This work is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License.13 Integration
We now come to the main use of measure theory: to define a general theory
of integration.
13.1 Measurable functions
From now on, by a measure space we shall mean a triple
(X,L,µ), where X is a set, L is a
σ-algebra on X, and µ is a σ-additive
measure defined on L. We say that the members of
L are measurable, or
L-measurable, if necessary to avoid confusion.
Definition 1
A function f:
X→ℝ
is measurable
if
Ec(f)={x∈ X: f(x)<c} that is Ec(f)=f−1((−∞,c))
|
is in L (that is Ec(
f)
is a measurable set) for any c∈ℝ
.A complex-valued function is measurable if its real and imaginary
parts are measurable.
Lemma 2
The following are equivalent:
-
A function f is measurable;
-
For any a<b the set f−1((a,b)) is measurable;
-
For any open set U⊂ ℝ the set f−1(U) is
measurable.
Proof.
To show
1 ⇒
2 we note that
f−1((a,b)) = Eb(f)∖ | ⎛
⎜
⎜
⎝ | | Ea+1/n(f) | ⎞
⎟
⎟
⎠ | .
|
For
2 ⇒
3
use that any open set
U⊂ ℝ is a union of countable
set of intervals (
a,
b), cf. proof of Cor.
25.
The final implication 3 ⇒ 1 directly follows from openness of (−∞,a).
□
Corollary 3
Let f:
X → ℝ
be measurable and
g: ℝ → ℝ
be continuous, then the
composition g(
f(
x))
is measurable.
Proof.
The preimage of the open set (−∞,
c) under a continuous
g is an open set, say
U. The preimage of
U under
f is measurable by Lem.
3. Thus, the preimage of (−∞,
c) under the composition
g ∘
f is measurable, thereafter
g ∘
f is a measurable function.
□
Theorem 4
Let f,g:X→ℝ be measurable. Then af
(a∈ℝ), f+g, fg, max(f,g) and
min(f,g) are all measurable. That is measurable functions form
an algebra and this algebra is closed under convergence a.e.
Proof.
Use Cor.
3 to show measurability of λ
f, |
f | and
f2. The measurability of a sum
f1 +
f2 follows from the relation
Ec(f1+f2)=⋃r∈ℚ (Er(f1)⋂
Ec−r(f2)).
|
Next use the following identities:
If (fn) is a non-increasing sequence of measurable functions
converging to f. Than Ec(f)=∪n Ec(fn).
Moreover any limit can be replaced by two monotonic limits:
| | fn(x)=
| | | | max
(fn(x), fn+1(x),…,fn+k(x)).
(70) |
Finally if f1 is measurable and f2=f1 almost everywhere,
then f2 is measurable as well.
□
We can define several types of convergence for measurable functions.
Definition 5
We say that sequence (
fn)
of functions converges-
uniformly to f (notated fn⇉
f) if
- almost everywhere to f (notated fn→a.e.f) if
fn(x)→ f(x) for all x∈ X∖ A,
µ(A)=0;
|
- in measure µ to f (notated
fn→µf) if for all ε>0
µ({x∈ X: | ⎪
⎪ | fn(x)−f(x) | ⎪
⎪ | >ε }) → 0.
(71) |
Clearly uniform convergence implies both convergences a.e and in
measure.
Theorem 6
On finite measures convergence a.e. implies convergence in measure.
Proof.
Define
An(ε)={
x∈
X: |
fn(
x)−
f(
x) |≥
ε}. Let
Bn(ε)=∪
k≥ n
Ak(ε). Clearly
Bn(ε)⊃
Bn+1(ε), let
B(ε)=∩
1∞Bn(ε). If
x∈
B(ε)
then
fn(
x)↛
f(
x). Thus µ(
B(ε))=0,
but µ(
B(ε))=lim
n→
∞µ(
Bn(ε)), cf. (
67). Since
An(ε)⊂
Bn(ε) we see that
µ(
An(ε))→ 0 as required for (
71)
□
Note, that the construction of sets Bn(ε) is just
another implementation of the “two monotonic limits”
trick (70) for sets.
Exercise 7
Present examples of sequences (
fn)
and functions
f such that:
-
fn→ µf but not
fn→ a.e.f.
-
fn→ a.e.f but not fn⇉ f.
However we can slightly “fix” either the set or the sequence to
“upgrade” the convergence as shown in the following two theorems.
