Measure and Integration Notes
Measure and Integration Notes
John K. Hunter
Department of Mathematics, University of California at Davis
Contents
Chapter 1. Measures
1.1. Sets
1.2. Topological spaces
1.3. Extended real numbers
1.4. Outer measures
1.5. -algebras
1.6. Measures
1.7. Sets of measure zero
1
1
2
2
3
4
5
6
9
10
12
14
18
19
20
22
27
30
33
33
34
36
37
38
Chapter 4. Integration
4.1. Simple functions
4.2. Positive functions
4.3. Measurable functions
4.4. Absolute continuity
4.5. Convergence theorems
4.6. Complex-valued functions and a.e. convergence
4.7. L1 spaces
4.8. Riemann integral
4.9. Integrals of vector-valued functions
39
39
40
42
45
47
50
50
52
52
55
55
iii
iv
CONTENTS
5.2.
5.3.
5.4.
5.5.
5.6.
5.7.
Premeasures
Product measures
Measurable functions
Monotone class theorem
Fubinis theorem
Completion of product measures
56
58
60
61
61
61
Chapter 6. Differentiation
6.1. A covering lemma
6.2. Maximal functions
6.3. Weak-L1 spaces
6.4. Hardy-Littlewood theorem
6.5. Lebesgue differentiation theorem
6.6. Signed measures
6.7. Hahn and Jordan decompositions
6.8. Radon-Nikodym theorem
6.9. Complex measures
63
64
65
67
67
68
70
71
74
77
Chapter 7. Lp spaces
7.1. Lp spaces
7.2. Minkowski and H
older inequalities
7.3. Density
7.4. Completeness
7.5. Duality
79
79
80
81
81
83
Bibliography
89
CHAPTER 1
Measures
Measures are a generalization of volume; the fundamental example is Lebesgue
measure on Rn , which we discuss in detail in the next Chapter. Moreover, as
formalized by Kolmogorov (1933), measure theory provides the foundation of probability. Measures are important not only because of their intrinsic geometrical and
probabilistic significance, but because they allow us to define integrals.
This connection, in fact, goes in both directions: we can define an integral
in terms of a measure; or, in the Daniell-Stone approach, we can start with an
integral (a linear functional acting on functions) and use it to define a measure. In
probability theory, this corresponds to taking the expectation of random variables
as the fundamental concept from which the probability of events is derived.
In these notes, we develop the theory of measures first, and then define integrals.
This is (arguably) the more concrete and natural approach; it is also (unarguably)
the original approach of Lebesgue. We begin, in this Chapter, with some preliminary definitions and terminology related to measures on arbitrary sets. See Folland
[4] for further discussion.
1.1. Sets
We use standard definitions and notations from set theory and will assume the
axiom of choice when needed. The words collection and family are synonymous
with set we use them when talking about sets of sets. We denote the collection
of subsets, or power set, of a set X by P(X). The notation 2X is also used.
If E X and the set X is understood, we denote the complement of E in X
by E c = X \ E. De Morgans laws state that
!c
!c
\
[
[
\
c
=
E ,
E
=
Ec .
E
I
1. MEASURES
is open;
(c) if {Ui T : i = 1, 2, . . . , N } is a finite collection of open sets, then their
intersection
N
\
Ui T
i=1
is open.
The complement of an open set in X is called a closed set, and T is called a topology
on X.
1.3. Extended real numbers
It is convenient to use the extended real numbers
R = {} R {}.
This allows us, for example, to talk about sets with infinite measure or non-negative
functions with infinite integral. The extended real numbers are totally ordered in
the obvious way: is the largest element, is the smallest element, and real
numbers are ordered as in R. Algebraic operations on R are defined when they are
unambiguous e.g. + x = for every x R except x = , but is
undefined.
We define a topology on R in a natural way, making R homeomorphic to a
compact interval. For example, the function : R [1, 1] defined by
if x =
1
(x) =
x/ 1 + x2 if < x <
1
if x =
is a homeomorphism.
A primary reason to use the extended real numbers is that upper and lower
bounds always exist. Every subset of R has a supremum (equal to if the subset
contains or is not bounded from above in R) and infimum (equal to if the
subset contains or is not bounded from below in R). Every increasing sequence
of extended real numbers converges to its supremum, and every decreasing sequence
converges to its infimum. Similarly, if {an } is a sequence of extended real-numbers
then
lim inf an = sup inf ai
lim sup an = inf sup ai ,
n
nN
in
nN
in
P
Every sum i=1 xi with non-negative terms xi 0 converges in R (to if
xi = for some i N or the series diverges in R), where the sum is defined by
(
)
X
X
xi = sup
xi : F N is finite .
i=1
iF
(xi + yi ) =
xi +
i=1
i=1
yi ;
i=1
X
X
xij : F N N is finite
xij = sup
i,j=1
(i,j)F
X
X
xij
=
=
i=1
j=1
j=1
i=1
xij
Our use of extended real numbers is closely tied to the order and monotonicity
properties of R. In dealing with complex numbers or elements of a vector space,
we will always require that they are strictly finite.
1.4. Outer measures
As stated in the following definition, an outer measure is a monotone, countably
subadditive, non-negative, extended real-valued function defined on all subsets of
a set.
Definition 1.2. An outer measure on a set X is a function
: P(X) [0, ]
such that:
(a) () = 0;
(b) if E F X, then (E) (F );
(c) if {Ei X : i N} is a countable collection of subsets of X, then
!
X
[
(Ei ).
Ei
i=1
i=1
1. MEASURES
1.5. -algebras
A -algebra on a set X is a collection of subsets of a set X that contains and
X, and is closed under complements, finite unions, countable unions, and countable
intersections.
Definition 1.3. A -algebra on a set X is a collection A of subsets of X such
that:
(a) , X A;
(b) if A A then Ac A;
(c) if Ai A for i N then
\
[
Ai A.
Ai A,
i=1
i=1
Among the most important -algebras are the Borel -algebras on topological
spaces.
Definition 1.8. Let (X, T ) be a topological space. The Borel -algebra
B(X) = (T )
1.6. MEASURES
1.6. Measures
A measure is a countably additive, non-negative, extended real-valued function
defined on a -algebra.
Definition 1.9. A measure on a measurable space (X, A) is a function
: A [0, ]
such that
(a) () = 0;
(b) if {Ai A : i N} is a countable disjoint collection of sets in A, then
!
X
[
(Ai ).
Ai =
i=1
i=1
|E (A E) = (A E).
As we will see, the construction of nontrivial measures, such as Lebesgue measure, requires considerable effort. Nevertheless, there is at least one useful example
of a measure that is simple to define.
Example 1.11. Let X be an arbitrary non-empty set. Define : P(X)
[0, ] by
(E) = number of elements in E,
where () = 0 and (E) = if E is not finite. Then is a measure, called counting measure on X. Every subset of X is measurable with respect to . Counting
measure is finite if X is finite and -finite if X is countable.
A useful implication of the countable additivity of a measure is the following
monotonicity result.
Proposition 1.12. If {Ai : i N} is an increasing sequence of measurable
sets, meaning that Ai+1 Ai , then
!
[
Ai = lim (Ai ).
(1.1)
i=1
\
Ai = lim (Ai ).
(1.2)
i=1
1. MEASURES
X
[
[
(Bi ) .
Bi =
Ai =
Moreover, since Aj =
Sj
i=1
i=1
i=1
i=1
Bi ,
(Aj ) =
j
X
(Bi ) ,
i=1
i=1
[
Bi = lim (Bi ) = (A1 ) lim (Ai ).
i=1
Since
i=1
Bi = A1 \
i=1
Ai ,
i=1
Bi
= (A1 )
i=1
Ai
,
Example 1.13. To illustrate the necessity of the condition (A1 ) < in the
second part of the previous proposition, or more generally (An ) < for some
n N, consider counting measure : P(N) [0, ] on N. If
An = {k N : k n},
\
\
An = ,
An = 0.
n=1
n=1
Note that completeness depends on the measure , not just the -algebra
A. Any measure space (X, A, ) is contained in a uniquely defined completion
(X, A, ), which the smallest complete measure space that contains it and is given
explicitly as follows.
Theorem 1.15. If (X, A, ) is a measure space, define (X, A, ) by
A = {A M : A A, M N where N A satisfies (N ) = 0}
(A M )c = Ac M c ,
M c = N c (N \ M ).
(A M )c = (Ac N c ) (Ac (N \ M )) A,
since Ac N c A and Ac (N \ M ) N . Moreover, A is closed under countable
unions because if Ai A and Mi Ni where (Ni ) = 0 for each i N, then
!
!
[
[
[
Mi A,
Ai
Ai Mi =
i=1
Ai A,
i=1
i=1
i=1
since
i=1
Mi
i=1
Ni ,
i=1
Ni
= 0.
CHAPTER 2
Lebesgue Measure on Rn
Our goal is to construct a notion of the volume, or Lebesgue measure, of rather
general subsets of Rn that reduces to the usual volume of elementary geometrical
sets such as cubes or rectangles.
If L(Rn ) denotes the collection of Lebesgue measurable sets and
: L(Rn ) [0, ]
denotes Lebesgue measure, then we want L(Rn ) to contain all n-dimensional rectangles and (R) should be the usual volume of a rectangle R. Moreover, we want
to be countably additive. That is, if
{Ai L(Rn ) : i N}
is a countable collection of disjoint measurable sets, then their union should be
measurable and
!
X
[
(Ai ) .
Ai =
i=1
i=1
The reason for requiring countable additivity is that finite additivity is too weak
a property to allow the justification of any limiting processes, while uncountable
additivity is too strong; for example, it would imply that if the measure of a set
consisting of a single point is zero, then the measure of every subset of Rn would
be zero.
