absolutely
continuous with respect to m, the derivative of the function f in (6.10) exists
almost everywhere, defines a Lebesgue integrable function f 0 : R R, and
that Z
(A) = f 0 dm
A
for all Lebesgue measurable sets A A. It will then follow that an absolutely
continuous function on R can be written as the integral of its derivative. This
is the fundamental theorem of calculus in measure theory (Theorem 6.19).
The starting point for this program is the assertion of Theorem 6.6. It
suggests the definition of the derivative of a signed measure
:AR
denote the Borel -algebra of Rn with the standard topology. Thus L1 (Rn )
denotes the space of Lebesgue integrable functions f : Rn R. An element of
L1 (Rn ) need not be Borel measurable but differs from a Borel measurable func-
tion on a Lebesgue null set by Theorem 2.14 and part (v) of Theorem 1.55.
For x Rn and r > 0 denote the open ball of radius r, centered at x, by
Br (x) := y Rn |x y| < r .
Here q
|| := 12 + + n2
denotes the Euclidean norm of = (1 , . . . , n ) Rn .
Definition 6.7 (HardyLittlewood Maximal Function).
Let : B [0, ) be a finite Borel measure. The maximal function of
is the function
M : Rn R
defined by
(Br (x))
(M )(x) := sup . (6.13)
r>0 m(Br (x))
The maximal function of a signed measure : B R is defined as the
maximal function
M := M || : Rn R
of its total variation || : B [0, ).
Theorem 6.8 (HardyLittlewood Maximal Inequality).
Let : B R be a signed Borel measure. Then the function M : Rn R
in Definition 6.7 is lower semi-continuous, i.e. the pre-image of the open in-
terval (t, ) under M is open for all t R. Hence M is Borel measurable.
Moreover,
kM k1, 3n ||(Rn ) (6.14)
and so M L1, (Rn ).
Proof. See page 195.
6.2. MAXIMAL FUNCTIONS 193
Choose y Rn such that |y x| < . Then Br (x) Br+ (y) and hence
This implies
(Br+ (y))
(M )(y) >t
m(Br+ (y))
and [
W B3ri (xi ). (6.17)
iS
Proof. Abbreviate Bi := Bri (xi ) and choose the ordering such that
r1 r2 r` .
Choose i1 := 1 and let i2 > 1 be the smallest index such that Bi2 Bi1 = .
Continue by induction to obtain a sequence
such that
Bij Bij0 = for j 6= j 0
and
Bi (Bi1 Bij ) 6= for ij < i < ij+1
(respectively for i > ik when j = k). Then
and hence
`
[ k
[
W = Bi B3rij (xij ).
i=1 j=1
Proof of Theorem 6.8. Fix a constant t > 0. Then the set Ut := A(t, M ) is
open by Lemma 6.9. Choose a compact set K Ut . If x K Ut then
(M )(x) > t and so there exists a number r(x) > 0 such that
||(Br(x) (x))
> t. (6.18)
m(Br(x) (x))
Here the second step follows from (6.18) with ri = r(xi ) and the last step
follows from the fact that the balls Bri (xi ) for i S are pairwise disjoint.
Take the supremum over all compact sets K Ut to obtain
3n
m(A(t, M )) = m(Ut ) ||(Rn ). (6.19)
t
(See Theorem 2.13.) Multiply the inequality (6.19) by t and take the supre-
mum over all real numbers t > 0 to obtain kM k1, 3n ||(Rn ). This
proves Theorem 6.8.
To see this, fix a function f L1 (Rn ) and assume without loss of gener-
ality that f is Borel measurable. (See Theorem 2.14 and part (v) of Theo-
rem 1.55.) By Theorem 4.15 there exists a sequence of continuous functions
gi : Rn R with compact support such that
1
kf gi k1 < for all i N.
2i
Since gi is continuous we have T gi = 0. Moreover, the function
hi := f gi
T hi M hi + |hi |
Tr f = Tr (gi + hi ) Tr gi + Tr hi
for all i and all r > 0. Take the limit superior as r tends to zero to obtain
T f T gi + T hi = T hi M hi + |hi |
for all i and all > 0. (See equation (6.1) for the notation A(, T f ) etc.)
Since hi and M hi are Borel measurable (see Theorem 6.8) the set
2 2 1
m(A(/2, hi )) khi k1, khi k1 i1
2
and, by Theorem 6.8,
2 2 3n 3n
m(A(/2, M hi )) kM hi k1, khi k1 i1 .
2
Thus
3n + 1
m(Ei ()) .
2i1
Since this holds for all i N it follows that the Borel set
\
E() := Ei ()
i=1
has Lebesgue measure zero for all > 0. Hence the Borel set
[
E := E(1/k)
k=1
A(1/k, T f ) E(1/k)
This shows that (T f )(x) = 0 for all x Rn \ E and hence every element of
Rn \ E is a Lebesgue point of f . This proves Theorem 6.14.
6.3. LEBESGUE POINTS 199
The right hand side denotes the Lebesgue integral of g over the interval [x, y].
If (iii) holds then there exists a Borel set E I such that m(E) = 0 and,
for all x I \ E, f is differentiable at x and f 0 (x) = g(x).
Proof. We prove that (iii) implies the last assertion of the theorem. Thus
assume that there exists a function g L1 (I) that satisfies (6.30) for all
x, y I with x < y. Then Theorem 6.14 asserts that there exists a Borel
set E I of Lebesgue measure zero such that every element of I \ E is a
Lebesgue point of g. By Theorem 6.16 with = 1/2, every element x I \ E
satisfies condition (ii) in Theorem 6.6 with A := g(x). Hence Theorem 6.6
asserts that the function f is differentiable at every point x I \ E and
satisfies f 0 (x) = g(x) for x I \ E.
202 CHAPTER 6. DIFFERENTIATION
We prove that (iii) implies (i). Thus assume f satisfies (iii) and define
the signed measure : B R by
Z
(B) := g dm (6.31)
B
for ever Borel set B I. Now let > 0. Since || m it follows from
Lemma 5.21 that there exists a constant > 0 such that ||(B) < for every
Borel set B PI with m(B) < . Choose a sequence s1 t1 s` t`
in I such that `i=1 |ti si | < and define Ui := (si , ti ) for i = 1, . . . , `. Then
the Borel set B := `i=1 Ui has Lebesgue measure m(B) = `i=1 |ti si | < .
S P
Hence ||(B) < . Since
Z Z
|f (ti ) f (si )| =
g dm |g| dm = ||(Ui )
Ui Ui
We prove that (ii) implies (iii). Assume f satisfies (ii). Then f has
bounded variation and under this assumption Exercise 6.20 below outlines
a proof that there exists a signed Borel measure : B R such that
f (y) f (x) = ((x, y]) for x, y I with x < y. Since + and are regular
by Theorem 3.18 and f is continuous, we have
f (y) f (x) = ([x, y]) = ((x, y]) = ([x, y)) = ((x, y)) (6.32)
for all x, y I with x < y. By the Lebesgue Decomposition Theorem 5.17
there exist two signed Borel measures a , s : B R such that
= a + s , a m, s m. (6.33)
Since a m it follows from Theorem 5.18 that there is an integrable
function g L1 (I) such that
Z
a (B) = g dm (6.34)
B
Then f = fa +fs by (6.32) and (6.33). Since (iii) implies (i) and (i) implies (ii)
(already proved) both functions f and fa satisfy (ii) and fa is absolutely
continuous. Moreover fs = f fa is continuous.
It remains to prove that fs 0. The proof given below follows an ar-
gument in Heil [7, Section 3.5.4]. By Theorem 6.17, there exists a Lebesgue
null set Es I such that a, b Es (without loss of generality) and
|s ((x r, x + r))|
|s |(I \ Es ) = 0, lim = 0 for all x I \ Es .
r0 r
This implies that every element x I \ Es satisfies condition (ii) in Theo-
rem 6.6 with A = 0 and f replaced by fs . Hence fs is differentiable at every
point x I \ Es and fs0 (x) = 0 for x I \ Es .
Since Es is a Lebesgue null set so are the sets f (Es ) and fa (Es ), because
the functions f and fa satisfy (ii). Since the difference of two Lebesgue null
sets in R is a Lebesgue null set, by Exercise 2.27, it follows that
fs (Es ) = {f (x) fa (x) | x Es } {f (x) fa (y) | x, y Es }
is a Lebesgue null set.
204 CHAPTER 6. DIFFERENTIATION
Here the last assertion follows from the fact that fs is differentiable on I \ Es
with derivative zero. We prove that the Lebesgue outer measure of the set
fs (I \ Es ) satisfies the estimate
(fs (I \ Es )) (b a + ). (6.35)
To see this, cover the set An by at most countably many open intervals Ui ,
each of length less than 1/n, such that
X
m(Ui ) (An ) +
i
Since the Lebesgue outer measure is continuous from below by part (iii) of
Theorem 2.13 and [
fs (I \ Es ) = fs (An ),
nN
it follows that
This proves (6.35). Since > 0 was chosen arbitrary, this implies that
fs (I \ Es ) is a Lebesgue null set as claimed. Since fs (Es ) is also a Lebesgue
null set, as noted above, it follows that fs (I) is a Lebesgue null set. Since fs
is continuous and fs (0) = 0 by definition, this implies fs 0. Hence f = fa
is absolutely continuous and this proves Theorem 6.19.
