Measure Theory Primer
Measure Theory Primer
Pramook Khungurn
January 23, 2022
This is a primer on measure theory and Lebesgue integration. The materials are taken from Bartle’s “The
Elements Of Integration And Lebesgue Measure” [Bartle, 1995], Billingsley’s “Probability and Measure”
[Billingsley, 1995], and Hunter’s note on measure theory [Hunter, 2011].
1 Introduction
Why do we care about measure theory and Lebesgue integration?
– They expand the class of functions for which integrations are defined compared to what can be
achieved by Riemann integration.
– Theorems relating to the intechange of limits and integrals are valid under less stringent conditions
(again, compared to Reimann integration).
– In particular, the dominated convergence theorem 1 is a very powerful tool. For examples, it can
be use to easily show that
Z ∞ −nx
e
lim √ dx = 0
n→∞ 0 x
and
Z ∞ Z ∞
d
x2 e−tx dx = − x3 e−tx dx.
dx 0 0
– Measure theory is the foundation of modern probability theory, and the dominated convergence
theorem shows up everywhere in it.
– Riemann integrals are defined in terms of approximating a function with constant functions over
intevals.
– An interval is a subset of the real line which is of one of the following forms:
[a, b] = {x ∈ R : a ≤ x ≤ b},
[a, b) = {x ∈ R : a ≤ x < b},
(a, b] = {x ∈ R : a < x ≤ b},
(a, b) = {x ∈ R : a < x < b}.
The real number a and b are said to be the endpoints of the interval, and b − a is the length of
the interval.
1 https://en.wikipedia.org/wiki/Dominated_convergence_theorem
1
– A step function ϕ is a linear combination of a finite number of characteristic functions of intevals.
n
X
ϕ(x) = cj χEj (x)
i=1
– If f is a bounded function on [a, b], then the Reimann integral is defined to the limit of the
integrals of step functions that approximate f .
– The lower Rieman integral is defined to be the supremum of integrals of all step functions φ
such that φ(x) ≤ f (x) for all x ∈ [a, b] and φ(x) = 0 for all x 6∈ [a, b].
The Lebesgue integral is defined similarly, with some differences.
When studying integration, it is convenient to work with the extended real number system R =
R ∪ {−∞, ∞}.
2
– The arithematic operations between the infiniites and real numbers are as follows:
If the limit superior and limit inferere of a sequence both exist and are equal, then the limit of the
sequence exists and is equal to that value.
We can show that a σ-algebraSis closed under countableSintersections as well. To see this, we note that
∞ ∞ c c
Acn ∈ X for all n ∈ N, and so i=1 Acn ∈ X As result, A
i=1 n ∈ X . Applying de Morgan’s law, we
T∞ S∞ c c
have that i=1 An = A
i=1 n ∈ X .
Definition 2. A measurable space (X, X ) is a non-empty set X equipped with a σ-algebra X on X.
Definition 3. Let A be a non-empty collection of subsets of X. The σ-algebra generated by A,
denoted by σ(A) is the smallest σ-algebra that contains A. In other words,
\n o
σ(A) = à ⊆ P(X) : A ⊆ à and à is a σ-algebra .
Definition 4. The Borel algebra is the σ-algebra B generated by all the open intervals (a, b) in R.
Any set in B is called a Borel set.
3
Observe that we can write any open interval (a, b) as a countable unions of closed intervals:
∞
[ 1 1
(a, b) = a + ,b −
n n
n≥N
where N is an integer such that b − a − N2 > 0. As a result, B is also generated by the collection of
close intervals [a, b] in R. The same is also true for half-open intervals of the form (a, b] and [a, b).
Let X be the set R of extended real numbers. If E is a Borel set, then define
E1 = E ∪ {∞}
E2 = E ∪ {−∞}
E3 = E ∪ {−∞, ∞}
Let B the collection of all sets E, E1 , E2 , and E3 as E varies over B. We have that B is a σ-algebra,
and it is called the extended Borel algebra.
A “measure” encapsulates the notion of length, area, volume, mass, etc. of a set.
Definition 5. Let (X, X ) be a measurable space. A measure is a function µ : X → [0, ∞] with the
following properties.
1. µ(∅) = 0.
2. µ is countably additive. That is, for a sequence (En ) of disjoint sets, it holds that
∞
[ X∞
µ En = µ(En ).
n=1 n=1
Here is an example of a measure that is σ-finite but not finite. Let X = N, and X = P(N). Define µ(E)
to be the number of elements in E with the convention that µ(E) = ∞ when E is infinite. Obviously,
µ(N) = ∞. However, N = {1} ∪ {2} ∪ · · · , and µ({n}) = 1 for all n ∈ N. The measure µ is called the
counting measure on N.
Lemma 6. Let µ be a measure defined on a σ-algebra X . Let E, F ∈ X be such that E ⊆ F , then
µ(E) ≤ µ(F ). If µ(E) < ∞, then µ(F − E) = µ(F ) − µ(E).
Proof. Since F = E ∪ (F − E) and E ∩ (F − E) = ∅, it follows that
Because µ(F − E) ≥ 0, it follows that µ(F ) ≥ µ(E). If µ(E) < ∞, we can subtract from both sides of
the equation.
4
Lemma 7. If (En ) is an increasing sequence of measurable sets, then
[ ∞
µ En = lim µ(En ).
n→∞
n=1
Proof. Let (En ) be increasing. Set F0 = E1 , and Fn = En+1 − En for all n ≥ 1. We have that (Fn ) is
a sequence of disjoint sets. So,
[∞ [ ∞ X ∞
µ En = µ Fn = µ(Fn )
n=1 n=0 n=0
Sn
Also because En = i=0 Fi , we have that
n
X
µ(En ) = µ(Fi ).
i=0
So,
∞
[ X∞ n
X
µ En = µ(Fn ) = lim µ(Fi ) = lim µ(En ).
n→∞ n→∞
n=1 n=0 i=0
Next, let (En ) be decreasing and µ(E1 ) < ∞. Let Fn = E1 − En . We have that (Fn ) is increasing and
µ(Fn ) = µ(E1 ) − µ(En ). If follows that
[ ∞
µ Fn = lim µ(Fn ) = lim µ(E1 ) − µ(En ) = µ(E1 ) − lim µ(En )
n→∞ n→∞ n→∞
n=1
Now,
∞
\ ∞
[
En = E1 − Fn
n=1 n=1
So,
∞
\ ∞
[
µ En = µ(E1 ) − µ Fn = µ(E1 ) − µ(E1 ) − lim µ(En ) = lim µ(En )
n→∞ n→∞
n=1 n=1
are required.