It is not possible to define the Lebesgue measure of all subsets of Rn in a
geometrically reasonable way. Hausdorff (1914) showed that for any dimension
n 1, there is no countably additive measure defined on all subsets of Rn that is
invariant under isometries (translations and rotations) and assigns measure one to
the unit cube. He further showed that if n 3, there is no such finitely additive
measure. This result is dramatized by the Banach-Tarski paradox: Banach and
Tarski (1924) showed that if n 3, one can cut up a ball in Rn into a finite number
of pieces and use isometries to reassemble the pieces into a ball of any desired volume
e.g. reassemble a pea into the sun. The construction of these pieces requires the
axiom of choice.1 Banach (1923) also showed that if n = 1 or n = 2 there are
finitely additive, isometrically invariant extensions of Lebesgue measure on Rn that
are defined on all subsets of Rn , but these extensions are not countably additive.
For a detailed discussion of the Banach-Tarski paradox and related issues, see [10].
The moral of these results is that some subsets of Rn are too irregular to define
their Lebesgue measure in a way that preserves countable additivity (or even finite
additivity in n 3 dimensions) together with the invariance of the measure under
1Solovay (1970) proved that one has to use the axiom of choice to obtain non-Lebesgue
measurable sets.
9
10
2. LEBESGUE MEASURE ON Rn
isometries. We will show, however, that such a measure can be defined on a algebra L(Rn ) of Lebesgue measurable sets which is large enough to include all set
of practical importance in analysis. Moreover, as we will see, it is possible to define
an isometrically-invariant, countably sub-additive outer measure on all subsets of
Rn .
There are many ways to construct Lebesgue measure, all of which lead to the
same result. We will follow an approach due to Caratheodory, which generalizes
to other measures: We first construct an outer measure on all subsets of Rn by
approximating them from the outside by countable unions of rectangles; we then
restrict this outer measure to a -algebra of measurable subsets on which it is countably additive. This approach is somewhat asymmetrical in that we approximate
sets (and their complements) from the outside by elementary sets, but we do not
approximate them directly from the inside.
Jones [5], Stein and Shakarchi [8], and Wheeler and Zygmund [11] give detailed
introductions to Lebesgue measure on Rn . Cohn [2] gives a similar development to
the one here, and Evans and Gariepy [3] discuss more advanced topics.
2.1. Lebesgue outer measure
We use rectangles as our elementary sets, defined as follows.
Definition 2.1. An n-dimensional, closed rectangle with sides oriented parallel
to the coordinate axes, or rectangle for short, is a subset R Rn of the form
R = [a1 , b1 ] [a2 , b2 ] [an , bn ]
where < ai bi < for i = 1, . . . , n. The volume (R) of R is
(R) = (b1 a1 )(b2 a2 ) . . . (bn an ).
If n = 1 or n = 2, the volume of a rectangle is its length or area, respectively.
We also consider the empty set to be a rectangle with () = 0. We denote the
collection of all n-dimensional rectangles by R(Rn ), or R when n is understood,
and then R 7 (R) defines a map
: R(Rn ) [0, ).
The use of this particular class of elementary sets is for convenience. We could
equally well use open or half-open rectangles, cubes, balls, or other suitable elementary sets; the result would be the same.
Definition 2.2. The outer Lebesgue measure (E) of a subset E Rn , or
outer measure for short, is
)
(
X
S
n
(Ri ) : E i=1 Ri , Ri R(R )
(2.1)
(E) = inf
i=1
where the infimum is taken over all countable collections of rectangles whose union
contains E. The map
: P(Rn ) [0, ],
: E 7 (E)
11
P
In this definition, a sum i=1 (Ri ) and (E) may take the value . We do
not require that the rectangles Ri are disjoint, so the same volume may contribute
to multiple terms in the sum on the right-hand side of (2.1); this does not affect
the value of the infimum.
Example 2.3. Let E = Q [0, 1] be the set of rational numbers between 0
and 1. Then E has outer measure zero. To prove this, let {qi : i N} be an
i
enumeration of the points in E.
SGiven > 0, let Ri be an interval of length /2
which contains qi . Then E i=1 (Ri ) so
0 (E)
(Ri ) = .
i=1
Hence (E) = 0 since > 0 is arbitrary. The same argument shows that any
countable set has outer measure zero. Note that if we cover E by a finite collection
of intervals, then the union of the intervals would have to contain [0, 1] since E is
dense in [0, 1] so their lengths sum to at least one.
The previous example illustrates why we need to use countably infinite collections of rectangles, not just finite collections, to define the outer measure.2 The
countable -trick used in the example appears in various forms throughout measure
theory.
Next, we prove that is an outer measure in the sense of Definition 1.2.
Theorem 2.4. Lebesgue outer measure has the following properties.
(a) () = 0;
(b) if E F , then (E) (F );
(c) if {Ei Rn : i N} is a countable collection of subsets of Rn , then
!
X
[
(Ei ) .
Ei
i=1
i=1
X
j=1
(Rij ) (Ei ) +
,
2i
Ei
Rij .
j=1
Ei
i=1
2The use of finitely many intervals leads to the notion of the Jordan content of a set, introduced by Peano (1887) and Jordan (1892), which is closely related to the Riemann integral; Borel
(1898) and Lebesgue (1902) generalized Jordans approach to allow for countably many intervals,
leading to Lebesgue measure and the Lebesgue integral.
2. LEBESGUE MEASURE ON Rn
12
and therefore
(E)
i,j=1
(Rij )
n
X
o X
(Ei ) + .
=
2i
i=1
(Ei ) +
i=1
(E)
(Ei )
i=1
Ni
[
Ii,j .
j=1
Then
(R) =
N1
X
j1 =1
Nn
X
(Sj1 j2 ...jn ) .
jn =1
Proof. Denoting the length of an interval I by |I|, using the fact that
|Ii | =
Ni
X
j=1
|Ii,j |,
3As a partial justification of the need to prove this fact, note that it would not be true if we
allowed uncountable covers, since we could cover any rectangle by an uncountable collection of
points all of whose volumes are zero.
13
Nn
N2
N1
X
X
X
|In,jn |
|I2,j2 | . . .
|I1,j1 |
=
=
j1 =1 j2 =1
N2
N1 X
X
j1 =1 j2 =1
jn =1
j2 =1
j1 =1
N2
N1 X
X
Nn
X
jn =1
Nn
X
jn =1
Proposition 2.6. If a rectangle R is an almost disjoint, finite union of rectangles {R1 , R2 , . . . , RN }, then
(2.3)
(R) =
N
X
(Ri ).
N
X
(Ri ).
i=1
(R)
i=1
where the ci,j are obtained by ordering the left and right ith coordinates of all faces
of rectangles in the collection {R1 , R2 , . . . , RN }, and define rectangles Sj1 j2 ...jn as
in (2.2).
Each rectangle Ri in the collection is an almost disjoint union of rectangles
Sj1 j2 ...jn , and their union contains all such products exactly once, so by applying
Lemma 2.5 to each Ri and summing the results we see that
N
X
i=1
(Ri ) =
N1
X
j1 =1
Nn
X
(Sj1 j2 ...jn ) .
jn =1
Similarly, R is an almost disjoint union of all the rectangles Sj1 j2 ...jn , so Lemma 2.5
implies that
Nn
N1
X
X
(Sj1 j2 ...jn ) ,
(R) =
j1 =1
jn =1
2. LEBESGUE MEASURE ON Rn
14
M
[
M
X
Si ,
i=1
i=1
(Si )
N
X
(Ri ).
i=1
M
X
i=1
(Si )
N
X
(Ri ),
i=1
(Si ) (Ri ) + i .
2
Then {Si : i N} is an open cover of the compact set R, so it contains a finite
N
X
i=1
(Si )
N n
X
o X
(Ri ) + i
(Ri ) + .
2
i=1
i=1
(R)
and it follows that (R) (R).
(Ri )
i=1
2.3. Carath
eodory measurability
We will obtain Lebesgue measure as the restriction of Lebesgue outer measure
to Lebesgue measurable sets. The construction, due to Caratheodory, works for any
outer measure, as given in Definition 1.2, so we temporarily consider general outer
measures. We will return to Lebesgue measure on Rn at the end of this section.
The following is the Caratheodory definition of measurability.
Definition 2.8. Let be an outer measure on a set X. A subset A X is
Caratheodory measurable with respect to , or measurable for short, if
(2.5)
(E) = (E A) + (E Ac )
2.3. CARATHEODORY
MEASURABILITY
15
(E) = (E A) + (E Ac )
= (E A B) + (E A B c )
+ (E Ac B) + (E Ac B c ).
The use of this inequality and the relation Ac B c = (A B)c in (2.6) implies that
(E) (E (A B)) + (E (A B)c )
so A B is measurable.
Moreover, if A is measurable and A B = , then by taking E = A B in
(2.5), we see that
(A B) = (A) + (B).
Thus, the outer measure of the union of disjoint, measurable sets is the sum of
their outer measures. The repeated application of this result implies that the finite
union of measurable sets is measurable and is finitely additive on the collection
of measurable sets.
Next, we we want to show that the countable union of measurable sets is
measurable. It is sufficient to consider disjoint unions. To see this, note that if
2. LEBESGUE MEASURE ON Rn
16
Bj =
Ai ,
for j 1
i=1
for j 2,
C1 = B1
Ai =
Cj .
j=1
i=1
Suppose that {Ai : i N} is a countably infinite, disjoint collection of measurable sets, and define
j
[
[
Ai ,
B=
Ai .
Bj =
i=1
i=1
= (E Aj ) + (E Bj1 ).
j
X
i=1
(E Ai ).
(E)
j
X
i=1
(E Ai ) + (E B c ).
X
i=1
(E Ai ) + (E B c )
i=1
E Ai
+ (E B c )
(E B) + (E B c )
(E).
2.3. CARATHEODORY
MEASURABILITY
X
[
(E Ai ),
E Ai =
17
i=1
Ai is mea-
i=1
i=1
= |L(Rn )
X
(Ri ).
(E) +
i=1
i +
Ri = R
N
[
j=1
From (2.3),
Si,j ,
i = Ri R R,
R
i ) +
(Ri ) = (R
N
X
Si,j Rc .
(Si,j ).
j=1
Using this result in the previous sum, relabeling the Si,j as Si , and rearranging the
resulting sum, we get that
X
X
i) +
(E) +
(R
(Si ).
i=1
i=1
2. LEBESGUE MEASURE ON Rn
18
X
X
i ),
(Si ).
