Lebesgue Theory
Lebesgue Theory
Lebesgue Theory
A Brief Overview
Ralf Pihlström
Examensarbete i matematik, 15 hp
Handledare: Anders Öberg
Examinator: Jörgen Östensson
Juni 2016
Department of Mathematics
Uppsala University
Abstract
We present different concepts in measure theory. We construct the
Lebesgue measure on R. We define the Lebesgue integral and prove
some famous convergence theorems. We compare the Riemann integral
with the Lebesgue integral and prove some famous results. We give
some examples of how to use the Lebesgue integral.
By doing the above, we hope to give a brief overview of the Lebesgue
theory.
Contents
1 Introduction 2
2 Outline 4
3 Measure on a σ-algebra 5
3.1 σ-algebra of sets . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.2 Limits of sequences of sets . . . . . . . . . . . . . . . . . . . . 6
3.3 Generated σ-algebras . . . . . . . . . . . . . . . . . . . . . . . 7
3.4 Borel σ-algebras . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.5 Measure on a σ-algebra . . . . . . . . . . . . . . . . . . . . . 8
3.6 Measures of a sequence of sets . . . . . . . . . . . . . . . . . . 10
3.7 Measurable space and measure space . . . . . . . . . . . . . . 12
4 Outer measures 13
4.1 Construction of outer measures . . . . . . . . . . . . . . . . . 14
5 Lebesgue measure on R 15
5.1 The Borel σ-algebra of R . . . . . . . . . . . . . . . . . . . . 17
5.2 The invariance of the Lebesgue measure space . . . . . . . . . 17
5.3 Existence of non Lebesgue measurable sets . . . . . . . . . . . 19
6 Measurable functions 20
6.1 Measurability of functions . . . . . . . . . . . . . . . . . . . . 20
6.2 Measurability of sequences of functions . . . . . . . . . . . . . 21
6.3 Measurability of the positive and negative part of f . . . . . 22
6.4 Measurability of the restriction and extension of f . . . . . . 23
6.5 Almost everywhere . . . . . . . . . . . . . . . . . . . . . . . . 23
2
7.6 Convergence theorems . . . . . . . . . . . . . . . . . . . . . . 30
1 Introduction
In 1904 Henri Lebesgue invented a new way of integrating functions. His
theory of integration was a generalization of that of Riemann’s—a larger set
of functions could be integrated and the problem of limits interacting badly
with integrals was solved.
In the center of the Lebesgue integral stands the following idea: the limit
of an integral should equal the integral of the limit. In other words,
Z Z
lim fn = lim fn . (1)
n→∞ n→∞
(This equality holds true in the Riemann sense too, but under less mild
conditions.) The aim of this paper is to establish (1). The following is some
background that will provide a starting point.
Suppose we have (1), and consider any nonnegative function f . Let
0 ≤ f1 ≤ f2 ≤ . . . be any sequence of simple functions1 approximating f
from below. (Such a sequence is guaranteed to exist.) Since integrals should
be sensitive to size,
Z Z
lim fn = sup{ fn }, n = 1, 2, 3, . . . .
n→∞
But Z Z
lim fn = f,
n→∞
1
A simple function is a function with finite range
3
where the supremum is taken over all simple function ϕ such that 0 ≤ ϕ ≤ f .
(For a negative functionR we simply consider −f .)
It remains to define ϕ where ϕ is simple. As we know, Riemann defined
his integral by partitioning the domain. Let us try something different.
Instead of partitioning the domain, let us partition the range. This was
Lebesgue’s breakthrough idea. It is the Lebesgue integral in a nutshell.
See Figure 1.
Figure 1: Riemann integration (blue) and Lebesgue integration (red). The Rie-
mann integral partitions the domain of f , while the Lebesgue integral partitions
the range of f
where
• ai is the different values ϕ assumes,
4
• m(Di ) is the size or measure of Di .
2 Outline
Section 3, Measure on a σ-algebra
We define a σ-algebra. We define sequences of sets and their limit.
We define generated σ-algebras. We define the Borel σ-algebra. We
define the concept of a measure on a σ-algebra. We define measure
sequences and their limits, and some important results regarding them.
We define the concept of measurable space and measure space.
5
Section 9, Comparison with the Riemann integral
We compare the Riemann integral with the Lebesgue integral. We
show that if a bounded function is Riemann integrable over an interval
[a, b], it is Lebesgue integrable. We show that f is Riemann integrable
if and only if the set of discontinuity points has measure 0. We prove
the counterpart to the Lebesgue dominated convergence theorem in
the Riemann sense. We demonstrate in a few examples how to use the
Lebesgue integral.
