2017-01-20-015145 @R Basicmeasuretheory - Skript PDF
2017-01-20-015145 @R Basicmeasuretheory - Skript PDF
2017-01-20-015145 @R Basicmeasuretheory - Skript PDF
Heinrich v. Weizscker
Revised Translation
2004/2008
Fachbereich Mathematik
Technische Universitt Kaiserslautern
Preface
The aim of this text is to give a short introduction into the basic concepts
of the Lebesgue theory of measure and integration on general spaces in order to
enable the reader to follow Master courses in Probability and in Functional analysis. The main advantage of Lebesgues theory over more elementary concepts
like the Riemann integral is its power in dealing with countable limit operations,
both for the measure (viz. generalized volume) of sets, and for the integral of
functions. These become essential features in higher real analysis and probability. The price is some additional work in constructing suitable domains of
definition for measure and integral.
There are two main approaches to this theory. The first approach starts with
measures and then constructs the integrals, in analogy to the classical passage
from volume to integration. Conversely, the second approach constructs in
several steps the space of integrable functions and then a set has finite measure
simply if the indicator function of this set is integrable. In this text we follow
the first line, even though it takes a little longer. The main reason for this choice
is that in Probability the measurable sets (events) have an important role by
themselves and this path of arguments gives additional information about them.
The common goal of both approaches are the limit theorems of chapter 4 (in
particular the dominated convergence theorem). A reader who knows these limit
theorems can start directly with chapter 5 and check the definitions and results
of the first four chapters only when needed.
Up to minor modifications this text is a translation of the original German
version [?]. Only chapter 2 has been revised more thoroughly. For most mathematical words and phrases we introduce both the English and German term.
Contents
1 Measures
1.1 -Algebras . . . . . . . . . . . . . . .
1.2 Measures . . . . . . . . . . . . . . .
1.3 Null-Sets . . . . . . . . . . . . . . . .
1.4 An example of a non-measurable set
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
5
5
7
8
9
2 Construction of measures
2.1 Set algebras and contents . . . . . . . . .
2.2 Outer measures . . . . . . . . . . . . . . .
2.3 Proof of the Measure Extension Theorem
2.4 Monotone classes and more on uniqueness
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
11
11
14
16
19
3 Measurable functions
.
.
.
.
.
.
.
.
21
4 The Integral
25
4.1 The elementary integral . . . . . . . . . . . . . . . . . . . . . . . 25
4.2 The Integral for non-negative measurable and for integrable functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.3 The Limit Theorems . . . . . . . . . . . . . . . . . . . . . . . . . 30
5 Fubinis Theorem
33
6 Convergence in measure
37
6.1 Comparison with other types of convergence . . . . . . . . . . . . 37
6.2 A Metric for Local Convergence in Measure . . . . . . . . . . . . 39
7 The Lp -spaces
43
Chapter 1
Measures
1.1
-Algebras
A measure measures the size of sets. Not all sets can be measured in this sense
as we shall see in section 1.4. On the other hand the class of sets which are measurable should be sufficiently rich. In particular we want to keep measurability
of sets if we perform simple operations like taking the complement or taking
(countable) unions and intersections. This leads to the following definition.
Definition 1.1 Let S be a set and let F be a collection of subsets of S. Then
F is called a -algebra over S, if the following holds:
a) S F,
b) If A F then Ac F. 1
S
c) For every sequence A1 , A2 , of elements of F the union n=1 An is also a
member (i.e. an element) of F.
Remark: 1. Properties a) and b) imply that the empty set belongs also
to F, since = S c .
2. Properties b) and
c) imply that for every sequence A1 , A2 , of elements of
T
F the S
intersection n=1 An is in F, because first Acn F for each n due to b),
hence n Acn F due to c) and finally b) yields
An =
n=1
!c
Acn
F.
n=1
S one can consider their intersection, i.e. the system of subsets of S which are
members of each of the given -algebras.
Theorem 1.1 LetT(Fi )iI be an arbitrary family of -algebras over S. Then
their intersection iI Fi is again a -algebra over S.
Proof. We should verify that the intersection has the three properties a)-c)
in Definition 1.1. We restrict
ourselves to condition c). Let A1 , A2 , be a
T
sequence of elements of S
F
. Thus An Fi for all n N and all i I. For
i
iI
fixed i I it follows
that
A
nN
S
T n Fi since Fi is a -algebra. This applies to
each i, and hence nN An iI Fi .
As a consequence for every system E of subsets of S there is a smallest
-algebra which contains E.
Definition 1.2 Let E be any system of subsets of S. The intersection of the
family of all -algebras which contain E is called the -algebra generated by
E. It will be denoted by (E) and E is called a generating system (Erzeugendensystem) of this -algebra.
Remarks 1. Thus a set A belongs to (E) if and only every -algebra which
contains all elements of E must also contain A. This concept is similar to the
concept of the linear hull of a set E of vectors in a vector space V , which is the
smallest vector space containing all vectors in E or, the intersection of all linear
subspaces of V which contain E as a subset. But while the elements of the linear
hull can be described explicitly as the linear combinations of the elements of E,
an analogous explicit representation of the elements of (E) by the elements of
E is not possible in general.
2. The definition shows: If two systems of sets E1 , E2 satisfy E1 (E2 ) then
we have even (E1 ) (E2 ), because the -algebra (E2 ) is by assumption one
of the -algebras whose intersection is (E1 ). In particular the larger a system
of sets, the larger the -algebra it generates.
Example 1.1 Let Q be the system of all compact boxes Q = [r1 , s1 ]
[rm , sm ] in Rm . Then every open set U Rm is an element of the -algebra
(Q).
Proof. Each point x U is in a cube of the form Q = [r1 , s1 ] [rm , sm ]
with rational numbers ri , si , such that even the whole cube Q is a subset of U .
The set of all cubes with rational sides which are contained inSU is countable,
1.2
Measures
[
X
Fn =
(Fn ).
(1.1)
n=1
n=1
[
X
An
(An ).
n=1
n=1
[
[
X
(A) =
An =
Fk =
(Fk )
=
n=1
n
X
lim
k=1
k=1
k=1
1.3
Null-Sets
Null-sets are very useful because for most purposes one can ignore everything
that happens on them.
Definition 1.6 We call a set A S null-set (Nullmenge) or, more precisely,
a -null-set if there is some N F with (N ) = 0 and A N .
(1.3)
The symmetric difference consists of precisely those points at which the two
sets differ, i.e. which are in one of the two sets but not in the other. If the
symmetric difference A M B of two sets is a -null-set then A and B are called
-equivalent. If moreover both sets are measurable then (AB) = (AB) =
(A) = (B).
Definition 1.7 A measure , and the associated measure space (S, F, ) are
complete (vollstndig), if every -null set belongs to the -algebra F.