(R
(E Rc )
(E R)
i=1
i=1
Hence,
(E) + (E R) + (E Rc ).
Since > 0 is arbitrary, it follows that
which proves the result.
(E) (E R) + (E Rc ),
[a1 +
k=1
1
1
1
1
1
1
, b1 ] [a2 + , b2 ] [an + , bn ].
k
k
k
k
k
k
so (E N ) = 0. Therefore, since E E N c ,
(E) (E N c ) = (E N ) + (E N c ),
i=1
Ri ,
X
i=1
(Ri ) < .
19
The argument in Example 2.3 shows that every countable set has Lebesgue
measure zero, but sets of measure zero may be uncountable; in fact the fine structure
of sets of measure zero is, in general, very intricate.
Example 2.14. The standard Cantor set, obtained by removing middle thirds
from [0, 1], is an uncountable set of zero one-dimensional Lebesgue measure.
Example 2.15. The x-axis in R2
A = (x, 0) R2 : x R
has zero two-dimensional Lebesgue measure. More generally, any linear subspace of
Rn with dimension strictly less than n has zero n-dimensional Lebesgue measure.
2.5. Translational invariance
An important geometric property of Lebesgue measure is its translational invariance. If A Rn and h Rn , let
A + h = {x + h : x A}
i=1
Ei [1, 2].
20
2. LEBESGUE MEASURE ON Rn
i=1
X
i=1
(Ei ) 3.
is the open ball of radius r centered at x Rn and | | denotes the Euclidean norm.
Remark 2.19. This definition is not constructive, since we start with the power
set of Rn and narrow it down until we obtain the smallest -algebra that contains
the open sets. It is surprisingly complicated to obtain B(Rn ) by starting from
the open or closed sets and taking successive complements, countable unions, and
countable intersections. These operations give sequences of collections of sets in Rn
(2.8)
G G G G . . . ,
F F F F . . . ,
where G denotes the open sets, F the closed sets, the operation of countable
unions, and the operation of countable intersections. These collections contain
each other; for example, F G and G F . This process, however, has to
be repeated up to the first uncountable ordinal before we obtain B(Rn ). This is
because if, for example, {Ai : i N} is a countable family of sets such that
A1 G \ G,
A2 G \ G , A3 G \ G , . . .
S
T
and so on, then there is no guarantee that
i=1 Ai or
i=1 Ai belongs to any of
the previously constructed families. In general, one only knows that they belong to
the + 1 iterates G... or G... , respectively, where is the ordinal number
21
Ri
i=1
is a decomposition of an open set G into an almost disjoint union of closed rectangles, then
[
Ri
G
i=1
X
X
(Ri ).
(Ri ) (G)
i=1
i=1
Since
(Ri )
(G) =
X
i=1
(Ri )
2. LEBESGUE MEASURE ON Rn
22
for any such decomposition and that the sum is independent of the way in which
G is decomposed into almost disjoint rectangles.
The Borel -algebra B is not complete and is strictly smaller than the Lebesgue
-algebra L. In fact, one can show that the cardinality of B is equal to the cardinality c of the real numbers, whereas the cardinality of L is equal to 2c . For example,
the Cantor set is a set of measure zero with the same cardinality as R and every
subset of the Cantor set is Lebesgue measurable.
We can obtain examples of sets that are Lebesgue measurable but not Borel
measurable by considering subsets of sets of measure zero. In the following example
of such a set in R, we use some properties of measurable functions which will be
proved later.
Example 2.22. Let f : [0, 1] [0, 1] denote the standard Cantor function and
define g : [0, 1] [0, 1] by
g(y) = inf {x [0, 1] : f (x) = y} .
Then g is an increasing, one-to-one function that maps [0, 1] onto the Cantor set
C. Since g is increasing it is Borel measurable, and the inverse image of a Borel
set under g is Borel. Let E [0, 1] be a non-Lebesgue measurable set. Then
F = g(E) C is Lebesgue measurable, since it is a subset of a set of measure zero,
but F is not Borel measurable, since if it was E = g 1 (F ) would be Borel.
Other examples of Lebesgue measurable sets that are not Borel sets arise from
the theory of product measures in Rn for n 2. For example, let N = E {0} R2
where E R is a non-Lebesgue measurable set in R. Then N is a subset of the
x-axis, which has two-dimensional Lebesgue measure zero, so N belongs to L(R2 )
since Lebesgue measure is complete. One can show, however, that if a set belongs
to B(R2 ) then every section with fixed x or y coordinate, belongs to B(R); thus, N
cannot belong to B(R2 ) since the y = 0 section E is not Borel.
As we show below, L(Rn ) is the completion of B(Rn ) with respect to Lebesgue
measure, meaning that we get all Lebesgue measurable sets by adjoining all subsets
of Borel sets of measure zero to the Borel -algebra and taking unions of such sets.
2.7. Borel regularity
Regularity properties of measures refer to the possibility of approximating in
measure one class of sets (for example, nonmeasurable sets) by another class of
sets (for example, measurable sets). Lebesgue measure is Borel regular in the sense
that Lebesgue measurable sets can be approximated in measure from the outside
by open sets and from the inside by closed sets, and they can be approximated
by Borel sets up to sets of measure zero. Moreover, there is a simple criterion for
Lebesgue measurability in terms of open and closed sets.
The following theorem expresses a fundamental approximation property of
Lebesgue measurable sets by open and compact sets. Equations (2.9) and (2.10)
are called outer and inner regularity, respectively.
Theorem 2.23. If A Rn , then
(2.9)
23
(Ri ) (A) + .
2
i=1
Si
i=1
X
i=1
(Si )
(Ri ) + ,
2
i=1
(G) (A) + .
(F ) (K) + (G),
(F ) = (A) + (F \ A).
(A) = (F ) (F \ A)
(F ) (G) +
(K) + ,
24
2. LEBESGUE MEASURE ON Rn
which implies (2.13) and proves the result for bounded, measurable sets.
Now suppose that A is an unbounded measurable set, and define
(2.14)
Ak = {x A : |x| k} .
(Ak ) (A)
as k .
(A) (Ak ) + .
2
Moreover, since Ak is bounded, there is a compact set K Ak such that
(Ak ) (K) + .
2
Therefore, for every > 0 there is a compact set K A such that
(A) (K) + ,
which gives (2.13), and completes the proof.
(G \ A) < .
Proof. First we assume that A is measurable and show that it satisfies the
condition given in the theorem.
Suppose that (A) < and let > 0. From (2.12) there is an open set G A
such that (G) < (A) + . Then, since A is measurable,
(G \ A) = (G) (G A) = (G) (A) < ,
which proves the result when A has finite measure.
25
(Gk \ Ak ) < k .
2
S
Then G =
k=1 Gk is an open set that contains A, and
!
[
X
X
Gk \ A
(Gk \ A)
(Gk \ Ak ) < .
(G \ A) =
k=1
k=1
k=1
This theorem states that a set is Lebesgue measurable if and only if it can be
approximated from the outside by an open set in such a way that the difference
has arbitrarily small outer Lebesgue measure. This condition can be adopted as
the definition of Lebesgue measurable sets, rather than the Caratheodory definition
which we have used c.f. [5, 8, 11].
The following theorem gives another characterization of Lebesgue measurable
sets, as ones that can be squeezed between open and closed sets.
Theorem 2.25. A subset A Rn is Lebesgue measurable if and only if for
every > 0 there is an open set G and a closed set F such that G A F and
(2.17)
(G \ F ) < .
(G \ A) < ,
(H \ Ac ) < .
2
2
Then, defining the closed set F = H c , we have G A F and
(G \ F ) (G \ A) + (A \ F ) = (G \ A) + (H \ Ac ) < .
Finally, suppose that (A) < and let > 0. From Theorem 2.23, since A is
measurable, there is a compact set K A such that (A) < (K) + /2 and
26
2. LEBESGUE MEASURE ON Rn
(G \ A) = (A \ F ) = 0.
Proof. For each k N, choose an open set Gk and a closed set Fk such that
Gk A Fk and
1
(Gk \ Fk )
k
Then
\
[
G=
Gk ,
F =
Fk
k=1
k=1
27
First, we consider the Lebesgue measure of rectangles whose sides are not paral we denote
lel to the coordinate axes. We use a tilde to denote such rectangles by R;
closed rectangles whose sides are parallel to the coordinate axes by R as before.
and R as oblique and parallel rectangles, respectively. We denote
We refer to R
by v(R),
i.e. the product of the lengths of its sides, to
the volume of a rectangle R
We know that (R) = v(R) for
avoid confusion with its Lebesgue measure (R).
parallel rectangles, and that R is measurable since it is closed, but we have not yet
= v(R)
for oblique rectangles.
shown that (R)
More explicitly, we regard Rn as a Euclidean space equipped with the standard
inner product,
(x, y) =
n
X
xi yi ,
x = (x1 , x2 , . . . , xn ),
y = (y1 , y2 , . . . , yn ).
i=1
and {
e1 , e2 , . . . , en } is another orthonormal basis, then we use R to denote rectangles
to denote rectangles whose sides are parallel
whose sides are parallel to {ei } and R
n
to {
ei }. The linear map Q : R Rn defined by Qei = ei is orthogonal, meaning
that QT = Q1 and
(Qx, Qy) = (x, y)
for all x, y Rn .
= QR
Since Q preserves lengths and angles, it maps a rectangle R to a rectangle R
= v(R).
such that v(R)
We will use the following lemma.
contains a finite almost disjoint collecLemma 2.29. If an oblique rectangle R
tion of parallel rectangles {R1 , R2 , . . . , RN } then
N
X
i=1
v(Ri ) v(R).
This result is geometrically obvious, but a formal proof seems to require a fuller
discussion of the volume function on elementary geometrical sets, which is included
in the theory of valuations in convex geometry. We omit the details.