Theorem 8 (Egorov)
If fn→
a.e.f on a finite measure
set X then for any σ>0
there is Eσ⊂
X
with µ(
Eσ)<σ
and fn⇉
f on
X∖
Eσ.
Proof.
We use
An(ε) and
Bn(ε) from the proof
of Thm.
6. Observe that |
f(
x)−
fk(
x) |< ε uniformly for all
x ∈
X∖
Bn(ε) and
k>
n.
For every ε>0 we seen
that µ(
Bn(ε))→ 0, thus for each
k there is
N(
k) such that µ(
BN(k)(1/
k))<σ/2
k. Put
Eσ=∪
k BN(k)(1/
k).
□
Theorem 9
If fn→ µf then there is a subsequence
(nk) such that fnk→
a.e.f for k→ ∞.
Proof.
In the notations of two previous proofs: for every natural
k
take
nk such that µ(
Ank(1/
k))< 1/2
k, which is possible since µ(
An(ε))→ 0. Define
Cm=∪
k=m∞Ank(1/
k) and
C=∩
Cm. Then,
µ(
Cm)=1/2
m−1 and, thus, µ(
C)=0 by (
67). If
x∉
C
then there is such
N that
x∉
Ank(1/
k) for all
k>
N. That means that |
fnk(
x)−
f(
x) |<1/
k for all
such
k, i.e
fnk(
x)→
f(
x). Thus, we have the point-wise convergence everywhere except the zero-measure set
C.
□
It is worth to note, that we can use the last two theorem subsequently
and upgrade the convergence in measure to the uniform convergence of a
subsequence on a subset.
Exercise 10
For your counter examples from
Exercise 7, find
-
a subsequence fnk of the sequence
from 1 which converges to f a.e.;
- a subset such that sequence
from 2 converges uniformly.
Exercise 11
Read about Luzin’s C-property.
13.2 Lebesgue Integral
First we define a sort of “basis” for the space of integral functions.
Definition 12
For A⊆
X, we define χ
A to be the indicator
function
of A, by
Then, if χA is measurable, then χA−1( (1/2,3/2) ) = A
∈ L; conversely, if A∈L, then
X∖ A∈L, and we see that for any
U⊆ℝ open, χA−1(U) is either
∅, A, X∖ A, or X, all of which are
in L. So χA is measurable if and only if
A∈L.
Definition 13
A measurable function f:X→ℝ is
simple if it attains only a countable number of values.
Lemma 14
A function f:
X→ℝ
is simple if and only if
for some (
tk)
k=1∞⊆ℝ
and
Ak∈
L. That is, simple functions are
linear combinations of indicator functions of measurable sets.Moreover in the above representation the sets Ak can be
pair-wise disjoint and all tk≠ 0 pair-wise different. In this case
the representation is unique.
Notice that it is now obvious that
Corollary 15
The collection of simple functions forms a vector space: this wasn’t
clear from the original definition.
Definition 16
A simple function in the form (72) with disjoint
Ak is called summable
if the following series converges:
| | ⎪
⎪ | tk | ⎪
⎪ | µ(Ak)
if f has the above unique representation f = | | tk
χAk .
(73) |
It is another combinatorial exercise to show that this definition is
independent of the way we write f.
Definition 17
We define the integral
of a simple function
f=∑
k tk χ
Ak (72) over a measurable set A by setting
Clearly the series converges for any simple summable function
f. Moreover
Lemma 18
The value of integral of a simple summable function is independent
from its representation by the sum of indicators (72). In particular, we can evaluate the integral taking the canonical representation over pair-wise
disjoint sets having pair-wise different values.
Proof.
This is another slightly tedious combinatorial exercise.
You need to prove that the integral of a simple function is
well-defined, in the sense that it is independent of the way we
choose to write the simple function.
□
Exercise 19
Let f be the function on [0,1] which take the value 1 in
all rational points and 0—everywhere else. Find the value of
the Lebesgue integral ∫[0,1] f,dµ with respect to the
Lebesgue measure on [0,1]. Show that the Riemann upper-
and lower sums for f converges to different values, so f is
not Riemann-integrable.
We will denote by S(X) the
collection of all simple summable functions on X.
Proposition 21
Let f, g:
X→ ℝ
be in S(
X)
(that is simple summable), let a, b∈ ℝ
and A
is a measurable set. Then:
-
∫A af+bg d µ = a∫A f d µ + b∫A g d µ,
that is S(X) is a linear space;
- The correspondence f→ ∫A f d µ is a linear
functional on S(X);
-
The correspondence A → ∫A f d µ is a charge;
-
If f≤ g then ∫X f d µ ≤ ∫X g d µ,
that is integral is monotonic;
-
The function
d1(f,g)= | ∫ | | ⎪
⎪ | f(x)−g(x) | ⎪
⎪ | d µ(x)
(74) |
has all properties of a metric (distance) on S(X) probably
except separation, but see the next item.