We note in passing that every measure can be completed.
Proposition 1.1 Let (S, F, ) be a measure space. Let
F = {A0 S : A0 is equivalent to some A F}.
Extend to F by letting (A0 ) = (A) if A and A0 are -equivalent. Then
(S, F , ) is a complete measure space. It is called the completion of (S, F, ),
and the -algebra F is the -completion of F.
Proof. The discussion preceding definition 1.7 shows that the extension of
to F is unambiguous. Two sets A0 and A F differ at exactly the same
points at which the complements A0c and Ac differ. This implies that F is
stable under complements. If (A0n ) is a sequence of sets in F then there are
sets An F such that A0n MSAn Nn F where (N
S n ) = 0 for each n. All
points at which the setSA0 = n A0n differs
from
A
=
n An F are contained
S
in the -null-set N = Nn and thus n A0n F . So F is a -algebra. If the
A0n are disjoint then the sets An N c = A0n N c are also disjoint and hence
X
n=1
(A0n ) =
X
n=1
(An ) =
n=1
1.4
The concept of a -algebra serves primarily for having suitable domains of definition for measures. Why not define a measure on all subsets A of the
underlying space? In this section we want to show that this is not possible for
Lebesgue measure on the real line, more precisely for a finite measure on an
interval with the very natural additional condition of translation-invariance.
10
Theorem 1.4 (Vitali set) There is a set E [0, 1] with the following property: Let be a measure on a -algebra L over [0, 1] such that ([0, 1]) = 1 and
for all F L we have F + x mod 1 L for x [0, 1] and
(F + x
Then E
/ L.
mod 1) = (F )
(1.4)
The right-hand side is positive and finite. But on the left we have added infinitely many times the same value (E). Both for (E) = 0 and for (E) > 0
we get a contradiction.
Remark 1.1 In the above situation every L-measurable subset F of E must be
a null set.
Proof. The sets Fr = F + r mod 1 are disjoint and therefore as above
[
X
X
(F ) =
(Fr ) =
Fr ([0, 1]) = 1
r
and hence (F ) = 0.
2 i.e.
E is not Lebesgue-measurable
Chapter 2
Construction of measures
Even though it is clear how to measure the size of simple sets by volume or
similar finitely additive functionals, it is not obvious that these definitions can
be extended to a measure on a whole -algebra of sets. The main result in this
chapter is Carathodorys extension theorem (theorem 2.2). As a special case
this contains the construction of Lebesgue measure in Euclidean space.
There are several approaches to this theorem. Our proof is based on an
approximation argument. It can easily be reframed to work also in the functional
analytic context.1
2.1
[
n=1
An = A1
(An (
n=2
n1
[
Am )c ).
m=1
This shows that an algebra is a -algebra if and only if it is closed under forming
countable disjoint sequences.
1 Carathodorys original proof is useful in the construction of Hausdorff measures in the
theory of fractals, cf. e.g. the script [?] Fractal sets and Preparation to Geometric Measure
Theory (http://www.mathematik.uni-kl.de/ wwwstoch/2002w/skriptrev.pdf). A proof via
approximation from inside is also possible, see the script [?]
11
12
Example 2.1 Let A be the set algebra on R which is generated by all intervals
which are closed on the right and open on the left. Then A consists precisely of
all sets of the form
n
[
A=
Ji
(2.1)
i=1
where each Ji is an interval of one of the three types (, b], (a, b] or (a, ).
It is called the interval algebra over R.
Sometimes it is convenient not to require that the underlying set S itself
belong to the system of sets. The local version of a set algebra is called a
ring:
Definition 2.2 Let S be a set and let R be a non-empty system of subsets of
S. Then R is called a ring ((Mengen-)Ring) on S, if the following hold:
b0 ) For A, B R the difference set A \ B = A B c also belongs to R.
c0 ) For A, B A the union A B is in R.
A ring R is also stable under the operation M (cf. 1.2). Moreover (1.3) shows
that A B = (A B) \ (A M B) and hence a ring is stable under intersections. If
one takes M as addition and as multiplication then a ring in our sense becomes
a commutative ring in the algebraic sense.2
Example 2.2 An example of a ring is the interval ring R of all bounded sets
in the interval algebra A of Example 2.1.
Obviously a ring is a set algebra if and only if it has the set S as an element.
Definition 2.3 Let S be a set and let R be a ring of subsets of S. A map
m : A [0, ] is a content (Inhalt), if m() = 0 and m is additive, i.e.
m(A B) = m(A) + m(B)
(2.2)
(2.3)
In fact both sides are equal to m(A) + m(B \ A) + m(A B). A set function
with (2.3) is called modular.
2 Here is a short way to see this: identify a set A with its indicator function 1
A and note
that 1AB = 1A 1B and 1AMB 1A + 1B (mod 2). A set algebra becomes a commutative
ring with 1 (the set S is the unit). In older literature set algebras are called fields (MengenKrper). The associated algebraic structures are called Boolean rings resp. algebras.
13
Example 2.3 Let A (resp. R) be the above algebra (resp. ring) of unions of
intervals. One can assume that in the representation (2.1) the intervals Ji are
disjoint. If F : R R is a non-decreasing function then we get a content mF
on A resp. R by the formula
mF (
n
[
Ji ) =
i=1
n
X
F (bi ) F (ai ),
(2.4)
i=1
where for unbounded intervals we use the limit values F () = limx F (x)
and F () = limx F (x) in the obvious way. It is called Stieltjes content.
For F (x) = x we get the usual geometric size of a set based on the length of
intervals.
Clearly if a content can be extended to a measure it must be -additive for
all sets for which it is defined.
Definition 2.4 A content m on a ring R is called -additive if for every
disjoint sequence (An ) in R whose union is in R, one has
m(
An ) =
n=1
m(An ).
(2.5)
n=1
n
X
m(Ak ) = m(A) m(
k=1
n
[
Ak ) = m(Rn )
k=1
N
[
n=1
Bnc
N
[
n=1
Rn \ Bn
14
N
X
m(Rn ) m(Bn )
n=1
N
X
2n < .
n=1
2.2
Outer measures
We saw that typically measures cannot be defined on all subsets of the underlying space. If we relax the condition of -additivity, this difficulty disappears.
Definition 2.6 Let S be a set. A map from the power set 2S to [0, ]
is called an outer measure (ueres Ma) on S, if it has the following
properties:
1. () = 0,
2. A B implies (A) (B) (monotonicity),
15
S
P
3. ( n=1 En ) n=1 (En ) (-sub-additivity).
Outer measures often are generated as follows.