5
Definition 10. A measure space (X, X , µ) is complete if every subset of a set of measure zero is
measureable.
Theorem 11. Let (X, X , µ) be a measure space. Define (X, X , µ) by
and
µ(A ∪ M ) = µ(A).
Then, (X, X , µ) is a complete measure space such that X ⊆ X and µ is the unique extension of µ to
X.
Proof (sketch). The hardest bit of the proof is to show that X is close under complementation. Let
A ∈ X , N ∈ X be a set of measure zero, and M ⊆ N . We have that (A ∪ M )c = Ac ∩ M c . Because
M c = N c ∪ (N − M ), we have that
6
2.3 Pi-Systems and Lambda-Systems
We take a detour to explore structures that are related to the σ-algebra that will be useful in further
studies.
Definition 12. A collection of sets A ⊆ P(X) is a π-system if it is closed under finite intersections.
In other words, if A, B ∈ A, then A ∩ B ∈ A.
Definition 13. A collection of sets A ⊆ P(X) is a λ-system it satisfies the following properties.
1. X ∈ A.
2. It is closed under complementation. That is, A ∈ A implies Ac ∈ A.
S∞ under countable disjoint unions. That is, if (An ) is a sequence of disjoint sets in A,
3. It is closed
then i=n An ∈ A.
Note that the π-system and the σ-system both have weaker conditions than that of the σ-algebra.
Lemma 14. A λ-system is closed under proper set differences. In other words, If A is a λ-system,
A, B ∈ A and A ⊆ B, then B − A ∈ A.
Proof. We have that B c ∈ A, and A ∩ B c = ∅. As a result, A ∪ B c ∈ A and so is (A ∪ B c )c = B − A.
Lemma 15. A collection A ⊆ P(X) that is both a π-system and a λ-system is a σ-algebra.
Proof. We only need to show that A is closed under countable (general) unions. Let (An ) be a sequence
of sets in A. Let Bn = An ∩ Acn−1 ∩ Acn−2 ∩ · · · Ac1 . We have that B
Sn∞∈ A because
S∞ it is formed by finite
intersections. Moreover, the Bn ’s are disjoint from one other, so n=1 Bn = n=1 An is a member of
A as well.
Definition 16. Let A ⊆ P(X) be a collection of sets. The λ-system generated by A, denoted by
λ(A) is the smallest λ-system that contains A. In other words,
\
λ(A) = L ⊆ P(X) : L is a λ-system and A ⊆ L .
.
Note that since a σ-algebra is a λ-system, we have that λ(A) ⊆ σ(A) for all A ⊆ P(X).
The following theorem goes in the opposite direction and is useful in proving many uniqueness theorems.
Proof. We will show that λ(A) is a σ-algebra. If this is the case, then σ(A) ⊆ λ(A) ⊆ B, and the
theorem holds.
By the last lemma, it is sufficient to show that λ(A) is closed under intersection. Doing this is rather
convoluted. First, for any A ⊆ X, define
GA = {B ⊆ X : A ∩ B ∈ λ(A)}.
7
S∞
Next, we show that GA is closed under countable disjoin unions.
S∞ Let B = n=1 Bn where each
Bn ∈ GA . We
S∞have that A ∩ Bn ∈ λ(A) for all n, and so is n=1 (A ∩ Bn ) = A ∩ ∪∞
n=1 Bn = A ∩ B.
Hence, B = n=1 Bn .
We have now established that A ∈ λ(A) =⇒ GA is a λ-system.
We shall now show that A ∈ A =⇒ λ(A) ⊆ GA . To see this, let B ∈ A. It follows that A ∩ B ∈ A
because A is a π-system. Because A ⊆ λ(A), it follows that A ∩ B ∈ λ(A). In other words, B ∈ GA ,
and so A ⊆ GA . Because A ∈ A ⊆ λ(A), it follows that GA is a λ-system that contains A. Thus,
λ(A) ⊆ GA .
Now,
A ∈ A =⇒ λ(A) ⊆ GA
A ∈ A =⇒ (B ∈ λ(A) =⇒ B ∈ GA )
(A ∈ A) ∧ (B ∈ λ(A)) =⇒ B ∈ GA .
Because B ∈ GA iff A ∈ GB , it follows that:
(A ∈ A) ∧ (B ∈ λ(A)) =⇒ A ∈ GB
B ∈ λ(A) =⇒ (A ∈ A =⇒ A ∈ GB )
B ∈ λ(A) =⇒ A ⊆ GB .
Because GB is a λ-system, it follows that λ(A) ⊆ GB . So,
B ∈ λ(A) =⇒ λ(A) ⊆ GB .
Hence,
A ∈ λ(A) ∧ B ∈ λ(A) =⇒ A ∈ λ(A) ∧ λ(A) ⊆ GB
A ∈ λ(A) ∧ B ∈ λ(A) =⇒ A ∈ GB
A ∈ λ(A) ∧ B ∈ λ(A) =⇒ A ∩ B ∈ λ(A)
A, B ∈ λ(A) =⇒ A ∩ B ∈ λ(A).
It follows that λ(A) is a π-system, and thus a σ-algebra.
So, E ∈ E as well.
8
2.4 Algebras and Monotone Classes
In this section, we take another detour on a structure that is similar to the σ-algebra and a theorem
about it that is similar to the Dynkin’s π-λ theorem.
Definition 19. A family A of subsets of a set X is said to be an algebra or a field on X if the
following properties are satisfied.
1. ∅, X ∈ A.
2. If E ∈ A, then E c = X − E ∈ A.
Sn
3. If E1 , E2 , . . . , En ∈ A, then i=1 Ei ∈ A.
Note that an algebra is a π-system because A ∩ B = (Ac ∪ B c )c . However, it is not a λ-system.