3 Measure on a σ-algebra
Notations. N = {1, 2, 3, . . .}, Z = {0, ±1, ±2, . . .}, R is the real line. P(X)
is the collection of all subsets of a set X.
Let X be any set. Suppose we want a function that measures subsets of
X. Let us call such a function a measure, and denote it by
2 ◦ A ∈ A ⇒ Ac ∈ A,
3 ◦ A, B ∈ A ⇒ A ∪ B ∈ A.
Lemma 1. If A is an algebra of subsets of X, then
6
(1) A, B ∈ A ⇒ A ∩ B ∈ A,
(2) A, B ∈ A ⇒ A \ B ∈ A.
Proof. A ∩ B = (Ac ∪ B c )c ∈ A, A \ B = A ∩ B c ∈ A.
4◦ (An : n ∈ N) ⊂ A ⇒
S
n∈N An ∈ A.
Lemma
T 2. If A is a σ-algebra of subsets of set X, then (An : n ∈ N) ⊂ A
⇒ n∈N An ∈ A.
T S c
Proof. n∈N An = n∈N An ∈ A.
In order to define the limit of sequences of sets, we must define the limit
inferior and the limit superior of sequences of sets.
Definition 4 (Limit inferior, limit superior). The limit inferior and the
limit superior of a sequence (An : n ∈ N) of subsets of a set X is defined by
[ \
lim inf An = Ak ,
n→∞
n∈N k≥n
\ [
lim sup An = Ak .
n→∞
n∈N k≥n
7
Definition 5 (Convergent sequence). Let (An : n ∈ N) be an arbitrary
sequence of subsets of a set X. We say that the sequence converges if
lim inf n→∞ An = lim supn→∞ An , and we set
Remark 1. It can be shown that for any sequence (An : n ∈ N), both
lim inf n→∞ An and lim supn→∞ An lies in A. In particular if (An : n ∈ N)
converges, then limn→∞ An ∈ A.
1 ◦ ∅ ∈ D,
2 ◦ X ∈ D,
8
3 ◦ {Eα : α ∈ A} ⊂ D ⇒
S
α∈A Eα ∈ D,
4 ◦ E1 , E2 ∈ D ⇒ E1 ∩ E2 ∈ D.
The pair (X, D) is called a topological space, and the sets in D are called the
open sets of the topological space.
9
(g) γ is countable subadditive on C if γ(∪nn∈N En ) ≤ n∈N γ(En ) for every
P
sequence (En : n ∈ N) in C such that ∪n∈N En ∈ C.
Lemma 3. For any sequence (En : n ∈ N) in an algebra A of subsets of X,
there exists a disjoint sequence (Fn : n ∈ N) in A such that
N
[ N
[ [ [
(1) En = Fn and (2) En = Fn .
n=1 n=1 n∈N n∈N
S
In particular, if A is a σ-algebra, then n∈N Fn ∈ A.
Proof. Let F1 = E1 and Fn = En \ (∪n−1
k=1 Ek ) ∈ A, and use induction.
10
3.6 Measures of a sequence of sets
Theorem 2 (Monotone convergence theorem for sequences of measurable
sets). Let µ be a measure on a σ-algebra A of subsets of a set X. Let (En :
n ∈ N) be a monotone sequence in A.
where, since En ↓,
! !
\ \
µ E1 \ En = µ(E1 ) − µ En = µ(E1 ) − µ( lim En ). (6)
n→∞
n∈N n∈N
11
Now
!
[ X X
µ Fn = µ(Fn ) = µ(En \ En+1 ) (7)
n∈N n∈N n∈N
X n
X
= [µ(En ) − µ(En+1 )] = lim [µ(En ) − µ(En+1 )] (8)
n→∞
n∈N k=1
= lim [µ(E1 ) − µ(En+1 )] = µ(E1 ) − lim µ(En+1 ). (9)
n→∞ n→∞
Substituting (3) and (4) into (2), we arrive at the desired result.
(d) If lim En exists and if there exists A ∈ A with µ(A) < ∞ such that
n→∞
En ⊂ A for n ∈ N then lim µ(En ) and
n→∞
since ∪k≥n Ek ⊂ En ⊂ A.
3. By (a),
12
4. By (b) and (a),
lim sup µ(En ) ≤ µ(lim sup En ) = µ(lim inf En ) ≤ lim inf µ(En ).
n→∞ n→∞ n→∞ n→∞
But lim inf n→∞ µ(En ) ≤ lim supn→∞ µ(En ) for any sequence (En : n ∈
N), and the result follows.
4 Outer measures
In this section we will introduce the outer measure. By its help, we will
connect the two concepts measure and σ-algebra.