2. LEBESGUE MEASURE ON Rn
28
X
i=1
+ .
v(Ri ) v(R)
Ri .
i=1
X
i=1
v(Ri ) v(S),
v(R)
+ .
v(Ri ) v(S)
X
i ) v(R) + .
(2.18)
v(R
i=1
X
+ .
v(Ri ) (E)
2
i=1
i,j : j N} of Ri by oblique
From (2.18), for each i N we can choose a cover {R
rectangles such that
X
i,j ) v(Ri ) + .
v(R
2i+1
i=1
i,j : i, j N} is a countable cover of E
by oblique rectangles, and
Then {R
i,j=1
i,j )
v(R
X
i=1
v(Ri ) +
+ .
(E)
2
29
v(Ri,j ).
i,j=1
i,j=1
v(Ri,j ) =
i,j=1
i,j ) (E)
+ ,
v(R
(E) (E).
we get the
By applying the same argument to the inverse mapping E = QT E,
reverse inequality, and it follows that (E) = (E).
Since is invariant under Q, the Caratheodory criterion for measurability is
invariant, and E is measurable if and only if QE is measurable.
It follows from Theorem 2.31 that Lebesgue measure is invariant under rotations
and reflections.4 Since it is also invariant under translations, Lebesgue measure is
invariant under all isometries of Rn .
Next, we consider the effect of dilations on Lebesgue measure. Arbitrary linear
maps may then be analyzed by decomposing them into rotations and dilations.
Proposition 2.32. Suppose that : Rn Rn is the linear transformation
(2.19)
: (x1 , x2 , . . . , xn ) 7 (1 x1 , 2 x2 , . . . , n xn )
2. LEBESGUE MEASURE ON Rn
30
In that case, according to the polar decomposition, the map T may be written
as a composition
T = QU
The use of half-open intervals is significant here because a Lebesgue-Stieltjes measure may assign nonzero measure to a single point. Thus, unlike Lebesgue measure,
we need not have F ([a, b]) = F ((a, b]). Half-open intervals are also convenient
because the complement of a half-open interval is a finite union of (possibly infinite) half-open intervals of the same type. Thus, the collection of finite unions of
half-open intervals forms an algebra.
The right-continuity of F is consistent with the use of intervals that are halfopen at the left, since
\
(a, a + 1/i] = ,
i=1
or
lim [F (a + 1/i) F (a)] = lim F (x) F (a) = 0.
xa+
Conversely, as we state in the next theorem, any such function F defines a Borel
measure on R.
Theorem 2.34. Suppose that F : R R is an increasing, right-continuous
function. Then there is a unique Borel measure F : B(R) [0, ] such that
F ((a, b]) = F (b) F (a)
for every a < b.
31
and restrict
to its Caratheodory measurable sets, which include the Borel sets.
See e.g. Section 1.5 of Folland [4] for a detailed proof.
The following examples illustrate the three basic types of Lebesgue-Stieltjes
measures.
Example 2.35. If F (x) = x, then F is Lebesgue measure on R with
F ((a, b]) = b a.
Example 2.36. If
then F
1 if x 0,
0 if x < 0,
is the -measure supported at 0,
1 if 0 A,
F (A) =
0 if 0
/ A.
F (x) =
CHAPTER 3
Measurable functions
Measurable functions in measure theory are analogous to continuous functions
in topology. A continuous function pulls back open sets to open sets, while a
measurable function pulls back measurable sets to measurable sets.
3.1. Measurability
Most of the theory of measurable functions and integration does not depend
on the specific features of the measure space on which the functions are defined, so
we consider general spaces, although one should keep in mind the case of functions
defined on R or Rn equipped with Lebesgue measure.
Definition 3.1. Let (X, A) and (Y, B) be measurable spaces. A function
f : X Y is measurable if f 1 (B) A for every B B.
Note that the measurability of a function depends only on the -algebras; it is
not necessary that any measures are defined.
In order to show that a function is measurable, it is sufficient to check the
measurability of the inverse images of sets that generate the -algebra on the target
space.
Proposition 3.2. Suppose that (X, A) and (Y, B) are measurable spaces and
B = (G) is generated by a family G P(Y ). Then f : X Y is measurable if
and only if
f 1 (G) A
for every G G.
Proof. Set operations are natural under pull-backs, meaning that
f 1 (Y \ B) = X \ f 1 (B)
and
f
i=1
It follows that
Bi
(Bi ) ,
i=1
i=1
Bi
f 1 (Bi ) .
i=1
M = B Y : f 1 (B) A
is a -algebra on Y . By assumption, M G and therefore M (G) = B, which
implies that f is measurable.
It is worth noting the indirect nature of the proof of containment of -algebras
in the previous proposition; this is required because we typically cannot use an
explicit representation of sets in a -algebra. For example, the proof does not
characterize M, which may be strictly larger than B.
If the target space Y is a topological space, then we always equip it with the
Borel -algebra B(Y ) generated by the open sets (unless stated explicitly otherwise).
33
34
3. MEASURABLE FUNCTIONS
f :X R
f : X R.
We will consider one case or the other as convenient, and comment on any differences. A positive extended real-valued function is a function
f : X [0, ].
B {},
B {},
B {, }
where B is a Borel subset of R. As Example 2.22 shows, sets that are Lebesgue
measurable but not Borel measurable need not be well-behaved under the inverse
of even a monotone function, which helps explain why we do not include them in
the range -algebra on R or R.
By contrast, when the domain of a function is a measure space it is often
convenient to use a complete space. For example, if the domain is Rn we typically
equip it with the Lebesgue -algebra, although if completeness is not required
we may use the Borel -algebra. With this understanding, we get the following
definitions. We state them for real-valued functions; the definitions for extended
real-valued functions are completely analogous
Definition 3.3. If (X, A) is a measurable space, then f : X R is measurable
if f 1 (B) A for every Borel set B B(R). A function f : Rn R is Lebesgue
measurable if f 1 (B) is a Lebesgue measurable subset of Rn for every Borel subset
B of R, and it is Borel measurable if f 1 (B) is a Borel measurable subset of Rn
for every Borel subset B of R
This definition ensures that continuous functions f : Rn R are Borel measurable and functions that are equal a.e. to Borel measurable functions are Lebesgue
measurable. If f : R R is Borel measurable and g : Rn R is Lebesgue (or
Borel) measurable, then the composition f g is Lebesgue (or Borel) measurable
since
1
(f g) (B) = g 1 f 1 (B) .
Note that if f is Lebesgue measurable, then f g need not be measurable since
f 1 (B) need not be Borel even if B is Borel.
We can give more easily verifiable conditions for measurability in terms of
generating families for Borel sets.
35
Proposition 3.4. The Borel -algebra on R is generated by any of the following collections of intervals
{(, b) : b R} ,
{(, b] : b R} ,
{(a, ) : a R} ,
{[a, ) : a R} .
[a, b +
n=1
1
).
n
From Proposition 2.20, the Borel -algebra B(R) is generated by the collection of
closed rectangles [a, b], so
({(, b) : b R}) = B(R).
The proof for the other collections is similar.
The properties given in the following proposition are sometimes taken as the
definition of a measurable function.
Proposition 3.5. If (X, A) is a measurable space, then f : X R is measurable if and only if one of the following conditions holds:
{x X : f (x) < b} A
for every b R;
{x X : f (x) > a} A
for every a R;
{x X : f (x) b} A
{x X : f (x) a} A
for every b R;
for every a R.
If any one of these equivalent conditions holds, then f 1 (B) A for every set
B B(R). We will often use a shorthand notation for sets, such as
{f < b} = {x X : f (x) < b} .
The Borel -algebra on R is generated by intervals of the form [, b), [, b],
(a, ], or [a, ] where a, b R, and exactly the same conditions as the ones
in Proposition 3.5 imply the measurability of an extended real-valued functions
f : X R. In that case, we can allow a, b R to be extended real numbers
in Proposition 3.5, but it is not necessary to do so in order to imply that f is
measurable.
Measurability is well-behaved with respect to algebraic operations.
Proposition 3.6. If f, g : X R are real-valued measurable functions and
k R, then
kf, f + g,
f g,
f /g
are measurable functions, where we assume that g 6= 0 in the case of f /g.
36
3. MEASURABLE FUNCTIONS
It follows that
fg =
1
(f + g)2 f 2 g 2
2
is measurable. Finally, if g 6= 0
if b < 0,
{1/b < g < 0}
{1/g < b} =
{ < g < 0}
if b = 0,
nN
nN
nN
inf fn ,
nN
lim sup fn ,
n
lim inf fn
n
37
\
sup fn b =
{fn b} ,
nN
inf fn < b
nN
n=1
n=1
{fn < b}
nN kn
Perhaps the most important way in which new functions arise from old ones is
by pointwise convergence.
Definition 3.9. A sequence {fn : n N} of functions fn : X R converges
pointwise to a function f : X R if fn (x) f (x) as n for every x X.
Pointwise convergence preserves measurability (unlike continuity, for example).
This fact explains why the measurable functions form a sufficiently large class for
the needs of analysis.
Theorem 3.10. If {fn : n N} is a sequence of measurable functions fn :
X R and fn f pointwise as n , then f : X R is measurable.
Proof. If fn f pointwise, then
f = lim sup fn = lim inf fn
n
(x) =
N
X
cn En (x)
n=1
where c1 , . . . , cN R and E1 , . . . , EN A.
Note that, according to this definition, a simple function is measurable. The
representation of in (3.1) is not unique; we call it a standard representation if the
constants cn are distinct and the sets En are disjoint.
38
3. MEASURABLE FUNCTIONS
k = 0, 1, . . . , 22n 1,
Fn = f 1 (Jn ).
22n
1
X
k2n Ek,n + 2n Fn
k=0
CHAPTER 4
Integration
In this Chapter, we define the integral of real-valued functions on an arbitrary
measure space and derive some of its basic properties. We refer to this integral as
the Lebesgue integral, whether or not the domain of the functions is subset of Rn
equipped with Lebesgue measure. The Lebesgue integral applies to a much wider
class of functions than the Riemann integral and is better behaved with respect to
pointwise convergence. We carry out the definition in three steps: first for positive
simple functions, then for positive measurable functions, and finally for extended
real-valued measurable functions.