- For f≥ 0 we have ∫X f d µ=0 if and only if
µ( { x∈ X : f(x)≠0 } ) = 0. Therefore for the function d1 (74):
d1(f,g)=0 if and only if f | | g.
|
- The integral is uniformly continuous with respect the above metric d1 (74):
| ⎪
⎪
⎪
⎪
⎪
⎪ | ∫ | | f(x) d µ(x)− | ∫ | | g(x) d µ(x) | ⎪
⎪
⎪
⎪
⎪
⎪ | ≤ d1(f,g).
|
Proof.
The proof is almost obvious, for example the
Property
1 easily follows from
Lem.
18.
We will outline 3 only. Let f is an indicator
function of a set B, then A→ ∫A
f d µ=µ(A∩ B) is a σ-additive measure (and
thus—a charge). By the Cor. 35 the
same is true for finite linear combinations of indicator functions
and their limits in the sense of distance d1.
□
We can identify functions which has the same values a.e. Then
S(X) becomes a metric space with the distance
d1 (74). The space may be incomplete and we
may wish to look for its completion. However, if we will simply try
to assign a limiting point to every Cauchy sequence in
S(X), then the resulting space becomes so huge that it
will be impossible to realise it as a space of functions on X.
Exercise 22
Use ideas of Ex. 1 to present a sequence of simple functions which has the Cauchy property in metric d1 (74) but does not have point-wise limits anywhere.
To
reduce the number of Cauchy sequences in S(X) eligible
to have a limit, we shall ask an additional condition. A convenient
reduction to functions on X appears if we ask both the convergence
in d1 metric and the point-wise convergence on X a.e.
Definition 23
A function f is summable
by a measure µ
if there is a sequence
(
fn)⊂
S(
X)
such that
-
the sequence (fn) is a Cauchy sequence in
S(X);
- fn→a.e. f.
Clearly, if a function is summable, then any equivalent function is
summable as well. Set of equivalent classes of summable functions will be denoted by
L1(X).
Lemma 24
If the measure µ is finite then any bounded measurable
function is summable.
Proof.
Define
Ekn(
f)={
x∈
X:
k/
n≤
f(
x)< (
k+1)/
n} and
fn=∑
k k/
n χ
Ekn (note that the sum is
finite due to boundedness of
f).
Since | fn(x)−f(x) |<1/n we have uniform convergence
(thus convergence a.e.) and (fn) is the Cauchy sequence:
d1(fn,fm)=∫X| fn−fm | d µ≤
(1/n+1/m)µ(X).
□
Another simple result, which is useful on many occasions is as
follows.
Lemma 26
If the measure µ
is finite and fn⇉
f then
d1(
fn,
f)→ 0
.
Corollary 27
For a convergent sequence fn→
a.e.
f, which admits the uniform bound |
fn(
x) |<
M for all
n and x, we have d1(
fn,
f)→ 0
.
Proof.
For any ε>0, by the Egorov’s theorem
8
we can find
E, such that
- µ(E)< ε/2M; and
- from the uniform convergence on X∖ E there exists
N such that for any n>N we have
| f(x)−fn(x) |<ε /2µ(X).
Combining this we found that for
n>
N,
d1(
fn,
f)<
M
ε/2
M + µ(
X) ε /2µ(
X) <
ε .
□
Exercise 28
Convergence in the metric d1 and a.e. do not imply each other:
-
Give an example of fn→a.e.
f such that d1(fn ,f)↛0.
- Give an example of the sequence (fn) and function f
in L1(X)
such that d1(fn ,f)→ 0 but fn does not
converge to f a.e.
To build integral we need the following
Lemma 29
Let (
fn)
and (
gn)
be two Cauchy sequences in
S(
X)
with the same limit a.e., then
d1(
fn,
gn)→ 0
.
Proof.
Let φ
n=
fn−
gn, then this is a Cauchy sequence with zero
limit a.e. Assume the opposite to the statement: there exist
δ>0 and sequence (
nk) such that
∫
x| φ
nk |
d µ>δ. Rescaling-renumbering
we can obtain ∫
x| φ
n |
d µ>1.