Definition 20. A class M of subsets of X is called a monotone class if it is closed under the
formation of monotone unions and intersections. In other words,
S∞
– If A1 , A2 , . . . ∈ M and Ai ⊆ Ai+1 for all i, then i=1 Ai ∈ M, amd
T∞
– If A1 , A2 , . . . ∈ M and Ai ⊇ Ai+1 for all i, then i=1 Ai ∈ M.
Note that a σ-algebra is a monotone class because it is already closed under arbitrary countable unions
and intersections.
Lemma 21. Let A be a subset of P(X). If A is both an algebra and a monotone class, then it is a
σ-algebra.
Proof. We only have to show Snthat A is closed under arbitrary countable unions. Let (An ) be a sequence
of sets in A. Define Bn = i=1 Ai . We have that S∞Bn ∈ A for all n because A is an algebra. Moreover,
because
S∞ B n ⊆
S∞Bn+1 for all n, it follows
S∞ that n=1 Bn ∈ A because A is a monotone class. Because
n=1 A n = n=1 B n , we have that n=1 A n ∈ A too, so A is a σ-algebra.
Definition 22. Let A be a subset of P(X). The monotone class generated by A, denoted by
m(A), is intersection of all monotone classes that contains A. That is,
\
m(A) = {M : M is a monotone class and A ⊆ M}.
9
S∞
which implies that n=1 An ∈ G1 . The condition involving countable monotone intersections can be
shown similarly.
Because A is an algebra, it follows that
A ∈ A ∧ B ∈ A =⇒ A ∪ B ∈ A.
Because A ⊆ m(A), we can also say that
A ∈ A ∧ B ∈ A =⇒ A ∪ B ∈ m(A)
A ∈ A =⇒ B ∈ A =⇒ A ∪ B ∈ m(A)
A ∈ A =⇒ ∀B∈A [A ∪ B ∈ m(A)],
A ∈ A =⇒ A ∈ G1 .
In other words, A ⊆ G1 . Because G1 is a monotone class and G1 contains A, it follows that m(A) ⊆ G1 .
Now,
m(A) ⊆ G1 ≡ A ∈ m(A) =⇒ A ∈ G1
≡ A ∈ m(A) =⇒ ∀B∈A [A ∪ B ∈ m(A)]
≡ A ∈ m(A) =⇒ B ∈ A =⇒ A ∪ B ∈ m(A)
≡ A ∈ m(A) ∧ B ∈ A =⇒ A ∪ B ∈ m(A).
Define G2 = {B : A ∪ B ∈ m(A) for all A ∈ m(A)}. It follows from the above statement that A ⊆ G2 .
Moreover, we can show again that G2 is a monotone class by repeating the same argument we used for
G1 . As a result, m(A) ⊆ G2 . In other words,
B ∈ m(A) =⇒ B ∈ G2
B ∈ m(A) =⇒ ∀A∈m(A) [A ∪ B ∈ m(A)]
B ∈ m(A) =⇒ A ∈ m(A) =⇒ A ∪ B ∈ m(A)
A ∈ m(A) ∧ B ∈ m(A) =⇒ A ∪ B ∈ m(A),
which means that m(A) is closed under finite unions. So, m(A) is an algebra.
3 Lebesgue Measures
The Lebesgue measure on R is a measure that corresponds to the notion of “length” on the real line. We
will construct it in this section.
10
Let F be the collection of subsets of R that contains all intervals of the forms
By the notion above, we have that ` is a function of signature F → [0, ∞]. However, we cannot claim
that it is a measure (1) we have not defined how to deal with countable unions of such intervals yet,
and (2) we have not identified the σ-algebra upon which it acts.
Furthermore, we cannot claim that F is a σ-algebra. However, F is an algebra.
We now turn back to the length function `. It turns out that we can show that it has more nice
properties that are worthy of being encapsulated with an abstraction.
Definition 25. Let A be an algebra on X. A premeasure on A is an extended real valued function
µ defined on A that satisfies the following properties.
1. µ(∅) = 0.
2. µ(E) ≥ 0 for all E ∈ A.
S∞
3. If (En ) is any disjoint sequence of sets in A such that i=n En belongs to A, then
∞
[ X∞
µ En = µ(En ).
n=1 n=1
11
A piece can be any of the 4 types. We will only deal with the (a, b] type in this proof as the proof of
other types are similar. Suppose, then, that
∞
[
(a, b] = (aj , bj ]
j=1
where the intervals are disjoint. Consider the first n intervals. We may assume that
We have that
n
X n
X
`((ai , bi ]) = (bi − ai ) ≤ bn − a1 ≤ b − a = `((a, b]).
i=1 i=1
For
P the other direction, let ε > 0 be arbitrary. Let (εj ) be a sequence of positive numbers with
εj < ε/2. Consider the interval Ij = (aj − εj , bk + εj ). The collection {Ij } of open sets is a cover of
the interval [a, b]. Since [a, b] is compact, it has a finite subcover, say, I1 , I2 , . . . , Im . By reordering
and discarding some intervals, we may assume that
a1 − ε1 < a
b < bm + εm
aj − εj < bj−1 + εj−1 .
If follows that
m
X m
X ∞
X
b − a ≤ (bm + εm ) − (a1 − ε1 ) ≤ [(bj + εj ) − (aj − εj )] ≤ ε + (bj − aj ) ≤ ε + (bj − aj ).
j=1 j=1 j=1
12
Definition 27. Given a premeasure µ defined on an algebra A on set X, define µ∗ : P(X) → [0, ∞]
to be
X∞
∗
µ (B) = inf µ(Ej ) : (Ej ) ∈ C(B)
j=1
S∞
where C(B) is the set of all sequences (Ej ) of sets in A such that B ⊆ j=1 Ej .
P∞ 28. For any set B ⊆ X and any real number ε > 0, there exists a sequence (Ej ) ∈ C(B) such
Lemma
that j=1 µ(Ej ) ≤ µ∗ (B) + ε.