Proposition 2.1 Let R be any system of subsets of S which contains , and
let m : R [0, ] be a set function with m() = 0. Let
(A) := inf{
m(Ri ) : Ri R, A
i=1
Ri }
(2.6)
m(Rin ) (En ) + 2n .
i=1
S
n=1
En )
n=1
m(Rin )
i,n=1
S
i,n=1
(En ) + .
n=1
Thus all properties of an outer measure are verified. The final inequality in the
assertion is clear since every R R forms an admissible covering of itself on the
right hand side of (2.6).
In the special case of a -additive content on a ring R the restriction of the
outer measure defined in (2.6) to R actually coincides with m. This will be
important below.
Lemma 2.1 Let R be a ring and let m be a -additive content. Then
|R = m.
S
Proof.
Let A R be given. Suppose A i Ri and Ri R. Then
S
A = i A Ri . As m is -sub-additive and monotone we get
m(A)
X
i=1
m(A Ri )
m(Ri ).
i=1
Taking the infimum over these coverings gives m(A) (A). The converse
inequality was part of the proposition.
Finally every outer measure induces a distance (Abstand) between subsets
of S.
16
(2.7)
(2.8)
holds. Moreover the function and the set operations , , \ are continuous
for d: If both d(An , A) 0 and d(Bn , B) 0 then (An ) (A), d(An
Bn , A B) 0, and similarly for and \.
Proof. Obviously for three sets A, B, C one has A M C (A M B) (B M
C). Since is monotone and sub-additive we get the triangular inequality.
The triangular inequality yields in particular the inequality in
| (A) (B)| = |d(, A) d(, B)| d(A, B) = (A M B)
(2.9)
which in turn implies the continuity of . Let now denote one of the above
set operations. Let A, B and A0 , B 0 be two pairs of sets. If a point x belongs to
A B but not to A0 B 0 or conversely, then x is at least in one of the symmetric
difference sets A M A0 or B M B 0 . More formally
(A B) M (A0 B 0 ) (A M A0 ) (B M B 0 ).
Then again the fact that is monotone and sub-additive implies
d(A B, A0 B 0 ) d(A, A0 ) + d(B, B 0 ).
This proves the continuity.
2.3
Actually the proof of theorem 2.2 yields more detailed information which sometimes is useful. The existence proof does not need -finiteness. It yields a
complete measure with a useful approximation property. The uniqueness part
of theorem 2.2 will be shown in remark 2.2 below. A much stronger uniqueness
result is established in the next (optional) section.
Theorem 2.3 Let m be a -additive content on the ring R. Let be the
associated outer measure given by proposition 2.1. Let N = {N S : (N ) =
0} be the class of -null-sets. Let M = (R N ) and let = |M be the
restriction of to the -algebra M. Then
a) is a complete measure which extends m i.e. |R = m.
b) For every A M with (A) < there is a sequence (An ) in R with
(A M An ) 0.
Proof. I 1. Let d denote the distance on the power set of S which is
associated to , cf. (2.7). Let Rf = {R R : m(R) < } and let Mf
be the system of all sets A S which are in the closure of Rf with respect
to d, i.e. A Mf iff there is a sequence (Rn ) in Rf with d(Rn , A) 0; it
17
=
=
=
=
lim m(An Bn )
lim m(An ) + lim m(Bn ) lim m(An Bn )
lim (An ) + lim (Bn ) lim (An Bn )
(A) + (B) + () = (A) + (B).
4. We claim that |Mf is -additive, and moreover for every disjoint seS
quence
(An ) in Mf the union A = An also belongs to Mf if and only if
P
n=1 (An ) < . Clearly if A Mf (or more generally if (A) < ) then
by step 2 the partial sums of the series are all dominated by (A) and the
series converges. Conversely, assume that the series converges. Then
d(A,
N
[
An ) = (A M
n=1
N
[
An ) = (
n=1
An )
n=N +1
(An ) 0
n=N +1
N
[
An ) = lim
n=1
N
X
n=1
(An ) =
(An ).
n=1
6. We note in passing that in the case where R is an algebra and m(S) <
the proof is now complete: The class Mf then is an algebra according to step
2. By step 3, the series in the statement of step 4 always are bounded and thus
Mf is closed under countable disjoint unions. Hence Mf is a -algebra which
contains both R = Rf and N , i.e. (R N ) Mf . Further is -additive
on Mf , in particular it induces a measure extension of m on (R) which is
complete because N is closed under taking subsets.
3
II. 7. The remainder of the proof is devoted to the extension to the unbounded case. Consider the system of sets
= {A S : A B Mf f or all B Mf }.
M
(2.10)
18
is a -algebra. Clearly S M.
Also M
is a ring since
8. We claim that M
is an algebra. Let now a disjoint sequence (An ) in M
be
Mf is a ring. So M
S
given with union A = An . For every B Mf the additivity of on Mf
(step 3) gives
X
n=1
(An B) = lim
N
N
X
(An B) = lim (
N
n=1
N
[
An B) (B) <
n=1
We can now give the uniqueness argument for theorem 2.2: Let m be -finite
and let (Ek ) be an increasing sequence in R of finite content whose union is the
whole set S. Let A (R). Then
(A) = lim (A Ek ) = lim (A Ek ) = (A).
k
19
1
k
k
k
many
rectangles
R
such
that
N
R
and
i
i i
i vol(Ri ) < k . Since N
T S k
d
k=1
i=1 Ri (R) = B(R ) the class N consists precisely of the null-sets in
the sense of definition 1.6 for the measure d|B(Rd ) . The -algebra L = (R N )
is then easily seen to be the completion of the Borel sets for Lebesgue measure in
the sense of proposition 1.1. It is called the -algebra of Lebesgue measurable
sets. Every Lebesgue measurable set differs only by a Lebesgue null-set from a
Borel set.
Example 2.5 (Counting measure) Let R be the family of all finite subsets of
a set S. Let m assign to each finite set its cardinality. The corresponding outer
measure assigns to each infinite set the value +. The empty set is the only
null-set and (R) = (R N ) consists of all countable subsets of S and their
complements. Let us look into the proof of theorem 2.3 in this case. The system
Mf defined there agrees with the system of finite sets whereas the -algebra
defined in (2.10) coincides with the full power set. We see that the proof
M
actually sometimes produces a measure extension on a larger -algebra than
just on (R N ).
2.4
This section can be read independently from the rest of this chapter. The
monotone class theorem below sometimes is quite convenient in order to extend
a statement from a certain class of sets to the -algebra generated by this class.
We shall use it e.g. in the proof of Fubinis theorem (theorem 5.1).4
As a consequence of theorem 2.2 two measures on some -algebra F agree if
they agree on a ring R which generates F and on which they are -finite. Even
for finite measures one might ask: Is a finite measure on a -algebra F already
uniquely determined by its values on an arbitrary generating system E of F?