13
Definition 18 (µ∗ -measurable set). Let µ∗ be an outer measure on a set
X. We say that E ∈ P(X) is measurable with respect to µ∗ (or simply µ∗ -
measurable) if it satisfies the so-called Caratheodory condition:
Proof. Let us prove (b). Let E1 ∪ E2 be the testing set in the Caratheodory
condition. Since E1 and E2 are disjoint, µ∗ (E1 ∪ E2 ) = µ∗ ((E1 ∪ E2 ) ∩ E1 ) +
µ∗ ((E1 ∪ E2 ) ∩ E1c ) = µ∗ (E1 ) + µ∗ (E2 ).
µ∗ (A ∩ E0 ) + µ∗ (A ∩ E0c ) ≤ µ∗ (A),
14
This is take-away: an outer measure induces a σ-algebra. And since the
measure is just the restriction of the outer measure, one may say that a
measure induces a σ-algebra. This way, the concept of a σ-algebra becomes
more clear.
2 ◦ ∅ ∈ B.
S
For every E ∈ P(X), if (Vn : n ∈ N) ∈ B such that E ⊂ n∈N Vn then
(Vn : n ∈ N) is a covering sequence for E.
2 ◦ γ(∅) = 0.
15
5 Lebesgue measure on R
In this section we construct the Lebesgue measure on R (using the Lebesgue
outer measure) and demonstrate some of its properties.
and call it the Lebesgue outer measure on R. We write ML for the σ-algebra
M(µ∗L ) of µ∗L -measurable sets E ∈ B(R) and call it the Lebesgue σ-algebra
of subsets of R. Members of the σ-algebra ML are called ML -measurable or
Lebesgue measurable sets. We call (R, ML ) the Lebesgue measurable space.
We write µL for the restriction of µ∗L to ML and call it the Lebesgue measure
on R. We call (R, ML , µL ) the Lebesgue measure space on R.
If any of the intervals is infinite, (10) holds. Thus consider the case
where every member is finite. Let us drop those members in the covering
sequence that is disjoint from I and contained in any other member of the
sequence. The resulting sequence Jn is a covering sequence of I, and since I
is compact, Jn has a finite subcover. Renumber the members of Jn so that
Jk = (ak , bk ) for k = 1, . . . , N and a1 ≤ a2 ≤ . . . ≤ aN . In fact since none of
the members are contained in another, we have a1 < a2 < . . . < aN . Let us
show that a2 < b1 . Assume not, then since J1 and J2 lies in I, there exist
x1 ∈ (a1 , b1 ) ∩ I and x2 ∈ (a2 , b2 ) ∩ I such that a1 < x1 < b1 ≤ a2 < x2 < b2 .
Note that [x1 , x2 ] ⊂ I. Since b1 ≤ a2 , there exists at least on point in [x1 , x2 ]
16
that is not covered by (J1 , . . . , Jn ). Thus we have a2 < b1 . Similarly,
a1 < a2 < b1
a2 < a3 < b2
...
aN −1 < aN < bN −1
aN < bN .
We get
N
X
l(Jk ) = (b1 − a1 ) + (b2 − a2 ) + . . . + (bN − aN )
k=1
> (a2 − a1 ) + (a3 − a2 ) + . . . + (bN − an )
= bN − a1 ≥ b − a = l(I).
µ∗L ((a, b)) ≤ µ∗L ([a, b]) ≤ µ∗L ({a}) + µ∗L ((a, b)) + µ∗L ({b}) = µ∗L ((a, b)),
since the Lebesgue measure of a singleton is 0. Thus µ∗L ((a, b)) = µ∗L ([a, b]) =
l([a, b]) = l((a, b)).
3. If I is finite and I = (a, b], µ∗L ((a, b]) ≤ µ∗L ((a, b)) and µ∗L ((a, b]) ≥
µ∗L ((a, b)) by monotonicity. Similarly if I = [a, b).
4. If I is infinite of the type I = (a, ∞), a ∈ R, then (a, ∞) ⊂ (a, n) for
every n ∈ N and thus µ∗L ((a, ∞)) ≥ µ∗L ((a, n)) = n − a. By the arbitrariness
of n ∈ N, µ∗L ((a, ∞)) = ∞ = l(a, ∞). Similarly for other types.
17
Proof. Let D be the collection of all open sets in R. Since any open set
in R can be written as a union of countable many open intervals, we have
D ⊂ ML and thus σ(D) = B R ⊂ ML .
18
Theorem 10. For E ∈ B(R) and a ∈ R, let αE = {y ∈ R : y =
ax for some x ∈ E}. Then µ∗L (αE) = |α|µ∗L (E).