P∞ is false. Then, there exists ε > 0 such that, for all (Ej ) ∈ C(B),
Proof. Suppose that the lemma
itPmust be the case that j=1 µ(Ej ) > µ∗ (B) + ε. It follows that µ∗ (B) + ε is a lower bound of
{ µ(Ej ) : Ej ∈ C(B)}. So,
∞
X
∗
µ (B) = inf µ(Ej ) : (Ej ) ∈ C(B) ≥ µ∗ (B) + ε,
j=1
and so
X X
µ∗ (A) = inf µ(Ej ) : (Ej ) ∈ C(A) ≤ inf µ(Ej ) : (Ej ) ∈ C(B) = µ∗ (B).
13
S
Let (Ej ) be a cover of A. We have that A = (A ∩ Ej ). Because µ is a measure,
∞
X ∞
X
µ(A) ≤ µ(A ∩ Ej ) ≤ µ(Ej ).
j=1 j=1
Definition 30. An outer measure µ∗ on a set X is a function µ∗ : P(X) → [0, ∞] such that the
following properties hold.
1. µ∗ (∅) = 0.
2. If E ⊆ F ⊆ X, then µ∗ (E) ≤ µ∗ (F ).
3. µ∗ is countably subadditive. In other words, if {Ei ⊆ X : i ∈ N} is a countable collection of
subsets of X, then
∞
[ ∞
X
µ∗ Ei ≤ µ∗ (Ei ).
i=1 i=1
µ∗ (A) = µ∗ (A ∩ E) + µ∗ (A ∩ E c )
A µ∗ -measurable set E splits any set A into pieces whose output measures add up to the outer measure
of A. In other words, a set is µ∗ -measurable if it splits other sets in a “nice” way.
Theorem 32 (Carathéodory’s extension theorem). Let µ∗ be an outer measure on X. The col-
lection of µ∗ -measurable sets is a σ-algebra on X. Moreover, µ∗ is a measure on this collection.
14
Proof. Let X ∗ denote the set of µ∗ -measurable sets on X. We shall show that X ∗ is a σ-algebra, and
µ∗ is a measure on it.
(∞, X ∈ X ∗ ) Because µ∗ is a premeasure, we have that µ∗ (∅) = 0. For any set A ⊆ X, we have that
Because ∅ and X are complements of each other, it follows that they are µ∗ -measurable and so belong
to X ∗ .
(Closure under complementation) Let E be a µ∗ -measurable set. It follows that
µ∗ (A) = µ∗ (∅) + µ∗ (A ∩ E) = µ∗ (A ∩ E c )
µ∗ (A) = µ∗ (A ∩ (E ∪ F )) + µ∗ (A ∩ (E ∪ F )c ).
µ∗ (A) ≤ µ∗ (A ∩ (E ∪ F )) + µ∗ (A ∩ (E ∪ F )c ).
Thus, it remains to show that µ∗ (A) ≥ µ∗ (A∩(E ∪F ))+µ∗ (A∩(E ∪F )c ). Because E is µ∗ -measurable,
we have that
µ∗ (A) = µ∗ (A ∩ E) + µ∗ (A ∩ E c ).
µ∗ (A ∩ E) = µ∗ (A ∩ E ∩ F ) + µ∗ (A ∩ E ∩ F c )
µ∗ (A ∩ E c ) = µ∗ (A ∩ E c ∩ F ) + µ∗ (A ∩ E c ∩ F c ).
As a result,
µ∗ (A) = µ∗ (A ∩ E ∩ F ) + µ∗ (A ∩ E ∩ F c ) + µ∗ (A ∩ E c ∩ F ) + µ∗ (A ∩ E c ∩ F c )
= µ∗ (A ∩ E ∩ F ) + µ∗ (A ∩ E ∩ F c ) + µ∗ (A ∩ E c ∩ F ) + µ∗ (A ∩ (E ∪ F )c ).
A ∩ (E ∪ F ) = (A ∩ E ∩ F ) ∪ (A ∩ E ∩ F c ) ∪ (A ∩ E c ∩ F ).
So,
µ∗ (A ∩ (E ∪ F )) ≤ µ∗ (A ∩ E ∩ F ) ∪ µ∗ (A ∩ E ∩ F c ) ∪ µ∗ (A ∩ E c ∩ F ).
Thus,
µ∗ (A) = µ∗ (A ∩ E ∩ F ) + µ∗ (A ∩ E ∩ F c ) + µ∗ (A ∩ E c ∩ F ) + µ∗ (A ∩ (E ∪ F )c )
≥ µ∗ (A ∩ (E ∪ F )) + µ∗ (A ∩ (E ∪ F )c ).
15
(Closure under disjoint countable unions implies closure under countable S unions) We need
to show closure under countable union. However, it suffices to only show that Ej is measurable for
all sequences (Ej ) where the sets are measurable and disjoint. To see this, let Fj denote the union of
the first j sets:
j
[
Fj = Ej ,
i=1
(Closure under disjoint countable unions) Let (Ej ) be a sequence of disjoint measurable sets.
Let
j
[ ∞
[
Fj = Ej , F = Ej .
i=1 i=1
We need to show show that, for any set A ⊆ X, it is true that µ∗ (A) = µ∗ (A ∩ F ) + µ∗ (A ∩ F c ).
Because µ∗ is subadditive, we already know that µ∗ (A) ≤ µ∗ (A ∩ F ) + µ∗ (A ∩ F c ). So, we only need
to show that µ∗ (A) ≥ µ∗ (A ∩ F ) + µ∗ (A ∩ F c ).
For any j, we have that Fj is measurable because of closure under finite unions. So
µ∗ (A) = µ∗ (A ∩ Fj ) + µ∗ (A ∩ Fjc )
[j
∗
=µ A∩ Ei + µ∗ (A ∩ Fjc )
i=1
n
[
= µ∗ (A ∩ Ei ) + µ∗ (A ∩ Fjc ).
i=1
j
X
µ∗ (A) ≥ µ∗ (A ∩ Ei ) + µ∗ (A ∩ F c ).
i=1
16
Taking the limit as j → ∞, we have that
∞
X
µ∗ (A) ≥ µ∗ (A ∩ Ei ) + µ∗ (A ∩ F c )
i=1
∞
[
∗
≥µ (A ∩ Ei ) + µ∗ (A ∩ F c )
i=1
∞
[
= µ∗ A ∩ Ei + µ∗ (A ∩ F c )
i=1
= µ∗ (A ∩ F ) + µ∗ (A ∩ F c ),
which implies that F is µ∗ -measurable.