Example 2.6 Here is a counter-example even if the total mass of the measures
is fixed. Let S = {1, 2, 3, 4} and let the system E consist of the two sets {1, 2}
and {2, 3}. Then all four singletons {1}, {2}, {3}, {4} can easily be produced
via algebra operations from these two sets, and hence (E) is the power set of
S. The two weight assignments {1/4, 1/4, 1/4, 1/4} and {1/2, 0, 1/2, 0} clearly
induce two different measures and on the power set S with (S) = (S) = 1
and (A) = (A) = 1/2 for both sets A E.
We shall see in a minute that the essential requirement is that E be stable
under intersection. The proof uses the following concept.
Definition 2.7 A monotone class (monotone Klasse) over the set S is a
system D of subsets of S with the following three properties:
4 In
20
Chapter 3
Measurable functions
Definition 3.1 Let (S, F) and (Y, G) be measurable spaces. A map f : S Y
is called F-G-measurable (mebar) (or just: measurable), if
f 1 (G) F f or all G G.
In the special case when Y is a metric space a map f : S Y is called Fmeasurable or measurable, if f is F-B(Y )-measurable.
For functions with one-dimensional range often the values occur in limit
procedures. Therefore it is convenient to consider also extended real valued
functions. The following notation is convenient.
Notation: Let f, g : S [, ] be two functions and let a R. Then
denote by {f a} the set {x S|f (x) a} = f 1 ((; a]), by {f = g} the
set {x S|f (x) = g(x)} and introduce similarly the notations
{f < a}, {f a}, {f > a}, {f < g} etc.
Moreover we recall that the extended real line [, ] is a metric space
when endowed with the metric
d : [, ] [, ] [0, 1],
d(x, y) =
|x y|
.
1 + |x y|
22
2. The (extended) real-valued measurable functions are - with certain restrictions to be discussed later - the natural candidates for the definition of the
integral.
The verification of measurability is often much simplified by the following
criterion.
Lemma 3.1 A map f : S Y is F-G-measurable if f 1 (E) F for all E E
where E is any system of subsets of the space Y such that G = (E).
Proof. Consider the following class of subsets of Y
A = {A Y |f 1 (A) F}.
According to our assumption E A. Moreover since f 1 (Y ) = S F we
have Y A. Let A A. Then f 1 (Ac ) = S\f 1 (A) = (f 1 (A))c F, and
therefore Ac A. Finally let An A for all n N. Then
[
[
f 1 (
An ) =
f 1 (An ) F
nN
nN
23
F for a c
{f < a} =
S F for a > c
So the constant functions are measurable.
2. If f : Rm Rn is continuous, Then f 1 (U ) is open and hence Borel for all
open sets U . So by lemma 3.1 all continuous functions are measurable.
Lemma 3.3 Let (S, F ), (Y, G), (Z, H) be measurable spaces. Let f : Y Z, g :
S Y be measurable maps. Then the composition f g : S Z is measurable.
Proof. For H H we get f 1 (H) G and (f g)1 (H) = g 1 (f 1 (H)) F.
Lemma 3.4 A vector-valued function f = (f1 , . . . , fn ) : S Rm is F-measurable
if and only if the components fi : S R are F-measurable for all i {1, . . . , m}.
Proof. Let f be measurable. Then for each a R the set A = [, a]
Rm1 Rm is closed, i.e. A B(Rm ). Thus {f1 a} = f 1 (A) F for all
a R. Therefore f1 is measurable by lemma 3.2. A similar argument works for
the other components.
Conversely assume that fi is measurable for all i. Let a box Q = [a1 , b1 ] . . .
[an , bn ]. be given. Then
f 1 (Q) =
n
\
fi1 ([ai , bi ]) F
i=1
This holds for all boxes and lemma 3.1 together with theorem 1.2 implies the
measurability of f .
Corollary 3.1 If f1 , f2 : S R are measurable then the same is true for
f1 + f2 , f1 f2 , f1 f2 , min(f1 , f2 ) and max(f1 , f2 ).
Proof. The maps 0 +0 ,0 0 ,0 0 , min, max : R2 R are continuous. By remark
3.12 they are measurable. The map f = (f1 , f2 ) : S R2 is measurable
according to lemma 3.4. Since f1 +f2 = 0 +0 f the function f1 +f2 is measurable
by lemma 3.3. The same arguments work also for difference and multiplication.
Theorem 3.2 Let (S, F) be a measurable space and let {fn }nN be a sequence of
measurable extended real valued functions on S. a) Then the (pointwise defined)
functions supn fn , inf n fn , lim supn fn , lim inf n fn are measurable.
b) The set F = {x : lim fn (x) exists} is in F and limn 1F fn is measurable.
Proof. For all a R the set {sup fn a} = nN {fn a} is in F.
So sup fn is measurable. By symmetry inf fn is measurable. Therefore also the
functions lim supn fn = inf m (supnm fm ) and lim sup fn = supm (inf nm fn )
are measurable and the set F = {lim supn fn = lim inf n fn } is in F.
Simple examples of real-valued measurable functions are the indicator functions 1A , A F. In fact the sets {1A > a} are in F because they are either
empty (if a 1), equal to A (if 0 a < 1), or equal to S (if a < 0).
24
Definition
Pn 3.3 Let (S, F) be a measurable space. A function of the form
f (x) =
i=1 i 1Ai (x) with i R, Ai F is called an elementary function or measurable step function (mebare Treppenfunktion).
Remark 3.2 Clearly the measurable step functions form a vector space. A
measurable step function takes only finitely many different values S
1 , . . . , m .
The sets Bj = {f = j } are disjoint P
and measurable and S = Bj . So f
also admits the representation f (x) =
j 1Bj (x), where the sets Bj form a
(disjoint) partition of the set S.
Theorem 3.3 For every measurable function f : S [0, ] there is a sequence (fn ) of non-negative elementary functions with fn f, i.e. f1 f2 . . .
and f = supn fn .
Proof. Let
fn (x) =
n
n2
1
X
(3.1)
k=0
i.e. if f (x) < n, then fn (x) is the largest member of the grid 2n N0 which is
situated on the left of f (x), and fn (x) = n if f (x) n. Because these grids
are successively contained in each other we get fn (x) fn+1 (x). Moreover
fn f . Finally f (x) fn (x) 2n , as soon as f (x) 2n n 2n = n and
fn (x) = n = f (x) if f (x) = . Therefore (fn ) converges point-wise to
f.
The next result is the only place in this section where a measure appears.
Corollary 3.2 Let be a measure on the -algebra F. Let f be extended realvalued F -measurable. Then f is -almost everywhere equal to an F-measurable
function g.
