measure and integration contents · 2019-01-18 · theory is almost not measurable. the lecture...

52
MEASURE AND INTEGRATION Lecture notes by Mark Veraar Contents Introduction 2 1. σ-algebras 3 2. Measures 6 3. Construction of measures 9 4. Measurable functions 13 5. Construction of the integral 18 5.1. Integral for simple functions 18 5.2. Integral for positive measurable functions 19 5.3. Integral for measurable functions 21 6. Convergence theorems and applications 25 6.1. The three main convergence results 25 6.2. Consequences and applications 26 7. L p -spaces 30 7.1. Minkowski and H¨ older’s inequalities 30 7.2. Completeness of L p 31 7.3. L p -spaces on intervals 32 8. Applications to Fourier series 36 8.1. Fourier coefficients 36 8.2. Weierstrass’ approximation result and uniqueness 37 8.3. Fourier series in L 2 (0, 2π) 39 8.4. Fourier series in C([0, 2π]) 41 Appendix A. Dynkin’s lemma 46 Appendix B. Carath´ eodory’s extension theorem 48 Appendix C. Non-measurable sets 51 References 52 Date : January 18, 2019. 1

Upload: others

Post on 29-May-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION

Lecture notes by Mark Veraar

Contents

Introduction 21. σ-algebras 32. Measures 63. Construction of measures 94. Measurable functions 135. Construction of the integral 185.1. Integral for simple functions 185.2. Integral for positive measurable functions 195.3. Integral for measurable functions 216. Convergence theorems and applications 256.1. The three main convergence results 256.2. Consequences and applications 267. Lp-spaces 307.1. Minkowski and Holder’s inequalities 307.2. Completeness of Lp 317.3. Lp-spaces on intervals 328. Applications to Fourier series 368.1. Fourier coefficients 368.2. Weierstrass’ approximation result and uniqueness 378.3. Fourier series in L2(0, 2π) 398.4. Fourier series in C([0, 2π]) 41Appendix A. Dynkin’s lemma 46Appendix B. Caratheodory’s extension theorem 48Appendix C. Non-measurable sets 51References 52

Date: January 18, 2019.

1

Page 2: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

2 MEASURE AND INTEGRATION

Introduction

These notes have been created for the “Measure and integration theory” part of a course onreal analysis at the TU Delft. Together with the first part of the course on metric spaces, thesenotes form the mathematical basis for several bachelor courses and master courses in appliedmathematics at TU Delft.

I would like to thank Emiel Lorist and Bas Nieraeth for their support in the preparation ofthese notes. All the figures have been created by Emiel Lorist. I would also like to thank thestudents of the course on real analysis for pointing out the typos in the manuscript.

In Section 1 and 2 we introduce σ-algebras and measures. The Lebesgue measure is constructedin Section 3 and is based on Appendix B on Caratheodory’s theorem. Uniqueness questions areaddressed in Appendix A on Dynkin’s monotone class theorem. The amount of books on measuretheory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A verycomplete treatment of measure theory is given in the impressive works [5].

In Sections 5, 6, 7 we introduce the integration theory and the Lebesgue spaces Lp. This theoryis fundamental in modern (applied) mathematics. There are many excellent books which givemore detailed treatments on the subject. See for instance [1], [4], [6], [15] for detailed treatments.

In Section 8 we give a brief introduction to the theory of Fourier series. More thorough treat-ments can be found in for example [9], [10], [13], [14] and [19]. A full course on Fourier Analysisis offered as 3rd elective course based on the lecture notes [10]. The theory of Fourier series willbe used in the 2nd year bachelor course on Partial Differential Equations [7], but also in severalother parts of Mathematical Physics and Numerical Analysis.

We end this brief introduction with a quote from a historical note of Zygmund [18]:

“The Lebesgue integral did not arise via the theory of Fourier series but was created through thenecessities of measuring geometric figures. But once it was introduced, it had an enormous impacton Analysis through Fourier series”:

• The Riesz-Fischer theorems 7.5, 8.7, in its initial formulation primarily a theorem onFourier series.

• The M. Riesz interpolation theorem.• Structure of sets of measure 0.• The theory of trigonometric series has become a workshop of new methods in analysis, a

place where new methods are first discovered before they are generalized and applied inother contexts.

• The theory of Fourier series gave a fresh impulse to problems of the differentiability offunctions (Sobolev spaces etc.).

Page 3: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 3

1. σ-algebras

For a set S we write P(S) for its power set.

Definition 1.1. Let S be a set. A collection R ⊆ P(S) is called a ring if

(i) ∅ ∈ R;(ii) A,B ∈ R =⇒ B \A ∈ R;

(iii) n ∈ N, A1, A2, . . . , An ∈ R =⇒n⋃j=1

Aj ∈ R.

Remark 1.2.

(1) An equivalent definition is obtained if one replaces (iii) by A,B ∈ R =⇒ A ∪ B ∈ R. Thisfollows by induction.

(2) If R is a ring, then for all A,B ∈ R one has A∩B ∈ R. Indeed, this follows from the identityA ∩B = A \ (A \B).

Definition 1.3. Let S be a set. A family A ⊆ P(S) is called a σ-algebra1 if

(i) ∅, S ∈ A;(ii) A ∈ A =⇒ Ac ∈ A;

(iii) A1, A2, . . . ,∈ A =⇒∞⋃j=1

Aj ∈ A.

The sets A ∈ A are often called the measurable subsets of S.

Remark 1.4.

(1) If A is a σ-algebra, then for all A1, A2, . . . ,∈ A one has∞⋂j=1

Aj ∈ A. This follows from the

identity∞⋂j=1

Aj =( ∞⋃j=1

Acj

)c.

(2) Every σ-algebra is a ring. This follows from the identity B \A = B ∩Ac.Example 1.5. Let S be a set.

(a) A = {S,∅} is the smallest possible σ-algebra on S.(b) A = P(S) is the largest possible σ-algebra on S.

Example 1.6. Let S = {1, 2, 3}.(a) Let A = {∅, S, {1}, {2, 3}}. Then A is a σ-algebra.(b) Let B = {∅, S, {1}, {2, 3}, {1, 2}}. Then B is not a σ-algebra.

Example 1.7. Let S be a set.

(a) The set R = {A ⊆ S : A is finite} is a ring.(b) Let2 A = {A ⊆ S : A is countable or Ac is countable}. Then A is a σ-algebra (see Exercise

1.2). It is called the countable-cocountable σ-algebra and is useful for counterexamples.

We continue with a more serious example which plays a crucial role in later constructions.

Example 1.8.

(a) Let S = R. Let I 1 be the collection of all sets of the form (a, b] with a ≤ b. These intervalswill be called half-open intervals. Then I 1 is not a ring since for instance (0, 3] \ (1, 2] =(0, 1] ∪ (2, 3] is not in I 1.

(b) Let S = R. Let F1 be the collection of sets which can be written as a finite union of half-openintervals (thus of the form (a, b] with a ≤ b). Then F1 is not a σ-algebra for several reasons.3

We check that F1 is a ring. (i) follows from ∅ = (1, 1] ∈ F1. (iii) is clear since a finite unionof a finite union of intervals of the form (a, b] is again a finite union. It remains to check (ii).

1In part of the literature a σ-algebra is also called a σ-field2Recall that a set A ⊆ S is countable if A is finite or there is a bijection f : N→ A.

3For instance∞⋃n=1

(0, 1− 1n

] = (0, 1) is not in F1.

Page 4: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

4 MEASURE AND INTEGRATION

For this we first note that it is simple to check that for A,B ∈ F1 one has A∩B ∈ F1 and byinduction this extends to the intersections of finitely many sets. For two intervals (a, b] and(c, d] using B \A = B ∩Ac and R \ (a, b] = (−∞, a] ∪ (b,∞) we find

(c, d] \ (a, b] = (c, d] ∩ (R \ (a, b])

=((c, d] ∩ (−∞, a]

)∪((c, d] ∩ (b,∞)

)= (c, a ∧ d] ∪ (b ∨ c, d].

This is in F1 again. Now if A =m⋃i=1

(ai, bi] and B =n⋃j=1

(cj , dj ] are in F1, then

B \A =

n⋃j=1

(cj , dj ] \A =

n⋃j=1

m⋂i=1

(cj , dj ] \ (ai, bi].

and by the previous observations this is in F1 again.(c) Let S = Rd. For a, b ∈ Rd with a = (α1, . . . , αd) and b = (β1, . . . , βd) with αj ≤ βj for

j ∈ {1, . . . , k} the half-open rectangles are given by

(a, b] = (α1, β1]× . . .× (αd, βd].