Take quickly convergent subsequence using the Cauchy property:
Renumbering agian assume d1(φk,φk+1)≤ 1/2k+2.
Since φ1 is a simple, take the canonical presentation φ1=∑k tk χAk,
then ∑k | tk | µ(Ak)=∫X
| φ1 | d µ≥ 1. Thus, there exists N, such that
∑k=1N | tk | µ(Ak)≥ 3/4. Put
A=⊔k=1N Ak and C=max1≤ k ≤
N| tk |=maxx∈ A| φ1(x) |.
By the Egorov’s Theorem 8 there is E⊂ A such
that µ(E)<1/(4C) and φn⇉ 0 on B=A∖
E. Then
| ∫ | | | ⎪
⎪ | φ1 | ⎪
⎪ | d µ= | ∫ | |
| ⎪
⎪ | φ1 | ⎪
⎪ | d µ− | ∫ | | | ⎪
⎪ | φ1 | ⎪
⎪ | d µ≥
| | − | | · C= | | .
|
By the triangle inequality for d1:
| ⎪
⎪
⎪
⎪
⎪
⎪ | ∫ | | | ⎪
⎪ | φn | ⎪
⎪ | d µ− | ∫ | |
| ⎪
⎪ | φn+1 | ⎪
⎪ | d µ | ⎪
⎪
⎪
⎪
⎪
⎪ | ≤
d1(φn,φn+1)≤ | |
we get
| ∫ | | | ⎪
⎪ | φn | ⎪
⎪ | d µ≥ | ∫ | | | ⎪
⎪ | φ1 | ⎪
⎪ | d µ−
| | ⎪
⎪
⎪
⎪
⎪
⎪ | ∫ | | | ⎪
⎪ | φn | ⎪
⎪ | d µ− | ∫ | |
| ⎪
⎪ | φn+1 | ⎪
⎪ | d µ | ⎪
⎪
⎪
⎪
⎪
⎪ | ≥
| | − | | | > | | .
|
But this contradicts to the fact ∫B | φn | d µ
→ 0, which follows from the uniform convergence
φn⇉ 0 on B.
□
It follows from the Lemma that we can use any Cauchy sequence of simple functions for the extension of integral.
Corollary 30
The functional IA(f)=∫A f(x) d µ(x), defined on any
A∈ L on the space of simple functions
S(X) can be extended by continuity to the functional
on L1(X,µ).
Definition 31
For an arbitrary summable f∈
L1(
X)
, we define
the Lebesgue integral
where the Cauchy sequence fn of summable simple functions
converges to f a.e.
Theorem 32
-
L1(X) is a linear space.
- For any measurable set A⊂ X the correspondence f↦
∫A f d µ is a linear functional on L1(X).
- For any f∈L1(X) the value ν(A)=∫A f
d µ is a charge.
- d1(f,g)=∫A | f−g | d µ is a distance on
L1(X).
Proof.
The proof is follows from Prop.
21 and
continuity of extension.
□
Summing up: we build L1(X) as a completion of
S(X) with respect to the distance d1 such that
elements of L1(X) are associated with (equivalence
classes of) measurable functions on X.
13.3 Properties of the Lebesgue Integral
The space L1 was defined from dual convergence—in
d1 metric and point-wise a.e. Can we get the continuity of the integral from
the convergence almost everywhere alone? No, in general. However,
we will state now some results on continuity of the integral under
convergence a.e. with some additional assumptions. Finally, we show
that L1(X) is closed in d1 metric.
Theorem 33 (Lebesgue on dominated convergence)
Let (
fn)
be a
sequence of µ
-summable functions on X, and there is
φ∈
L1(
X)
such that |
fn(
x) |≤ φ(
x)
for all x∈
X, n∈ℕ
. If fn→a.e. f, then
f∈L1(X) and for any measurable A:
Proof.
For any measurable
A the expression ν(
A)=∫
A φ
d µ defines a finite measure on
X due to non-negativeness
of φ and Thm.
32.
Lemma 34 (Change of variable)
If g is measurable and bounded then f=φ
g is µ
-summable
and for any µ
-measurable set A we have
Proof.[Proof of the Lemma]
Let
M be the set of all
g such that the Lemma is
true.
M includes any indicator functions
g=χ
B of a
measurable
B:
| ∫ | | f d µ= | ∫ | | φχB d µ = | ∫ | | φ
d µ =ν(A⋂ B)= | ∫ | | χB d ν= | ∫ | | g d ν.
|
Thus
M contains also finite linear combinations of indicators.