Now µ∗L (E) = inf S∈SE λ(S) and µ∗L (αE) = inf T ∈SαE λ(T ). But Mα (S E )
is one-to-one, so inf T ∈SαE λ(T ) = inf S∈SE λ(Mα (S)) = inf S∈SE |α|λ(S) =
|α|µ∗L (E). Thus µ∗L (αE) = |α|µ∗L (E) when α 6= 0. For α = 0 the equality is
trivial.
1 1 1
µ∗L ( A) = µ∗L ( A ∩ E) + µ∗L ( A ∩ E c ). (11)
α α α
By Theorem 10,
1 1 ∗
µ∗L ( A) = µ (A),
α |α| L
1 1 ∗ α
µ∗L ( A ∩ E) = µL (A ∩ ),
α |α| E
1 1 1
µ∗L ( A ∩ E c ) = µ∗ ( A ∩ (αE c )).
α |α| L α
Substituting these into (11), we get the Caratheodory condition for αE, so
that αE ∈ ML . Then by Theorem 10, µL (αE) = |α|µL (E). Since αE ∈ ML
for every E ∈ ML , we have αML ⊂ ML . But ML = α α1 ML ⊂ α ML , and
we conclude.
19
5.3 Existence of non Lebesgue measurable sets
Not all sets in R are Lebesgue measurable.
20
Finally, let us show that P ∈
/ ML . Assume the contrary, then by Remark
◦
9, Pn = P + rn ∈ ML and µL (Pn ) = µL (P ). Thus
[ X X
1 = µL ([0, 1)) = µL Pn = µL (Pn ) = µL (P ). (13)
n∈Z+ n∈Z+ n∈Z+
6 Measurable functions
It is natural to develop a concept of measurable functions, that is, functions
whose preimage is measurable.
A measurable function pulls back measurable sets to measurable sets,
much like a continuous function pulls back open sets to open sets.
21
Proof. 1. (a) ⇔ (b). D1 and D2 partition D. Thus if D1 ∈ A then D2 =
D \ D1 ∈ A. Similarly, if D2 ∈ A then D1 ∈ A.
2. (c) ⇔ (d) as above.
3. (d) ⇒ (a). Note that
\ 1
D1 = {D : f < α + }. (14)
n
n∈N
22
3. By the definition of lim sup,
g3 (x) = lim sup fm (x) .
n→∞ m≥n
4. Similarly as above,
g4 (x) = lim inf fm (x) .
n→∞ m≥n
are measurable.
Tn
Proof.
Tn {minn=1,...,N fn > α} = n=1 {fi > α}, {maxn=1,...,N fn < α} =
{f
n=1 i < α}.
Proof. 1. Consider the sequence fn (x) = (f1 (x) = f (x), 0, 0, . . .). By Theo-
rem 17,
max fn (x) = max{f (x), 0} = f +
n=1,...,N
23
6.4 Measurability of the restriction and extension of f
Lemma 6. Let (X, A) be a measurable space.
(a) If f is an extended real-valued measurable function on a set D ∈ A,
then for every D0 ⊂ D such that D0 ∈ A, the restriction of f to D0 is a
measurable function on D0 .
(b) Let (Dn : n ∈ N) be a sequence in A and let D = ∪n∈N Dn . Let
f be an extended real-valued function on D. If the restriction of f to Dn is
A-measurable on Dn for every n ∈ N, then f is A-measurable on D.
Proof. 1. {D0 : f ≤ α} = {D : f ≤ α} ∩ D0 ∈ A.
2. {D : f ≤ α} = {∪n∈N Dn : f ≤ α} = ∪n∈N {Dn : f ≤ α} ∈ A.
Proof. Let us prove (b). Suppose f = g a.e. on D. Then there exists a null
set N such that N ⊂ D and f = g on D \ N . Since f is measurable on D,
it is measurable on the subset D \ N by (a) of Lemma 6. Since f = g on
D \ N , g is measurable on D \ N . But since N is a null set and the measure
space is complete, g is measurable on N by (a). Thus g is measurable on
D \ N and on N and therefore measurable on (D \ N ) ∪ N = D according
to (b) of Lemma 6.
24
real-valued measurable function on D. We say that (fn : n ∈ N) converges
almost uniformly on D to f if for every δ > 0 there exists a measurable subset
E of D such that µ(E) < δ and (fn : n ∈ N) converges uniformly on D \ E
to f .
25
(i ) If t ∈ [0, n), then
ϕn (t) = b2n tc/2n = 2b2n tc/2n+1 ≤ b2n+1 tc/2n+1 = ϕn+1 (t),
where we have used that nbtc ≤ bntc. (This can be proven by induction and
using the fact that if a < b then bac < bbc. Idea proof: bac < a < b, and
bbc < b. Thus bac and bbc are two integers not exceeding b, and so by the
definition of floor, bac < bbc.)