Let Fd be the collection of sets which can be written as a finite unions of half-open rectangles.Then Fd is a ring (see Exercise 1.7).

The proof of the next result is Exercise 1.3.

Proposition 1.9 (Intersection of σ-algebras). Suppose that Ai is a σ-algebra on S for every i ∈ I.Then

⋂i∈IAi is a σ-algebra.

Definition 1.10. Let S be a set and let F ⊆ P(S). We write σ(F) for the smallest σ-algebrawhich contains F . Then σ(F) is called the σ-algebra generated by F . More precisely:4

σ(F) =⋂{A : A is a σ-algebra on S and F ⊆ A}.

Example 1.11. Let S = {1, 2, 3}, F = {{1, 2}} and G = {{2, 3}, {1, 2}}.(a) σ(F) = {∅, S, {1, 2}, {3}}.(b) σ(G) = P(S). Indeed, {2, 3}c = {1}, {1, 2}c = {3} and {2, 3} ∩ {1, 2} = {2}. Thus the

singletons {1}, {2} and {3} are in σ(G). Therefore, the required result follows since we canform every subset of S by taking suitable finite unions.

Example 1.12. Let S = N and F = {{n} : n ∈ N}. Then σ(F) = P(N).

Definition 1.13. Let (S, d) be a metric space.5 Let B(S) be the σ-algebra generated by the opensets in S. Thus

B(S) = σ{open sets in S}.The σ-algebra B(S) is called6 the Borel σ-algebra of S. The sets of B(S) are called the Borelsubsets of S.

Example 1.14. One of the most important σ-algebras is the Borel σ-algebra of R which is usuallydenoted by B(R). Later on we will show that B(R) 6= P(R).

The following lemma will be useful in some of the exercises on the Borel σ-algebra of R and Rd.

Lemma 1.15 (Lindelof). Let A ⊆ Rd. Assume that for each i ∈ I let Oi ⊆ Rd be open. IfA ⊆ ⋃

i∈IOi, then there exists a countable J ⊆ I such that A ⊆ ⋃

i∈JOi

4Note that Proposition 1.9 ensures that σ(F) is indeed a σ-algebra. The intersection makes sure we obtain thesmallest possible one

5More generally one could take any topological space here6Named after the French mathematician Felix Borel 1871-1956

Page 5: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 5

Proof. Choose for each x ∈ A, ix ∈ I and rx > 0 such that B(x, rx) ⊆ Oix . For each x ∈ A chooseax ∈ Qd and sx ∈ Q ∩ (0,∞) such that x ∈ B(ax, sx) ⊆ B(x, rx) ⊆ Oix . Let F = {B(ax, sx) : x ∈A}. Then clearly A ⊆ ⋃

x∈AB(ax, sx). Moreover, F has at most countably many sets. Indeed, this

follows from the fact that it is a subset of {B(q, r) : q ∈ Qd, r ∈ Q ∩ (0,∞)} which is countable.Therefore, we can write F = {B(axn , sxn) : n ∈ N} with xn ∈ A for each n ∈ N.

Now let J = {ixn ∈ I : n ∈ N}. Then A ⊆ ⋃i∈J

Oi. Indeed, if x ∈ A, then x ∈ B(ax, sx) and

choosing n ∈ N such that axn = ax and sxn = sx we find that x ∈ Oixn ⊆⋃i∈J

Oi ,

Exercises

Exercise 1.1. Let S = R and F = {A ⊆ R : A ⊆ [0, 1] or Ac ⊆ [0, 1]}. Is F a ring?

Exercise∗ 1.2. Prove that the collection in Example 1.7 (b) is a σ-algebra.

Exercise 1.3.

(a) Prove Proposition 1.9.(b) Give an example of two σ-algebras A and B on S = {1, 2, 3} such that A∪B is not a σ-algebra.

Exercise∗ 1.4. Let S be a set and let F = {{s} : s ∈ S} be the collection consisting of allsets which contain one element of S. Show that σ(F) coincides with the countable-cocountableσ-algebra of Example 1.7.Hint: Use the followsing (well-known) facts: The subset of a countable set is again countable;The countable union of countable sets is again countable.

Exercise 1.5. Show that N,Q,R \Q ∈ B(R). That is N, Q and R \Q are Borel subsets of R.

Exercise∗ 1.6. Consider the following collection B0 = {(−∞, x) : x ∈ R}) of subsets of R.

(a) Show that σ(B0) contains all open intervals.(b) Show that every open set in R can be written as the union of countably many open intervals.

Hint: Use Lindelof’s Lemma 1.15.(c) Conclude that σ(B0) = B(R).

Exercise∗ 1.7. Let I d ⊆ Fd be the collection of half-open rectangles of Example 1.8. Prove thefollowing assertions:

(a) If I, J ∈ I d then I ∩ J ∈ I d.(b) If I, J ∈ I d, then I \J is the union of finitely many disjoint sets from I d, and thus I \J ∈ Fd.

Hint: Use induction on the dimension d. Use Example 1.8 (b) for d = 1.(c) Each A ∈ Fd can be written as union of finitely many disjoint sets in I d.

Hint: Use induction on n to prove this for all sets of the form A =n⋃k=1

Ik with I1, . . . , In ∈ I d.

(d) Fd is a ring.

Exercise∗∗ 1.8. Prove that a σ-algebra is either finite or uncountable.7

Hint: Recall that P(N) is uncountable.

7This shows that σ-algebras are either easy finite sets or quite complicated

Page 6: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

6 MEASURE AND INTEGRATION

2. Measures

Definition 2.1. Let S be a set and R ⊆ P(S) a ring. Let µ : R → [0,∞] be a function with8

µ(∅) = 0

(i) µ is called additive if for all disjoint9 A1, . . . , An ∈ R one has

µ( n⋃j=1

Aj

)=

n∑j=1

µ(Aj).

(ii) µ is called σ-additive if for each disjoint sequence (An)n≥1 in R which satisfies∞⋃n=1

An ∈ Rit holds that

µ( ∞⋃j=1

Aj

)=

∞∑j=1

µ(Aj).

Remark 2.2.

(1) By an induction argument it suffices to consider n = 2 in the definition of additive. (SeeExercise 2.1).

(2) If µ is σ-additive, then it is additive as follows by taking Am = ∅ for m ≥ n+ 1.

Definition 2.3. Let S be a set and let A be a σ-algebra on S.

(i) The pair (S,A) is called a measurable space.(ii) A function µ : A → [0,∞] which satisfies µ(∅) = 0 and which is σ-additive on A is called a

measure. In this case the triple (S,A, µ) is called a measure space.(iii) If additionally to (ii) µ(S) < ∞, then µ is called a finite measure. If moreover, µ(S) = 1,

then µ is called a probability measure and (S,A, µ) is called a probability space. 10

Example 2.4 (Counting measure). Let S = N and A = P(N). We write #A for the number ofelements of a finite set A, and we set #A = ∞ if A is infinite. Let µ : A → [0,∞] be given byµ(A) = #A. Then µ is a measure. Often µ is denoted by τ and called the counting measure.

Example 2.5 (Dirac measure/Dirac’s delta function). Let S = R, A = P(R). Let x ∈ R. Letδx : A → [0,∞] be given by δx(A) = 1 if x ∈ A and δx(A) = 0 if x ∈ R \A. Then δx is a measure.It is usually called the Dirac measure11 at x.

Example 2.6. Let S be a set and let µ : P(S) → [0,∞] be given by µ(∅) = 0 and µ(A) = 1 ifA 6= ∅. If S contains at least two elements, then µ is not a measure.

Example 2.7 (Length of an interval). 12 Let R = F1 as in Example 1.8. For a ≤ b let λ((a, b]) =b − a (length of the interval). If A = (a1, b1] ∪ . . . ∪ (an, bn] is a union of disjoint sets such thataj ≤ bj for j = 1, . . . , n, then we let13 λ(A) =

∑nj=1 bj − aj . Then λ is additive. Later we will see

that λ is σ-additive on F1 and has an extension to a measure on σ(F1) = B(R).