For any
n∈ℕ and a bounded
g two functions
g−(
x)=1/
n[
ng(
x)] and
g+(
x)=
g−+1/
n are
finite linear combinations of indicators and are in
M. Since
g−(
x)≤
g(
x)≤
g+(
x) we have
| ∫ | | g− d ν= | ∫ | | φ g− d µ≤ | ∫ | |
φ g d µ≤ | ∫ | | φ g+ d µ= | ∫ | |
g+ d ν.
|
By squeeze rule for
n→ ∞ we have the middle
term tenses to ∫
Ag d ν, that is
g∈
M.
Note, that formula (75) is a change of
variable in the Lebesgue integral of the type: ∫f(sinx)
cosx d x = ∫f(sinx) d (sinx).
□
For the proof of the theorem define:
gn(x) | = | ⎧
⎨
⎩ | fn(x)/φ(x), | if φ(x)≠ 0, |
0, | if φ(x)= 0,
|
|
|
|
g(x) | = | ⎧
⎨
⎩ | f(x)/φ(x), | if φ(x)≠ 0, |
0, | if φ(x)= 0.
|
|
|
|
|
|
Then
gn is bounded by 1 and
gn→
a.e. g. To show the theorem it
will be enough to show lim
n→ ∞∫
A
gn d ν=∫
A g d ν. For the uniformly bounded functions on
the finite measure set this can be derived from the Egorov’s
Thm.
8, see an example of this in the proof of
Lemma
29.
□
Note, that in the above proof summability of φ was used to obtain the
finiteness of the measure ν, which is required for Egorov’s
Thm. 8.
Exercise 35
Give an example of fn→a.e.
f such that ∫X fn d µ ≠ ∫X f d µ. For such an example, try to find a function φ such that | fn | ≤ φ for all n and check either φ is summable.
Exercise 36 (Chebyshev’s inequality)
Show that: if f is non-negative and summable, then
µ{x∈ X: f(x)>c} < | | ∫ | | f d µ.
(76) |
Theorem 37 (B. Levi’s, on monotone convergence)
Let (
fn)
be monotonically increasing sequence of µ
-summable
functions on X. Define f(
x)=lim
n→∞ fn(
x)
(allowing the value +∞
).
-
If all integrals ∫X fn d µ are bounded by the same
value C<∞ then f is summable and ∫X
f d µ=limn→∞∫X fn d µ.
- If limn→∞∫X fn d µ=+∞ then
function f is not summable.
Proof.
Replacing
fn by
fn−
f1 and
f by
f−
f1 we can
assume
fn≥ 0 and
f≥ 0. Let
E be the set where
f is infinite, then
E=⋂N⋃n ENn, where
ENn={x∈ X: fn(x)≥ N}.
|
By Chebyshev’s
inequality (
76) we have
Nµ(ENn) < | ∫ | | fn d µ ≤ | ∫ | | fn d µ≤ C,
|
then µ(
ENn)≤
C/
N . Thus
µ(
E)=lim
N→∞lim
n→∞
µ(
ENn)=0.
Thus f is finite a.e.
Lemma 38
Let f be a measurable non-negative function attaining only
finite values. f is summable
if and only if sup∫A f d µ<∞, where the supremum
is taken over all finite-measure set A such that f is
bounded on A.
Proof.[Proof of the Lemma]
Necessity: if
f is summable then for any set
A⊂
X we have
∫
A f d µ≤ ∫
X f d µ<∞, thus the supremum is
finite.
Sufficiency: let sup∫A f d µ=M<∞, define B={x∈ X:
f(x)=0} and
Ak={x∈ X:
2k≤ f(x)<2k+1, k∈ℤ}, by (76) we have
µ(Ak)<M/2k and
X=B⊔(⊔k=0∞Ak). Define
g(x) | = | |
fn(x) | = | ⎧
⎨
⎩ | f(x), | if x∈ ⊔−nn An, |
0, | otherwise.
|
|
|
|
|
|
Then g(x)≤ f(x) < 2g(x). Function g is a simple
function, its summability follows from the estimate
∫⊔−nn Ak g d µ≤∫⊔−nn Ak
f d µ≤ M which is valid for any n, taking
n→ ∞ we get summability of g. Furthermore,
fn →a.e. f and fn(x)≤
f(x) <2g(x), so we use the Lebesgue
Thm. 33 on dominated convergence to
obtain the conclusion.