(ii ) If t ∈ [n, n + 1), then
ϕn (t) = n ≤ btc = 2n+1 btc/2n+1 ≤ b2n+1 tc/2n+1 ≤ ϕn+1 (t).
(iii ) If t ∈ [n + 1, ∞], then
ϕn (t) = n < n + 1 = ϕn+1 (t).
This shows that for all t ∈ [0, ∞] we have 0 ≤ ϕ1 (t) ≤ ϕ2 (t) ≤ . . . ≤ t.
We show next that φn (t) → t for t ∈ [0, ∞]. If t ∈ [0, ∞), then 2n t − 1 <
n n
< b22ntc , implying t − 21n < b22ntc . If n > t, then
n
b2n tc and therefore 2 2t−1
n
t − 21n < ϕn (t) ≤ t. This implies ϕn (t) → t for t ∈ [0, ∞). Lastly, if t = ∞
then ϕn (t) = ∞. Thus ϕn (t) → t for t ∈ [0, ∞].
Step 2. sn : X → [0, ∞) defined by sn = ϕ◦f is simple and measurable,
and 0 ≤ sn (x) = ϕn (f (x)) ≤ ϕn+1 (f (x)) = sn+1 (x). Moreover sn (x) =
ϕn (f (x)) ≤ f (x), since f (x) ∈ [0, ∞], proving the first part of the theorem.
Lastly, sn (x) = ϕn (f (x)) → f (x).
or, briefly,
Z n
X
ϕ dµ = ai µ(Di ).
D i=1
26
7.4 Integration of measurable functions
Definition 35 (LebesgueR integral). Let R f be measurable on a set D. If at
least one of the integrals D f + dµ and D f − dµ are finite, we define
Z Z Z
f dµ = +
f dµ − f − dµ.
D D D
Proof. Let us prove (c). Firstly, let f1 and f2 be nonnegative and measurable
functions on a set D with f1 ≤ f2 . We claim that
Z Z
f1 dµ ≤ f2 dµ. (20)
D D
27
Indeed, for i = 1 and 2, let Φi be the collection of all nonnegative simple
functions ϕ on D such that 0 ≤ ϕ ≤ fi . Since 0 ≤ ϕ ≤ f1 ≤ f2 on D, we
have Φ1 ⊂ Φ2 . Then
Z Z
{ ϕ dµ : ϕ ∈ Φ1 } ⊂ { ϕ dµ : ϕ ∈ Φ2 }.
D D
Let
Z
S1 = { ϕ dµ : ϕ ∈ Φ1 },
ZD
S2 = { ϕ dµ : ϕ ∈ Φ2 }.
D
sup S1 ≤ sup S2 ,
or, Z Z
f1 dµ ≤ f2 dµ.
D D
This proves (20). Note that the result now follows since
Z Z Z Z Z Z
+ − + −
f dµ = f dµ − f dµ ≤ g dµ − g dµ = g dµ.
D D D D D D
28
Suppose f is simple, and let f = m
P
i=1 ai 1Di be its canonical represen-
tation.
Since {Di ∩A, : i = 1, . . . , m} is a disjoint collection with m
S
i=1 (Di ∩A) =
A, the restriction of ϕ to A is given by
m
X
ϕ= ai 1Di ∩A .
i=1
Therefore
Z m
X m
X X
v(A) = f dµ = ai µ (Di ∩ A) = ai µ(Di ∩ An )
A i=1 i=1 n∈N
XX m XZ X
= [ ai µ(Di ∩ An )] = f dµ = v(An ),
n∈N i=1 n∈N An n∈N
29
Hence
Z Z Z Z
v(A1 ∪ A2 ) = f dµ ≥ ϕ dµ = ϕ dµ + ϕ dµ
A1 ∪A2 A1 ∪A2 A1 A2
≥ v(A1 ) + v(A2 ) − 2,
2. Let us write
Z Z Z
v(A) = f dµ = f + dµ − f − dµ,
A A A
30
Theorem 21. If f ∈ L(u) on E, then |f | ∈ L(u) on E and
Z Z
f dµ ≤ |f | dµ.
E E
Proof. Write E = A ∪ B, where f (x) ≥ 0 on A and f (x) < 0 on B. Then A
and B partitions E so by (b) of Theorem 20,
Z Z Z Z Z
|f | dµ = |f | dµ + |f | dµ = f + dµ + f − dµ.
E A B A B
But A ⊂ E and A ∈ A, so f ∈ L(u) on A by (f) of Remark 11, implying
that Z
f + dµ < ∞.
A
Similarly, Z
f − dµ < ∞.
B
This shows that |f | ∈ L(µ).