6.5. EXERCISES 205
6.5 Exercises
Exercise 6.20. Let I = [a, b] R be a compact interval and let B 2I
be the Borel -algebra. A function f : I R is said to be of bounded
variation if
`
X
V (f ) := sup |f (ti ) f (ti1 )| < . (6.36)
a=t0 <t1 <<t` =b i=1
Exercise 6.21. Let (R, A, m) be the Lebesgue measure space and fix a con-
stant 0 < < 1/2. Prove that there does not exist a Lebesgue measurable
set E R such that
m(E I)
< <1
m(I)
for every nonempty bounded open interval I R. Hint: Consider the
function f := E[1,1] and define the measure f : B R by
Z
f (B) := f dm = m(B E [1, 1]).
B
Examine the Lebesgue points of f . (Compare this with Exercise 2.32.)
Exercise 6.22. Prove the Theorem of VitaliCarath eodory:
Let (X, U) be a locally compact Hausdorff space and let B 2X be its Borel
-algebra. Let : B [0, ] be an outer regular Borel measure that is
inner regular on open sets. Let f L1 () and let > 0. Then there exists
an upper semi-continuous function u : X R that is bounded above and a
lower semi-continuous function v : X R that is bounded below such that
Z
u f v, (v u) d < . (6.40)
X
Hint: Assume first that f 0. Use Theorem 1.26 to find a sequence of
measurable sets Ei A, not necessarily disjoint, and a sequence of real
numbers ci > 0 such that (Ei ) < for all i and
X
f= ci Ei .
i=1
Thus Z
X
ci (Ei ) = f d < .
i=1 X
is open for all t R), and u is upper semi-continuous (i.e. u1 ((, t))
is open for all t R).
6.5. EXERCISES 207
Exercise 6.23. Fix two real numbers a < b and prove the following.
(i) If f : [a, b] R is everywhere differentiable then f 0 : [a, b] R is Borel
measurable.
Rb
(ii) If f : [a, b] R is everywhere differentiable and a |f 0 (t)| dt < then f
is absolutely continuous.
Hint: Fix a constant > 0. By the VitaliCaratheodory Theorem in Exer-
cise 6.22 there is a lower semi-continuous function g : [a, b] R such that
Z b Z b
0
g>f, g(t) dt < f 0 (t) dt + .
a a
for a x b. Consider a point a x < b. Since g(x) > f 0 (x) and g is lower
semi-continuous, find a number x > 0 such that
f (t) f (x)
g(t) > f 0 (x), < f 0 (x) + for x < t < x + x .
tx
Deduce that
F (t) > F (x) for x < t < x + x .
Since F (a) = 0 there exists a maximal element x [a, b] such that F (x) = 0.
If x < b it follows from the previous discussion that F (t) > 0 for x < t b.
In either case F (b) 0 and hence
Z b Z b
f (b) f (a) g(t) dt + (b a) < f 0 (t) dt + + (b a).
a a
Since this holds for all > 0 and all > 0 it follows that
Z b
f (b) f (a) f 0 (t) dt.
a
Rb
Replace f by f to obtain the equation f (b) f (a) = a
f 0 (t) dt. Now
deduce that Z x
f (x) f (a) = f 0 (t) dt
a
for all x [a, b].
208 CHAPTER 6. DIFFERENTIATION
Example 6.24. (i) The Cantor function is the unique monotone function
f : [0, 1] [0, 1] that satisfies
X X
ai ai
f 2 i
=
i=1
3 i=1
2i
for all sequences ai {0, 1}. It is continuous and nonconstant and its deriva-
tive exists and vanishes on the complement of the standard Cantor set
[ X n n
\ ai X ai 1
C := 2 i
,2 i
+ n .
n=1 i=1
3 i=1
3 3
ai {0,1}
This Cantor set has Lebesgue measure zero. Hence f is almost everywhere
differentiable and its derivative is integrable. However, f is not equal to the
integral of its derivative and therefore is not absolutely continuous.
(ii) The following construction was explained to me by Theo Buehler. Define
the homeomorphisms g : [0, 1] [0, 2] and h : [0, 2] [0, 1] by
g(x) := f (x) + x, h := g 1 .
The image g([0, 1] \ C) is a countable union of disjoint open intervals of
total length one and hence has Lebesgue measure one. Thus its complement
K := g(C) [0, 2] is a modified Cantor set of Lebesgue measure one. Hence,
by Theorem 6.19, g is not absolutely continuous. Moreover, by Lemma 2.15
there exists a set E K which is not Lebesgue measurable. However, its
image F := h(E) [0, 1] under h is a subset of the Lebesgue null set C
and hence is a Lebesgue null set. Thus F is a Lebesgue measurable set
and E = h1 (F ) is not Lebesgue measurable. This shows that the function
h : [0, 2] [0, 1] is not measurable with respect to the Lebesgue -algebras on
both domain and target (i.e. it is not Lebesgue-Lebesgue measurable).
(iii) Let I, J R be intervals. Then it follows from Lemma 2.15 and
Theorem 6.19 that every Lebesgue-Lebesgue measurable homeomorphism
h : I J has an absolutely continuous inverse.
(iv) Let h : [0, 2] [0, 1] and F C [0, 1] be as in part (ii). Then the
characteristic function F : R R is Lebesgue measurable and h : [0, 2] R
is continuous. However, the composition F h : [0, 2] R is not Lebesgue
measurable because the set (F h)1 (1) = E is not Lebesgue measurable.
(v) By contrast, if I, J R are intervals, f : J R is Lebesgue measurable,
and h : I J is a C 1 diffeomorphism, then f h : I R is again Lebesgue
measurable by Theorem 2.17.
Chapter 7
Product Measures
209
210 CHAPTER 7. PRODUCT MEASURES
(E c )x = {y Y | (x, y)
/ E} = (Ex )c B
M A B.
is a monotone class.
This follows immediately from the definition of monotone class.
Step 2. Let P, Q X Y . Then Q (P ) if and only if P (Q).
This follows immediately from the definition of (P ) in Step 1.
Step 3. If P, Q E then P Q, P \ Q, P Q E.
For the intersection this follows from the fact that
Lemma 7.6. Let (X, UX ) and (Y, UY ) be topological spaces, let UXY be
the product topology on X Y (see Appendix B), and let BX , BY , BXY be
the associated Borel -algebras. Then
BX BY BXY . (7.3)
BX BY = BXY . (7.4)
Lemma 7.7. Let (X, A) be a measurable space such that the cardinality of
X is greater than that of 2N . Then the diagonal := {(x, x) | x X} is not
an element of A A.
Proof. The proof has three steps.
Step 1. Let Y be a set. For E 2Y denote by (E) 2Y the smallest
-algebra containing E. If D (E) then there exists a sequence Ei E for
i N such that D ({Ei | i N}).
The union of the sets (E 0 ) over all countable subsets E 0 E is a -algebra
that contains E and is contained in (E). Hence it is equal to (E).
Step 2. Let Y be a set, let E 2Y , and let DS (E). T ThenTthere is a
sequence Ei E and a set I 2N such that D = II iI Ei iN\I Eic .
By Step 1 there exists T Ei Ec such that D ({Ei | i N}). For
T a sequence
I N define EI := iI Ei iN\I Ei . These sets form a partition of Y .
S
Hence the collection F := II E I | I 2 N
is a -algebra on Y . Since
Ei F for each i N it follows that D F. This proves Step 2.
Step 3. / A A.
Let E 2XX be the collection of all sets of the form A B with A, B A.
Let D A A.SBy Step 2 there are sequences Ai , Bi ATand a set I 2N
(A Bi )c .
T
such that D = II EI , where EI := iI (ATi Bi )
S TiN\I i
Thus EI = JN\I AIJ BIJ , where AIJ := iI Ai jJ (X \ Aj ) and
T T
BIJ := iI Bi jN\(IJ) (X \ Bj ). If D then, for all I and J, we
have AIJ BIJ and so AIJ BIJ is either empty or a singleton. Thus
the cardinality of D is at most the cardinality of the set of pairs of disjoint
subsets of N, which is equal to the cardinality of 2N . Since the cardinality of
the diagonal is bigger than that of 2N it follows that / A A as claimed.
This proves Lemma 7.7.
Example 7.8. Let X be an uncountable set, of cardinality greater than
that of 2N , and equipped with the discrete topology so that BX = UX = 2X .
Then is an open subset of X X with respect to the product topology
(which is also discrete because points are open). Hence BXX = 2XX .
However, / BX BX by Lemma 7.7. Thus the product BX BX of the
Borel -algebras is not the Borel -algebra of the product. In other words,
the inclusion (7.3) in Lemma 7.6 is strict in this example. Note also that
the distance function d : X X R defined by d(x, y) := 1 for x 6= y and
d(x, x) := 0 is continuous with respect to the product topology but is not
measurable with respect to the product of the Borel -algebras.