□
Let A be a finite measure set such that f is bounded on A, then
| ∫ | | f d µ | | | ∫ | | fn d µ≤
| | ∫ | | fn d µ≤ C.
|
This show summability of f by the previous Lemma. The rest of
statement and (contrapositive to) the second part follows from the Lebesgue
Thm. 33 on dominated convergence.
□
Now we can extend this result dropping the monotonicity assumption.
Lemma 39 (Fatou)
If a sequence (
fn)
of µ
-summable non-negative functions
is such that:
-
∫X fn d µ≤ C for all n;
- fn →a.e. f,
then f is µ
-summable and ∫
X f d µ≤
C.
Proof.Let us replace the limit
fn→
f by two monotonic
limits. Define:
gkn(x) | = | min(fn(x),…,fn+k(x)), |
gn(x) | = | |
|
Then
gn is a non-decreasing sequence of functions and
lim
n→ ∞ gn(
x)=
f(
x) a.e. Since
gn≤
fn, from monotonicity of integral we get ∫
X gn d µ≤
C for all
n. Then Levi’s Thm.
37 implies
that
f is summable and ∫
X f d µ≤
C.
□
Exercise 41
Give an example such that under the Fatou’s lemma condition we get
limn→∞∫X fn d µ ≠ ∫X f d µ.
Now we can show that L1(X) is complete:
Theorem 42
L1(
X)
is a Banach space.
Proof.
It is clear that the distance function
d1 indeed define a norm
||
f||
1=
d1(
f,0). We only need to demonstrate the
completeness. We again utilise the three-step procedure from
Rem.
7.
Take a Cauchy sequence (fn)
and building a subsequence if necessary, assume that its
quickly convergent that is
d1(fn,fn+1)≤ 1/2k. Put
φ1=f1 and
φn=fn−fn−1 for n>1. Then fn= | | φk .
|
The sequence
ψn(x)=∑1n | φk(x) | is monotonic, integrals
∫X ψn d µ are bounded by the same constant
||f1||1+1. Thus, by the B. Levi’s
Thm. 37 and its proof, ψn→
ψ for a summable essentially bounded function ψ. Therefore, the series ∑φk(x) converges as well to a value f(x) of a
function f. But, this means that fn
→a.e. f (the
first step is completed).
We also notice
| fn(x) |≤| ψ(x) |. Thus by the Lebesgue
Thm. 33 on dominated convergence f∈
L1(X) (the second step is completed).
Furthermore,
0≤ | | | ∫ | | ⎪
⎪ | fn−f | ⎪
⎪ | d µ≤
| | | | ⎪⎪
⎪⎪ | φk | ⎪⎪
⎪⎪ | =0. |
That
is, fn→ f in the norm of L1(X). (That
completes the third step and the whole proof).
□
The next important property of the Lebesgue integral is its
absolute continuity.
Theorem 43 (Absolute continuity of Lebesgue integral)
Let f∈
L1(
X)
. Then for any ε>0
there
is a δ>0
such that | ∫
A f d µ |<ε
if
µ(
A)<δ
.
Proof.
If
f is essentially bounded by
M, then it is enough to set
δ=ε/
M. In general let:
An | = | {x∈ X: n≤ | ⎪
⎪ | f(x) | ⎪
⎪ | < n+1}, |
|
Bn | = | ⊔0n Ak, |
Cn | = | X∖ Bn.
|
|
Then ∫
X|
f |
d µ=∑
0∞∫
Ak|
f |
d µ, thus there is an
N such that
∑
N∞∫
Ak|
f |
d µ=∫
CN|
f |
d µ<ε/2.
Now put δ =ε/2
N+2, then for any
A⊂
X
with µ(
A)<δ:
| ⎪
⎪
⎪
⎪
⎪
⎪ | ∫ | | f d µ | ⎪
⎪
⎪
⎪
⎪
⎪ | ≤ | ∫ | | | ⎪
⎪ | f | ⎪
⎪ | d µ=
| ∫ | | | ⎪
⎪ | f | ⎪
⎪ | d µ+ | ∫ | |
| ⎪
⎪ | f | ⎪
⎪ | d µ
< | | + | | =ε.
|
□
13.4 Integration on Product Measures
It is well-known geometrical interpretation of an integral in calculus
as the “area under the graph”. If we advance from “area” to a
“measure” then the Lebesgue integral can be
treated as theory of measures of very special shapes created by graphs
of functions. This shapes belong to the product spaces of the function
domain and its range. We introduced product measures in
Defn. 40, now we will study them in same details
using the Lebesgue integral. We start from the following
Theorem 44
Let X and Y be spaces, and let S and T be semirings
on X and Y respectively and µ
and ν
be measures
on S and T respectively. If µ
and ν
are
σ
-additive, then the product measure ν× µ
from
Defn. 40 is σ
-additive as well.