For the last part, note that since f ≤ |f | and −f ≤ |f |, for all x ∈ E,
Z Z Z Z Z
f dµ ≤ |f | dµ, − f dµ = (−f ) dµ ≤ |f | dµ,
E E E E E
where we have used (c) and (d) of Remark 11.
Theorem 22. Suppose f is measurable on E, |f | ≤ g, and g ∈ L(µ) on E.
Then f ∈ L(µ) on E.
Proof. Note that |f | ≤ g on E implies f + ≤ g and f − ≤ g on E. Then
Z Z
f + dµ < g dµ < ∞,
E E
and Z Z
f − dµ < g dµ < ∞,
E E
implying that f ∈ L(µ) on E.
31
In particular, if fn ∈ L(µ) is nondecreasing and fn → f pointwise, then
f ∈ L(µ). As we will se, this is not true in the Riemann sense.
lim fn (x)
n→∞
En = {x ∈ E : fn (x) ≥ cϕ(x)}, n ∈ N.
32
Let us define a set function v on A by setting
Z
v(A) = ϕ dµ,
A
where the last equality comes from (29). Then by Theorem 2, we have
33
0 00
f2 . By Theorem 19, this is possible. Put sn = sn + sn . Then by Theorem
20 Z Z Z
0 00
sn dµ = sn dµ + sn dµ,
E E E
and letting n → ∞ and applying Theorem 23 the result follows.
Next, suppose f1 ≥ 0, f2 ≤ 0. Put
or,
Z Z Z
f1 dµ = f dµ − f2 dµ. (33)
B B B
In the general case, E can be decomposed into four pairwise disjoint sets
Ei where f1 (x) and f2 (x) have constant sign, and we proceed as above.
then Z ∞ Z
X
f dµ = fn dµ.
E n=1 E
34
Proof. By Theorem 24, for any finite n ∈ N, we have
Z X n Xn Z
fj (x) dµ = fj (x) dµ.
E j=1 j=1 E
then Z Z
f dµ ≤ inf fn dµ.
E n→∞ E
Proof. By definition,
inf fn = lim inf fk .
n→∞ n→∞ k≥n
fn (x) → f (x), (x ∈ E)
then Z Z
lim fn dµ = f dµ.
n→∞ E E
35
Proof. Since fn + g ≥ 0, Fatou’s theorem shows that
Z Z
(f + g) dµ ≤ lim inf (fn + g) dµ,
E n→∞ E
or, Z Z
f dµ ≤ lim inf fn dµ,
E n→∞ E
Similarly g − fn ≥ 0, so by Fatou’s,
Z Z
(g − f ) dµ ≤ lim inf (g − fn ) dµ,
E n→∞ E
or, Z Z
− f dµ ≤ lim inf − fn dµ ,
E n→∞ E
or Z Z
f dµ ≥ lim sup fn dµ.
E n→∞ E
Thus Z Z Z Z
f dµ ≥ lim sup fn dµ ≥ lim inf fn dµ ≥ f dµ.
E n→∞ E n→∞ E E
and in particular
Z Z
lim fn dµ = f dµ. (35)
n→∞ D D
36
Since this holds for all n ≥ N , we have
Z
sup |fn − f | dµ ≤ µ(D) + 2M δ
n≥N D
which implies
Z
lim sup |fn − f | dµ ≤ µ(D) + 2M δ,
n→∞ m≥n D
or,
Z
lim sup |fn − f | dµ ≤ µ(D) + 2M δ.
n→∞ D
Then Z Z
0 ≤ lim inf |fn − f | dµ ≤ lim sup |fn − f | dµ = 0,
n→∞ D n→∞ D
showing (60).
Let us note that (61) can be derived from (60). Indeed,
Z Z Z Z
| fn dµ − f dµ| = | (fn − f ) dµ| ≤ |fn − f | dµ. (36)
D D D D
The equalities in (36) may need justification. For the first one, note
that since fn and f are measurable and bounded, and µ(D) < ∞, we have
fn , f ∈ L(µ) by Remark 11 (a). Then −fn ∈ L(µ) by (d) of the same
Remark. We then use Theorem 24. For the last equality, we use Theorem
21.
Applying (60) to the right most member of (36), we get
Z Z
lim | fn dµ − f dµ| = 0.
n→∞ D D
37
Proof. 1. Suppose f ≥ 0 on D. Let
so that X 1
0 < µ(D1 ) ≤ µ{D : f ≥ }.
k
k∈N
38
R
This shows that f ≥ 0 on D \ E and D\E f dµ = 0. But then f = 0 a.e.
on D \ E by our result above. Thus there exists a null set F such that
F ⊂ D \ E and such that f = 0 on (D \ E) \ F = D \ (E ∪ F ). Thus f = 0
a.e. on D. R R
2.R If f ≤ g a.e. on D, then g−f ≥ 0 a.e. on D. If also D f dµ = D g dµ,
then D (g − f ) dµ = 0. Then by (a), g − f = 0 a.e. on D.