(Countable additivity) In the proof of closure under countable union, we established that
∞
X
µ∗ (A) ≥ µ∗ (A ∩ Ei ) + µ∗ (A ∩ F c )
i=1
∞
[
∗
≥µ (A ∩ Ei ) + µ∗ (A ∩ F c )
i=1
= µ∗ (A ∩ F ) + µ∗ (A ∩ F c )
≥ µ∗ (A).
Therefore, it must be the case that
∞
X ∞
[
∗ ∗ ∗
µ (A ∩ Ei ) + µ (A ∩ F ) = µ c
(A ∩ Ei ) + µ∗ (A ∩ F c ),
i=1 i=1
In the proof of Theorem 32, there is a useful fact that we will use later, so let us extract it into a
lemma here.
Lemma 33. For any sequence of sets (Ej ) where each set belongs to an algebra A, there exists a
sequence of disjoint sets (Hj ) such that, for each j, we have that Hj ∈ A, and Hj ⊆ Ej . Moreover,
∞
[ ∞
[
Hj = Ej
j=1 j=1
Proof. Choose Hj to be Gj−1 as defined in the proof of Theorem 32. We have that all the properties
we want have already been proven except for Hj ⊆ Ej . However, this should be clear because
c c c
Hj = Gj−1 = Fj ∩ Fj−1 = (Ej ∪ Fj−1 ) ∩ Fj−1 = Ej ∩ Fj−1 ⊆ Ej .
We are done.
17
Proposition 34. The measure space (X, X ∗ , µ∗ ) in Theorem 32 is complete.
Proof. Let E be a µ∗ -measurable set of measure zero. Let B ⊆ E. We need to show that B is also
µ∗ -measurable, and that µ∗ (B) = 0. The latter statement is immediate because µ∗ is subadditive.
Let A ⊆ X. First, 0 ≤ µ∗ (A ∩ B) ≤ µ∗ (A ∩ E) ≤ µ∗ (E) = 0, so µ∗ (A ∩ B) = 0. Now,
A ∩ Bc ⊆ A
µ∗ (A ∩ B c ) ≤ µ∗ (A)
µ∗ (A ∩ B) + µ∗ (A ∩ B c ) ≤ µ∗ (A).
Theorem 35. Let µ be a premeasure on an algebra A on X. Let µ∗ be the outer measure generated
by µ, and let A∗ be the collection of µ∗ -measurable sets. Then, A ⊆ A∗ , and (X, A∗ , µ∗ ) is complete
measure space with the property that µ∗ (E) = µ(E) for all E ∈ A.
Proof. That µ∗ is a σ-algebra on A∗ follows from Theorem 32. That (X, A∗ , µ∗ ) is a complete measure
space follows from Proposition 34. That µ∗ (E) = µ(E) for all E ∈ A follows from Lemma 29. It
remains to show that A ⊆ A∗ .
Let E ∈ A. Let A ⊆ X. Let (Ej ) be a collection of sets in A that covers A. According to Lemma 28,
we can choose (Ej ) so that
∞
X ∞
X
∗
µ (Ej ) = µ(Ej ) ≤ µ∗ (A) + ε
j=1 j=1
for any arbitrary ε > 0. Because (Ej ) covers A, it follows that (Ej ∩ E) covers A ∩ E, and (Ej ∩ E c )
covers A ∩ E c . As a result,
∞
X ∞
X
µ∗ (A ∩ E) ≤ µ(Ej ∩ E) = µ∗ (Ej ∩ E),
j=1 j=1
∞
X X∞
µ∗ (A ∩ E c ) ≤ µ(Ej ∩ E c ) = µ∗ (Ej ∩ E c ).
j=1 j=1
c
Because Ej ∩ E and Ej ∩ E are disjoint, we have that
∞
X ∞
X
µ∗ (A ∩ E) + µ∗ (A ∩ E c ) ≤
∗
µ (Ej ∩ E) + µ∗ (Ej ∩ E c ) = µ(Ej ) ≤ µ∗ (A) + ε.
j=1 j=1
18
Proof. Let ν be an extension of µ that is a measure on A∗ . We have that ν(E) = µ(E) = µ∗ (E) for
all E ∈ A. We need to show that ν(E) = µ∗ (E) for all E ∈ A∗ .
Let E ⊆ A∗ . Let (Ej ) be a sequence of sets that covers E such that Ej ∈ A for all j. We have that
∞
[ ∞
X ∞
X
ν(E) ≤ ν Ej ≤ ν(Ej ) = µ(Ej ).
j=1 j=1 j=1
It follows that ν(E) is a lower bound of the set { µ(Ej ) : Ej ∈ C(E)}. As a result, ν(E) ≤ µ∗ (E) for
P
any E ∈ A∗ .
S
Since µ is σ-finite, there exists a sequence (Ej ) with each Ej ∈ A such that j Ej = X and µ(Ej ) =
µ∗ (Ej ) = ν(Ej ) is finite for all j. Applying Lemma 33 to (Ej ), we have a disjoint sequence (Hj ) that
covers X. Morever, µ(Hj ) ≤ µ(Ej ) < ∞.
For each Hj , we have that
Because µ∗ (Hj ∩ E) ≥ ν(Hj ∩ E) and µ∗ (Hj ∩ E c ) ≥ ν(Hj ∩ E c ), the equality can hold only if
µ∗ (Hj ∩ E) = ν(Hj ∩ E) and µ∗ (Hj ∩ E c ) = ν(Hj ∩ E c ).
Lastly,
∞
[ ∞
[ X∞
∗ ∗ ∗ ∗
µ (E) = µ (E ∩ X) = µ E∩ Hj =µ (E ∩ Hj ) = µ∗ (E ∩ Hj ).
j=1 j=1 j=1
Similarly,
∞
X
ν(E) = ν(E ∩ Hj ).
j=1
Because we just show that µ∗ (Hj ∩ E) = ν(Hj ∩ E) for all j, it follows that µ∗ (E) = ν(E).