Theorem 2.8. Let R be a ring and µ : R → [0,∞] be additive. The following assertions hold:

(i) If A,B ∈ R and A ⊆ B, then µ(A) ≤ µ(B) (monotonicity).

(ii) If A1, A2, . . . ∈ R are disjoint and∞⋃j=1

Aj ∈ R, then

µ( ∞⋃j=1

Aj

)≥∞∑j=1

µ(Aj).

8This assumption will always be made.9Here we mean Ai ∩Aj = ∅ if i 6= j10Measure theory is at the very heart of probability theory. See the third year elective course11Named after the English theoretical physicist Paul Dirac 1902-1984. The “delta function” is actually not a

function, but can be interpreted as a generalized function or as measure.12See Section 3 for a further construction of the so-called Lebesgue measure13One can check that this does not depend on the way we write the set A. See below Definition 3.6.

Page 7: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 7

(iii) If A1, A2, . . . ∈ R and∞⋃j=1

Aj ∈ R and µ is σ-additive on R, then

µ( ∞⋃j=1

Aj

)≤∞∑j=1

µ(Aj) (σ-subadditivity).

Proof. (i): Write B = A ∪ (B \A). Then

µ(B) = µ(A ∪ (B \A)) = µ(A) + µ(B \A) ≥ µ(A).

(ii): Let n ∈ N. Thenn⋃j=1

Aj ⊆∞⋃j=1

Aj and therefore by additivity of µ

n∑j=1

µ(Aj) = µ( n⋃j=1

Aj

) (i)

≤ µ( ∞⋃j=1

Aj

).

The result follows by letting n→∞.(iii): Let

(2.1) B1 = A1, B2 = A2 \A1, B3 = A3 \ (A1 ∪A2), etc.

Then (Bn)n≥1 is a disjoint sequence and∞⋃j=1

Bj =∞⋃j=1

Aj . Therefore, by the σ-additivity of µ

µ( ∞⋃j=1

Aj

)= µ

( ∞⋃j=1

Bj

)=

∞∑j=1

µ(Bj)(i)

≤∞∑j=1

µ(Aj).

,

Let (an)n≥1 be a sequence of real numbers. We write an ↑ a if (an)n≥1 is an increasing sequencewhich converges to a. Similarly, we write an ↓ a if it decreases and converges to a. This notationwill now be extended to sets.

Definition 2.9. Let S be a set.

(i) A sequence (An)n≥1 of subsets of S will be called increasing if An ⊆ An+1 for all n ∈ N.

In this case we write An ↑ A, where A =∞⋃n=1

An.

(ii) A sequence (An)n≥1 of subsets of S will be called decreasing if An ⊇ An+1 for all n ∈ N.

In this case we write An ↓ A, where A =∞⋂n=1

An.

Theorem 2.10. Let (S,A, µ) be a measure space and let (An)n≥1 be a sequence in A.

(i) If An ↑ A, then µ(An) ↑ µ(A).(ii) If An ↓ A and µ(A1) <∞, then µ(An) ↓ µ(A).

Proof. (i): Define (Bn)n≥1 as in (2.1). Then (Bn)n≥1 is a disjoint sequence and the following

identities hold:∞⋃j=1

Bj = A andn⋃j=1

Bj = An. Therefore, the σ-additivity of µ gives

µ(A) =

∞∑j=1

µ(Bj) = limn→∞

n∑j=1

µ(Bj) = limn→∞

µ(An).

(ii): See Exercise 2.4. ,

The following result will help us to check σ-additivity on rings. It will be used in the constructionof the Lebesgue measure in Lemma 3.9.

Lemma 2.11 (Sufficient condition for σ-additivity). Let R be a ring and let µ : R → [0,∞] besuch that µ(∅) = 0 and µ is additive. Suppose that for each sequence (An)n≥1 with An ↓ ∅ onehas µ(An)→ 0. Then µ is σ-additive on R.

Page 8: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

8 MEASURE AND INTEGRATION

Proof. Let (Bj)j≥1 be a disjoint sequence in R with B :=∞⋃j=1

Bj ∈ R. We need to show that

(2.2) µ(B) =

∞∑j=1

µ(Bj).

Let An =∞⋃j=n

Bj = B \ (B1 ∪ . . . ∪Bn−1). Then An ∈ R and An ↓ ∅. Now the assumption yields

µ(An)→ 0. On the other hand

µ(B) = µ(An ∪B1 ∪B2 ∪ . . . ∪Bn−1) = µ(An) +

n−1∑j=1

µ(Bj).

Therefore,∣∣∣µ(B)−

n−1∑j=1

µ(Bj)∣∣∣ = µ(An)→ 0 and (2.2) follows. ,

Exercises

Exercise 2.1. Let R be a ring on a set S. Assume µ : R → [0,∞] satisfies µ(∅) = 0 andµ(A ∪B) = µ(A) + µ(B) for all sets A,B ∈ R with A ∩B = ∅. Show that µ is additive.

Exercise 2.2. Let R be a ring on a set S. Let µ : R → [0,∞] be additive.

(a) Prove that for A,B ∈ R with µ(A) <∞ and A ⊆ B one has

µ(B \A) = µ(B)− µ(A).

(b) Prove that for A,B ∈ R with µ(A) <∞ one has

µ(A ∪B) = µ(A) + µ(B)− µ(A ∩B).

(c) Show that for any n ∈ N and any sets (Aj)nj=1 in R,

finite subadditivity µ(A1 ∪ . . . ∪An) ≤ µ(A1) + . . .+ µ(An).

Exercise 2.3. Let (an)n≥1 be numbers in [0,∞). Set µ(∅) = 0 and define µ : P(N) → [0,∞] by

µ(A) =∑n∈A

an. Prove that µ is a measure on N.

Exercise∗ 2.4.

(a) Prove Theorem 2.10 (ii).(b) Give an example of a measure space (S,A, µ) and sets An ∈ A such that An ↓ ∅ and

µ(An) =∞ for all n ∈ N.Hint: Use the counting measure.

Exercise∗ 2.5. Let (S,A, µ) be a measure space. For A1, A2, . . . ⊆ S define

lim supn→∞

An =

∞⋂k=1

∞⋃n=k

An.

(a) Show that s ∈ lim supn→∞

An if and only if there are infinitely many n ∈ N such that s ∈ An.

(b) Assume A1, A2, . . . ∈ A. Show that lim supn→∞

An ∈ A.

(c) Assume A1, A2, . . . ∈ A satisfy

∞∑n=1

µ(An) <∞. Show that µ(lim supn→∞

An) = 0.14

Exercise∗ 2.6. Let A be the σ-algebra from Example 1.7 (b) with S = R. Define µ : A → [0,∞]by µ(A) = 0 if A is countable and µ(A) = 1 if Ac is countable. Show that µ is a measure.

14This is called the Borel-Cantelli lemma in probability theory.

Page 9: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 9

3. Construction of measures

It is not a simple task to construct a measure. In this section we will construct the Lebesguemeasure on Rd of which we have previously shown it is an additive mapping on the ring F1 inExample 2.7. To extend it to the Borel σ-algebra we use a deep result of Caratheodory.15 Hisresult basically says that it is enough to check that a measure is σ-additive on a ring generatingthe desired σ-algebra. A detailed proof can be found in Theorem B.6 in the appendix, but it willdo no harm if one takes the result for granted.16

Theorem 3.1 (Caratheodory’s extension theorem). Let S be a set and let R ⊆ P(S) be a ring.Suppose that µ : R → [0,∞] is σ-additive on R and µ(∅) = 0. Then µ extends to a measure µ onσ(R). More precisely, there exists a measure µ on (S, σ(R)) such that µ(A) = µ(A) for all A ∈ R.

Remark 3.2. The measure µ is often unique.17 When there is no danger of confusion we will writeµ again for the extension to σ(R). However, in general one has to be careful about uniqueness.For instance if we define µ on the ring F1 by µ(A) = ∞ if A ∈ F1 is nonempty, then µ has atleast two extensions: the counting measure on B(R) is an extension of µ, but also the measureν : B(R)→ [0,∞] given by ν(A) =∞ if A 6= ∅ is an extension of µ.

We continue with a uniqueness result which will be proved in the appendix.

Definition 3.3 (π-system). A collection E ⊆ P(S) is called a π-system if for all A,B ∈ E onehas A ∩B ∈ E.