Let
S(f ) = sup S(f, P) and S(f ) = inf S(f, P).
P∈B P∈B
We call S(f ) and S(f ) the lower and upper Darboux integrals of f on I.
Definition 37. A bounded real-valued function f on I = [a, b] ⊂ R is Rie-
mann integrable on I if and only if
S(f ) = S(f ).
In this case we set Z b
f (x) dx := S(f ) = S(f ).
a
Theorem 28. If f is a continuous real-valued function on I = [a, b], then
S(f ) = S(f ) and consequently f is Riemann integrable on I.
Proof. If f is continuous on the compact set I, it is uniformly continuous
on I. Thus for every > 0 there exists δ > 0 such that for all x, x00 ∈ [a, b],
|x − x00 | ≤ δ ⇒ |f (x0 ) − f (x00 )| ≤ . (38)
b−a
Let P0 ∈ B be such that |P0 | < δ. Let P0 = {x0 , . . . , xn } where a =
x0 < . . . < xn = b. In preceding, let us note that (38) implies
sup f − inf f ≤ . (39)
[xk−1 ,xk ] [xk−1 ,xk ] b−a
39
Indeed, let α, β ∈ [xk−1 , xk ]. By (38),
f (α) − f (β) ≤ ,
b−a
or,
f (α) ≤ + f (β).
b−a
Then
f
[xk , xk−1 ] ≤ + f (β). (40)
b−a
Thus the right hand side of (40) is an upper bound to [xk , xk−1 ] f . But
sup[xk ,xk−1 ] f is the smallest upper bound to [xk , xk−1 ] f , so
sup f≤ + f (β). (41)
[xk ,xk−1 ] b−a
Thus the left hand side of (42) is a lower bound to [xk , xk−1 ] f , and since
inf[xk , xk−1 ] f is the largest such lower bound,
sup f− ≤ inf f,
[xk ,xk−1 ] b − a [xk ,xk−1 ]
40
8.2 The upper and lower envelope integrals
Definition 38 (Limit inferior, limit superior of f ). For x0 ∈ R and δ > 0,
let
U (x0 , δ) = (x0 − δ, x0 + δ),
and
U0 (x0 , δ) = (x0 − δ, x0 + δ) \ {x0 }.
Let f be an extended real-valued function on a set D ∈ R and let x0 ∈ D.
The limit inferior and limit superior of f as x approaches x0 are defined by
lim inf f (x) = lim inf f and lim sup f (x) = lim sup f.
x→x0 δ→0 U0 (x0 ,δ)∩D x→x0 δ→0 U0 (x0 ,δ)∩D
We have
f (x0 ) ∈ R,
f is continuous at x0 ⇔
lim inf f (x) = lim sup f (x) = f (x0 ).
x→x0 x→x0
(
f (x0 ) ∈ R,
f is continuous at x0 ⇔ (48)
f∗ (x0 ) = f ∗ (x0 ).
Proof. Let us prove (48), since (46) and (47) follows directly from the defi-
nition.
Suppose f is continuous at x0 ∈ D. Then for every > 0 there exists
δ > 0 such that for all x ∈ U (x0 , δ) ∩ D,
41
Now since inf U (x0 ,δ)∩D f (x) ↑ as δ ↓ 0 (as the infimum is taken over a
smaller set), we have
If we then combine (49) and (50), we get f ∗ (x0 ) − f∗ (x0 ) ≤ 2. Since
> 0 was chosen arbitrarily, we have f∗ (x0 ) = f ∗ (x0 ).
Conversely, if f (x0 ) ∈ R and f∗ (x0 ) = f ∗ (x0 ), then
1
If necessary, let us add partition points to Pm so that |Pm | ≤ m . Note that
this not affect the validity of (51) and (52), since refining a partition does
not increase the upper Darboux sum.
Let Pm = {xm1 , xm2 , . . . , xmn } and let Imk = [xmk−1 , xmk ] for k =
1, . . . , n. For each m ∈ N, define a function ψm on I by
n
X
ψm = sup f · 1Imk . (53)
k=1 Imk
42
Let E be the countable set consisting of all partition points in (Pm : m ∈
N), that is,
(Pm : m ∈ N) = (x11 , x22 , . . .).
Let us show that
where each inequality comes from the fact that we take the sup over succes-
sively greater sets, and the last equality from how we defined ψm .