19
Proposition 37. Let A be a Lebesgue measurable subset of R. There exists a Borel measurable subset
B of R such that A ⊆ B, and `∗ (B − A) = 0.
Proof. Let A be a Lebesgue measurable subset of R. Let us assume for now that `∗ (A) is finite. For
arbitrary ε > 0 and for each n ∈ N, let S (En,j ) be a sequence of sets in F such that (En,j ) covers A and
`(En,j ) ≥ `∗ (A) + ε/n. Let Bn = EP
P
n,j . We have that Bn is a Borel set because
T∞ it is a countable
union of intervals. Moreover, `∗ (Bn ) ≤ `(Bn,j ) ≤ `∗ (A) + ε/n. Take B = n=1 Bn . We have that
B is a Borel set because it is a countable Snintersection of Borel sets. We also have that A ⊆ B because
each Bn covers A. Moreover, let Bn0 = i=1 Bi . We have that `∗ (B10 ) is finite, and the sequence Bn0 is
decreasing, and `∗ (Bn0 ) ≤ `∗ (Bn ) ≤ µ∗ (A) + ε/n By Lemma 7, we have that
ε
`∗ (B) = lim `∗ (Bn0 ) ≤ lim `∗ (Bn ) ≤ lim `∗ (A) + = `∗ (A).
n→∞ n→∞ n→∞ n
However, because A ⊆ B, it follows that `∗ (B) ≥ `∗ (A), and thus `∗ (B) = `∗ (A). Because `∗ (B) is
finite, it follows that `∗ (B − A) = `∗ (B) − `∗ (A) = 0.
Let us now remove the assumption that `∗ (A) is finite. Let Ak = A ∩ [k, k + 1) for any n ∈ Z. We have
that the Ak ’s are disjoint. Moreover `∗ (Ak ) ≤ `∗ ([k, k + 1)) = 1. Because Ak is finite,Sthere exists a
∞
Borel set Bk such that Ak ⊆ Bk and `∗ (Bk ) = `∗ (Ak ) and `∗ (Bk − Ak ) = 0. Take B = k=−∞ Bk . It
Sj union of Borel sets. Moreover, A ⊆ B obviously.
follows that B is aTBorel set because it is a countable
∞
Lastly, B − A = k=−∞ (Bk − Ak ). Let Cj = k=−j (Bj − Aj ). We have that Cj is an increasing
sequence of sets. By Lemma 7,
∞
[
∗ ∗
` (B − A) = ` Cj = lim `∗ (Cj ) = 0.
j→∞
j=0
We are done.
Here, the length of intervals becomes the volume of “rectangles” of the form
I1 × I2 × · · · × Ik
The measure generated in this way is called the Lebesgue measure on Rk . There are also Borel
measure on Rk and Borel algebra on Rk . All the results proved thus far do hold on them.
4 Measurable Functions
An important component of measure theory is the definition of integrals of functions and the study of their
properties. Measurable functions are nice functions upon which integrals can be defined.
20
Definition 38. A function f : X → R is said to be X -measurable (or simply measurable) if, for
every real number α, the set {x ∈ X : f (x) > α} is measurable (in other words, belongs to X ).
Proposition 39. For a function f : X → R, the following statements are equivalent.
are also measurable. For the case of 1/f , we assume that f (x) 6= 0 for all x.
The RHS is measurable, so cf is measurable. The case where c < 0 can be handled similarly.
(b) If α < 0, then {x ∈ X : (f (x))2 > α} = X ∈ X . If α ≥ 0, then
√ √
{x ∈ X : (f (x))2 > α} = {x ∈ X : f (x) > a} ∪ {x ∈ X : f (x) < − a}.
21
(c) We have that
[
{x ∈ X : f (x) + g(x) < α} = {x ∈ X : f (x) < q} ∩ {x ∈ X : g(x) < r}.
q+r<b;q,r∈Q
In all cases, the RHS is a measurable set. It follows that 1/f is also measurable if f (x) 6= 0 for all x.
(g) and (h) We have that
Definition 41. For any function f : X → R, let f + and f − be non-negative functions defined by:
22
Lemma 43. An extended real-valued function f is measurable if and only if (1) the sets
A = {x ∈ X : f (x) = ∞},
B = {x ∈ X : f (x) = −∞}
are measurable, and (2) the real-valued function f1 defined by
(
f (x), x ∈ A ∪ B,
f1 (x) =
0, x 6∈ A ∪ B
is measurable.
Proof. TODO
The consequence to the last lemma and the lemmas in the last section is that, if f is M (X, X ), then
the functions
cf, f 2 , |f |, f + , f −
are also in M (X, X ).
For cf , we use the convention that 0(±∞) = 0.
For f + g, note that the sum is not well-defined when f (x) and g(x) are infinities with different signs.
Lemma 44. Let (fn ) be a sequence of functions in M (X, X ). Define
f (x) = inf {fn (x)},
n≥1
Corollary 45. If (fn ) is a sequence in M (X, X ) which converges to f on X, then f is also in M (X, X ).
Proof. We have that f (x) = lim fn (x) = lim inf fn (x).
The following lemma establishes the fact that a measurable non-negative function can be approximated
by an increasing sequence of functions, all of which takes on a finite number of real values.
Lemma 47. If f is non-negative function in M (X, X ), then there exists a sequence (ϕn ) in M (X, X )
such that:
(a) 0 ≤ ϕn (x) ≤ ϕn+1 (x) for all x ∈ X and n ∈ N.
(b) f (x) = lim ϕn (x) for each x ∈ X.
(c) Each ϕn has only a fininte number of real values.
Proof. TODO
23
5 Integration
In this section, we consider a fixed measure space (X, X , µ).
Denote the set of all X -measurable functions by M = M (X, X ). Denote the set of all non-negative
X -measurable functions by M + = M + (X, X ).
Definition 48. A real-valued function is simple if it has only a finite number of values.
A simple measurable function ϕ can be represented by
n
X
ϕ= aj χEj
j=1
Note that the value of the integral is always well defined because all the aj ’s are positive, so we never
encounter an expression of the form ∞ + (−∞).
Lemma 50. If ϕ and ψ are simple functions in M + (X, X ) and c ≥ 0, then
Z Z
cϕ dµ = c ϕ dµ,
Z Z Z
(ϕ + ψ) dµ = ϕ dµ + ψ dµ.