Example 3.4.

(a) Every ring is a π-system.(b) Let S = Rd. The half-open rectangles Id are a π-system.

The following result will be proved in Proposition A.7.

Proposition 3.5 (Uniqueness). Let µ1 and µ2 both be measures on measurable space (S,A).Assume the following conditions:

(i) E ⊆ A is a π-system with σ(E) = A;(ii) µ1(S) = µ2(S) <∞ and µ1(E) = µ2(E) for all E ∈ E.

Then µ1 = µ2 on A.

We continue with the construction of the Lebesgue 18 measure λ. In Example 1.8 we intro-duced the half-open intervals (a, b] ∈ I1 with a ≤ b. Also recall that F1 denotes the collectionof all finite unions of half-open intervals. We have seen that F1 is a ring. Moreover, in of courseevery set in F1 can be written as a finite union of disjoint half-open intervals.

Definition 3.6 (on unions of half-open intervals). For A ∈ F1 of the form A = (a1, b1] ∪ . . . ∪(an, bn] with disjoint ((aj , bj ])

nj=1 in I1 define λ : F1 → [0,∞] as the sum of the lengths:

λ(A) =

n∑j=1

(bj − aj).

The above is well-defined. To see this assume A = (c1, d1]∪. . . (cm, dm] is another representationof A as a union of disjoint intervals. Let Iij = (ci, di] ∩ (aj , bj ]. Then either Iij is empty or a halfopen interval,

m⋃i=1

Iij = (aj , bj ] and

n⋃j=1

Iij = (cj , dj ].

15Caratheodory 1873-1950 was a Greek mathematician working in Analysis, but also on Thermodynamics.16The appendix is not part of the exam17For instance when µ is a finite measure. See Proposition A.7.18Henri Lebesgue 1875–1941 was a French mathematician well-known for his integration theory. See Section 5.

Page 10: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

10 MEASURE AND INTEGRATION

From the definition and the disjointness of the (Iij)m,ni,j=1 we obtain

n∑j=1

λ((aj , bj ]

)=

n∑j=1

λ( m⋃i=1

Iij

)=

n∑j=1

m∑i=1

λ(Iij) =

m∑i=1

n∑j=1

λ(Iij) =

m∑i=1

λ( n⋃j=1

Iij

)=

m∑i=1

λ((ci, di]

),

which proves the well-definedness. Alternatively, one can observe that λ(A) coincides with theRiemann integral of 19 1A. Indeed, fix an interval I such that A ⊆ I. By linearity of the Riemannintegral ∫

I

1A dx =

n∑j=1

∫I

1(aj ,bj ] dx =

n∑j=1

(bj − aj) = λ(A).

In Example 1.8 we introduced the half-open rectangles (a, b] ∈ Id with a = (α1, . . . , αd) andb = (β1, . . . , βd) and αj ≤ βj for j ∈ {1, . . . , d}. Also recall that Fd denotes the collection of allfinite unions of half-open rectangles. By Exercise 1.7 (d), Fd is a ring. Moreover, in Exercise1.7 (c) it was shown that every set in Fd can be written as a finite union of disjoint half-openrectangles.

Definition 3.7 (on unions of half-open rectangles). For a half open rectangle I = (a, b] ∈ Id witha = (α1, . . . , αd) and b = (β1, . . . , βd) let its volume be denoted by

|I| =d∏j=1

(βj − αj).

For A ∈ Fd of the form A = I1 ∪ . . . ∪ In with disjoint (Ij)nj=1 in Id define λd : Fd → [0,∞] by

λd(A) =

n∑j=1

|Ij |.

This is well-defined since for a rectangle R ⊆ Rd with A ⊆ R, we again have

λd(A) =

∫R

1A dx,

where the latter is a d-dimensional Riemann integral.

Remark 3.8.

(1) When the dimension is fixed and there is no danger of confusion we will write λ for λd. It iscommon to write |A| for A ∈ Fd as well.

(2) For A ∈ Fd, λ(A) = λd(A) = |A| equals the volume of A.

Next we want to extend λd to σ(Fd) = B(Rd) (see Exercise 3.1 for this identity). To applyTheorem 3.1, we first need to check the σ-additivity of λ on Fd. This will be done via Lemma2.11.

Lemma 3.9. The function λ : Fd → [0,∞] is σ-additive on Fd.

Proof. By Lemma 2.11 it suffices to prove each sequence (An)n≥1 in Fd with An ↓ ∅, satisfiesµ(An)→ 0. Fix ε > 0. We have to find N ∈ N such that λ(An) < ε for all n ≥ N .

Step 1: For each n ∈ N choose a Bn ∈ Fd such that Bn ⊆ An and λ(An \Bn) ≤ 2−nε.20 Since

Bn ⊆ An also∞⋂n=1

Bn = ∅. It follows that {(Bn)c : n ∈ N} is an open cover of the set A1 which is

compact by the Heine-Borel theorem. Therefore, there exists an N such that A1 ⊆N⋃n=1

(Bn)c. It

follows thatN⋂n=1

Bn ⊆ Ac1. Since all for all n ≥ 1, Bn ⊆ A1, we must have thatN⋂n=1

Bn = ∅.

19Recall that 1A(x) = 1 if x ∈ A and 1A = 0 if x ∈ Ac.20So we just choose a set Bn which is slightly smaller than An.

Page 11: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 11

Step 2: Let Cn =n⋂j=1

Bj for n ≥ 1. For every n ≥ 1, An \ Cn =n⋃j=1

(An \ Bj) ⊆n⋃j=1

(Aj \ Bj).

Therefore, using Theorem 2.8 (i) in (∗) and Exercise 2.2 (c) in (∗∗), we find

λ(An \ Cn)(∗)≤ λ

( n⋃j=1

(Aj \Bj)) (∗∗)≤

n∑j=1

λ(Aj \Bj) ≤n∑j=1

2−jε < ε.

Since Cn = ∅ for all n ≥ N , we can conclude that λ(An) = λ(An \ Cn) < ε for every n ≥ N . ,

We can now deduce the main result of this section.

Theorem 3.10 (Lebesgue measure). There exists a unique measure λ on (Rd,B(Rd)) such thatfor all half-open rectangles I, one has λ(I) = |I|, where |I| is the volume of I. Moreover, for allh ∈ Rd and A ∈ B(Rd), λ(A+ h) = λ(A).21

In the above A+ h := {x+ h : x ∈ A}.Proof. Step 1: Existence. In Lemma 3.9 we have shown that λ is σ-additive on the ring Fd.Therefore, by Theorem 3.1 λ extends to a measure on σ(Fd) = B(Rd) (see Exercise 3.1).

Step 2: Uniqueness. Let µ be another measure such that µ(I) = |I| for half-open rectanglesI ∈ Id. Fix n ∈ N and let Sn = (−n, n]d. Define λ(n) and µ(n) on B(Rd) by

λ(n)(A) = λ(A ∩ Sn) and µ(n)(A) = µ(A ∩ Sn).

Then λ(n) and µ(n) are measures and λ(n)(Rd) = λ(Sn) = |Sn| and similarly µ(n)(Rd) = |Sn|.Since λ(n) and µ(n) coincide on Id, it follows from Example 3.4 (b) and Proposition 3.5 thatλ(n) = µ(n) on B(Rd). Therefore, for any A ∈ B(Rd), since A ∩ Sn ↑ A Theorem 2.10 yields

λ(n)(A) = λ(A ∩ Sn)→ λ(A) and µ(n)(A) = µ(A ∩ Sn)→ µ(A).

Thus λ(A) = µ(A).

Step 3: Translation invariance: Let h ∈ Rd. We claim that for every A ∈ B(Rd) one hasA + h ∈ B(Rd). For this let Ah = {A ∈ B(Rd) : A + h ∈ B(Rd)}. By definition Ah ⊆ B(Rd).One can check that Ah is a σ-algebra. For each open set A one has A + h is open and henceA+ h ∈ B(Rd). Therefore, B(Rd) = σ({open sets}) ⊆ Ah, and the claim follows.