Since supU (x0 ,δ)∩I f ”shrinks” from above to f ∗ (x0 ) as δ ↓ 0, we have
that for an arbitrary > 0 there exists δ > 0 such that
Let N = max(M, M 0 ), then by (55), (57), and (56), we have that for all
m≥N
f ∗ (x0 ) ≤ ψm (x0 ) = sup f ≤ sup f < f ∗ (x0 ) + ,
Im k U (x0 ,δ)∩I
proving (54).
Consider the sequence of simple functions (ψm : m ∈ N) defined by (53).
Then Z X n
ψm dµ= sup f · l(Imk ) = S(f, Pm );
I k=1 Imk
43
the upper Darboux sum of f corresponding to Pm . Then by (52),
Z
lim ψm dµ = S(f ). (58)
n→∞ I
44
8.4 Lebesgue’s integrability condition
Theorem 30. Let f be bounded real-valued function on I = [a, b] and let E
be the set of all points of discontinuity of f in I. Then
f is Riemann integrable on I
⇔
∗
f∗ = f , a.e. on I
⇔
µL (E) = 0.
Proof. Since the fn ’s are continuous and converge uniformly to f , the limit
function must be continuous. By Theorem 28, f is Riemann-integrable and
Z b Z b Z b Z b
fn (x) dx − f (x) dx ≤ fn (x) dx − f (x) dx
a a a a
Z b
≤ |fn (x) − f (x)| dx
a
≤ (b − a) sup |fn (x) − f (x)| → 0.
45
Note that Theorem 31 does not say ¬(uniform) ⇒ ¬(interchange). How-
ever in practice, to be sure that interchanging is justified it is uniform con-
vergence that we check.
Even when the sequence is uniformly bounded, pointwise convergence is
not enough. Recall the Bounded convergence theorem:
Theorem 32 (Bounded convergence theorem). Let (fn : n ∈ N) be a
uniformly bounded sequence of real-valued measurable functions a on a set
D ∈ A with µ(D) < ∞. Let f a bounded real-valued measurable function on
D. If (fn : n ∈ N) converges to f a.e. on D, then
Z
lim |fn − f | dµ = 0, (60)
n→∞ D
and in particular
Z Z
lim fn dµ = f dµ. (61)
n→∞ D D
Between any two points xj and xj+1 there exists a rational number.
Therefore the sup over [xj , xj+1 ] must be 1. This means that
n
X
S(f, P) = 1 · (xk − xk−1 )
k=1
= (x1 − x0 ) + (x2 − x1 ) + . . . + (xn − xn−1 )
= xn − x0 = 1 − 0 = 1,
so that
S(f ) = inf S(f, P) = 1.
P∈B
Then by Definition 37, f is not Riemann integrable.
46
Proposition 5. There exists a uniformly bounded sequence (fn : n ∈ N)
of real-valued functions defined on the set [0, 1] that converges to a bounded
function f , but where
Z 1 Z 1
lim fn dx 6= lim fn dx. (62)
n→∞ 0 0 n→∞
Proof. Let f be the Dirichlet function on [0, 1]. Then f is clearly bounded.
Let {rn }∞
n=1 be any enumeration of the rationals in [0, 1], and define
(
1, if x = rk , for some 1 ≤ k ≤ n,
fn (x) =
0, otherwise.
Remark 13. Note that in the Riemann sense, we could not so easily be
sure of the equality
Z 1 Z 1
1 1
lim dµ = lim dµ.
n→∞ 0 nx + 1 0 n→∞ nx + 1
One way of being sure, is to know that (fn (x)) converges uniformly to 0
and then use [0, 1] Theorem 31. The problem is, it doesn’t:
1
sup | | = 1 6= 0,
nx + 1
(To see this, note that |fn (x)| ≤ 1, so 1 is an upper bound to |fn (x)|. Now
since fn (0) = 1, it is indeed the smallest upper bound.)
47
x n
Proposition 6. Let fn (x) = 1 + n , x ∈ [0, ∞).
Then (fn ) is non-decreasing.
Example 3. Compute
Z ∞
x n −2x
lim 1+ e dµ.
n→∞ 0 n
n
Solution. Let fn (x) = 1 + nx e−2x , then each fn is measurable by Theo-
rem 13, and
0 ≤ f1 (x) ≤ f2 (x) ≤ . . . , x ∈ [0, ∞),
n
since 1 + nx → ex , monotonically, from below, by Proposition 6.
References
[1] Yeh, J. Real Analysis, Theory of Measure and Integration (2nd. ed.),
World Scientific Publishing Co. Pte. Ltd. 2006
[2] Burkill, J.C. The Lebesgue Integral, Cambridge University Press 1971
48