214 CHAPTER 7. PRODUCT MEASURES
Definition 7.10. Let (X, A, ) and (Y, B, ) be -finite measure spaces. The
product measure of and is the map : A B [0, ] defined by
Z Z
( )(Q) := (Qx ) d(x) = (Qy ) d(y) (7.8)
X Y
Step 5. = A B.
Since (X, A, ) and (Y, B, ) are -finite, there exist sequences of measurable
sets Xn A and Yn B such that
Xn Xn+1 , Yn Yn+1 , (Xn ) < , (Yn ) <
for all n N and X =
S S
n=1 Xn and Y = n=1 Yn . Define
M := Q A B Q (Xn Yn ) for all n N .
Then M is a monotone class by Steps 3 an 4, E M by Steps 1 and 2, and
M AB by definition. Hence it follows from Lemma 7.5 that M = AB.
In other words Q (Xn Yn ) for all Q A B. By Step 3 this implies
[
Q= Q (Xn Yn ) for all Q A B.
n=1
Then the set Qy = {x [0, 1] | j(x) j(y)} is countable for all y [0, 1] and
the set [0, 1] \ Qx = {y [0, 1] | j(y) 4 j(x)} is countable for all x [0, 1].
Proof of Claim 2. By Zorns Lemma every set admits a well ordering.
Choose any well ordering on A := [0, 1] and define
B := {b A | the set {a A | a b} is uncountable} .
If B = choose W := A = [0, 1] and j = id. If B 6= then, by the well
ordering axiom, B contains a smallest element b0 . Since b0 B, the set
W := B \ A = {w A | w b0 } is uncountable. Since W B = the set
{w W | w z} is countable for all z W . Since W is an uncountable
subset of [0, 1], the continuum hypothesis asserts that there exists a bijection
j : [0, 1] W . This proves Claim 2.
218 CHAPTER 7. PRODUCT MEASURES
Exercise 7.15. Let (X, A, ), (Y, B, ) be -finite measure spaces and let
: X X, :Y Y
( ) (A B) = A B, ( ) ( ) = .
n : Bn [0, ]
Bk B` = Bn
k ` = n .
Then
n n+1 , n n+1 for all n N
by part (i) of Theorem 1.35. Moreover, it follows from the Lebesgue Mono-
tone Convergence Theorem 1.37 that
(x) = lim n (x), (y) = lim n (y)
n n
Define : X [0, ] by
Z
(x) := fx d for x X
Y
and let g Lq (). Then the function X Y [0, ] : (x, y) 7 f (x, y)|g(x)|
is A B-measurable. Hence it follows from Theorem 7.17 that
Z Z Z
|g| d = f (x, y)|g(x)| d(y) d(x)
X X Y
Z Z
= f (x, y)|g(x)| d(x) d(y)
Y X
Z
kf y kLp () kgkLq () d(y)
Y
= c kgkLq () .
Here the third step follows from Holders inequality in Theorem 4.1. Since
(X, A, ) is semi-finite by part (ii) of Lemma 4.30, it follows from Lemma 4.34
that kkLp () c. This proves Theorem 7.19.
222 CHAPTER 7. PRODUCT MEASURES
Proof. We prove part (i) and the first equation in (7.16). The functions
f := max{f, 0} : X Y [0, ) are AB-measurable by Theorem 1.24.
Hence the functions fx := max{fx , 0} = f (x, ) : Y [0, ) are B-
measurable by Lemma 7.2. Define : Y [0, ] by
Z
(x) := fx d for x X.
Y
xE / L1 ().
fx
7.3. FUBINIS THEOREM 223
Define : X [0, ) by
(x), if x
/ E,
(x) := for x X.
0, if x E,
Then it follows from (7.17) that L1 () and
Z Z
d = f d( ).
X XY
Hence = + L1 () and
Z Z Z
d = +
d d
X ZX X Z
= +
f d( ) f d( )
ZXY XY
= f d( ).
XY
This proves (i) and the first equation in (7.16). An analogous argument
proves (ii) and the second equation in (7.16). This proves Theorem 7.20.
Example 7.21. Let (X, A, ) = (Y, B, ) be the Lebesgue measure space
in the unit interval [0, 1] as in Example 7.12. Let gn : [0, 1] [0, ) be a
sequence of smooth functions such that
Z 1
gn (x) dx = 1, gn (x) = 0 for x [0, 1] \ [2n1 , 2n ]
0
The sum on the right is finite for every pair (x, y) [0, 1]2 . Then
Z Z
X
f (x, y) dx = 0, f (x, y) dy = gn (x) gn+1 (x) = g1 (x),
X Y n=1
and hence
Z 1 Z 1 Z 1 Z 1
f (x, y) dx dy = 0 6= 1 = f (x, y) dy dx.
0 0 0 0
Example 7.22. This example shows that the product measure is typically
not complete. Let (X, A, ) and (Y, B, ) be two complete -finite measure
spaces. Suppose (X, A, ) admits a nonempty null set A A and B 6= 2Y .
Choose B 2Y \ B. Then A B / A B. However, A B is contained in
the -null set A Y and so belongs to the completion (A B) .
In the first version of Fubinis Theorem integrability was not an issue. In
the second version integrability of fx was only guaranteed for almost all x.
In the third version the function fx may not even be measurable for all x.
Theorem 7.23 (Fubini for the Completion). Let (X, A, ) and (Y, B, )
be complete -finite measure spaces, let (X Y, (A B) , ( ) ) de-
note the completion of the product space, and let f L1 (( ) ). Define
fx (y) := f y (x) := f (x, y) for x X and y Y . Then the following holds.
(i) fx L1 () for -almost every x X and the map : X R defined by
R
f d, if fx L1 (),
Y x
(x) := (7.18)
0, / L1 (),
if fx
is -integrable.
(ii) f y L1 () for -almost every y Y and the map : Y R defined by
R y
X
f d, if f y L1 (),
(y) := (7.19)
0, if f y
/ L1 (),
is -integrable.
(iii) Let L1 () and L1 () be as in (i) and (ii). Then
Z Z Z
d = f d( ) = d. (7.20)
X XY Y
gx L1 () for all x X \ E 0 ,
g y L1 () for all y Y \ F 0 .
f d, for x X \ (E E 0 ),
R
(x) := Y x
0, for x E E 0 ,
f y d, for y Y \ (F F 0 ),
R
(y) := X
0, for y F F 0 .
orem 7.20 for g that L1 (). The same argument, using part (ii) of Theo-
rem 7.20 for g, shows that L1 (). Moreover, the three integrals in (7.20)
for f agree with the corresponding integrals for g because
(E E 0 ) = (F F 0 ) = ( )(Q) = 0.
Hence equation (7.20) for f follows from part (iii) of Theorem 7.20 for g.
This proves Theorem 7.23.
Example 7.24. Assume (X, A, ) is not complete. Then there exists a set
E 2X \ A and a set N A such that E N and (N ) = 0. In this case
the set E Y is a null set in the completion (X Y, (A B) , ( ) ).
Hence f := EY L1 (( ) ). However, the function f y = E is not
measurable for every y Y . This shows that the hypothesis that (X, A, )
and (Y, B, ) are complete cannot be removed in Theorem 7.23.
Exercise 7.25. Continue the notation of Theorem 7.23 and suppose that
f : X Y [0, ] is (A B) -measurable. Prove that fx is B-measurable
for -almost all x X, that f y is A-measurable for -almost all y Y , and
that equation (7.11) continues to hold. Hint: The proof of Theorem 7.23
carries over verbatim to nonnegative measurable functions.
226 CHAPTER 7. PRODUCT MEASURES
We close this section with two remarks about the construction of product
measures in the non -finite case, where the story is considerably more subtle.
These remarks are not used elsewhere in this book and can safely be ignored.
Remark 7.26. Let (X, A, ) and (Y, B, ) be two arbitrary measure spaces.
In [4, Chapter 251] Fremlin defines the function : 2XY [0, ] by
( )
X An A, Bn B for n N
(W ) := inf (An ) (Bn ) (7.21)
and W
S
n=1 n=1 An Bn
If the measure spaces (X, A, ) and (Y, B, ) are both -finite then the
measures 0 and 1 agree and are equal to the completion of the product
measure on A B (see [4, Proposition 251K]).
Remark 7.27. For topological spaces yet another approach to the product
measure is based on the Riesz Representation Theorem 3.15. Let (X, UX )
and (Y, UY ) be two locally compact Hausdorff spaces, denote by BX and BY
their Borel -algebras, and let X : BX [0, ] and Y : BY [0, ] be
Borel measures. Define : Cc (X Y ) R by
Z Z
(f ) := f (x, y) d(y) d(x)
X Y
Z Z (7.23)
= f (x, y) d(x) d(y)
Y X
for f Cc (X Y ). That the two integrals agree for every continuous function
with compact support follows from Fubinis Theorem 7.20 for finite measure
spaces. (To see this, observe that every compact set K X Y is contained
in the product of the compact sets KX := {x X | ({x} Y ) K 6= } and
KY := {y Y | (X {y}) K 6= }.) Since is a positive linear functional,
the Riesz Representation Theorem 3.15 asserts that there exists a unique
outer regular Borel measure 1 : BXY [0, ] that is inner regular on
open sets and a unique Radon measure 0 : BXY [0, ] such that
Z Z
(f ) = f d0 = f d1
XY XY
for all f Cc (X Y ). It turns out that in this situation the Borel -algebra
BXY is contained in the -algebra C 2XY of Remark 7.26 and
0 = 0 |BXY , 1 = 1 |BXY .