N
X
ci E i
i=1
d =
N
X
ci (Ei ) .
i=1
40
4. INTEGRATION
In this definition, we approximate the function f from below by simple functions. In contrast with the definition of the Riemann integral, it is not necessary to
approximate a measurable function from both above and below in order to define
its integral.
If A X is a measurable set and f : X [0, ] is measurable, we define
Z
Z
f d = f A d.
A
Unlike the Riemann integral, where the definition of the integral over non-rectangular
subsets of R2 already presents problems, it is trivial to define the Lebesgue integral
over arbitrary measurable subsets of a set on which it is already defined.
The following properties are an immediate consequence of the definition and
the corresponding properties of simple functions.
Proposition 4.5. If f, g : X [0, ] are positive, measurable, extended realvalued function on a measure space X, then:
Z
Z
kf d = k f d
if k [0, );
Z
Z
0 f d g d
if 0 f g.
The integral is also linear, but this is not immediately obvious from the definition and it depends on the measurability of the functions. To show the linearity,
we will first derive one of the fundamental convergence theorem for the Lebesgue
integral, the monotone convergence theorem. We discuss this theorem and its applications in greater detail in Section 4.5.
41
then
lim
fn d =
f d.
Proof. The pointwise limit f : X [0, ] exists since the sequence {fn }
is increasing. Moreover, by the monotonicity of the integral, the integrals are
increasing, and
Z
Z
Z
fn d fn+1 d f d,
An
Moreover, if
=
N
X
ci E i
i=1
i=1
i=1
42
4. INTEGRATION
In particular, this theorem implies that we can obtain the integral of a positive
measurable function f as a limit of integrals of an increasing sequence of simple
functions, not just as a supremum over all simple functions dominated by f as in
Definition 4.4. As shown in Theorem 3.12, such a sequence of simple functions
always exists.
Proposition 4.7. If f, g : X [0, ] are positive, measurable functions on a
measure space X, then
Z
Z
Z
(f + g) d = f d + g d.
f = f + f ,
f + = max{f, 0},
f = max{f, 0}.
|f | = f + + f .
Note that f is measurable if and only if f + and f are measurable.
Definition 4.8. If f : X R is a measurable function, then
Z
Z
Z
f d = f + d f d,
R
R
provided that at least Rone of theR integrals f + d, f d is finite. The function
f is integrable if both f + d, f d are finite, which is the case if and only if
Z
|f | d < .
43
This Lebesgue integral has all the usual properties of an integral. We restrict
attention to integrable functions to avoid undefined expressions involving extended
real numbers such as .
Proposition 4.9. If f, g : X R are integrable functions, then:
Z
Z
kf d = k f d
if k R;
Z
Z
Z
(f + g) d = f d + g d;
Z
Z
f d g d
if f g;
Z
Z
f d |f | d.
Proof. These results follow by writing functions into their positive and negative parts, as in (4.3), and using the results for positive functions.
If f = f + f and k 0, then (kf )+ = kf + and (kf ) = kf , so
Z
Z
Z
Z
Z
Z
kf d = kf + d kf d = k f + d k f d = k f d.
Similarly, (f )+ = f and (f ) = f + , so
Z
Z
Z
Z
(f ) d = f d f + d = f d.
If h = f + g and
f = f + f ,
g = g+ g,
h = h+ h
are the decompositions of f, g, h into their positive and negative parts, then
h+ h = f + f + g + g .
f + g + h+ = f + + g + + h .
The linearity of the integral on positive functions gives
Z
Z
Z
Z
Z
Z
f d + g d + h+ d = f + d + g + d + h d,
+
h d h d = f d f d + g d g d,
R
R
R
or (f + g) d = f d + g d.
It follows that if f g, then
Z
Z
Z
0 (g f ) d = g d f d,
R
R
so f d g d. The last result is then a consequence of the previous results
and |f | f |f |.
Let us give two basic examples of the Lebesgue integral.
44
4. INTEGRATION
X
f d =
xn ,
N
n=1
where the integral is finite if and only if the series is absolutely convergent. Thus,
the theory of absolutely convergent series is a special case of the Lebesgue integral.
Note that a conditionally convergent series, such as the alternating harmonic series,
does not correspond to a Lebesgue integral, since both its positive and negative
parts diverge.
Example 4.11. Suppose that X = [a, b] is a compact interval and : L([a, b])
R is Lesbegue measure on [a, b]. We note in Section 4.8 that any Riemann integrable function f : [a, b] R is integrable with respect to Lebesgue measure , and
its Riemann integral is equal to the Lebesgue integral,
Z b
Z
f d.
f (x) dx =
[a,b]
Thus, all of the usual integrals from elementary calculus remain valid for the
Lebesgue integral on R. We will write an integral with respect to Lebesgue measure
on R, or Rn , as
Z
f dx.
1
1
1
sin + cos
x
x
x
dx
1
dx
x
45
The inability of the Lebesgue integral to deal directly with the cancelation
between large positive and negative parts in oscillatory or singular integrals, such
as the ones in the previous examples, is sometimes viewed as a defect (although the
integrals above can still be defined as an appropriate limit of Lebesgue integrals).
Other definitions of the integral such as the Henstock-Kurzweil integral, which is a
generalization of the Riemann integral, avoid this defect but they have not proved
to be as useful as the Lebesgue integral. Similar issues arise in connection with
Feynman path integrals in quantum theory, where one would like to define the
integral of highly oscillatory functionals on an infinite-dimensional function-space.
4.4. Absolute continuity
The following results show that a function with finite integral is finite a.e. and
that the integral depends only on the pointwise a.e. values of a function.
R
N
X
ci E i ,
i=1
R
and Definition 4.4 implies thatR f d = 0.
Conversely, suppose that f d = 0. For n N, let
En = {x X : f (x) 1/n} .
[
{x X : f (x) > 0} =
En ,
n=1
46
4. INTEGRATION
For integrable functions we can strengthen the previous result to get the following property, which is called the absolute continuity of the integral.
Proposition
4.16. Suppose that f : X R is an integrable function, meaning
R
|f | d < . Then, given any > 0, there exists > 0 such that
Z
(4.5)
0
|f | d <
that
(f fn ) d + n(A).
X
0
(f fn ) d < ,
2
X
and then choose
.
2n
If (A) < , we get (4.5), which proves the result.
=
47
By modifying this example, and the following ones, we can obtain a sequence fn that
converges pointwise to zero but whose integrals converge to infinity; for example
2
n if 0 < x < 1/n,
fn (x) =
0
otherwise.
Example 4.19. Define fn : R R by
1/n if 0 < x < n,
fn (x) =
0
otherwise.
Then fn 0 as n pointwise on R, and even uniformly, but
Z
fn dx = 1
for every n N.
Example 4.20. Define fn : R R by
1 if n < x < n + 1,
fn (x) =
0 otherwise.
48
4. INTEGRATION
Then {gn } is a monotone increasing sequence which converges pointwise to lim inf fn
as n , so by the monotone convergence theorem
Z
Z
(4.8)
lim
gn d = lim inf fn d.
n
so that
lim
gn d lim inf
n
fn d.
We may have strict inequality in (4.7), as in the previous examples. The monotone convergence theorem and Fatous Lemma enable us to determine the integrability of functions.
49
if x = 0.
For n N, let
x if 1/n x 1,
n
if 0 x < 1/n.
Then {fn } is an increasing sequence of Lebesgue measurable functions (e.g since
fn is continuous) that converges pointwise to f . We denote the integral of f with
R1
respect to Lebesgue measure on [0, 1] by 0 f (x) dx. Then, by the monotone convergence theorem,
Z 1
Z 1
f (x) dx = lim
fn (x) dx.
fn (x) =
fn (x) dx
1
1
f d
as n .
which gives
f d lim inf
n
fn d.
Similarly, g fn 0, so
Z
Z
Z
Z
(g f ) d lim inf (g fn ) d g d lim sup fn d,
n
which gives
f d lim sup
n
fn d,
An alternative, and perhaps more illuminating, proof of the dominated convergence theorem may be obtained from Egoroffs theorem and the absolute continuity
of the integral. Egoroffs theorem states that if a sequence {fn } of measurable functions, defined on a finite measure space (X, A, ), converges pointwise to a function
f , then for every > 0 there exists a measurable set A X such that {fn } converges
uniformly to f on A and (X \ A) < . The uniform integrability of the functions
50
4. INTEGRATION
and the absolute continuity of the integral imply that what happens off the set A
may be made to have arbitrarily small effect on the integrals. Thus, the convergence
theorems hold because of this almost uniform convergence of pointwise-convergent
sequences of measurable functions.
4.6. Complex-valued functions and a.e. convergence
In this section, we briefly indicate the generalization of the above results to
complex-valued functions and sequences that converge pointwise almost everywhere. The required modifications are straightforward.
If f : X C is a complex valued function f = g + ih, then we say that f is
measurable if and only if its real and imaginary parts g, h : X R are measurable,
and integrable if and only if g, h are integrable. In that case, we define
Z
Z
Z
f d = g d + i h d.
Note that we do not allow extended real-valued functions or infinite integrals here.
It follows from the discussion of product measures that f : X C, where C is
equipped with its Borel -algebra B(C), is measurable if and only if its real and
imaginary parts are measurable, so this definition is consistent with our previous
one.
The integral of complex-valued functions satisfies the properties given in Proposition 4.9, where we allow k C and the condition
f Rg is only relevant for
R
real-valued functions. For example, to show that | f d| |f | d, we let
Z
Z
f d = f d ei
for a suitable argument , and then
Z
Z
Z
Z
Z
f d = ei f d = [ei f ] d |[ei f ]| d |f | d.
4.7. L1 SPACES
51
as n .
0 if 0 x < 1/4,
1 if 0 x 1/4,
1 if 1/4 x 1/2,
f4 (x) =
f5 (x) =
1 if 1/4 < x 1,
0 if 1/2x < x 1,
Thus, f has compact support if and only if it vanishes outside a bounded set.