Proof.
For any
C=
A×
B∈
S×
T let us define
fC(
x)=χ
A(
x)ν(
B). Then
(µ×ν)(C)=µ(A)ν(B)= | ∫ | | fC d µ.
|
If the same set
C has a representation
C=⊔
k Ck for
Ck∈
S×
T, then σ-additivity of ν implies
fC=∑
k fCk. By the Lebesgue
theorem
33 on dominated
convergence:
Thus
□
The above correspondence C↦ fC can be extended to the ring
R(S× T) generated by S× T by the formula:
fC= | | fCk, for C=⊔k Ck∈ R(S× T).
|
We have the uniform continuity of this correspondence:
| ⎪⎪
⎪⎪ | fC1−fC2 | ⎪⎪
⎪⎪ | 1≤ (µ×ν)(C1▵ C2)=d1(C1,C2)
|
because from the representation C1=A1⊔ B and
C2=A2⊔ B, where B=C1∩ C2 one can see that
fC1−fC2=fA1−fA2, fC1▵
C2=fA1+fA2 together with | fA1−fA2 |≤
fA1+fA2 for non-negative functions.
Thus the map C↦ fC can be extended to the map of
σ-algebra L(X× Y) of
µ×ν-measurable set to L1(X) by the formula
flimn Cn=limn fCn.
Exercise 45
Describe topologies where two limits from the last formula are taken.
The following lemma provides the geometric interpretation of the
function fC as the size of the slice of the set C along
x=const.
Lemma 46
Let C∈
L(
X×
Y)
. For almost every x∈
X the
set Cx={
y∈
Y: (
x,
y)∈
C}
is ν
-measurable and
ν(
Cx)=
fC(
x)
.
Proof.
For sets from the ring
R(
S×
T) it is true by the
definition. If
C(n) is a monotonic sequence of sets, then
ν(lim
n Cx(n))=lim
n ν(
Cx(n)) by
σ-additivity of measures. Thus the property
ν(
Cx)=
fx(
C) is preserved by monotonic limits. The following
result of the separate interest:
Lemma 47
Any measurable set can be received (up to a set of zero measure)
from elementary sets by two monotonic limits.
Proof.[Proof of Lem.
47]
Let
C be a measurable set, put
Cn∈
R(
S×
T)
to approximate
C up to 2
−n in µ×ν. Let
C′=∩
n=1∞∪
k =1∞Cn+k, then
(µ× ν) | ⎛
⎝ | C∖ ⋃k=1∞Cn+k | ⎞
⎠ | =0 and
(µ× ν) | ⎛
⎝ | ⋃k=1∞Cn+k∖ C | ⎞
⎠ | =21−n.
|
|
|
Then (µ×ν)(
C′▵
C)≤ 2
1−n for
any
n∈ℕ.
□
Coming back to Lem.
46 we notice that (in the above notations)
fC=
fC′ almost everywhere. Then:
fC(x) | | fC′(x)=ν(C′x)=ν(Cx).
|
□
The following theorem generalizes the meaning of the integral as “area under the graph”.
Theorem 48
Let µ
and ν
are σ
-finite measures and C be
a µ×ν
measurable set X×
Y. We define
Cx={
y∈
Y: (
x,
y)∈
C}
. Then for µ
-almost every
x∈
X the set Cx is ν
-measurable, function
fC(
x)=ν(
Cx)
is µ
-measurable and
where both parts may have the value +∞
.
Proof.
If
C has a finite measure, then the statement is reduced to
Lem.
46 and a passage to limit
in (
77).
If C has an infinite measure, then there exists a sequence of
Cn⊂ C, such that ∪n Cn=C and
(µ×ν)(Cn)→ ∞. Then fC(x)=limn
fCn (x) and
Thus fC is measurable and non-summable.
□
This theorem justify the well-known technique to calculation of areas
(volumes) as integrals of length (areas) of the sections.
Theorem 50 (Fubini)
Let f(
x,
y)
be a summable function on the product of spaces
(
X,µ)
and (
Y,ν)
. Then:-
For µ-almost every x∈ X the function f(x,y) is
summable on Y and fY(x)=∫Y f(x,y) d ν(y) is a
µ-summable on X.