Recall from Lemma 7.6 that the product -algebra BX BY agrees with the
Borel -algebra BXY whenever one of the spaces X or Y is second countable.
If they are both second countable then so is the product space (X Y, UXY )
(Appendix B). In this case
0 = 1 = X Y
Proof. The assertion follows directly from (7.24) and Theorem 7.17.
For Lebesgue measurable functions f : Rn [0, ] the analogous state-
ment is considerably more subtle. In that case the function fx , respec-
tively f y , need not be Lebesgue measurable for all x, respectively all y.
However, they are Lebesgue measurable for almost all x Rk , respectively
almost all y R` , and the three integrals in (7.25) can still be defined and
agree. The key result that one needs to prove this is that the Lebesgue
measure on Rn = Rk R` is the completion of the product of the Lebesgue
measures on Rk and R` . Then the assertion follows from Exercise 7.25.
Theorem 7.29. Let k, ` N, define n := k + `, and identify Rn with the
product space Rk R` in the canonical way. Denote the completion of the
product space (Rk R` , Ak A` , mk m` ) by (Rk R` , (Ak A` ) , (mk m` ) ).
Then An = (Ak A` ) and mn = (mk m` ) .
Proof. Define
n o
Cn := [a1 , b1 ) [an , bn ) ai , bi R and ai < bi for i = 1, . . . , n
n
so that Cn Bn An 2R for all n. We prove the assertion in three steps.
7.4. FUBINI AND LEBESGUE 229
and
Step 2. Ak A` An .
We prove that
E Ak = E R` An . (7.26)
To see this, fix a set E Ak . Then there exist Borel sets A, B Bk such that
A E B and mk (B \ A) = 0. Let : Rn Rk denote the projection onto
the first k coordinates. This map is continuous and hence Borel measurable
by Theorem 1.20. Thus the sets A R` = 1 (A) and B R` = 1 (B) are
Borel sets in Rn . Moreover, by Step 1
mn ((B R` ) \ (A R` )) = mn ((B \ A) R` )
= (mk m` )((B \ A) R` )
= mk (B \ A) m` (R` )
= 0.
Since A R` E R` B R` it follows that E R` An . This
proves (7.26). A similar argument shows that
F A` = Rk F An .
Hence E F = (E R` ) (Rk F ) An for all E Ak and all F A` .
Thus Ak A` An and this proves Step 2.
Step 3. (Ak A` ) = An and (mk m` ) = mn .
Let A An . Then there are Borel sets B0 , B1 Bn such that B0 A B1
and mn (B1 \B0 ) = 0. By Step 1, B0 , B1 Ak A` and (mk m` )(B1 \B0 ) = 0.
Hence A (Ak A` ) and
(mk m` ) (A) = (mk m` )(B0 ) = mn (B0 ) = mn (A).
Thus we have proved that
An (Ak A` ) , (mk m` ) |An = mn .
Since Ak A` An by Step 2 it follows that
mn |Ak A` = (mk m` ) |Ak A` = mk m` .
Now let A (Ak A` ) . Then there are sets A0 , A1 Ak A` such that
A0 A A1 and (mk m` )(A1 \ A0 ) = 0. Hence A0 , A1 An by Step 2 and
mn (A1 \ A0 ) = 0. Since (Rn , An , mn ) is complete it follows that A \ A0 An
and so A = A0 (A \ A0 ) An . Hence An = (Ak A` ) . This proves Step 3
and Theorem 7.29.
7.5. CONVOLUTION 231
7.5 Convolution
An application of Fubinis Theorem is the convolution product on the space
of Lebesgue integrable functions on Euclidean space. Fix an integer n N
and let (Rn , A, m) be the Lebesgue measure space. The convolution of two
Lebesgue integrable functions f, g L1 (Rn ) is defined by
Z
(f g)(x) := f (x y)g(y) dm(y) for almost all x Rn .
Rn
Here the function Rn R : y 7 f (x y)g(y) is Lebesgue integrable for al-
most every x Rn and the resulting almost everywhere defined function f g
is again Lebesgue integrable. This is the content of Theorem 7.33. The con-
volution descends to a bilinear map : L1 (Rn ) L1 (Rn ) L1 (Rn ). This
map is associative and endows L1 (Rn ) with the structure of a Banach alge-
bra. Throughout we use the notation f g for two Lebesgue measurable
functions f, g : Rn R to mean that they agree almost everywhere with
respect to the Lebesgue measure.
232 CHAPTER 7. PRODUCT MEASURES
We prove (iii). By (ii) and Theorem 1.55 it suffices to assume that f and
g are Borel measurable. Now define F, G : R2n R and : R2n R2n by
F (x, y) := f (x y)g(y),
G(x, y) := f (x)g(y),
(x, y) := (x y, y)
F =G
is Borel measurable by Fubinis Theorem 7.28. Thus the set E(f, g) where
this function takes on the value is a Borel set. Moreover, the functions
F := max{F, 0}
are Borel measurable and so are the functions Fe : R2n [0, ) defined by
F (x, y), if x Rn \ E(f, g),
Fe (x, y) := for (x, y) R2n .
0, if x E(f, g),
Since Z Z
(f g)(x) = Fe+ (x, y) dm(y) Fe (x, y) dm(y)
Rn Rn
for all x Rn it follows from Theorem 7.28 that f g is Borel measurable.
This proves (iii).
We prove (iv). Since gx f = (fx g) x it follows from Theorem 2.17 that
and
Z Z Z
(f g)(x) = fx g dm = (fx g) x dm = gx f dm = (g f )(x)
Rn Rn Rn
The integral on the right is infinite if and only if x E(|f |, |g| |h|). This
proves the first equivalence in (7.32). The proof of the second equivalence is
analogous with y and z interchanged.
Now let x Rn \ E. Then Fx L1 (R2n ) and x Rn \ E(f, g h).
Moreover, for z Rn , the function Rn R : y 7 Fx (y, z) is integrable if
and only x z / E(g, h) and in that case its integral is equal to
Z Z
Fx (y, z) dm(y) = f (z) g(x y z)h(y) dm(y) = f (z)(g h)(x z).
Rn Rn
Since E(g, h) is a Lebesgue null set, it follows from Theorem 7.30 that
Z Z
Fx dm2n = f (z)(g h)(x z) dm(z) = (f (g h))(x)
R2n Rn
R x
The last equation holds because / E(f, g h). A similar argument with y
and z interchanged shows that R2n Fx dm2n = ((f g)h)(x) for all x Rn \E.
This proves (v) and Theorem 7.32.
7.5. CONVOLUTION 235
Theorem 7.33. Let 1 p, q, r such that 1/p + 1/q = 1 + 1/r and let
f Lp (Rn ) and g Lq (Rn ). Then m(E(f, g)) = 0 and
kf gkr kf kp kgkq . (7.33)
Thus f g Lr (Rn ). The estimate (7.33) is called Youngs inequality.
Proof. Define the function h : Rn [0, ] by
Z
h(x) := |f (x y)g(y)| dm(y) for x Rn .
Rn
Then |f g| h and E(f, g) = {x Rn | h(x) = }. Hence it suffices to
prove that khkr kf kp kgkq . For r = this follows from Holders inequality.
So assume r < . Then 1 p, q < . Define
p p p
:= 1 = p , q 0 := .
r q
0
Then 0 < 1 and 1/q + 1/q = 1. Also = 0 if and only if q = 1. If > 0
then Holders inequality in Theorem 4.1 shows that
Z
h(x) = |fx | |fx |1 |g| dm
|fx |
q0
|fx |1 |g|
q ,
Rn
where fx (y) := f (x y). Since q 0 = p this implies
Z q/q0 Z
q q 0
h(x) |fx | dm |fx |(1)q |g|q dm
n n
R
Z R (7.34)
q
= kf kp |f (x y)|(1)q |g(y)|q dm(y)
Rn
n
for all x R . This continues to hold for = 0. Now it follows from
Minkowskis inequality in Theorem 7.19 with the exponent s := r/q 1 that
Z q/r Z 1/s
q r qs
khkr = h dm = h dm
Rn Rn
Z Z s 1/s
kf kq
p |f (x y)| (1)q q
|g(y)| dm(y) dm(x)
Rn Rn
Z Z 1/s
kf kq
p |f (x y)| (1)qs qs
|g(y)| dm(x) dm(y)
RnRn
q (1)q
= kf kp kf kp kgkqq .
Here the last equation follows from the fact that (1 )qs = (1 )r = p.
This proves Theorem 7.33.
236 CHAPTER 7. PRODUCT MEASURES
It follows from Theorem 7.33 and part (ii) of Theorem 7.32 that the
convolution descends to a map
Proof. We prove (i). Assume first that f Lp (Rn ) and fix a constant > 0.
By Theorem 4.15 there is a function g Cc (Rn ) such that kf gkp < 1/p /3.