Pkn
Proof. Consider elementary functions fn =
k=1 kn 1Bnk with Bnk
0
F such that fn f + . Each of the sets Bnk is -equivalent to a set Bnk
0
+
F. Outside the countably many null-sets Bnk M Bnk f is equal to the Fmeasurable function
kn
X
0 .
nk 1Bnk
g + = sup
n
k=0
Chapter 4
The Integral
In this chapter (S, F, ) is a measure space. We define the integral with respect
to and study some of its properties.
4.1
The simplest form of the integral is for elementary functions. We use the convention 0 = 0.
Definition 4.1 Let (S, F, ) be a measure space. Let f be a non-negative elementary function with the representation
f=
n
X
i 1Ai
(4.1)
i=1
Z
f d :=
f (x) d(x) :=
n
X
i (Ai )
(4.2)
i=1
The representation 4.1 is not unique, because the sets of the form {f = }
can be split in many different ways into disjoint measurable parts. But
Lemma 4.1 The integral in Definition 4.1 does not depend on the choice of the
representation in (4.1).
Pm
Proof. Let f = j=1 j 1Bj be a second representation of S
f with disjoint
S
sets Bj F . Without loss of generality we may assume S = i Ai = j Bj
since the coefficient 0 is admitted. Moreover i = j for all pairs (i, j) of indices
with Ai Bj 6= , because both coefficients agree with the common value which
the function f takes on this intersection. The additivity of shows
n
X
i=1
i (Ai ) =
n X
m
X
i=1 j=1
i (Ai Bj ) =
m X
n
X
j=1 i=1
25
j (Ai Bj ) =
m
X
j=1
j (Bj ).
26
For the sake of simplicity of this lemma we have assumed in Definition 4.1
that the sets Ai in the representation (4.1) are disjoint. Part a) of the following
theorem shows in particular that the formula (4.2) is also valid if the Ai are not
disjoint.
Theorem 4.1 The integral as a functional on the set of non-negative elementary functions is
R
R
R
a) additive and positively homogeneous, i.e. af + bg d = a f d + b g d
if a [0, ),
R
R
b) monotone increasing, i.e. f g implies f d g d,
c) -continuous
R
R from below, i.e. if fn is [0, ]-valued and fn f point-wise
then fn d f d.
Proof. The proofs for a) and b) follow from the fact that for any two
elementary functions there is a common partition of S into disjoint measurable
subsets on which
Pm both functions are constant.
c) Let f = j=1 j 1Aj , and assume without loss of generality that j > 0 for
Z
lim
n
fn d lim
n
m
X
X
1
1
(j ) (Aj ).
(j ) (Bjn,k ) =
k
k
j=1
R
R
P
As k we get limn fn d
j (Aj ) = f dx. The converse inequality
follows from the monotonicity of the integral.
4.2
We extend the definition of the integral from elementary functions in two steps:
First for all non-negative measurable functions and then for those measurable
functions whose positive and negative parts have a finite integral.
R
Definition 4.2 For a measurable function f : S 7 [0, ] we define
f d by
Z
Z
f d = sup{ g d : g f, g is an elementary f unction}.
Theorem 3.3 says that a measurable f 0 can be approximated from below
by an increasing sequence of elementary functions. One reaches the sup in the
definition by any such sequence.
Lemma
R 4.2 LetR(fn ) be any sequence of elementary functions such that fn f .
Then
fn d f d.
R
R
by definition f d fn d for each n and thus
R Proof. Obviously
R
f d limn fn d. For the converse inequality let g be an elementary
27
n
n
R
limn
min(fn , g) d = g d. The definition of f d yields the desired
inequality.
From Definition 4.2 and lemma 4.2 it is straightforward to see that the
integral for non-negative measurable functions is again additive, monotone and
positively homogeneous.
We now admit functions which also take negative values. Every function
f : S [, ] has the decomposition f = f + f , where f + = max(f, 0)
and f = max(f, 0). Then |f | = f + + f . All these functions are measurable
if f is measurable (cf. corollary 3.1).
R
Definition 4.3 a) A measurable function f : S [, ] with |f | d <
is called -integrable (-integrierbar). For a -integrable function the integral is defined by
Z
Z
Z
+
f d := f d f d.
The set of finite-valued -integrable functions is denoted by L1 ().
b) If A F and f is either
R measurable andR 0 or integrable the same is true
for 1A f and one writes A f d instead of 1A f d.
Example 4.1 As in Example 2.4 the letter L is a reference to the name Lebesgue.
In the special case where S = Rd and F is the -algebra of Lebesgue measurable
sets and = d the elements
of L1 (d ) are called Lebesgue-integrable.
One
R
R
1
d
often writes L (R ) and Rd f (x) dx instead of L1 (d ) and f dd . In Example
4.3 below we study the relation to the Riemann integral.
We collect some elementary facts
Proposition 4.1 Let f be measurable. Then
a) (Markovs inequality). For all a > 0
Z
1
{|f | a}
|f | d.
a
R
b) |f |d = 0 if and only if {f 6= 0} = 0.
R
R
c) If f is integrable then {|f | = } = 0 and | f d| |f | d.
d) f is integrable if there is an integrable function h with |f | h.
(4.3)
28
c) The first assertion follows by letting a tend to + in (4.3). For the second
assertion note that
Z
Z
Z
Z
Z
Z
( f + d + f d) f + d f d f + d + f d.
d) Clearly if |f | h and h is integrable then
|f | d
hd < .
= ( f d f d) = f d.
2. If f, g L1 (), then (f + g)+ (f + g) = f + g = f + + g + f g , i.e.
(f + g)+ + f + g = (f + g) + f + + g + . The additivity of the integral for
non-negative functions and subtraction give
Z
Z
Z
f + g d = f d + g d.
3. If f g, then
g d =
f d +
g f d
f d since g f 0.
Example 4.2 Consider the counting measure on the power set of N. Then
the integral is just a series:
Z
X
f d =
f (n).
(4.4)
n=1
29
f n (x)dx = inf
n
b
a
f n (x)dx.
The two functions f and f are -according to theorem 3.2- Borel measurable.
Since they are bounded and on a bounded interval the constants are integrable
they areRLebesgue-integrable
by proposition
4.1 d). Monotonicity of the integral
R
R
implies f d1 = f d1 . Thus f f d1 = 0 and by proposition 4.1 d)
f f = 0 a.e.. Therefore f = f a.e. and proposition 4.2 shows that f is
Lebesgue-integrable and
Z
Z
Z b
Z b
1
1
f d = f d =
f (x) dx =
f (x) dx.
a
R
However a so-called improper Riemann integral like in the formula 0 sinx x dx =
R
/2 typically is not a Lebesgue-integral, e.g. in this example 0 | sinx x | dx = ,
i.e. the integrand is not 1 -integrable in the sense of Definition 4.3.