- For ν-almost every y∈ Y the function f(x,y) is
summable on X and fX(y)=∫X f(x,y) d µ(x) is a
ν-summable on Y.
- There are the identities:
| | = |
| ∫ | | ⎛
⎜
⎜
⎜
⎜
⎝ | ∫ | | f(x,y) d ν(y) | ⎞
⎟
⎟
⎟
⎟
⎠ | dµ(x) |
| (79) |
| = | | ∫ | | ⎛
⎜
⎜
⎜
⎜
⎝ | ∫ | | f(x,y) d µ(x) | ⎞
⎟
⎟
⎟
⎟
⎠ | dν(y).
|
| |
|
- For a non-negative functions the existence of any repeated
integral in (79) implies summability of f on
X× Y.
Proof.
From the decomposition
f=
f+−
f− we can reduce our consideration
to non-negative functions. Let us consider the product of three
spaces (
X,µ), (
Y,ν), (ℝ,λ), with
λ=
dz being the Lebesgue measure on ℝ. Define
C={(x,y,z)∈ X× Y× ℝ: 0≤ z≤ f(x,y)}.
|
Using the relation (
78) we get:
Cxy | = | {z∈ ℝ: 0≤ z≤ f(x,y)},
λ(Cxy)=f(x,y) |
Cx | = | {(y,z)∈ Y× ℝ: 0≤ z≤ f(x,y)},
(ν× λ)(Cx)= | ∫ | | f(x,y) d ν(y).
|
|
|
the theorem follows from those relations.
□
Exercise 51
-
Show that the first three conclusions of the Fubini Theorem may
fail if f is not summable.
- Show that the fourth conclusion of the Fubini Theorem may
fail if f has values of different signs.
13.5 Absolute Continuity of Measures
Here, we consider another topic in the measure theory which benefits
from the integration theory.
Definition 52
Let X be a set with σ-algebra R and
σ-finite measure µ and finite charge ν on
R. The charge ν is absolutely continuous
with respect to µ if µ(A)=0 for A∈ R
implies ν(A)=0. Two charges ν1 and ν2 are
equivalent if two conditions
| ν1 |(A)=0 and | ν2 |(A)=0 are equivalent.
The above definition seems to be not justifying “absolute
continuity” name, but this will become clear from the following
important theorem.
Theorem 53 (Radon–Nikodym)
Any charge ν
which absolutely continuous with respect to a
measure µ
has the form
where f is a function from L1. The function
f∈
L1 is uniquely defined by the charge ν
.
Proof.[Sketch of the proof]
First we will assume that ν is a
measure. Let
D be the collection of measurable
functions
g:
X→[0,∞) such that
Let α = sup
g∈D ∫
X g d µ ≤ ν(
X) <
∞. So we can find a sequence (
gn) in
D
with ∫
X gn d µ → α.
We define f0(x) = supn gn(x). We can show that
f0=∞ only on a set of µ-measure zero, so if we
adjust f0 on this set, we get a measurable function
f:X→[0,∞). There is now a long argument to show
that f is as required.
If ν is a charge, we can find f by applying the previous
operation to the measures ν+ and ν− (as it is easy to
verify that ν+,ν−⋘µ).
We show that f is essentially unique. If g is another
function inducing ν, then
| ∫ | | f−g d µ = ν(E) − ν(E) = 0 (E∈L).
|
Let E = {x∈ X : f(x)−g(x)≥ 0}, so as f−g is
measurable, E∈L. Then ∫E f−g d µ =0 and
f−g≥0 on E, so by our result from integration theory, we
have that f−g=0 almost everywhere on E. Similarly, if F =
{x∈ X : f(x)−g(x)≤ 0}, then F∈L and
f−g=0 almost everywhere on F. As E∪ F=X, we conclude
that f=g almost everywhere.
□
Corollary 54
Let µ be a measure on X, ν be a finite charge, which
is absolutely continuous with respect to µ. For any
ε>0 there exists δ>0 such that
µ(A)<δ implies | ν |(A)<ε .
Proof.
By the Radon–Nikodym theorem there is a function
f∈
L1(
X,µ) such that ν(
A)=∫
A
f d µ. Then | ν |(
A)=∫
A
|
f |
d µ ad we get the statement from
Theorem
43 on absolute continuity
of the Lebesgue integral.
□
Last modified: November 6, 2024.