Since g is uniformly continuous there is a > 0 such that, for all Rn ,
1/p
|| < = sup |g(x + ) g(x)| <
xRn 3p m(supp(g)
Take Rn such that || < . Then
Z 1/p
p
|f (x + ) f (x)| dm(x)
Rn
Z 1/p
p
2 kf gkp + |g(x + ) g(x)| dm(x)
Rn
1/p
21/p
p
+ m(supp(g)) sup |g(x + ) g(x)| < 1/p .
3 xRn
This proves (i) for f Lp (Rn ). To prove the result in general choose a
compact set K Rn such that B1 (x) K for all x B and multiply f by a
smooth compactly supported cutoff function to obtain a function f 0 Lp (Rn )
that agrees with f on K. Then (i) holds for f 0 and hence also for f .
We prove (ii). Assume first that f Lp (Rn ) and g Lq (Rn ) and fix a
constant > 0. By part (i) there exists a > 0 such that, for all Rn ,
Z !p
p
|| < = |f (y + ) f (y)| dm(y) <
Rn kgkq
Fix two elements x, Rn such that || < and denote fx (y) := f (x y).
Then, by Holders inequality in Theorem 4.1,
Z
|(f g)(x + ) (f g)(x)| = (fx+ fx )g dm
n R
kfx+ fx kp kgkq
Z 1/p
p
= |f (y + ) f (y)| dm(y) kgkq
Rn
< .
This shows that f g is uniformly continuous. If f is locally p-integrable and
g Lq (Rn ) has compact support continuity follows by taking the integral
over a suitable compact set. In the converse case continuity follows by taking
the Lq -norm of g over a suitable compact set. This proves (ii).
238 CHAPTER 7. PRODUCT MEASURES
Define : Rn R by
1 x
(x) := n
for > 0 and x Rn . Then
Z
supp( ) B , dm = 1
Rn
Proof. The inequality (7.36) was established in the proof of Lemma 6.2. We
prove (7.37) in four steps.
R
Step 1. tp f (t) X |f |p d for all t 0.
R R
Since tp A(t,f ) |f |p it follows that tp f (t) = X tp (A(t,f ) d X |f |p d for
all t 0. This proves Step 1.
R R
Step 2. If f (t) = for some t > 0 then X |f |p d = = 0 tp1 f (t) dt.
R
By Step 1, we have X |f |p d p1
R =p1. Moreover, t f (t) = for t > 0
sufficiently small and hence 0 t f (t) dt = . This proves Step 2.
Step 3. Assume (X, A, ) is -finite and f (t) < for all t > 0. Then
equation (7.37) holds.
Let B 2[0,) be the Borel -algebra and denote by m : B [0, ] the
restriction of the Lebesgue measure to B. Let (X [0, ), A B, m) be
the product measure space of Definition 7.10. We prove that
Q(f ) := (x, t) X [0, ) 0 t < |f (x)| A B.
To see this, assume first that f is an A-measurable step-function. Then
there exist finitely many pairwise disjoint measurable P sets A1 , . . . , A` A
`
and positive real numbers 1 , . . . , ` such that |f | = i=1 i Ai . In this
S`
case Q(f ) = i=1 Ai [0, i ) A B. Now consider the general case. Then
Theorem 1.26 asserts that there is a sequence of A-measurable step-functions
fi : X [0, ) such that 0 f1 f2 andSfi converges pointwise
to |f |. Then Q(fi ) A B for all i and so Q(f ) = i=1 Q(fi ) A B.
Now define h : X [0, ) [0, ) by h(x, t) := ptp1 . This function is
A B-measurable and so is hQ(f ) . Hence, by Fubinis Theorem 7.17,
!
Z Z Z |f (x)|
|f |p d = ptp1 dt d(x)
X X 0
Z Z
= (hQ(f ) )(x, t) dm(t) d(x)
X 0
Z Z
= (hQ(f ) )(x, t) d(x) dm(t)
0 X
Z
= ptp1 (A(t, f )) dt.
0
Step 4. Assume f (t) < for all t > 0. Then (7.37) holds.
Define X0 := {x X | f (x) 6= 0}, A0 := A A A X0 , and 0 := |A0 .
Then the measure space (X0 , A0 , 0 ) is -finite because Xn := A(1/n, f ) is
a sequence S of An -measurable sets such that 0 (Xn ) = f (1/n) < for all n
and X0 = n=1 Xn . Moreover, f0 := f |X0 : X0 R is A0 -measurable and
f = f0 . Hence it follows from Step 3 that
Z Z Z Z
p p p1
|f | d = |f0 | d0 = t f0 (t) dt = tp1 f (t) dt.
X X0 0 0
L1 () Lq () Lp ().
Here the last equation follows with the choice of c := (2c1 )1/(q1) /(2cq )q/(q1) .
This proves Theorem 7.37.
f = K f. (7.45)
u = f, i u = Ki f for i = 1, . . . , n. (7.47)
By Theorem 7.35 the space C0 (Rn ) is dense in L2 (Rn ). Thus Lemma 7.44
shows that the linear operator f 7 k (Kj f ) extends uniquely to a bounded
linear operator from L2 (Rn ) to L2 (Rn ). The heart of the proof of the
CalderonZygmund inequality is the following delicate argument which shows
that this operator also extends to a continuous linear operator from the Ba-
nach space L1 (Rn ) to the topological vector space L1, (Rn ) of weakly inte-
grable functions introduced in Section 6.1. This argument occupies the next
six pages. Recall the definition
kf k1, := sup tf (t),
t>0
where
A(t, f ) := x Rn |f (x)| > t .
f (t) := m(A(t, f )),
(See equation (6.1).)
Lemma 7.45. Fix an integer n 2. Then there is a constant c = c(n) > 0
such that
kk (Kj f )k1, c kf k1 (7.54)
for all f C0 (Rn ) and all indices j, k = 1, . . . , n.
Proof. Fix two integers j, k {1, . . . , n} and let T : L2 (Rn ) L2 (Rn ) be
the unique bounded linear operator that satisfies
T f = k (Kj f ) (7.55)
for f C0 (Rn ). This operator is well defined by Lemma 7.44. We prove in
three steps that there is a constant c = c(n) > 0 such that kT f k1, c kf k1
for all f L1 (Rn )L2 (Rn ). Throughout we abuse notation and use the same
letter f to denote a function in L2 (Rn ) and its equivalence class in L2 (Rn ).
Step 1. There is a constant c = c(n) 1 with the following significance. If
B Rn is a countable union of closed cubes Qi Rn with pairwise disjoint
interiors and if h L2 (Rn ) L1 (Rn ) satisfies
Z
h|Rn \B 0, h dm = 0 for all i N (7.56)
Qi
then
1
T h (t) c m(B) + khk1 (7.57)
t
for all t > 0.
7.7. THE CALDERONZYGMUND INEQUALITY 249
For i N define hi : Rn R by
h(x), if x Qi ,
hi (x) :=
0, if x
/ Qi .
Denote by qi Qi the center of the cube Qi and by 2ri > 0 its side length.
Then |x qi | nri for all x Qi . Fix an element x Rn \ Qi . Then Kj
is smooth on x Qi and so Theorem 7.35 asserts that
(T hi )(x) = (k Kj hi )(x)
Z
(7.58)
= k Kj (x y) k Kj (x qi ) hi (y) dm(y).
Qi
This identity is more delicate than it looks at first glance. To see this, note
that the formula (7.55) only holds for compactly supported smooth func-
tions but is not meaningful for all L2 functions f because Kj f may not be
differentiable. The function hi is not smooth so care must be taken. Since
x / Qi = supp(hi ) one can approximate hi in L2 (Rn ) by a sequence of com-
pactly supported smooth functions that vanish near x (by using the mollifier
method in the proof of Theorem 7.35). For the approximating sequence
part (iii) of Theorem 7.35 asserts that the partial derivative with respect to
the kth variable of the convolution with Kj is equal to the convolution with
k Kj near x. Now the first equation in (7.58) follows by taking the limit.
The second equation follows from (7.56). It follows from (7.58) that
Z
|(T hi )(x)| |k Kj (x y) k Kj (x qi )||hi (y)| dm(y)
Qi
sup |k Kj (x y) k Kj (x qi )| khi k1
yQi
nri sup |k Kj (x y)| khi k1
yQi
1
c1 ri sup khi k1
yQi |x y|n+1
c1 r i
= khi k1 .
d(x, Qi )n+1
Here d(x, Qi ) := inf yQi |x y| and
n(n + 3)
c1 = c1 (n) := max sup |y|n+1 |k Kj (y)| .
j,k yRn \{0} n
Here the last inequality follows by differentiating equation (7.43).
250 CHAPTER 7. PRODUCT MEASURES
Now define
Pi := x Rn |x qi | < 2 nri Qi .
Then d(x, Qi ) |x qi | nri for all x Rn \ Pi . Hence
Z Z
1
|T hi | dm c1 ri n+1 dm(x) khi k1
Rn \Pi Rn \Pi (|x qi | nri )
Z
1
= c1 r i n+1 dm(y) khi k1
|y|>2 nri (|y| nri )
Z
n sn1 ds
= c1 ri n+1 khi k1
2 nri (s nri )
Z n1
(s + nri ) ds
= c1 n ri n+1
khi k1
nri s
Z
n1 ds
c1 2 n ri 2
khi k1
nri s
= c2 khi k1 .