Definition 4.4 We denote by L1 () the set of -equivalence of integrable functions.
Part a) of the following lemma follows from proposition 4.1 c). Part b) is
obvious.
30
4.3
The limit results of this section are a crucial feature of the Lebesgue integral.
They all treat in one way or another the question whether the integral can be
exchanged with a limit.
Theorem 4.3 (Theorem of monotone convergence). Let (fn ) be a nondecreasing sequence of non-negative measurable functions. Let f = supn fn .
Then
Z
Z
f d = lim
fn d.
Proof. Theorem 3.3 shows that for every n N there exists a sequence
(fnk )kN of elementary functions such that fnk k fn . Let us consider the
elementary function gk = max(f1k , . . . , fkk ). For each k we have gk+1 gk and
f gk . Moreover supk gk supkn fnk = fn for each n. So f supk gk
supn fn = f , i.e. gk f . Since the gn are elementary lemma 4.2 gives
Z
Z
Z
f d = lim
gn d lim
fn d.
n
f d limn
Remark 4.1 The additivity of the integral and the theorem of monotone convergence together show that every non-negative measurable function f gives rise
to a new measure f : F [0, ], defined by
Z
f (A) = 1A f d.
In fact for every sequence (An ) of pairwise disjoint sets with union A
f (A) =
Z X
X
X
(
1An )f d =
1An f d =
f (An ).
n=1
n=1
n=1
31
Theorem 4.4 (Fatous lemma). Let (fn ) be a sequence of non-negative measurable functions. Then
Z
Z
lim inf fn d lim inf fn d.
n
m nm
Example 4.4 In general the inequality in Fatous lemma is strict even if the
functions converge point-wise and are in L1 (). We consider the standard
Lebesgue measure space ([0, 1], B([0, 1]), |B([0, 1])). Let f = 0 and fn = n1(0, n1 ] .
Then fn f point-wise. On the other hand
Z 1
Z 1
1
lim
fn (x) dx = lim n = 1 6= 0 =
f (x)dx.
n
n
n
0
0
Lebesgues result below shows that domination by an integrable function is
sufficient to avoid the effect in this example. Note that the function x 7 x1
which is a natural upper bound for our sequence fails to be integrable.
The following theorem is perhaps the most useful single result in measure
theory.
Theorem 4.5 (Lebesgues dominated convergence theorem, Satz von
der majorisierten Konvergenz) Let (fn ) be a sequence of measurable functions such that there is an integrable function h with |fn | h -a.e.. Let f be
such that f (x) = limn fn (x) -a.e.. Then f is integrable as well and
Z
Z
f d = lim
fn d
n
and
Z
lim
|f fn | d = 0.
Proof. The union N of the countably many null-sets {fn 9 f }, {|fn | > h}
and {|h| = } is a null-set. Without changing the integrals or the integrability
of the functions involved we can set all function to vanish on N . Let us consider
the auxiliary functions gn = 2h (f fn ). Then limn gn = 2h and gn 0, so
by Fatous lemma
Z
Z
Z
2h d =
lim gn d lim inf gn d
n
n
Z
Z
Z
f d fn d .
=
2h d lim sup
n
32
R
Therefore lim supn
f d fn d 0. Similarly we can apply Fatous
lemma to the non-negative
R
2h + f fn , and get in the same way the
Rfunctions
estimate lim inf n
f d fn d 0. Together this gives the first of the two
limit assertions. The second one follows if we apply the first one to the sequence
(|f fn |) which is dominated by 2h and converges everywhere to 0.
As a corollary one gets the following rule for differentiation under the integral
sign.
Theorem 4.6 Let (a, b) an open interval and let f : (a, b)S R be a function
such that
i) for fixed t (a, b) is the function f (t, ) is -integrable on S,
ii) for fixed x S the function f (, x) is continuously differentiable on (a, b),
f (t, x) d(x) =
f (t, x) d(x).
(4.6)
t S
t
S
Proof. Let (tn ) be sequence of points which converges to some t (a, b).
(t,x)
. By the
Consider the measurable functions gn defined by gn (x) = f (tn ,x)f
tn t
mean-value theorem for each x there is a point tn (x) (a, b) such that gn (x) =
t f (tn (x), x). By hypothesis we can conclude |gn (x)| h(x). Furthermore
gn (x) t
f (t, x) for all x S. The theorem of dominated convergence gives
R
R
Z
Z
f (tn , x) d(x) S f (t, x) d(x)
S
= gn (x) d
f (t, x)d.
tn t
t
i.e. the relation (4.6). Similarly the continuity of the right-hand side in (4.6)
follows because iii) implies
Z
Z
Z
lim
f (tn , x) d(dx) = lim f (tn , x) d(dx) =
f (t, x) d(dx).
n
n t
t
t
Chapter 5
Fubinis Theorem
Fubinis theorem is the main tool for multiple integrals.
Theorem 5.1 (Fubini) Let (X, F, ) and (Y, G, ) two -finite measure spaces.
Let F G be the -algebra on the cartesian product X Y which is generated
by the measurable rectangles F G with F F, G G. Then
a) There is a unique measure on the -algebra F G, such that
(F G) = (F )(G)
for all F F, G G.
b) Let f : X Y [0, ] be a F G-measurable function. For every x X the
sectional
function f (x, ) : y 7 f (x, y) is G-measurable and the function x 7
R
f
(x,
y)
d(y) is F-measurable. Similarly the
Y
R sectional function f (, y) is Fmeasurable for each y Y and the map y 7 X f (x, y) d(x) is G-measurable.
Moreover
Z
Z Z
f (x, y) d (x, y) =
f (x, y) d(y) d(x)
(5.1)
XY
X Y
Z Z
=
f (x, y) d(x) d(y).
Y
34
35
Qn Rn
36
1 (B) =
Z
(B x ) d(x),
0
1
2 (B) =
(By ) d(y) =
(By ).
1 (D) =
(Dx ) d(x) =
1 dx = 1,
whereas
2 (D) =
X
y
(Dy ) =
X
y
0 = 0.
Chapter 6
Convergence in measure
Besides point-wise convergence or a.e. convergence there are several other ways
to express that a sequence of measurable functions converges. The Lp -spaces
of the next chapter will give a full scale of such concepts. Here we discuss
convergence in measure which is particularly useful in probability theory. The
reader who wants to pass directly to the next chapter will need the completeness
result of theorem 6.4 b) which can be understood independently of the other
parts of this section.
6.1
38
Theorem 6.1 Between the three concepts in Definition 6.1 the following implications hold:
a) If fn f in the mean then fn f in measure.
b) If fn f a.e. then fn 1A f 1A in measure for every A F with
(A) < .
The converse implications are wrong even for finite measure spaces.