Proof. TODO
Lemma 51. Let ϕ be a simple function in M + (X, X ). For each E ∈ X , define λ(E) to be
Z
λ(E) = ϕ χE dµ.
Then, λ is a measure on X .
Proof. TODO
Definition 52. If f belongs to M + (X, X ), define the (Lebesgue) integral of f with respect to µ
to be the extended real number
Z Z
f dµ = sup ϕ dµ ϕ ∈ M + (X, X ) is simple, and ϕ(x) ≤ f (x) for all x ∈ S
For any E ∈ X , define the (Lebesgue) integral of f over E with respect to µ to be the extended
real number
Z Z
f dµ = f χE dµ.
E
24
Lemma 53. If f, g ∈ M + (X, X ), then
R R
f dµ ≤ g dµ.
Proof. TODO
Proof. TODO
Proof. TODO
Proof. TODO
Proof. TODO
then λ is a measure.
Proof. TODO
Corollary
R 60. Let f belong to M + (X, X ). Then, f (x) = 0 µ-almost everyone on X if and only if
f dµ = 0.
Proof. TODO
Corollary 61. Let f belong to M + (X, X ), and let λ be defined as in Corollary 59. Then, the measure
λ is absolutely continuous with respect to µ in the sense that if E ∈ X , then µ(E) = 0, then λ(E) = 0.
Proof. TODO
25
Corollary 62. If (fn ) is a monotonically increasing sequence of functions in M + (X, X ) which con-
verges µ-almost everywhere on X to a function f in M + (X, X ), then
Z Z
f dµ = lim fn dµ.
n→∞
Proof. TODO
Proof. TODO
6 Integrable Functions
Definition 64. The collection L = L(X, X , µ) of integrable functions consists of all real-valued
X -measurable functions f defined on X, such that both the positive and negative parts (f + and f − ) of
f have finite integrals with respective to µ.
Definition 65. If f ∈ L(X, X , µ), the integral of f with respect to µ is defined to be
Z Z Z
f dµ = f + dµ + f − dµ.
If E ∈ X , define
Z Z Z
f dµ = f + dµ + f − dµ.
E E E
The difference between a charge and a measure is that a measure is always non-negative, but a charge
can be negative. Moreover, a charge cannot take infinite values.
Lemma 67. If f belongs to L, and λ(E) = E f dµ, then λ is a charge.
R
Proof. TODO
The function λ defined as in the above lemma is often called the indefinite integral of f with
respect to µ.
Since λ is a charge, if (En ) is a disjoint sequence of sets in X whose union is E, then
Z X∞ Z
f dµ = f dµ.
E n=1 En
26
Theorem 68. A measurable function f belongs fo L if and only if |f | belons to L. Moreover,
Z Z
f dµ ≤ |f | dµ.
as required.
Theorem 70. A constant multiple cf and a sum f + g of integrable functions are integrable, and the
integrals are given by:
Z Z
cf dµ = c f dµ,
Z Z Z
(f + g) dµ = f dµ + g dµ.
Proof. TODO
Proof. TODO
for each x ∈ X, and that there exists an integrable function g such that |f (x, t)| ≤ g(x). Then,
Z Z
f (x, t0 ) dµ(x) = lim f (x, t) dµ(x).
t→t0
27
Proof. Just apply the dominated convergence theorem.
Corollary 73. Let the function t 7→ f (x, t) be continuous on [a, b] for all x ∈ X. If there is an
integrable function g on X such that |f (x, t)| ≤ g(x), then the function F defined by
Z
F (t) = f (x, t) dµ(x)
Corollary 74. Suppose that for some t0 ∈ [a, b], the function x 7→ f (x, t0 ) is integrable on X. Suppose
that ∂f /∂t exists on X × [a, b]. Moreover, let there be an integrable function g on X such that
∂f
(x, t) ≤ g(x).
∂t
Corollary 75. Let the function t 7→ f (x, t) be continuous on [a, R b] for all x ∈ X. Let there be an
integrable function g on X such that |f (x, t)| ≤ g(x). Let F (t) = f (x, t) dµ(x). We have that
Z b Z b Z Z Z b
F (t) dt = f (x, t) dµ(x) dt = f (x, t) dt dµ(x)
a a a
– If we force the measure to agree on rectangles, though, the measure would agree with the Lebesgue
measure on the Borel sets in Rk by the Hahn’s extension theorem.
Definition 76. If (X, X ) and (Y, Y) are measurable spaces, then the set of the form A × B with A ∈ X
and B ∈ B are called a measurable rectangle or simply rectangle in Z = X × Y . Let Z0 denote
the set of all finite unions of rectangles in Z.
It can be shown that each element of Z0 can be written as a finite union of disjoin rectangles in Z.
(Just use the inclusion–exclusion principle.)
28
Lemma 77. The collection Z0 is an algebra on Z.
Proof. (∅ ∈ Z0 ) We have that ∅ ∈ X and ∅ ∈ Y. So, ∅ = ∅ × ∅ ∈ Z0 .
(Z ∈ Z0 ) Since X ∈ X and Y ∈ Y, we have that Z = X × Y ∈ Z0 .
(Closure under complementation) If C ∈ Z0 , then C = A × B for some A ∈ X and B ∈ Y. It
follows that C c = (Ac × B) ∪ (A × B c ) ∪ (Ac × B c ), which is a finite unions of rectangles in Z, so
C c ∈ Z0 .
Closure under finite union follows from definition of Z0 .
Definition 78. Let (X, X ) and (Y, Y) be measurable spaces. Let Z = X × Y denote the σ-algebra of
subsets of Z = X × Y generated by rectangles A × B with A ∈ X and B ∈ Y. A set in Z is called a
Z-measurable set or as a measurable subset of Z.
Theorem 79 (Product measure theorem). Let (X, X , µ) and (Y, Y, ν) be measure spaces. There
exists a measure π defined on Z = X × Y with
π(A × B) = µ(A) ν(B)
for all A ∈ X and B ∈ Y. If the measure is σ-finite, then then it is unique. The measure π is called
the product of µ and ν.