Here c2 = c2 (n) := c1 (n)2n1 n n 2n1 n3/2 (n + 3). The third step in
the above computation follows from Fubinis Theorem in polar coordinates
(Exercise 7.47). Thus we have proved that
Z
|T hi | dm c2 khi k1 for all i N. (7.59)
Rn \Pi
Now define
[
A := Pi .
i=1
Moreover,
X
X
m(A) m(Pi ) = c3 m(Qi ) = c3 m(B),
i=1 i=1
where
m(B2n )
c3 = c3 (n) := n
= m(Bn ) = n nn/21 .
m([1, 1] )
Hence
x Rn \ A |T h(x)| > t
tT h (t) tm(A) + tm
Z
tm(A) + |T h| dm
Rn \A
c3 tm(B) + c2 khk1
c4 tm(B) + khk1
and
|f (x)| t for almost all x Rn \ B, (7.62)
S
where B := i=1 Qi .
For Zn and ` Z define
Q(, `) := x Rn 2` i xi 2` (i + 1) .
Let
Q := Q(, `) Zn , ` Z
= f (i j g) dm
Rn
Z
= f (i (Kj g)) dm
Rn
kf kp ki (Kj g)kq
c(n, q) kf kp kgkq .
Since C0 (Rn ) is dense in Lq (Rn ) by Theorem 7.35, and the Lebesgue measure
is semi-finite, it follows from Lemma 4.34 that ki (Kj f )kp c(n, q) kf kp
for all f C0 (Rn ). This proves Theorem 7.46.
Proof of Theorem 7.43. Fix an integer n 2 and a number 1 < p < .
Let c = c(n, p) be the constant of Theorem 7.46 and let u C0 (Rn ). Then
j u = j (K u) = Kj u by Theorem 7.41. Hence it follow from Theo-
rem 7.46 with f = u that ki j ukp = ki (Kj u)kp c(n, p) kukp for
i, j = 1, . . . , n. This proves Theorem 7.43.
7.8. EXERCISES 255
7.8 Exercises
Exercise 7.47 (Lebesgue Measure on the Sphere).
For n N let (Rn , An , mn ) the Lebesgue measure space, denote the open
unit ball by B n := {x Rn | |n| < 1}, and the unit sphere by
S n1 := B n = {x Rn | |x| = 1} .
p
For A S n1 define A := {x B n1 | (x, 1 |x|2 ) A}. Prove that
the collection
AS := A S n1 | A+ , A An1
then f M and kf k = kf k1 .
(v) If f, g L1 (R) then
f g = f g .
(vi) Let , , M. Then = if and only if
Z Z
f d = f (x + y)d( )(x, y)
R R2
for all bounded Borel measurable functions f : R R.
(vii) If , M and B B then
Z
( )(B) = (B t) d(t).
R
258 CHAPTER 7. PRODUCT MEASURES
Chapter 8
The purpose of this last chapter is to prove the existence and uniqueness of a
normalized invariant Radon measure on any compact Hausdorff group. In the
case of a locally compact Hausdorff group the theorem asserts the existence
of a left invariant Radon measure that is unique up to a scaling factor. An
example is the Lebesgue measure on Rn . A useful exposition is the paper by
Gert K. Pedersen [16] which also discusses the original references.
G G : x 7 x1 . (8.2)
U 2G
such that the group multiplication (8.1) and the inverse map (8.2) are con-
tinuous. Here the continuity of the group multiplication (8.1) is understood
with respect to the product topology on G G (see Appendix B). A lo-
cally compact Hausdorff group is a topological group (G, U) such that
the topology is locally compact and Hausdorff (see page 81).
259
260 CHAPTER 8. THE HAAR MEASURE
Example 8.1. Let G be any group and define U := {, G}. Then (G, U) is
a compact topological group but is not Hausdorff unless G = {1l}.
Example 8.2 (Discrete Groups). Let G be any group. Then the pair
(G, U) with the discrete topology U := 2G is a locally compact Hausdorff
group, called a discrete group. Examples of discrete groups (where the
discrete topology appears naturally) are the additive group Zn , the multi-
plicative group SL(n, Z) of integer n n-matrices with determinant one, the
mapping class group of isotopy classes of diffeomorphisms of any manifold,
and every finite group.
Example 8.4. If (V, kk) is a normed vector space (Example 1.11) then the
additive group V is a Hausdorff topological group. It is locally compact if
and only if V is finite-dimensional.
Example 8.5. The rational numbers Q with the additive structure form a
Hausdorff topological group with the relative topology as a subset of R. It is
totally disconnected (every connected component is a single point) but does
not have the discrete topology. It is not locally compact.
8.1. TOPOLOGICAL GROUPS 261
N0 := N {0}
This set is a commutative ring with unit. Addition and multiplication are
defined term by term, i.e.
for x = (xk )kN0 Zp and y = (yk )kN0 Zp . The ring of p-adic integers is
a compact metric space with
n o
k
dp (x, y) := inf p k N0 , xk = yk . (8.4)
(f Lx ) = (f ) (8.12)
(f Rx ) = (f ) (8.13)
for all f Cc (G) and all x G. It is called invariant if it is both left and
right invariant. It is called a left Haar integral if it is left invariant, posi-
tive, and nonzero. It is called a right Haar integral if it is right invariant,
positive, and nonzero.
Theorem 8.14 (Haar). Every locally compact Hausdorff group G admits a
left Haar integral, unique up to a positive factor. Moreover, if : Cc (G) R
is a left Haar integral and f Cc (G) is a nonnegative function that does not
vanish identically then (f ) > 0.
Proof. See page 268.
The proof of Theorem 8.14 given below follows the notes of Pedersen [16]
which are based on a proof by Weil. Our exposition benefits from elegant
simplifications due to Urs Lang [11]. In preparation for the proof it is con-
venient to introduce some notation. Let
Cc+ (G) := f Cc (G) f 0, f 6 0
(8.14)
Lemma 8.15. Let : Cc+ (G) [0, ) be an additive functional. Then there
is a unique positive linear functional on Cc (G) whose restriction to Cc+ (G)
agrees with . If is left invariant then so is its linear extension to Cc (G).
Proof. We prove that is monotone. Let f, g Cc+ (G) such that f g. If
f 6= g then g f Cc+ (G) and hence
(f ) (f ) + (g f ) = (g)
for all a, b Q with 0 < a < c < b, and this implies (cf ) = c(f ).
Now define (0) := 0 and, for f Cc (G), define (f ) := (f + ) (f ).
If f, g Cc (G) then f + + g + + (f + g) = f + g + (f + g)+ , hence
for f Cc (G) and a G. The next two lemmas establish some basic
properties of the operators L and R . Denote by
kf k := sup|f (x)|
xG
L L = L f Lx = Lx f, kL f k kk kf k ,
R R = R , f Rx = Rx f, kR f k kk kf k ,
L R = R L .
Lemma 8.17. Let f, g Cc+ (G). Then there exists a P such that
f L g.
Then
k
X f (xi ) + 1
f (a) g(yx1
i a) = (L g)(a)
i=1
g(y)
for all a supp(f ) and hence f L g. This proves Lemma 8.17.
Proof of Theorem 8.14. The proof has five steps. Step 1 is the main con-
struction of a subadditive functional Mg : Cc+ (G) (0, ) associated to a
function g Cc+ (G). Step 2 shows that Mg is asymptotically linear as g
concentrates near the unit 1l. The heart of the convergence proof is Step 3
and is due to Cartan. Step 4 proves uniqueness and Step 5 proves existence.
Step 1. For f Cc+ (G) define
Mg (f ) := M (f ; g) := inf kk P, f L g .
(8.19)
We prove part (ii). Monotonicity follows directly from the definition. Homo-
geneity follows from the identities Lc g = cL g and kck = c kk. To prove
left invariance observe that
(L g) Lx = Lx g, k x k = kk
M (f + f 0 ; g) k + 0 k = kk + k0 k < M (f ; g) + M (f 0 ; g) + .
(L f ) kk (f ) (8.21)
k
X k
X k
X
(L f ) (i (f Lxi )) = i (f Lxi ) = i (f ) = kk (f )
i=1 i=1 i=1
Here the first step follows from subadditivity, the second step follows from
homogeneity, the third step follows from left invariance, and the last step
follows from the definition of kk. This proves (8.21). Now let f, g Cc+ (G).
By Lemma 8.17 there is a P such that f L g. Since is monotone
this implies (f ) (L g) kk (g) by (8.21). Now take the infimum
over all P such that f L g to obtain (f ) M (f ; g)(g).
We prove parts (i) and (iv). Since the map Cc+ (G) (0, ) : f 7 kf k
is an element of L it follows from part (iii) that
kf k M (f ; g) kgk (8.22)
and hence M (f ; g) > 0 for all f, g Cc+ (G). Next observe that M (f ; f ) 1
by (8.22) and M (f ; f ) 1 because f = L1l f . This proves Step 1.