Theorem 4.27. The space Cc (Rn ) is dense in L1 (Rn ). Explicitly, if f
L (Rn ), then for any > 0 there exists a function g Cc (Rn ) such that
1
kf gkL1 < .
52
4. INTEGRATION
d(x, Gc )
d(x, K) + d(x, Gc )
4.8. Riemann integral
53
f d = lim
n d,
CHAPTER 5
Product Measures
Given two measure spaces, we may construct a natural measure on their Cartesian product; the prototype is the construction of Lebesgue measure on R2 as the
product of Lebesgue measures on R. The integral of a measurable function on
the product space may be evaluated as iterated integrals on the individual spaces
provided that the function is positive or integrable (and the measure spaces are
-finite). This result, called Fubinis theorem, is another one of the basic and most
useful properties of the Lebesgue integral. We will not give complete proofs of all
the results in this Chapter.
5.1. Product -algebras
We begin by describing product -algebras. If (X, A) and (Y, B) are measurable
spaces, then a measurable rectangle is a subset A B of X Y where A A and
B B are measurable subsets of X and Y , respectively. For example, if R is
equipped with its Borel -algebra, then Q Q is a measurable rectangle in R R.
(Note that the sides A, B of a measurable rectangle A B R R can be
arbitrary measurable sets; they are not required to be intervals.)
Definition 5.1. Suppose that (X, A) and (Y, B) are measurable spaces. The
product -algebra A B is the -algebra on X Y generated by the collection of
all measurable rectangles,
A B = ({A B : A A, B B}) .
Ex = {y Y : (x, y) E} ,
As stated in the next proposition, all sections of a measurable set are measurable.
Proposition 5.2. If (X, A) and (Y, B) are measurable spaces and E A B,
then Ex B for every x X and E y A for every y Y .
Proof. Let
M = {E X Y : Ex B for every x X and E y A for every y Y } .
Then M contains all measurable rectangles, since the x-sections of A B are either
or B and the y-sections are either or A. Moreover, M is a -algebra since, for
example, if E, Ei X Y and x X, then
!
[
[
(Ei )x .
Ei
=
(E c )x = (Ex )c ,
i=1
55
i=1
56
5. PRODUCT MEASURES
Proposition 5.3. Suppose that Rm , Rn are equipped with their Borel -algebras
B(Rm ), B(Rn ) and let Rm+n = Rm Rn . Then
B(Rm+n ) = B(Rm ) B(Rn ).
where R(Rm+n ) denotes the collection of rectangles in Rm+n . From Proposition 2.21, the rectangles generate the Borel -algebra, and therefore
B(Rm ) B(Rn ) B(Rm+n ).
M = {A Rm : A Rn B(Rm+n )} .
[
[
c
n
n c
(Ai Rn ) .
Ai Rn =
A R = (A R ) ,
i=1
i=1
By the repeated application of this result, we see that the Borel -algebra on
Rn is the n-fold product of the Borel -algebra on R. This leads to an alternative
method of constructing Lebesgue measure on Rn as a product of Lebesgue measures
on R, instead of the direct construction we gave earlier.
5.2. Premeasures
Premeasures provide a useful way to generate outer measures and measures,
and we will use them to construct product measures. In this section, we derive
some general results about premeasures and their associated measures that we use
below. Premeasures are defined on algebras, rather than -algebras, but they are
consistent with countable additivity.
Definition 5.4. An algebra on a set X is a collection of subsets of X that
contains and X and is closed under complements, finite unions, and finite intersections.
5.2. PREMEASURES
57
[
Ai E,
i=1
then
Ai
i=1
(Ai ) .
i=1
Proposition 5.7. The set function : P(X) [0, ] given by Definition 5.6.
is an outer measure on X. Every set A E is Caratheodory measurable and
(A) = (A).
X
X
(Aj ),
(Bj )
(A) =
j=1
j=1
58
5. PRODUCT MEASURES
X
(Bi ).
(E) +
i=1
X
i=1
(Bi A) +
X
i=1
(Bi Ac ) (E A) + (E Ac ),
59
X
i=1
(Ai Bi ).
(Ai Bi )
i=1
Ai Bi (x, y).
i=1
Ai (x)Bi (y).
Ai (x)(Bi ).
(Ai )(Bi ),
i=1
Integrating this equation over Y for fixed x X and using the monotone convergence theorem, we get
A (x)(B) =
i=1
i=1
N
X
i=1
(Ai )(Bi ),
E=
N
[
i=1
Ai Bi
SN
where E = i=1 Ai Bi is any representation of E E as a disjoint union of
measurable rectangles.
60
5. PRODUCT MEASURES
Proposition 5.11 implies that is countably additive on E, since we may decompose any countable disjoint union of sets in E into a countable common disjoint
refinement of rectangles, so is a premeasure as claimed. The outer product measure associated with , which we write as ( ) , is defined in terms of countable
coverings by measurable rectangles. This gives the following.
Definition 5.13. Suppose that (X, A, ) and (Y, B, ) are measure spaces.
Then the product outer measure
( ) : P(X Y ) [0, ]
on X Y is defined for E X Y by
)
(
X
S
( ) : A B [0, ],
( ) = ( ) |AB
( )(A B) = (A)(B)
for every A A, B B.
Moreover, if (X, A, ) and (Y, B, ) are -finite measure spaces, then ( ) is the
unique measure on A B with this property.
Note that, in general, the -algebra of Caratheodory measurable sets associated
with ( ) is strictly larger than the product -algebra. For example, if Rm and
Rn are equipped with Lebesgue measure defined on their Borel -algebras, then the
Caratheodory -algebra on the product Rm+n = Rm Rn is the Lebesgue -algebra
L(Rm+n ), whereas the product -algebra is the Borel -algebra B(Rm+n ).
5.4. Measurable functions
If f : X Y C is a function of (x, y) X Y , then for each x X we define
the x-section fx : Y C and for each y Y we define the y-section f y : Y C
by
fx (y) = f (x, y),
f y (x) = f (x, y).
Theorem 5.15. If (X, A, ), (Y, B, ) are measure spaces and f : X Y C
is a measurable function, then fx : Y C, f y : X C are measurable for every
x X, y Y . Moreover, if (X, A, ), (Y, B, ) are -finite, then the functions
g : X C, h : Y C defined by
Z
Z
g(x) = fx d,
h(y) = f y d
are measurable.
61
F1 F2 Fi . . . ,
then
i=1
Ei C,
i=1
Fi C.
X
X
|amn | <
m=1
n=1
then
m=1
n=1
amn
n=1
m=1
amn
62
5. PRODUCT MEASURES
Rm
Rm
Rn
Z
Rm
Rn
|f (x, y)| dy
Z
f (x, y) dx dy =
Rm
Z
dx
f (x, y) dy
Rn
dx,
CHAPTER 6
Differentiation
The generalization from elementary calculus of differentiation in measure theory
is less obvious than that of integration, and the methods of treating it are somewhat
involved.
Consider the fundamental theorem of calculus (FTC) for smooth functions of
a single variable. In one direction (FTC-I, say) it states that the derivative of the
integral is the original function, meaning that
Z x
d
f (y) dy.
(6.1)
f (x) =
dx a
In the other direction (FTC-II, say) it states that we recover the original function
by integrating its derivative
Z x
(6.2)
F (x) = F (a) +
f (y) dy,
f = F .
a
As we will see, (6.1) holds pointwise a.e. provided that f is locally integrable, which
is needed to ensure that the right-hand side is well-defined. Equation (6.2), however,
does not hold for all continuous functions F whose pointwise derivative is defined
a.e. and integrable; we also need to require that F is absolutely continuous. The
Cantor function is a counter-example.
First, we consider a generalization of (6.1) to locally integrable functions on
Rn , which leads to the Lebesgue differentiation theorem. We say that a function
f : Rn R is locally integrable if it is Lebesgue measurable and
Z
|f | dx <
K
for every compact subset K R ; we denote the space of locally integrable functions by L1loc (Rn ).
Let
(6.3)
Br (x) = {y Rn : |y x| < r}
denote the open ball of radius r and center x Rn . We denote Lebesgue measure
on Rn by and the Lebesgue measure of a ball B by (B) = |B|.
To motivate the statement of the Lebesgue differentiation theorem, observe
that (6.1) may be written in terms of symmetric differences as
Z x+r
1
f (y) dy.
(6.4)
f (x) = lim+
r0 2r xr
In other words, the value of f at a point x is the limit of local averages of f over
intervals centered at x as their lengths approach zero. An n-dimensional version of
63
64
6. DIFFERENTIATION
(6.4) is
(6.5)
f (x) = lim+
r0
1
|Br (x)|
f (y) dy
Br (x)
where the integral is with respect n-dimensional Lebesgue measure. The Lebesgue
differentiation theorem states that (6.5) holds pointwise -a.e. for any locally integrable function f .
To prove the theorem, we will introduce the maximal function of an integrable
function, whose key property is that it is weak-L1, as stated in the Hardy-Littlewood
theorem. This property may be shown by the use of a simple covering lemma, which
we begin by proving.
Second, we consider a generalization of (6.2) on the representation of a function
as an integral. In defining integrals on a general measure space, it is natural to
think of them as defined on sets rather than real numbers. For example, in (6.2),
we would write F (x) = ([a, x]) where : B([a, b]) R is a signed measure. This
interpretation leads to the following question: if , are measures on a measurable
space X is there a function f : X [0, ] such that
Z
f d.
(A) =
A
i=1
i=1
Moreover, if B1 , B2 are nondisjoint open balls and the radius of B1 is greater than
b1 B2 .
or equal to the radius of B2 , then B
65
i=1
It follows that
N
[
i=1
Bi
Bi
M
[
b .
B
j
j=1
M
M
X
X
b
Bj ,
Bj = 3n
i=1
i=1
1
|Br (x)|
Br (x)
|f (y)| dy.
The use of centered open balls to define the maximal function is for convenience.