Define µh on B(Rd) by µh(A) = λ(A + h). Then µh is a measure and for any half-openrectangle I, µh(I) = |I + h| = |I| = λ(I). By the uniqueness of step 2, we find µh(A) = λ(A) forall A ∈ B(Rd) and this proved the result. ,

Remark 3.11. From Theorem B.6 one can actually see that for any A ∈ B(Rd),

λ(A) = inf{ ∞∑j=1

|Ij | : A ⊆∞⋃j=1

Ij , where (Ij)j≥1 are disjoint half-open rectangles},

but we will not use this formula.

Exercises on the Lebesgue measureIf the dimension is fixed we write λ instead of λd for simplicity.

Exercise∗ 3.1. Let Fd be as in Example 1.8. Show that σ(Fd) is the Borel σ-algebra B(Rd).Hint: Use the Lindelof Lemma 1.15.

Exercise 3.2.

(a) Show that any countable subset A ⊆ Rd is in B(Rd).Hint: First show that {x} ∈ B(Rd) for every x ∈ Rd.

(b) Show that any countable subset A ⊆ Rd satisfies λ(A) = 0. In particular, λ(Qd) = 0.Hint: First show that λ({x}) = 0.

21This is called translation invariance. Up to a scaling factor λ is the only measure on B(Rd) which satisfiesthis property. See Exercise 3.6.

Page 12: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

12 MEASURE AND INTEGRATION

Exercise∗ 3.3. For a, b ∈ Rd with a = (α1, . . . , αd) and b = (β1, . . . , βd) with αj ≤ βj forj ∈ {1, . . . , k} let

(a, b) = (α1, β1)× . . .× (αd, βd) and [a, b] = [α1, β1]× . . .× [αd, βd]

be the open and closed rectangle, respectively. Prove that

λ((a, b)

)= λ

([a, b]

)= λ

((a, b]

)and thus all coincide with the volume of the rectangle.Hint: Use Theorem 2.10.

Exercise∗ 3.4 (Uncountable sets can have measure zero). Show that the Cantor set C ⊆ [0, 1] isin B(Rd) and satisfies λ(C) = 0.Hint: What can you say about λ(Cn) ?

Exercise∗ 3.5. For A ⊆ R and t ≥ 0 let tA = {tx : x ∈ A}. Show that for each A ∈ B(R),λ(tA) = tλ(A).Hint: Use the same method as in the proof of Theorem 3.10.

Exercise∗ 3.6. Let µ : B(R) → [0,∞] be a measure such that for all h ∈ R and A ∈ B(R),µ(A+ h) = µ(A). Let c = µ((0, 1]) and assume c ∈ (0,∞). Prove the following assertions: 22

(a) For each x ≥ 0 and q ∈ N, µ((0, qx]

)= qµ

((0, x]

).

(b) For each p, q ∈ N, µ((0, pq ]

)= cpq .

(c) For each x ≥ 0, µ((0, x]

)= cx.

(d) For each a ≤ b, µ((a, b]

)= c(b− a).

(e) For each A ∈ B(R), µ(A) = cλ(A).

Exercise∗∗ 3.7 (Lebesgue-Stieltjes23 measure). Let a < b and let F : R→ R be right-continuousand increasing. Show that µ((a, b]) = F (b)− F (a) for a ≤ b extends to a measure on B(R).

Exercises on general measures

Exercise∗ 3.8. Let A be a σ-algebra on S and let T ⊆ S. Define the restricted σ-algebra AT onT by

AT = {A ∩ T : A ∈ A}.(a) Show that AT is a σ-algebra.(b) If T ∈ A, show that AT = {A ⊆ T : A ∈ A}.(c) Let µ be a measure on (S,A). If T ∈ A show that the restriction of µ to AT is a measure

again.(d) If A is the Borel σ-algebra on metric space (S, d), then AT coincides with the Borel σ-algebra

on (T, d).

Exercise 3.9 (Non-uniqueness of extensions I). Let S = {1, 2, 3, 4} and let F = {{1, 2}, {1, 3}}.Define µ : F → [0,∞] by µ({1, 2}) = µ({1, 3}) = 1

2 . Find two different extensions of µ toσ(F) = P(S). Why does this not contradict Proposition 3.5?

Exercise∗ 3.10 (Non-uniqueness of extensions II). Let S = N and let F ={{n, n+1, . . .} : n ∈ N

}.

(a) Show that F is a π-system.(b) Show that σ(F) = P(N)(c) Let τ be the counting measure and let µ : P(N)→ [0,∞] be defined by µ(A) = 2τ(A). Show

that τ = µ on F . Why does this not contradict Proposition A.7?

22From this exercise we see that up to a scaling factor, λ is the only translation invariant measure on B(R). The

case for dimensions d ≥ 2 holds as well and can be proved in a similar way.23Thomas Stieltjes 1856-1894 was a Dutch mathematician working in Analysis. He has even worked in Delft.

Page 13: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 13

4. Measurable functions

One of the aims will be to integrate functions f : S → R with respect to a measure µ on ameasurable space (S,A). A way to do this is to use discretization in the range24 of f . So we wouldlike to know the measure of for instance the set Ay,ε = {s ∈ S : f(s) ∈ [y, y + ε]}. Knowing thisfor all y ∈ R and all ε > 0 makes it possible estimate the “area” under f . Of course we do needthe sets to be in A to make this work. This is the motivation of the definition of measurability.See Section 5 for details on integration.

The natural setting to introduce measurability of functions is a follows:

Definition 4.1. Let (S,A) and (T,B) be two measurable spaces. A function f : S → T is calledmeasurable if for each B ∈ B, one has25 f−1(B) ∈ A.

Remark 4.2.

(1) The composition of two measurable function is again measurable (see Exercise 4.1).(2) Instead of f−1(B) or {s ∈ S : f(s) ∈ B} one sometimes writes {f ∈ B} for the same set.(3) In probability theory measurable functions are called random variables.

It suffices to check measurability on a generating collection F ⊆ B:

Lemma 4.3. Let (S,A) and (T,B) be two measurable spaces and let f : S → T . Suppose F ⊆ Bis such that σ(F) = B. If f−1(F ) ∈ A for all F ∈ F , then f is measurable.

Proof. Define B = {B ∈ B : f−1(B) ∈ A}. Our aim is to show that B = B. We claim that Bis a σ-algebra. Indeed, since f−1(∅) = ∅ ∈ A also ∅ ∈ B. Similarly, since f−1(T ) = S ∈ A,

we find T ∈ B. If B ∈ B, then f−1(T \ B) = S \ f−1(B) ∈ A, which implies that Bc ∈ B. If

B1, B2, . . . ∈ B, then

f−1( ∞⋃n=1

Bn

)=

∞⋃n=1

f−1(Bn) ∈ A.

Therefore, the claim follows.

Now since F ⊆ B, the claim yields B = σ(F) ⊆ B ⊆ B. This implies B = B. ,

In the sequel a metric space X will always be equipped with its Borel σ-algebra B(X) (unlessotherwise stated).

Proposition 4.4 (Continuous mappings are measurable). Let (X, d) and (Y, ρ) be metric spaces.If f : X → Y is continuous, then f is measurable.26

Proof. By the continuity of f we find that for all open O ⊆ Y the inverse image is f−1(O) openin X and hence in B(X). Since the open sets of Y generate the Borel σ-algebra, the result followsfrom Lemma 4.3 with F = {O ⊆ Y : O is open}. ,

The most frequent case we will encounter is when f : S → R and (S,A) is a measurable space.Unless otherwise stated we consider the Borel σ-algebra B(R) on R. The following characterizationof measurability will be useful.

Proposition 4.5 (Real valued functions). Let (S,A) be a measurable space. For f : S → R thefollowing are equivalent:

(i) f is measurable.(ii) For all r ∈ R, one has f−1((−∞, r]) ∈ A.

(iii) For all r ∈ R, one has f−1((−∞, r)) ∈ A.

24In Riemann integration of functions f : Rd → R the discretization is always done in the domain of the function.

This is one of the major differences with Lebesgue integration25Recall that f−1(B) is called the inverse image of B by f and is defined by f−1(B) = {s ∈ S : f(s) ∈ B}26The same result holds for topological spaces and the proof is the same. In the setting of Borel-σ-algebras,

measurable functions are often called Borel measurable.

Page 14: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

14 MEASURE AND INTEGRATION

Proof. (i) ⇒ (ii) and (i) ⇒ (iii) are trivial.(iii) ⇒ (i): Let F = {(−∞, r) : r ∈ R}. In Exercise 1.6 we have seen that σ(F) = B(R).