270 CHAPTER 8. THE HAAR MEASURE
Step 2. Let f, f 0 Cc+ (G) and let > 0. Then there is an open neighborhood
U G of 1l such that every g Cc+ (G) with supp(g) U satisfies
Mg (f ) + Mg (f 0 ) < (1 + )Mg (f + f 0 ). (8.23)
By Urysohns Lemma A.1 there is a function Cc (G) such that (x) = 1
for all x supp(f ) supp(f 0 ). Choose a constant 0 < 1/2 such that
2 + 2 kf + f 0 k M (; f + f 0 ) < . (8.24)
Define
h := f + f 0 + kf + f 0 k .
Then f /h and f 0 /h extend to continuous functions on G with compact sup-
port by setting them equal to zero on G \ supp(). By Lemma 8.10 there
exists an open neighborhood U G of 1l such that
f (x) f (y) f 0 (x) f 0 (y)
1
x yU = h(x) h(y) + h(x) h(y) <
`
f (x1
i )
X
Mg (f ) i + .
i=1
h(x1
i )
`
f (x1 0 1
i ) + f (xi )
X
0
Mg (f ) + Mg (f ) i + 2 kk (1 + 2).
i=1
h(x1
i )
1 + 2 + 2 kf + f 0 k M (; f + f 0 ) Mg (f + f 0 )
(1 + )Mg (f + f 0 ).
Here we have used the inequalities 1 + 2 2 and (8.24). This proves Step 2.
8.2. HAAR MEASURES 271
Step 3. Let f Cc+ (G) and let > 0. Then there is an open neighborhood
U G of 1l with the following significance. For every g Cc+ (G) such that
supp(g) U, g(x) = g(x1 ) for all x G, (8.25)
there exists an open neighborhood W G of 1l such that every h Cc+ (G)
with supp(h) W satisfies the inequality
M (f ; g)Mh (g) (1 + )Mh (f ). (8.26)
This inequality continues to hold with Mh replaced by any left invariant pos-
itive linear functional : Cc (G) R.
By Urysohns Lemma A.1 there is a function Cc+ (G) such that (x) = 1
for all x K := supp(f ). Choose 0 and 1 such that
1 + 0 0
0 < 0 < 1, 1 + , 1 := . (8.27)
1 0 2M (; f )
By Lemma 8.10 there exists an open neighborhood U G of 1l such that
x1 y U = |f (x) f (y)| < 1 (8.28)
for all x, y G. We prove that the assertion of Step 3 holds with this
neighborhood U . Fix a function g Cc+ (G) that satisfies (8.25). Define
1
2 := . (8.29)
2M (f ; g)
Use Lemma 8.10 to find an open neighborhood V G of 1l such that
xy 1 V = |g(x) g(y)| < 2 (8.30)
for all x, y G. Then the sets xV for x K form an open coverSof K. Hence
there exist finitely many points x1 , . . . , x` K such that K `i=1 xi V . By
Theorem A.4 there exist functions 1 , . . . , ` Cc+ (G) such that
`
X
0 i 1, supp(i ) xi V, i |K 1. (8.31)
i=1
Here the second step uses (8.29) and the inequality Mh (f ) M (f ; g)Mh (g).
+ P
Thus (f 21 ) Mh (g) L g, where := i Mh (i f ) x1 i
. This implies
X
Mg ((f 21 )+ )Mh (g) Mh (i f ) (1 + 0 )Mh (f ).
i
for all g Cc+ (G) by (8.20). The same argument shows that 0 (g) > 0 for
all g Cc+ (G).
Now fix two functions f, f0 Cc+ (G) and a constant > 0. Choose an
open neighborhood U G of 1l that satisfies the requirements of Step 3 for
both f and f0 and this constant . By Urysohns Lemma A.1 there exists a
function g Cc+ (G) such that
supp(g) x G | x U and x1 U .
g(1l) > 0,
(f ) M (f ; g)(g) (1 + )(f )
and
(1 + )(f0 ) M (f0 ; g)(g) (f0 ).
Take the quotient of these inequalities to obtain
(f ) M (f ; g) (f )
(1 + )1 (1 + ) .
(f0 ) M (f0 ; g) (f0 )
2 (f ) 0 (f ) (f )
(1 + ) 0 (1 + )2 .
(f0 ) (f0 ) (f0 )
Since this holds for all f Cc+ (G) it follows that 0 and c agree on Cc+ (G).
Hence 0 = c by Lemma 8.15. This proves Step 4.
274 CHAPTER 8. THE HAAR MEASURE
L (f ) = L (f Lx )
(1 + )1 (f + f 0 ) (f ) + (f 0 ) (1 + ) (f + f 0 ) (8.41)
for all > 0. To prove the first inequality in (8.41) choose any functional
L (f + f 0 ). Then there exists an open neighborhood W G of 1l such
that (f + f 0 ) (1 + )h (f + f 0 ) for all h Cc+ (G) with supp(h) W .
Moreover, we have seen above that h Cc+ (G) can be chosen such that
supp(h) W and also h L (f ) L (f 0 ). Any such h satisfies
(1 + )1 (f + f 0 ) h (f + f 0 ) h (f ) + h (f 0 ) (f ) + (f 0 ).
(1 + )1 (f ) + 0 (f 0 ) h (f ) + h (f 0 ) h (f + f 0 ) (f + f 0 ).
is a directed set equipped with a map g 7 g that takes values in the space
M (f0 ; f )1 (f ) M (f0 ; f )
+
L := : Cc (G) R .
for all f Cc+ (G)
The map G L : g 7 g is a net. A net can be thought of as an uncountable
analogue of a sequence and a subnet as an analogue of a subsequence. The
existence of a universal subnet is guaranteed by the general theory and its
convergence for each f by the fact that the target space is compact. Instead
Step 3 in the proof of Theorem 8.14 implies that the original net g 7 g
converges and so there is no need to choose a universal subnet. That this
can be proved with a refinement of the uniqueness argument ( in Step 3) is
pointed out in Pedersen [16]. That paper also contains two further uniqueness
proofs. One is based on Fubinis Theorem and the other on the Radon
Nikodym Theorem. Another existence proof for compact second countable
Hausdorff groups is due to Pontryagin. It uses the Arzel`aAscoli theorem to
establish the existence of a sequence i P with ki k = 1 such that Li f
converges to a constant function whose value is then taken to be (f ).
Proof of Theorem 8.12. Existence and uniqueness in (i) follow directly from
Theorem 8.14 and the Riesz Representation theorem 3.15. That nonempty
open sets have positive measure follows from Urysohns Lemma A.1. To
prove (ii) consider the map : G G defined by (x) := x1 for x G.
Since is a homeomorphism it preserves the Borel -algebra B. Since
Rx = Lx1 , a measure : B [0, ] is a left Haar measure if and
only if the measure : B [0, ] defined by (B) := ((B)) = (B 1 ) is
a right Haar measure. Hence assertion (ii) follows from (i).
We prove (iii). Assume G is compact and let : B [0, 1] be the unique
left Haar measure such that (G) = 1. For x G define x : B [0, 1]
by x (B) := (Rx (B)) for B B. Since Rx commutes with Ly for all y
by (8.11), x is a left Haar measure. Since x (G) = (Rx (G)) = (G) = 1
it follows that x = for all x G. Hence is right invariant. Therefore
the map B [0, 1] : B 7 (B) := ((B)) = (B 1 ) is a left Haar measure
and, since (G) = 1, it agrees with . This proves Theorem 8.12.
8.2. HAAR MEASURES 277
In the noncompact case the left and right Haar measures need not agree.
The above argument then shows that the measure x differs from by a
positive factor. Thus there exists a unique map : G (0, ) such that
(Rx (B)) = (x)(B) (8.42)
for all x G and all B B. The map : G (0, ) in (8.42) is a continuous
group homomorphism, called the modular character. It is independent of
the choice of . A locally compact Hausdorff group is called unimodular
iff its modular character is trivial or, equivalently, iff its left and right Haar
measures agree. Thus every compact Hausdorff group is unimodular.
Exercise 8.18. Prove that is a continuous homomorphism.
Exercise 8.19. Prove that the group of all real 2 2-matrices of the form
a b
, a, b R, a > 0,
0 1
is not unimodular. Prove that the additive group Rn is unimodular. Prove
that every discrete group is unimodular.
Exercise 8.20. Let : B [0, ] be a right Haar measure. Show that the
modular character is characterized by the condition (Lx1 (B)) = (x)(B)
for all x G and all B B.
Haar measures are extremely useful tools in geometry, especially when the
group in question is compact. For example, if a compact Hausdorff group G
acts on a topological space X continuously via
G X X : (g, x) 7 g x, (8.43)
one can use the Haar measure to produce G-invariant continuous functions
by averaging. Namely, if f : X R is any continuous function, and is the
Haar measure on G with (G) = 1, then the function F : X R defined by
Z
F (x) := f (a x) d(a) (8.44)
G
Exercise 8.22. Show that the map F in (8.44) is continuous and G-invariant.
Exercise 8.24. Show that the Haar measure on a discrete group is a multiple
of the counting measure. Deduce that for a finite group the formula (8.44)
defines F (x) as the average (with multiplicities) of the values of f over the
group orbit of x.