Proof. 1. Markovs inequality (theorem 4.1 a) implies for all > 0
Z
1
({|fn f | })
|fn f | d.
39
S
P
N Am0 N Am l=m Bl and (N Am0 ) l=m 2l = 2 2m for all
m m0 . Hence (N Am0 ) = 0 for all m0 and finally (N ) = 0 as desired.
2. Conversely suppose the subsequence condition holds. Let A F be a set of
finite measure and > 0. Choose a sub-sequence such that
lim sup (A {|fn f | > }) = lim (A {|fnk f | > }).
n
c
m }) < . Let A = m=1 {gnm m }. Then (A ) < . For the verification of
1
uniform convergence let > 0. Choose m . Then for all x A and n nm
1
one gets |fn (x) f (x)| gnm (x) m
.
6.2
40
Theorem 6.4 Let be a functional which assigns to every non-negative Fmeasurable functions g a value (g) [0, ] with the following properties:
(g) = 0 if and only if
0f g
=
X
(
gi )
i=1
g = 0 a.e.,
(f ) (g),
X
(gi ),
(6.1)
(6.2)
(6.3)
i=1
(g) <
g < a.e..
(6.4)
Then the set L of all real valued measurable functions f with (|f |) < is
complete with respect to the pseudo-metric d(f, g) = (|f g|).
Proof. The triangular inequality d(f, h) d(f, g) + d(g, h) follows from
(6.2) and (6.3). Let (fn ) be a Cauchy sequence with respect to d, i.e. (|fn
fP
m |) converges to 0 as m, n tend to . Choose
Pa subsequence (nk ) such that
k1
X
for k > k0 ,
l=k0
the sequence (fnk )(x) converges for k to a finite real number f (x) for all
x A. Outside of A let f (x) = 0. The measurability
of the fn and of A imply
P
the measurability of f . Moreover |f fnk | l=k |fnl+1 fnl | everywhere and
by (6.2) and (6.3)
(|f fnk |)
l=k
that
c
(A
)
<
.
Then
h
=
|h|d =
n
n
n=1 cn 1An is measurable and
P
c
(A
)
<
and
of
course
h(x)
>
0
for
all
x
S.
n
n
n=1
Theorem 6.5 Let h L1 () be everywhere positive. Define for finite measurable functions f, g the distance d(f, g) by
Z
d(f, g) =
h |f g| d.
S
41
X
i=1
Z
gi ) =
X
i=1
gi d
Z X
i=1
(h gi ) d =
Z
X
i=1
h gi d =
(gi ).
i=1
all
indices
because
h(x)
>
0
and
therefore
the
series
i=1 gi (x) converges. Thus
P
g
is
finite
a.e..
Therefore
d
is
a
complete
pseudo-metric.
i=1 i
P
P
If n d(fn , f ) < then the series n fn (x) f (x) converges absolutely -a.e.
because of (6.4) and in particular fn f -a.e..
Assume that limn d(fn , f ) = 0. In order to prove local convergence in measure
we verify the sub-sequence criterion P
of theorem 6.2. Let (fnk ) be a sub-sequence.
Then every sub-sub-sequence with l d(fnk l , f ) < converges to f -a.e..
Conversely assume local convergence in measure of fn to f or |fn f | to 0. Then
the same is true for the functions |fn f | h. By corollary
6.1 these functions
R
converge also in mean, i.e. by definition d(fn , f ) = fn f | d 0.
42
Chapter 7
The Lp-spaces
Definition 7.1 Let (S, F, ) be a measure space and let f be an extended realvalued F-measurable function. We define for p [1, ] the number kf kp by
Z
1/p
kf kp :=
|f |p d
(1 p < )
kf k
1
r
1
p
+ 1q .
44
1
r
1
p
Proof. (of the lemma) The function (x) = x p is concave because of p > r.
Using the equation of the tangent at the point (1, 1) we get (x) pr (x1)+1 =
r
r
px + q.
r
p
Since (
) =
r
1 p
r
p
r
q
p
we get (
)
r
p
+ qr .
d
+
d = + = 1,
kf krp kgkrq
p kf kpp
kgkqq
p q
i.e. Hlders inequality.
b) (Minkowski) For p = 1 and p = the assertion is easy. Let now 1 < p < .
Let 0 < kf kp + kgkp < . For the proof of Minkowskis inequality we may
assume f, g 0. Because of kf + gk 2p max(f, g)p 2p (f p + g p ) one gets
p
kf + gkp < . Letting r = 1, q = p1
we get from a)
kf + gkpp
p1
kq =
(p1)q
q1
Z
p1
p
p
.
d
=
h d
= khkp1
p
(7.2)
Corollary 7.1 Let (S) < and p < q. Then kf kp kf kq (S) p q for all
measurable functions f and in particular Lq () Lq ().
Proof. Let s be the number for which
s
1
s
1
s
1
p
45
and hence (B) < is equivalent to 1B Lp (). Moreover the theorem 4.5 of
dominated convergence holds also in Lp : If a sequence of measurable functions
satisfies |fn | h Lp () and fn f point-wise, then kfn f kp 0. In fact
p
|fn f |p
can apply the original theorem 4.5 to these functions and
R h and we
conclude |fn f |p d 0.
Proposition 7.1 Let E be a -stable generating system of the -algebra F.
Assume that
S there is an increasing sequence Ek of sets in E with (Ek ) <
and S = k Ek . Then the linear span of all indicator functions 1E , E E is
dense in Lp () for 1 p < .
Proof. Let L be the smallest closed linear subspace of Lp () which contains
the indicator functions 1E , E E. We want to show L = Lp (). Fix a number
k N. We consider the class Dk of all sets F F such that 1F Ek L. For
every E E we have E Ek E and hence 1EEk L by assumption. So
E Dk .
We claim that Dk is a monotone class. Since S Ek = Ek E we have S
Dk . Let F, G Dk such that G F . Then 1(F \G)Ek = 1F Ek 1GEk L
since L is a linear space. Thus F \ G Dk . Let Fn be an increasing sequence
in Dk and Fn F . Then Fn Ek F Ek as n and (F Ek ) < .
Thus k1F Ek 1Fn Ek kpp = (F \ Fn ) Ek 0 as n and hence 1F L
since L is closed. Therefore F Dk . The monotone class theorem 2.4 shows
that F Dk for all k. Let now F F be any measurable set with (F ) < .
Then F Ek F and as above we see that 1F = limk 1F Ek L.
Finally let f Lp (). By theorem 3.3 there is a sequence of elementary
functions (fn )n which increases to f + . Each fn is in L as a linear combination
of indicators of sets of finite measure. Then fn f + in Lp () by the above
extended dominated convergence and hence f + L and similarly f L and
f L.