Therefore, (i) follows from Lemma 4.3.(ii) ⇒ (i): This can be proved as before by proving the required version of Exercise 1.6. ,

Example 4.6. Let S = {1, 2, 3} and A = {S,∅, {1, 2}, {3}}. Let f : S → R be given by f(s) = 2016if s = 1 and f(s) = 0 if s 6= 1. Then f is not measurable, because f−1({2016}) = {1} /∈ A. If wereplace A by (for example) A′ := {S,∅, {1}, {2, 3}}, then f becomes measurable.

Recall that x ∨ y = max{x, y} and x ∧ y = min{x, y}.Theorem 4.7. Let (S,A) be a measurable space. Let f, g : S → R be measurable functions andlet α ∈ R. Then the following functions are all measurable as well:

f + g, f − g, f · g, f ∨ g, f ∧ g, f+ := f ∨ 0, f− := (−f) ∨ 0, |f |, α · f, 1

f(if f 6= 0 on S).

Proof. We first claim that h : S → R2 given by h(s) = (f(s), g(s)) is measurable. To prove thisobserve that for all half-open rectangles I = I1 × I2 ⊆ R2, one has

h−1(I) = {s ∈ S : f(s) ∈ I1 and g(s) ∈ I2} = f−1(I1) ∩ g−1(I2) ∈ A.By Exercise 3.1 σ(half open rectangles) = σ(F2) = B(R2), we can use Lemma 4.3 to find that his measurable.

To prove the statements we use that continuous functions are measurable (see Proposition 4.4)and the fact that the composition of measurable functions is again measurable (see Exercise 4.1).For instance let ϕ : R2 → R be given by ϕ(x, y) = x + y. Then ϕ is continuous and thereforemeasurable. Now writing f + g = ϕ ◦ h the required measurability follows.

The proofs for the difference, product, maximum, minimum are similar. Note that the maximumx ∨ y = 1

2 (x + y) + 12 |x − y| and this is a continuous function from R2 to R. For the minimum

there is an analogue formula.The measurability of f±, |f |, α · f and 1

f all follow in the same way by rewriting them as φ ◦ ffor a suitable continuous function φ.

The case 1f requires some comment. Let φ : R\{0} → R be given by φ(x) = 1

x . Note that in this

case R \ {0} is a metric space on which φ is continuous, and therefore measurable by Proposition4.4. Now for all open sets B ∈ B(R), C := φ−1(B) is open in R \ {0} and hence in B(R). Thus(φ ◦ f)−1(B) = f−1(C) ∈ A. Therefore, the measurability of φ ◦ f follows from Lemma 4.3. ,

Let R = [−∞,∞]. It will be useful to introduce measurability of functions f : S → R as well.For this, we introduce an analogue of the Borel σ-algebra on R.

Definition 4.8. Let B(R) be the the σ-algebra generated by the sets {∞}, {−∞} and B ∈ B(R).The σ-algebra B(R) will be called the Borel σ-algebra of R.

From Lemma 4.3 we see that a function f : S → R is measurable if and only if {s ∈ S : f(s) =±∞} ∈ A and f−1(B) ∈ A for each B ∈ B(R). We extend addition, multiplication, etc. to R inthe following way:

∞+ a = a+∞ =∞, for all a ∈ (−∞,∞]

−∞+ a = a−∞ = −∞, for all a ∈ [−∞,∞)

∞ · a = a · ∞ =∞, for all a ∈ (0,∞]

∞ · a = a · ∞ = −∞, for all a ∈ [−∞, 0)

∞ · 0 = 0 · ∞ =a

∞ =a

−∞ = 0 for all a ∈ (−∞,∞)

In this setting Theorem 4.7 remains true27 for functions f, g : S → R.

27We do not define ∞−∞, so some cases need to be excluded. For the proof one additionally needs to checkin each of the cases that inverse images of {∞}, {−∞}, {0} are measurable. We leave this to the reader.

Page 15: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 15

The next result shows that measurability is preserved under taking countable suprema, count-able infimum, limits of sequences, etc. For a sequence of numbers (xn)n≥1 in R, let

lim supn→∞

xn = limk→∞

supn≥k

xn and lim infn→∞

xn = limk→∞

infn≥k

xn.

The limit of yk := supn≥k xn exists28 since (yk)k≥1 is decreasing. Moreover,

(4.1) lim supn→∞

xn = limk→∞

yk = infk≥1

yk = infk≥1

supn≥k

xn.

Similar formulas hold for the lim infn→∞ xn. Recall that the lim supn→∞ xn and lim infn→∞ xnalways exist. Moreover they both coincide with limn→∞ xn if and only if (xn)n≥1 converges in R.

Theorem 4.9. Let (S,A) be a measurable space. For each n ∈ N let fn : S → R be a measurablefunction. Then each of the following functions is measurable as well:29

supn≥1

fn, infn≥1

fn, lim supn→∞

fn, lim infn→∞

fn.

Moreover, if fn → f pointwise30, then f is measurable again.

Proof. Let g = supn≥1 fn. Then for each r ∈ R,

g−1([−∞, r]) = {s ∈ S : g(s) ≤ r} = {s ∈ S : fn(s) ≤ r for all n ∈ N} =

∞⋂n=1

f−1n ([−∞, r]) ∈ A.

Since σ({[−∞, r] : r ∈ R}) = B(R), the measurability of g follows from Lemma 4.3. The case ofinfima follows from infn≥1 fn = − supn≥1(−fn).

By (4.1) we can write lim supn→∞ fn = infk≥1 supn≥k fk. Therefore, the measurability followsfrom the previous cases. The remaining cases follow from lim infn→∞ fn = − lim supn→∞(−fn)and limn→∞ fn = lim supn→∞ fn. ,

Definition 4.10. A function f : S → R is called a simple function31 if f is measurable andtakes only finitely many values.

Letting x1, . . . , xn ∈ R denote the distinct values of f and Aj = {s ∈ S : f(s) = xj}, we canalways represent a simple function as

f =

n∑j=1

xj · 1Aj .

Of course if xj = 0 for some j ∈ {1, . . . , n}, we could leave it out from the sum.

Example 4.11. Let S = R and A = B(R). The following functions are simple functions:

(a) f = π · 1(0,1) − 4 · 1(13,14] + 5 · 1Z∩(−∞,0).(b) f = 1Q.

Next we show that measurable functions can be written as limits of simple functions. For thiswe discretize in the range space in a suitable way. The result plays a crucial role in the integrationtheory in Section 5.

Theorem 4.12. Let (S,A) be a measurable space. 32

(i) Let f : S → [0,∞] be measurable. Then there exists a sequence of simple functions (fn)n≥1such that 0 ≤ f1(s) ≤ f2(s) ≤ . . . and limn→∞ fn(s) = f(s) for each s ∈ S.

(ii) Let f : S → R be measurable. Then there exists a sequence of simple functions (fn)n≥1 suchthat limn→∞ fn(s) = f(s) for all s ∈ S.

28Here we allow divergence to ±∞29Here it is important what we work with countable suprema, infimum, etc.30Here we allow divergence to ±∞31In part of the literature this is called a step function, but we will use this name for a different class of functions32In the following we allow divergence to ±∞

Page 16: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

16 MEASURE AND INTEGRATION

Proof. 33 (i): For each n ∈ N and j ∈ {0, 1, . . . , 4n − 1}, let

An,j = {s ∈ S : j2−n ≤ f(s) < (j + 1)2−n} and An = {s ∈ S : f(s) ≥ 2n}.Then for each n, j one has An,j , An ∈ A. Define34

fn = 2n1An +

4n−1∑j=0

j

2n1An,j .

It is clear that each fn takes finitely many values. Moreover, by Exercise 4.5 and Theorem 4.7,each fn is measurable. Thus each fn is a simple function.

Now fix s ∈ S. We first prove 0 ≤ fn(s) ≤ fn+1(s) for each n ∈ N. First assume f(s) < 2n.Then selecting the unique j ∈ {0, . . . , 4n − 1} such that s ∈ An,j we find that fn(s) = j2−n.Similarly, we can select k ∈ {0, . . . , 4n+1 − 1} such that s ∈ An+1,k and we find that fn+1(s) =

k2−(n+1). Since, f(s) ≥ j2−n = 2j2−(n+1), we find that k ≥ 2j and therefore

fn(s) = j2−n ≤ k2−(n+1) = fn+1(s).