We could use non-centered balls or other sets, such as cubes, to define the maximal
function. Some restriction on the shapes on the sets is, however, required; for
example, we cannot use arbitrary rectangles, since averages over progressively longer
and thinner rectangles about a point whose volumes shrink to zero do not, in
general, converge to the value of the function at the point, even if the function is
continuous.
Note that any two functions that are equal a.e. have the same maximal function.
Example 6.3. If f : R R is the step function
1 if x 0,
f (x) =
0 if x < 0,
then
M f (x) =
1
if x > 0,
1/2 if x 0.
66
6. DIFFERENTIATION
|x|n Ba (0)
where C > 0. The function 1/|x|n is not integrable on Rn \ Ba (0), so if M f is
integrable then we must have
Z
|f (y)| dy = 0
Ba (0)
2x 0 y log2 y
1
2x| log x|
so M f
/ L1loc (R).
67
C
.
t
An estimate of this form arises for integrable function from the following, almost
trivial, Chebyshev inequality.
Theorem 6.7 (Chebyshevs inequality). Suppose that (X, A, ) is a measure
space. If f : X R is integrable and 0 < t < , then
(6.6)
1
kf kL1 .
t
1
x
for x 6= 0 satisfies
2
,
t
so f belongs to weak-L1 (R), but f is not integrable or even locally integrable.
{x R : |f (x)| > t} =
C
kf kL1
t
68
6. DIFFERENTIATION
(K)
C
t
Rn
|f (y)| dy.
balls {B1 , B2 , . . . , BM
} such that
(K)
N
X
i=1
3n
|Bi |
M
X
j=1
|Bj |
M Z
n X
3
t
3n
t
n
which proves the result with C = 3 .
j=1
Bj
|f | dx
|f | dx,
r0+
1
|Br (x)|
Br (x)
|f (y) f (x)| dy = 0.
Proof. Since
Z
Z
1
1
f (y) f (x) dy
|f (y) f (x)| dy,
|Br (x)| Br (x)
|Br (x)| Br (x)
|f (y) f (x)| dy .
f (x) = lim sup
|Br (x)| Br (x)
r0+
69
(f g) f + g = f ,
f = (f g + g) (f g) + g = (f g) ,
which shows that (f g) = f .
If f L1 (Rn ), then we claim that there is a constant C, depending only on n,
such that for every 0 < t <
({x Rn : f (x) > t})
(6.7)
C
kf kL1 .
t
#
Z
1
f (x) sup
|f (y) f (x)| dy
r>0 |Br (x)| Br (x)
#
"
Z
1
|f (y)| dy + |f (x)|
sup
r>0 |Br (x)| Br (x)
M f (x) + |f (x)|.
It follows that
{f > t} {M f + |f | > t} {M f > t/2} {|f | > t/2} .
By the Hardy-Littlewood theorem,
({x Rn : M f (x) > t/2})
2 3n
kf kL1 ,
t
2
kf kL1 .
t
Combining these estimates, we conclude that (6.7) holds with C = 2 (3n + 1).
70
6. DIFFERENTIATION
Finally suppose that f L1 (Rn ) and 0 < t < . From Theorem 4.27, for any
> 0, there exists g Cc (Rn ) such that kf g||L1 < . Then
({x Rn : f (x) > t}) = ({x Rn : (f g) (x) > t})
C
kf gkL1
t
C
.
t
Since > 0 is arbitrary, it follows that
({x Rn : f (x) > t}) = 0,
{x Rn : f (x) > 0} =
that
This proves the result.
k=1
The set of points x for which the limits in Theorem 6.10 exist for a suitable
definition of f (x) is called the Lebesgue set of f .
Definition 6.11. If f L1loc (Rn ), then a point x Rn belongs to the Lebesgue
set of f if there exists a constant c R such that
"
#
Z
1
lim
|f (y) c| dy = 0.
r0+ |Br (x)| Br (x)
If such a constant c exists, then it is unique. Moreover, its value depends only
on the equivalence class of f with respect to pointwise a.e. equality. Thus, we can
use this definition to give a canonical pointwise a.e. representative of a function
f L1loc (Rn ) that is defined on its Lebesgue set.
Example 6.12. The Lebesgue set of the step function f in Example 6.3 is
R \ {0}. The point 0 does not belong to the Lebesgue set, since
Z r
1
1
|f (y) c| dy = (|c| + |1 c|)
lim
+
2r r
2
r0
is nonzero for every c R. Note that the existence of the limit
Z r
1
1
lim
f (y) dy =
2
r0+ 2r r
71
X
[
(Ai ).
Ai =
i=1
i=1
We say
that
Sthat a signed measure is finite if it takes only finite values.PNote
(A
since ( i=1 Ai ) does not depend on the order of the Ai , the sum
i)
i=1
converges unconditionally if it is finite, and therefore it is absolutely convergent.
Signed measures have the same monotonicity property (1.1) as measures, with
essentially the same proof. We will always refer to signed measures explicitly, and
measure will always refer to a positive measure.
Example 6.14. If (X, A, ) is a measure space and + , : A [0, ] are
measures, one of which is finite, then = + is a signed measure.
Example 6.15. If (X, A, ) is a measure space and f : X R is an Ameasurable function whose integral with respect to is defined as an extended real
number, then : A R defined by
Z
(6.8)
(A) =
f d
A
We will show that any signed measure can be decomposed into a difference of
singular measures, called its Jordan decomposition. Thus, Example 6.14 includes all
signed measures. Not all signed measures have the form given in Example 6.15. As
we discuss this further in connection with the Radon-Nikodym theorem, a signed
measure of the form (6.8) must be absolutely continuous with respect to the
measure .
6.7. Hahn and Jordan decompositions
To prove the Jordan decomposition of a signed measure, we first show that a
measure space can be decomposed into disjoint subsets on which a signed measure
is positive or negative, respectively. This is called the Hahn decomposition.
Definition 6.16. Suppose that is a signed measure on a measurable space
X. A set A X is positive for if it is measurable and (B) 0 for every
measurable subset B A. Similarly, A is negative for if it is measurable and
(B) 0 for every measurable subset B A, and null for if it is measurable and
(B) = 0 for every measurable subset B A.
72
6. DIFFERENTIATION
Because of the possible cancelation between the positive and negative signed
measure of subsets, (A) > 0 does not imply that A is positive for , nor does
(A) = 0 imply that A is null for . Nevertheless, as we show in the next result, if
(A) > 0, then A contains a subset that is positive for . The idea of the (slightly
tricky) proof is to remove subsets of A with negative signed measure until only a
positive subset is left.
Lemma 6.17. Suppose that is a signed measure on a measurable space (X, A).
If A A and 0 < (A) < , then there exists a positive subset P A such that
(P ) > 0.
Proof. First, we show that if A A is a measurable set with |(A)| < ,
then |(B)| < for every measurable subset B A. This is because takes
at most one infinite value, so there is no possibility of canceling an infinite signed
measure to give a finite measure. In more detail, we may suppose without loss of
generality that : A [, ) does not take the value . (Otherwise, consider
.) Then (B) 6= ; and if B A, then the additivity of implies that
(B) = (A) (A \ B) 6=
since (A) is finite and (A \ B) 6= .
Now suppose that 0 < (A) < . Let
1 = inf {(E) : E A and E A} .
Then 1 0, since A. Choose A1 A such that 1 (A1 ) 1 /2
if 1 is finite, or (A1 ) 1 if 1 = . Define a disjoint sequence of subsets
{Ai A : i N} inductively by setting
o
n
S
i1
A
i = inf (E) : E A and E A \
j
j=1
S
i1
and choosing Ai A \
j=1 Aj such that
i (Ai )
1
i
2
if < i 0, or (Ai ) 1 if i = .
Let
[
Ai ,
P = A \ B.
B=
i=1
(Ai ).
i=1
As proved above, (B) is finite, so this negative sum must converge. It follows that
(Ai ) 1 for only finitely many i, and therefore i is infinite for at most finitely
many i. For the remaining i, we have
X
1X
(Ai )
i 0,
2
P
so
i converges and therefore i 0 as i .
73
Ai
i=1
so P \ P is both positive and negative for and therefore null, and similarly for
P \ P . Thus, the decomposition is unique up to -null sets.
To describe the corresponding decomposition of the signed measure into the
difference of measures, we introduce the notion of singular measures, which are
measures that are supported on disjoint sets.
Definition 6.19. Two measures , on a measurable space (X, A) are singular, written , if there exist sets M, N A such that M N = , M N = X
and (M ) = 0, (N ) = 0.
Example 6.20. The -measure in Example 2.36 and the Cantor measure in
Example 2.37 are singular with respect to Lebesgue measure on R (and conversely,
since the relation is symmetric).
Theorem 6.21 (Jordan decomposition). If is a signed measure on a measurable space (X, A), then there exist unique measures + , : A [0, ], one of
which is finite, such that
= +
and + .
74
6. DIFFERENTIATION
(A) = (A N )
The next result clarifies the relation between Definition 6.22 and the absolute
continuity property of integrable functions proved in Proposition 4.16.
Proposition 6.25. If is a finite signed measure and is a measure, then
if and only if for every > 0, there exists > 0 such that |(A)| <
whenever (A) < .
Proof. Suppose that the given condition holds. If (A) = 0, then |(A)| <
for every > 0, so (A) = 0, which shows that .
Conversely, suppose that the given condition does not hold. Then there exists
> 0 such that for every k N there exists a measurable set Ak with ||(Ak )
and (Ak ) < 1/2k . Defining
[
\
B=
Aj ,
k=1 j=k
we see that (B) = 0 but ||(B) , so is not absolutely continuous with respect
to .
75
The Radon-Nikodym theorem provides a converse to Example 6.24 for absolutely continuous, -finite measures. As part of the proof, from [4], we also show
that any signed measure can be decomposed into an absolutely continuous and
singular part with respect to a measure (the Lebesgue decomposition of ). In
the proof of the theorem, we will use the following lemma.
Lemma 6.26. Suppose that , are finite measures on a measurable space
(X, A). Then either , or there exists > 0 and a set P such that (P ) > 0
and P is a positive set for the signed measure .