Urysohns Lemma
K U.
f : X [0, 1]
such that
f |K 1, supp(f ) = x X f (x) 6= 0 U. (A.1)
279
280 APPENDIX A. URYSOHNS LEMMA
U := U X | U is open and y
/U .
K U := U1 Un .
K V V U. (A.2)
Proof. We first prove the assertion in the case where K = {x} consist of
a single element. Choose a compact neighborhood B X of x. Then
F := B \ U is a closed subset of B and hence is compact by part (i) of
Lemma A.2. Since x / F it follows from part (ii) of Lemma A.2 that there
exist open sets W, W 0 X such that x W , F W 0 and W W 0 = .
Hence V := int(B) W is an open neighborhood of x, its closure is a closed
subset of B and hence compact, and
V B W B \ W 0 B \ F U.
Proof of Theorem A.1. The proof has three steps. The first step requires the
Axiom of Dependent Choice.
Step 1. There exists a family of open sets Vr X with compact closure,
parametrized by r Q [0, 1], such that
K V1 V 1 V0 V 0 U (A.3)
and
s>r = V s Vr (A.4)
for all r, s Q [0, 1].
The existence of open sets V0 and V1 with compact closure satisfying (A.3)
follows from Lemma A.3. Now choose a bijective map N0 Q [0, 1] : i 7 qi
such that q0 = 0 and q1 = 1. Suppose by induction that the open sets
Vi = Vqi have been constructed for i = 0, 1, . . . , n such that (A.4) holds for
r, s {q0 , q1 , . . . , qn }. Choose k, ` {0, 1, . . . , n} such that
qk := max {qi | 0 i n, qi < qn+1 } ,
q` := min {qi | 0 i n, qi > qn+1 } .
Then V ` Vk . Hence it follows from Lemma A.3 that there exists an open set
Vn+1 = Vqn+1 X with compact closure such that V ` Vn+1 V n+1 Vk .
This completes the induction argument and Step 1 then follows from the
Axiom of Dependent Choice. (Denote by V the set of all open sets V X
such that K V V U . Denote by V the set of all finite sequences
v = (V0 , . . . , Vn ) in V that satisfy (A.3) and qi < qj = V j Vi for all i, j.
Define a relation on V by v = (V1 , . . . , Vn ) v0 = (V10 , . . . , Vn00 ) iff n < n0 and
Vi = Vi0 for i = 0, . . . , n. Then V is nonempty and for every v V there is
a v0 V such that v v0 . Hence, by the Axiom of Dependent Choice, there
exists a sequence vj = (Vj,0 , . . . , Vj,nj ) V such that vj vj+1 for all j N.
Define the map Q [0, 1] V : q 7 Vq by Vqi := Vj,i for i, j N with nj i.
This map is well and satisfies (A.3) and (A.4) by definition of V and .)
Step 2. Let Vr X be as in Step 1 for r Q [0, 1]. Then
f (x) := sup {r Q [0, 1] | x Vr }
(A.5)
= inf s Q [0, 1] | x / Vs
for all x X. (Here the supremum of the empty set is zero and the infimum
over the empty set is one.)
282 APPENDIX A. URYSOHNS LEMMA
Let (X, UX ) and (Y, UY ) be topological spaces, denote the product space by
X Y := (x, y) x X, y Y ,
and let X : X Y X and Y : X Y Y be the projections onto
the first and second factor. Consider the following universality property for
a topology U 2XY on the product space.
(P) Let (Z, UZ ) be any topological space and let h : Z X Y be any map.
Then h : (Z, UZ ) (X Y, U) is continuous if and only if the maps
f := X h : (Z, UZ ) (X, UX ),
(B.1)
g := Y h : (Z, UZ ) (Y, UY )
are continuous.
Theorem B.1. (i) There is a unique topology U on X Y that satisfies (P).
(ii) Let U 2XY be as in (i). Then W U if and only if there
S are open sets
Ui UX and Vi UY , indexed by any set I, such that W = iI (Ui Vi ).
(iii) Let U 2XY be as in (i). Then U is the smallest topology on X Y
with respect to which the projections X and Y are continuous.
(iv) Let U 2XY be as in (i). Then the inclusion
x : (Y, UY ) (X Y, U), x (y) := (x, y) for y Y,
is continuous for every x X and the inclusion
y : (X, UX ) (X Y, U), y (x) := (x, y) for x X,
is continuous for every y Y .
285
286 APPENDIX B. THE PRODUCT TOPOLOGY
This appendix contains a proof of the inverse function theorem. The result
is formulated in the setting of continuously differentiable maps between open
sets in a Banach space. Readers who are unfamiliar with bounded linear
operators on Banach spaces may simply think of continuously differentiable
maps between open sets in finite-dimensional normed vector spaces. The in-
verse function theorem is used on page 71 in the proof of Lemma 2.19, which
is a key step in the proof of the transformation formula for the Lebesgue mea-
sure (Theorem 2.17). Assume throughout that (X, kk) is a Banach space.
When : X X is a bounded linear operator denote its operator norm by
kxk
kk := kkL(X) := sup .
xX\{0} kxk
Then
B(1)r ((x0 )) (Br (x0 )) B(1+)r ((x0 )). (C.2)
Moreover, the map is injective, its image is open, the map 1 is contin-
uously differentiable, and d 1 (y) = d( 1 (y))1 for all y (Br (x0 )).
287
288 APPENDIX C. THE INVERSE FUNCTION THEOREM
1
k(x) (x0 ) (x x0 )k
1
(1 ) kx x0 k
ky y0 k .
Hence 1 is differentiable at y0 and d 1 (y0 ) = 1 = d( 1 (y0 ))1 . Thus
d 1 is continuous by Step 1. This proves Theorem C.1.
Bibliography
[1] R.B. Ash, Basic Probability Theory, John Wiley & Sons 1970, Dover Publications, 2008.
http://www.math.uiuc.edu/~r-ash/BPT/BPT.pdf
[2] V.I. Bogachev, Measure Theory, Volumes 1 & 2, Springer Verlag, 2006.
[3] W.F. Eberlein, Notes on integration I: The underlying convergence theorem. Comm. Pure Appl.
Math. 10 (1957), 357360.
[4] D.H. Fremlin, Measure Theory, Volumes 15, 2000.
https://www.essex.ac.uk/maths/people/fremlin/mt.htm
[5] D. Gilbarg and N.S. Trudinger, Elliptic Partial Differential Equations of the Second Order, Springer
Verlag, 1983.
[6] P.R. Halmos, Measure Theory, Springer-Verlag, 1974.
[7] C.E. Heil, Lecture Notes on Real Analysis, Georgia Institute of Technology, Fall 2007.
http://people.math.gatech.edu/~heil/6337/fall07/section3.4.pdf
http://people.math.gatech.edu/~heil/6337/fall07/section3.5b.pdf
[8] H. Konig, New versions of the RadonNikod ym theorem. Arch. Math. 86 (2006), 251260.
http://www.math.uni-sb.de/service/preprints/preprint132.pdf
[9] T.W. K orner, A Companion to Analysis. Graduate Studies in Mathematics 62, AMS 2004. https:
//www.dpmms.cam.ac.uk/~twk/
[10] U. Lang, Mass & Integral, Vorlesungsmanuskript, ETH Z urich, September 2006.
http://www.math.ethz.ch/~lang/mi.pdf
[11] U. Lang, Existence and uniqueness of Haar integrals. ETH Z urich, January 2015.
http://www.math.ethz.ch/~lang/HaarIntegral.pdf
[12] Dan Ma, The Lexicographic Order and The Double Arrow Space. 7 Oct. 2009.
http://dantopology.wordpress.com/2009/10/07/
[13] P. Malliavin, Integration and Probability (in cooperation with H el`
ene Airault, Leslie Kay, G
erard
Letac), Springer-Verlag, 1995.
[14] J.R. Munkres, Topology, Second Edition. Prentice Hall, 2000.
[15] G. Nagy, Real Analysis IV.6, Duals of Lp spaces.
http://www.math.ksu.edu/~nagy/real-an/4-06-dual-lp.pdf
[16] G.K. Pedersen, The existence and uniqueness of the Haar integral on a locally compact topological
group. Preprint, University of Copenhagen, November 2000.
http://www.math.ku.dk/kurser/2004-2/mat3re/haarintegral.pdf
[17] W. Rudin, Real and Complex Analysis, 3rd Ed. McGrawHill, 1987.
[18] D.A. Salamon, Das Riemannsche Integral. Lecture Notes, ETHZ, 2014.
http://www.math.ethz.ch/~salamon/PREPRINTS/ana1-int.pdf
[19] D.A. Salamon, Mehrfache Integrale. Lecture notes, ETHZ, 2014.
http://www.math.ethz.ch/~salamon/PREPRINTS/ana2-int.pdf
[20] D.A. Salamon, Funktionentheorie. Birkhauser, 2012.
http://www.math.ethz.ch/~salamon/PREPRINTS/cxana.pdf
[21] W.F. Trench, Introduction to Real Analysis, Prentice Hall, 2003.
http://ramanujan.math.trinity.edu/wtrench/misc/index.shtml
289
Index
290
INDEX 291