The case f(s) ≥ 2n can be treated similarly and is left to the reader.To prove that fn(s) → f(s), first assume f(s) < ∞. Let ε > 0. Choose N ∈ N so large that

f(s) < 2N and 2−N < ε. Let n ≥ N . Selecting j ∈ {0, . . . , 4n− 1} such that s ∈ An,j we find that

|f(s)− fn(s)| ≤ 2−n ≤ 2−N < ε.

Therefore, fn(s) → f(s) in this case. If f(s) = ∞, then fn(s) = 2n for every n ∈ N and thusfn(s)→∞ = f(s).

(ii): Write f = f+ − f−. Then by (i) we can find simple functions fn,+, fn,− : S → R suchthat fn,+ → f+ and fn,− → f−. Let fn = fn,+ − fn,− for n ∈ N. Define the sets A+ and A− byA± = {s ∈ S : ±f(s) ∈ [0,∞]}. Then A+ ∪A− = S. If s ∈ A+, then

fn(s) = fn,+(s)→ f+(s) = f(s).

The case s ∈ A− is similar.,

Exercises

Exercise 4.1. Let (Sj ,Aj) for j = 1, 2, 3 be measurable spaces. Assume f : S1 → S2 andg : S2 → S3 are both measurable. Show that the composition g ◦ f : S1 → S3 is measurable.

Exercise 4.2. Let (S,A) be a measurable space. Let f, g : S → R be measurable functions. Showthat {s ∈ S : f(s) = g(s)} ∈ A and {s ∈ S : f(s) < g(s)} ∈ A.

Exercise 4.3. Let (S,A) be a measurable space. Let f : S → R be a measurable function andp ∈ (0,∞). Show that the function |f |p is measurable.

Exercise 4.4. Let S be a set. For a function f : S → R let Af = {f−1(B) : B ∈ B(R)}. Showthat Af is a σ-algebra.35

Exercise 4.5. Let (S,A) be a measurable space and let A ⊆ S. Show that A ∈ A if and only ifthe function 1A : S → R is measurable.

Exercise 4.6. Let (S,A) be a measurable space. Let f1, f2, . . . : S → R be measurable functionsand A1, A2, . . . ∈ A be disjoint. Show that the function f =

∑∞n=1 1Anfn is measurable.

Exercise∗ 4.7 (Vector-valued functions). Let (S,A) be a measurable space. For each j ∈ {1, . . . , d}let fj : S → R be a function and let f : S → Rd be given by f = (f1, . . . , fd). Prove that f ismeasurable if and only if fj is measurable for each j ∈ {1, . . . , d}.Hint: For the “if part” one can use the same technique as in Theorem 4.7.

33For the proof make a picture where you make a partition into intervals of length 2−n on the y-axis. For thepicture put the set S is on the x-axis

34The idea is that for each n ∈ N we approximate f up to 2−n on the set {f < 2n}.35The σ-algebra Af is called the σ-algebra generated by f . It is the smallest σ-algebra for which f is measurable.

Page 17: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 17

Exercise∗ 4.8 (Monotone functions). Assume that f : R → R is increasing. Show that f ismeasurable, where as usually on R we consider the Borel σ-algebra.Hint: Use Proposition 4.5.

Exercise∗ 4.9 (Set of convergence). Let (S,A) be a measurable space. Let f1, f2, . . . : S → R bemeasurable functions. Define

A = {s ∈ S : (fn(s))n≥1 is convergent}.(a) Explain why A = {s ∈ S : (fn(s))n≥1 is a Cauchy sequence}.(b) Show that for each k, n,m ∈ N the set A(k, n,m) = {s ∈ S : |fn(s)− fm(s)| < 1

k} is in A.

(c) Show that A =∞⋂k=1

∞⋃N=1

∞⋂m=N

∞⋂n=N

A(k, n,m).

Hint: s ∈ A if and only if ∀k ∈ N ∃N ∈ N ∀n ∈ N ∀m ∈ N : |fn(s)− fm(s)| < 1k and connect

the universal quantifier with an intersection and the existential quantifier with a union.(d) Conclude that A ∈ A.

Page 18: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

18 MEASURE AND INTEGRATION

5. Construction of the integral

In this section (S,A, µ) is a measure space. Our goal is to construct an integral for measurablefunctions f : S → R. Notation: ∫

S

f dµ or

∫S

f(s) dµ(s).

The integral will be built in three steps:

(1) For simple functions f : S → [0,∞);(2) For measurable functions f : S → [0,∞];(3) For (certain) measurable functions f : S → R.

The advantage of this setting is that it works for any measure space (S,Σ, µ). Moreover, in thespecial case of the Lebesgue measure it extends the Riemann integral.

It will be convenient to use the following terminology.

Definition 5.1. For measurable functions f, g : S → R we say that f = g almost everywhereif µ({s ∈ S : f(s) 6= g(s)}) = 0. Notation: f = g a.e.36

Similarly, one can define f <∞ a.e., etc.

5.1. Integral for simple functions.

Definition 5.2 (Integral for simple functions). Let f : S → [0,∞] be a simple function given by

(5.1) f =

n∑j=1

xj · 1Aj ,

with x1, . . . , xn ∈ [0,∞] and (Aj)nj=1 disjoint sets in A. For E ∈ A let37∫E

f dµ =

n∑j=1

xj · µ(E ∩Aj).

This is called the integral of f over the set E.

Remark 5.3. By using a common refinement one checks that if a different representation for thefunction f from (5.1) is used, then the integral over E gives the same value (see Lemma 5.4 (ii)).Clearly

∫Ef dµ ∈ [0,∞] for every E ∈ A.

We continue with some basic properties of the integral.

Lemma 5.4. Let f, g : S → [0,∞] be simple functions. Then the following hold:

(i) For all E ∈ A,∫Ef dµ =

∫S

1Ef dµ.

(ii) (monotonicity I) If E ∈ A and f ≤ g on E, then∫Ef dµ ≤

∫Eg dµ.

(iii) (monotonicity II) If E,F ∈ A satisfy E ⊆ F , then∫Ef dµ ≤

∫Ff dµ.

(iv) (linearity) For all E ∈ A and α, β ∈ [0,∞),∫Eαf + βg dµ = α

∫Ef dµ+ β

∫Eg dµ.

(v) (additivity) For all disjoint sets E1, E2 ∈ A,∫E1∪E2

f dµ =∫E1f dµ+

∫E2f dµ.

(vi)∫Sf dµ = 0 if and only if f = 0 almost everywhere.38.

Proof. (i). This is immediate from the definition and the identity 1E∩A = 1E · 1A.For the proof of the remaining assertions we continue with a preliminary observation. Write

f =∑mi=1 xi · 1Ai with (Ai)

mi=1 disjoint sets in A with

m⋃i=1

Ai = S, and g =∑nj=1 yj · 1Bj with

36In probability theory this is usually called almost surely and this is abbreviated as a.s.37Here we use the convention 0 · ∞ = 038See Definition 5.1

Page 19: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 19

(Bj)nj=1 disjoint sets in A with

m⋃i=1

Bi = S. Let Ci,j = Ai ∩Bj for all i and j. Then (Ci,j)m,ni,j=1 are

disjoint sets in A. By additivity∫E

f dµ =

m∑i=1

xi · µ(E ∩Ai) =

m∑i=1

n∑j=1

xi · µ(E ∩ Ci,j)(5.2)

∫E

g dµ =

n∑j=1

yj · µ(E ∩Bj) =

n∑j=1

m∑i=1

yj · µ(E ∩ Ci,j).(5.3)

(ii): For disjoint sets A,B ⊆ S one has 1A∪B = 1A + 1B and hence

(5.4) 1Ef =

m∑i=1

n∑j=1

xi · 1E∩Ci,j and 1Eg =

n∑j=1

m∑i=1

yj · 1E∩Ci,j .