Proof. For each n N, let X = Pn Nn be a Hahn decomposition of X for
the signed measure n1 . If
P =
Pn
N=
n=1
Nn ,
n=1
1
(N )
n
for every n N, so (N ) = 0. Thus, either (P ) = 0, when , or (Pn ) > 0
for some n N, which proves the result with = 1/n.
0 (N )
where a and s .
C = A {x X : g(x) h(x)} ,
76
6. DIFFERENTIATION
and therefore
Z
A
max {g, h} d =
Let
g d +
m = sup
Z
g d : g F
(X).
so f F and
f d = m.
Define s : A [0, ) by
s (A) = (A)
f d.
A
ZA
f d + (A P )
ZA
(f + P ) d.
i=1
Ai
77
into a countable disjoint union of sets with (Ai ) < and (Ai ) < . We
decompose the finite measure i = |Ai as
i = ia + is
where ia i and is i .
ia ,
s =
ia
i=1
i=1
d
.
d
xA
There are generalizations of the Radon-Nikodym theorem which apply to measures that are not -finite, but we will not consider them here.
6.9. Complex measures
Complex measures are defined analogously to signed measures, except that they
are only permitted to take finite complex values.
Definition 6.29. Let (X, A) be a measurable space. A complex measure on
X is a function : A C such that:
(a) () = 0;
(b) if {Ai A : i N} is a disjoint collection of measurable sets, then
!
X
[
(Ai ).
Ai =
i=1
i=1
where a and s .
78
6. DIFFERENTIATION
for every A A.
To prove the result, we decompose a complex measure into its real and imaginary parts, which are finite signed measures, and apply the corresponding theorem
for signed measures.
CHAPTER 7
Lp spaces
In this Chapter we consider Lp -spaces of functions whose pth powers are integrable. We will not develop the full theory of such spaces here, but consider only
those properties that are directly related to measure theory in particular, density, completeness, and duality results. The fact that spaces of Lebesgue integrable
functions are complete, and therefore Banach spaces, is another crucial reason for
the success of the Lebesgue integral. The Lp -spaces are perhaps the most useful
and important examples of Banach spaces.
7.1. Lp spaces
For definiteness, we consider real-valued functions. Analogous results apply to
complex-valued functions.
Definition 7.1. Let (X, A, ) be a measure space and 1 p < . The space
Lp (X) consists of equivalence classes of measurable functions f : X R such that
Z
|f |p d < ,
where two measurable functions are equivalent if they are equal -a.e. The Lp -norm
of f Lp (X) is defined by
Z
1/p
p
.
kf kLp =
|f | d
X
|xn |p < .
n=1
X
p
k{xn }kp =
.
|xn |
n=1
The space L (X) is defined in a slightly different way. First, we introduce the
notion of esssential supremum.
79
7. Lp SPACES
80
Equivalently,
ess sup f = inf sup g : g = f pointwise a.e. .
X
Thus, the essential supremum of a function depends only on its -a.e. equivalence
class. We say that f is essentially bounded on X if
ess sup |f | < .
X
Definition 7.4. Let (X, A, ) be a measure space. The space L (X) consists
of pointwise a.e.-equivalence classes of essentially bounded measurable functions
f : X R with norm
kf kL = ess sup |f |.
X
so k kLp is not a norm in that case. Nevertheless, for 0 < p < 1 we have
|f + g|p |f |p + |g|p ,
7.4. COMPLETENESS
81
7.3. Density
Density theorems enable us to prove properties of Lp functions by proving them
for functions in a dense subspace and then extending the result by continuity. For
general measure spaces, the simple functions are dense in Lp .
Theorem 7.8. Suppose that (X, A, ) is a measure space and 1 p .
Then the simple functions that belong to Lp (X) are dense in Lp (X).
Proof. It is sufficient to prove that we can approximate a positive function
f : X [0, ) by simple functions, since a general function may be decomposed
into its positive and negative parts.
First suppose that f Lp (X) where 1 p < . Then, from Theorem 3.12,
there is an increasing sequence of simple functions {n } such that n f pointwise.
These simple functions belong to Lp , and
p
|f n | |f |p L1 (X).
n
X
c i Ai
i=1
belongs to Lp for 1 p < if and only if (Ai ) < for every Ai such that
ci 6= 0, meaning that its support has finite measure. On the other hand, every
simple function belongs to L .
For suitable measures defined on topological spaces, Theorem 7.8 can be used to
prove the density of continuous functions in Lp for 1 p < , as in Theorem 4.27
for Lebesgue measure on Rn . We will not consider extensions of that result to more
general measures or topological spaces here.
7.4. Completeness
In proving the completeness of Lp (X), we will use the following Lemma.
Lemma 7.9. Suppose that X is a measure space and 1 p < . If
{gk Lp (X) : k N}
X
kgk kLp < ,
k=1
p
X
gk = f
k=1
7. Lp SPACES
82
n
X
k=1
|gk | ,
h=
k=1
|gk | .
n
X
k=1
kgk kLp M
P
where k=1 kgk kLp = M . It follows that h Lp (X) withPkhkLp M , and in
particular that h is finite pointwise a.e. Moreover, the sum k=1 gk is absolutely
convergent pointwise a.e., so it converges pointwise a.e. to a function f Lp (X)
with |f | h. Since
p
!p
n
n
X
X
|gk |
(2h)p L1 (X),
gk |f | +
f
k=1
k=1
meaning that
k=1
gk converges to f in Lp .
The following theorem implies that Lp (X) equipped with the Lp -norm is a
Banach space.
Theorem 7.10 (Riesz-Fischer theorem). If X is a measure space and 1 p
, then Lp (X) is complete.
Proof. First, suppose that 1 p < . If {fk : k N} is a Cauchy sequence
in Lp (X), then we can choose a subsequence {fkj : j N} such that
fkj+1 fkj
p 1 .
L
2j
X
j=1
X
j=1
gj
7.5. DUALITY
83
X
X
gj = f
gi = fk1 +
lim fkj = lim fk1 +
j
j=1
i=1
1
m
c
for all j, k n and x Nj,k,m
Nj,k,m .
j,k,mN
Then N is a null set, and for every x N c the sequence {fk (x) : k N} is Cauchy
in R. We define a measurable function f : X R, unique up to pointwise a.e.
equivalence, by
f (x) = lim fk (x)
for x N c .
k
7. Lp SPACES
84
with Lp (X) for 1 < p < . Under a -finiteness assumption, it is also true that
1
H
olders inequality implies that functions in Lp define bounded linear funcp
tionals on L with the same norm, as stated in the following proposition.
Proposition 7.13. Suppose that (X, A, ) is a measure space and 1 < p .
If f Lp (X), then
Z
F (g) = f g d
In proving the reverse inequality, we may assume that f 6= 0 (otherwise the result
is trivial).
First, suppose that 1 < p < . Let
p /p
|f |
g = (sgn f )
.
kf kLp
7.5. DUALITY
85
It follows that
kF kL1 kf kL ,
(7.2)
J : L (X) L (X) ,
J(f ) : g 7 f g d,
is an isometry from Lp into Lp . The main part of the following result is that J is
onto when 1 < p < , meaning that every bounded linear functional on Lp arises
p < , then (7.2) defines an isometric isomorphism of Lp (X) onto the dual space
p
of L (X).
Proof. We just have to show that the map J defined in (7.2) is onto, meaning
i=1
Ai ,
i=1
X
X
X
(Ai ),
F (Ai ) =
Ai =
(A) = F (A ) = F
i=1
i=1
i=1
7. Lp SPACES
86
where M = kF kLp .
Taking = sgn f , which is a simple function, we see that f L1 (X). We may
then extend the integral of f against bounded functions by continuity. Explicitly,
if g L (X), then from Theorem 7.8 there is a sequence of simple functions {n }
with |n | |g| such that n g in L , and therefore also in Lp . Since
|f n | kgkL |f | L1 (X),
the dominated convergence theorem and the continuity of F imply that
Z
Z
F (g) = lim F (n ) = lim
f n d = f g d,
n
and that
(7.3)
Z
f g d M kgkLp
a priori that f Lp .
Let {n } be a sequence of simple functions such that
n f
pointwise a.e. as n
and |n | |f |. Define
gn = (sgn f )
|n |
kn kLp
p /p
M.
7.5. DUALITY
87
Defining B =
n=1
as n .
define Fn (N) by
Fn (x) =
1X
xi ,
n i=1
meaning that Fn maps a sequence to the mean of its first n terms. Then
kFn k = 1
7. Lp SPACES
88
for every n N, so by the Alaoglu theorem on the weak- compactness of the unit
ball, there exists a subsequence {Fnj : j N} and an element F (N) with
X
xi yi
for every x .
F (x) =
i=1
Then, denoting by ek the sequence with kth component equal to 1 and all
other components equal to 0, we have
1
=0
yk = F (ek ) = lim Fnj (ek ) = lim
j nj
j
Bibliography
[1] V. I. Bogachev, Measure Theory, Vol I. and II., Springer-Verlag, Heidelberg, 2007.
[2] D. L. Cohn, Measure Theory, Birkh
auser, Boston, 1980.
[3] L. C. Evans and R. F. Gariepy, Measure Theory and Fine Properties of Functions, CRC
Press, Boca Raton, 1992.
[4] G. B. Folland, Real Analysis, 2nd ed., Wiley, New York, 1999.
[5] F. Jones, Lebesgue Integration on Euclidean Space, Revised Ed., Jones and Bartlett, Sudberry, 2001.
[6] S. Lang, Real and Functional Analysis, Springer-Verlag, 1993.
[7] E. H. Lieb and M. Loss, Analysis, AMS 1997.
[8] E. M. Stein and R. Shakarchi, Real Analysis, Princeton University Press, 2005.
[9] M. E. Taylor, Measure Theory and Integration, American Mathematical Society, Providence,
2006.
[10] S. Wagon, The Banach-Tarski Paradox, Encyclopedia of Mathematics and its Applications,
Vol. 24 Cambridge University Press, 1985.
[11] R. L. Wheeler and Z. Zygmund, Measure and Integral, Marcel Dekker, 1977.
89