Proof. TODO
Definition 82. For any f : Z → R, and x ∈ X, the x-section of f is the function fx defined on Y as
fx (y) = f (x, y)
for all y ∈ Y . The y-section is defined as
f y (x) = f (x, y)
for all x ∈ X.
Lemma 83. If f : Z → R is a measurable function, then every section of f is measurable.
Proof. TODO
Lemma 84. Let (X, X , µ) and (Y, Y, ν) be σ-finite measure spaces. If E ∈ Z = X × Y, then the;
functions defined by
f (x) = ν(Ex ),
g(y) = µ(E y )
are measurable and
Z Z
f dµ = π(E) = g dν.
X Y
29
Proof. TODO. This requires the monotone class theorem (Theorem 23).
Theorem 85 (Tonelli’s theorem). Let (X, X , µ) and (Y, Y, ν) be σ-finite measure spaces. Let π be
the product of µ and ν, which is a measure on Z = X ×Y. Let F : Z → R be a non-negative measurable
function. Then, the functions defined on X and Y by
Z
f (x) = Fx dν,
Y
Z
g(y) = F y dµ
X
Equivalently,
Z Z Z Z Z
Fx dν dµ = F dπ = F y dµ dν.
X Y Z Y X
Proof. TODO. This requires the monotone class theorem (Theorem 23).
Theorem 86 (Fubini’s theorem). Let (X, X , µ) and (Y, Y, ν) be σ-finite measure spaces. Let π be
the product of µ and ν, which is a measure on Z = X × Y. Let F : Z → R be an integrable function
with respect to π. Then, the extended real value functions defined almost everywhere by
Z
f (x) = Fx dν,
Y
Z
g(y) = F y dµ
X
Equivalently,
Z Z Z Z Z
Fx dν dµ = F dπ = F y dµ dν.
X Y Z Y X
Proof. TODO
30
The condition (b) is introduced to avoid undefined quantities such as ∞ − ∞.
A signed measure is very similar to a charge, but it can take an infinite value while a charge cannot.
Note that a charge is a signed measure, but not the other way around.
It can be shown that, if (En ) is an increasing sequence of sets in X , then
∞
[
λ En = lim λ(En ).
n→∞
n=1
Definition 88. Let λ be a signed measure on (X, X ). A set P in X is said to be positive with respect
to λ if λ(E ∩ P ) ≥ 0 for any E ∈ X . A set N ∈ X is said to be negative with respect to λ if
λ(E ∩ N ) ≤ 0 for all E ∈ X . A set M ∈ X is said to be a null set for λ if λ(E ∩ M ) = 0 for all
E ∈ X.
It can be shown that a measurable subset of a positive set is positive, and the union of two positive
sets is also positive.
Theorem 89 (Hahn decomposition theorem). If λ is a signed measure on X , then there exist
sets P and N in X with X = P ∪ N , P ∩ N = ∅, and such that P is positive and N is negative with
respect to λ.
Proof. TODO
A pair (P, N ) of measurable sets in the above theorem is said to form a Hahn decomposition of X
with respect to λ.
In general, Hahn decompositions are not unique. In fact, if M is a null set for λ, then (P ∪ M, N − M )
and (P − M, N ∪ M ) are also Hahn decompositions of X w.r.t λ.
Lemma 90. If (P1 , N1 ) and (P2 , N2 ) are Hahn decompositions of X w.r.t λ, and E ∈ X , then
λ(E ∩ P1 ) = λ(N ∩ P2 ),
λ(E ∩ N1 ) = λ(N ∩ N2 ).
Proof. TODO
Definition 91. Let λ be a signed measure on X , and let (P, N ) be a Hahn decomponent w.r.t λ. The
positive and negative variations of λ are the non-negative measures λ+ , λ− defined by:
λ+ (E) = λ(E ∩ P ),
λ− (E) = −λ(E ∩ N ).
By Lemma 90, the positive and negative variations are well defined and do not depend on the particular
choice of Hahn decomposition.
31
We also have that
λ(E) = λ+ (E) − λ− (E),
and this result is encapsulated in the following theorem.
Theorem 92 (Jordan decomposition theorem). If λ is a signed measure on X , then there exists
measures λ+ and λ− on X , at least one of which is finite, such that the following conditions are
satisifed.
(a) λ = λ+ − λ− .
(b) There exists sets M and N such that M ∪ N = X, M ∩ N = ∅, λ+ (N ) = 0, and λ− (M ) = 0.
(c) If λ = µ − ν where µ and ν are measures, then µ(E) ≥ λ+ (E) and ν(E) ≥ λ− (E).
Proof. TODO
Proof. TODO
10 Radon–Nikodym Theorem
From Corollary 59, we can create a measure λ from a non-negative extended real-valued measurable
function f by integrating with respect to an existing measure.
The Radon–Nikodym theorem is the converse of the above corollary. It indicates when a measure λ
can be expressed as an integration of a function f with respect to µ. Hence, it is useful in defining
probability density functions.
A necessary and sufficient condition for the theorem to hold is given below.
Definition 94. A measure λ on X is said to be absolutely continuous with respect to a measure
µ on X if λ(E) = 0 for all set E ∈ X such that µ(E) = 0. In this case, we write λ µ. A signed
measure λ is absolsution continuous with respect to a signed measure µ if the total variation |λ| is
absolutely continuous with respect to |µ|.
Lemma 95. Let λ and µ be finite measures on X . Then, λ µ if and only if, for every ε > 0, there
exists a δ(ε) > 0 such that λ(E) < ε for all E such that µ(E) < δ(ε).
Proof. TODO
32
Proof. TODO
The function f above is often called the Random–Nikodym derivative of λ with respective to µ,
and it is denoted by dλ/dµ.
References
[Bartle, 1995] Bartle, R. G. (1995). The Elements of Integration and Lebesgue Measure. John Wiley & Sons,
New York.
[Billingsley, 1995] Billingsley, P. (1995). Probability and Measure. John Wiley and Sons, third edition.
[Hunter, 2011] Hunter, J. K. (2011). Measure theory. https://www.math.ucdavis.edu/~hunter/measure_
theory/measure_notes.pdf. Accessed: 2021-12-28.
33