Theorem 7.3 Let 1 p < . Let be a measure on B(Rd ) for which bounded
sets have finite measure. Then the continuous functions with compact support
are dense in Lp (). Similarly for a finite measure on the Borel sets of a
metric space the bounded continuous functions are dense in Lp ().
Proof. In the first situation let L be the closure of Cc (Rd ) in Lp (). Let
U be an open set of finite -measure. There is a sequence (fn ) in Cc (Rd ) such
46
Chapter 8
The Radon-Nikodym
Theorem
8.1
f d
(8.1)
d
d
d
d .
is uniquely determined by :
Proposition 8.1 Let , 0 be two measures with the the Radon-Nikodym den0
d
0
0
and f 0 = d
sities f = d
d . Then if and only if f f -a.e.. Two
measurable functions are Radon-Nikodym densities of the same measure if and
only if they are -equivalent.
Proof. Suppose that 0 , yet ({f > f 0 }) > 0. Then proposition 4.1
part b) applied to the function 1{f >f 0 } (f f 0 ) would give the contradiction
Z
Z
({f 0 > f }) =
f d >
f 0 d = 0 ({f 0 > f }).
{f >f 0 }
{f >f 0 }
48
(8.3)
Proof. 1. Note that b) holds for all pairs , for which a) holds. First
we get the equality in b) if g is a non-negative elementary function. Then by
monotone convergence one extends the equality to all non-negative functions g
and finally to all g L1 ().
2. For the existence proof assume first that and are finite measures with
0 . We use, following an idea of J. v. Neumann, the following basic tool
from functional analysis (theorem of Fischer-Riesz): Let H be a real Hilbert
space with the scalar product < , >. Let ` : H R be a linear functional for
which there is some positive constant c such that `(h) c khk for all h H.
Then there is an element f of H such that `(h) =< h, f > for all h H.
To apply this we work in the space L2 () and in the associated
R Hilbert space
2
L (). The norm is induced by the scalar product < f, h >= f h d. Then
by corollary 7.1 or by Cauchy-Schwarz we get
Z
Z
Z
1
1
1
1
h2 d 2 (S) 2
h2 d 2 = const khk2,
| h d| (S) 2
for all h L2 (). Moreover two -equivalent
functions are also -equivalent
R
and hence the value of the integral h d depends only on the -equivalence
49
R
class of h. Thus by `(h) = h d a linear functional on L2 () is defined which
fulfils the assumption
of the theorem
of Fischer-Riesz. So there is a function
R
R
f L2 () with h d = `(h) = hf d for all h L2 (). Let A F . Since
is finite we have 1A L2 () and so
Z
(A) = `(1A ) =
f d,
A
Z
(A) =
i.e.
g
hf
1
1A h d =
h
Z
1A
1
dh =
h
1
1A f dg =
h
Z
A
g
f d,
h
In order to illustrate the role of -finiteness we consider the Lebesgue measure on the unit interval and the measure with the same null-sets but
d
does not exist.
(A) = if (A) > 0. Then but d
8.2
(8.4)
where ac and s .
Proof. Again consider = + . Then is also -finite: If (En ) and (En0 )
are two increasing sequence with union S and (En ) < and (En0 ) < then
(En En0 ) is an increasing sequence with union S and (En En0 ) = (En
En0 ) + (En En0 ) (En ) + (En0 ) < . Thus there are RN-densities f = d
d
.
Let
D
=
{f
>
0}.
Then
where
(A)
=
(A
D)
and
and g = d
ac
s
ac
d
50
R
s (A) = (A Dc ). Then s since (Dc ) = {f =0} f d = 0 and s (D) = 0.
Finally
Z
Z
Z
g
g
ac (A) = (A D) =
1D d.
g d =
1D f d =
f
f
A
AD
A
Therefore ac has a Radon-Nikodym density with respect to , i.e. ac .
Sometimes it is useful to admit measures which can have also negative
values.
Definition 8.4 Let (S, F) be a measurable space. A function : F R is
called a signed measure (signiertes Ma), if there are two finite measures
1 , 2 with = 1 2 .
Remark. Actually, every real valued set function on F which is -additive
in the sense of equation (1.1) is a signed measure. This fact is interesting but
rarely used. So we omit the proof. Note also that contrary to common use of
language a signed measure in general is not a measure. The representation of
a signed measures as a difference of two (non-negative) measures is not unique.
The following result provides a minimal decomposition of this type.
Theorem 8.3 (Hahn-Jordan decomposition) Every signed measure on the algebra F has a unique decomposition = + where + and are two
orthogonal measures. This decomposition is minimal: If = 1 2 for two
measures then + 1 and 2 .
Proof. By assumption = 1 2 for some finite measures i . Consider
i
the measure = 1 + 2 . Let fi = d
d for i = 1, 2. The function f = f1 f2
and the set D = {f 0} satisfy for all A F
Z
Z
Z
(A) =
f1 d
f2 d =
f d
A
A
ZA
Z
=
f d +
f d = (A D) + (A Dc )
A{f 0}
A{f <0}
where clearly in the last two sums the first term is non-negative and the second
term is non-positive for all A. Thus + = (D ) and = (Dc ) provide
the indicated decomposition.
Every representation = + as difference of two singular measures is
minimal in the above sense: Suppose = 1 2 and let D be a measurable
set such that + (Dc ) = 0 = (D). Then
+ (A) = (A D) = 1 (A D) 2 (A D) 1 (A D) 1 (A)
for all A F, i.e. + 1 and hence also 2 = (1 ) ( + ) =
1 + 0. If moreover 1 2 we can argue by symmetry that 1 + and
2 . This proves the uniqueness.
Definition 8.5 The space of signed measures on the measurable space (S, F) is
denoted by M(S, F). For M(S, F) the measure || = + + is called the
51
(8.5)
The Hahn-Jordan decomposition allows to extend the Radon-Nikodym theorem to signed measures .
Corollary 8.1 (signed Radon-Nikodym) Let be a measure. Let the signed
measure have the property that every -null-set is a -null-set. Then + and
d
L1 (), uniquely determined
have the same property. There is a function d
R d
up to -equivalence, such that (A) = A d d for all A F. Moreover
(
d +
d +
d
d
d
d||
) =
, ( ) =
, | |=
.
d
d
d
d
d
d
(8.6)
d
Proof. For the uniqueness of d
note that the proof of proposition 8.1 carries
over to our situation. In order to show + and let (N ) = 0.
Then (N D) = (N Dc ) = 0 and hence + (N ) = (N D) = 0 and
+
d
d
d
(N ) = (N Dc ) = 0. The function f = d
defined by d
= d
d d
has the desired properties since according to the proof of the Hahn-Jordan
decomposition f 0 on D and f < 0 on Dc .