Therefore, xi ≤ yj whenever i, j satisfy E ∩ Ci,j 6= ∅. This together with (5.2), (5.3), yields (ii).(iii): This follows from µ(E ∩Ai) ≤ µ(F ∩Ai) which is immediate from the monotonicity of µ.(iv): By (5.4) with E = S, we can write αf + βg =

∑mi=1

∑nj=1(αxi + βyj) · 1Ci,j . Therefore,∫

E

αf + βg dµ =

m∑i=1

n∑j=1

(αxi + βyj)µ(E ∩ Ci,j)

= α

m∑i=1

n∑j=1

xiµ(E ∩ Ci,j) + β

n∑j=1

m∑i=1

yjµ(E ∩ Ci,j) = α

∫E

f dµ+ β

∫E

g dµ,

where we used (5.2) and (5.3).(v): Since 1E1∪E2f = 1E1f + 1E2f ,∫

E1∪E2

f dµ(i)=

∫S

1E1∪E2f dµ

(iv)=

∫S

1E1f dµ+

∫S

1E2f dµ

(i)=

∫E1

f dµ+

∫E2

f dµ.

(vi): By leaving out some of those k for which xk = 0, we can assume xi > 0 for all i. LetA = {s ∈ S : f(s) > 0} and observe that A =

⋃ni=1Ai. Now

∫Sf dµ =

∑mi=1 xiµ(Ai) with

µ(Ai) ≥ 0 for each i. Therefore, if the integral is zero, then µ(Ai) = 0 for each i and henceµ(A) =

∑ni=1 µ(Ai) = 0. Conversely, if f = 0 a.e., monotonicity yields µ(Ai) ≤ µ(A) = 0, and the

result follows. ,

Example 5.5. Let λ be the Lebesgue measure on R. Since Q ∈ B(R) it follows from Exercises 1.5and 4.5 that 1Q is measurable and by Exercise 3.2 for any E ∈ B(R),

∫E

1Q dλ = λ(E ∩ Q) = 0.Note that 1Q is not Riemann integrable on any interval [a, b] with a < b. More on the Riemannintegral can be found in Example 5.16.

5.2. Integral for positive measurable functions. In order to extend the definition of theintegral to arbitrary measurable functions f : S → [0,∞] we use the following lemma. In thesequel, we write fn ↑ f if for all s ∈ S, (fn(s))n≥1 is increasing and fn(s)→ f(s).

Lemma 5.6 (Consistency). Let f : S → [0,∞] be a measurable function. Suppose that (fn)n≥1is sequence of simple functions such that 0 ≤ fn ↑ f . Suppose g : S → [0,∞] is a simple function

such that 0 ≤ g ≤ f . Then

∫E

g dµ ≤ limn→∞

∫E

fn dµ for every E ∈ A.

Observe that limn→∞

∫E

fn dµ exists in [0,∞], because it is an increasing sequence of real numbers.

Proof. Let ε ∈ (0, 1) and set En = {s ∈ E : (1 − ε)g(s) ≤ fn(s)} for n ∈ N. Then using theindicated parts of Lemma 5.4 in the estimates below, we find

(1− ε)∫En

g dµ(iv)=

∫En

(1− ε)g dµ(ii)

≤∫En

fn dµ(iii)

≤∫E

fn dµ(ii)

≤ limn→∞

∫E

fn dµ.

Page 20: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

20 MEASURE AND INTEGRATION

It remains to prove∫Eng dµ→

∫Eg dµ. For this write g =

∑mi=1 xi ·1Ai with (Ai)

mi=1 in A disjoint

and x1, . . . , xm ≥ 0. Since En ∩Ai ↑ E ∩Ai as n→∞, Theorem 2.10 yields

(5.5)

∫En

g dµ =

m∑i=1

xi · µ(En ∩Ai)→m∑i=1

xi · µ(E ∩Ai) =

∫E

g dµ.

,

Definition 5.7 (Integral for positive functions). For a measurable function f : S → [0,∞] and asequence of simple functions (fn)n≥1 with 0 ≤ fn ↑ f we define the integral of f over E ∈ A as∫

E

f dµ = limn→∞

∫E

fn dµ in [0,∞].

The above limit exists in [0,∞] since the numbers an :=∫Efn dµ form an increasing sequence

(an)n≥1 in [0,∞]. Note that by Theorem 4.12 we can always find simple functions fn : S → [0,∞)such that 0 ≤ fn ↑ f . However, we need to check that the above definition does not depend on thechoice of the sequence (fn)n≥1. Let (gm)m≥1 be another sequence of simple functions such that0 ≤ gm ↑ f and let bm =

∫Egm dµ. It suffices to show that limn→∞ an = limm→∞ bm. By Lemma

5.6 for each m ≥ 1,

bm =

∫E

gm dµ ≤ limn→∞

∫E

fn dµ = limn→∞

an.

From this we obtain limm→∞ bm ≤ limn→∞ an. Reversing the roles of gm and fn, one sees thatthe converse holds as well.

Next we can extend the properties of the integral of Lemma 5.4.

Proposition 5.8. Let f, g : S → [0,∞] be measurable functions. Then the following hold:

(i) For all E ∈ A,∫Ef dµ =

∫S

1Ef dµ.

(ii) (monotonicity I) If E ∈ A and f ≤ g on E, then∫Ef dµ ≤

∫Eg dµ.

(iii) (monotonicity II) If E,F ∈ A satisfy E ⊆ F , then∫Ef dµ ≤

∫Ff dµ.

(iv) (linearity) For all E ∈ A and α, β ∈ [0,∞),∫Eαf + βg dµ = α

∫Ef dµ+ β

∫Eg dµ.

(v) (additivity) For all disjoint sets E1, E2 ∈ A,∫E1∪E2

f dµ =∫E1f dµ+

∫E2f dµ.

(vi)∫Sf dµ = 0 if and only if f = 0 almost everywhere.

Proof. (i): Let (fn)n≥1 be simple functions such that 0 ≤ fn ↑ f . Then by Lemma 5.4 (i),∫E

f dµ = limn→∞

∫E

fn dµ = limn→∞

∫S

1Efn dµ =

∫S

1Ef dµ.

(ii)–(v): See Exercise 5.2.(vi): Assume f = 0 a.e. Choose a sequence of simple functions (fn)n≥1 such that 0 ≤ fn ↑ f .

Then fn = 0 a.e. for each n ∈ N and thus by Lemma 5.4 (vi),∫S

f dµ = limn→∞

∫S

fn dµ = 0.

For the converse we use contraposition. Assume one does not have f = 0 a.e. Then E = {s ∈S : f(s) > 0} satisfies µ(E) > 0. Letting En = {s ∈ S : f(s) ≥ 1

n} we find En ↑ E, so by Theorem2.10, µ(En)→ µ(E). Therefore, there exists an n ∈ N such that µ(En) > 0 and thus∫

S

f dµ(iii)

≥∫En

f dµ(ii)

≥∫En

1n dµ = 1

nµ(En) > 0.

,

Example 5.9 (Series are integrals). Let S = N and A = P(N). Let τ : P(N) → [0,∞] denote thecounting measure. Let f : N→ [0,∞] be arbitrary. Then f is clearly measurable. Now define foreach n ≥ 1, fn =

∑nj=1 f(j)1{j}. Since each fn is a simple function and 0 ≤ fn ↑ f we find∫

Nf dτ = lim

n→∞

∫Nfn dτ = lim

n→∞

n∑j=1

f(j) =

∞∑j=1

f(j).

Page 21: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 21

5.3. Integral for measurable functions. The final step is to define the integral for measurablefunctions f : S → R by using the splitting f = f+ − f−. Recall that f+ = max{f, 0} andf− = max{−f, 0}. Note that |f | = f+ + f−.

Definition 5.10. A measurable function f : S → R is called integrable when both∫Sf+ dµ and∫

Sf− dµ are finite. In this case the integral of f over E ∈ A is defined as∫

E

f dµ =

∫E

f+ dµ−∫E

f− dµ.

Remark 5.11.

(1) If f : S → R is integrable, then by monotonicity∫Ef+ dµ ∈ [0,∞) and

∫Ef− dµ are finite for

every E ∈ A. Therefore,∫Ef dµ is a number in R.

(2) A measurable function f : S → [0,∞] is integrable if and only if∫Sf dµ <∞.

(3) If f : S → R is integrable, then µ({s ∈ S : f(s) = ±∞}} = 0 (see Exercise 5.1).

To check integrability the following characterization is very useful.

Proposition 5.12. For a measurable function f : S → R the following are equivalent:

(i) f is integrable;(ii) |f | is integrable.

Moreover, in this case for each E ∈ A,

(triangle inequality)∣∣∣ ∫E

f dµ∣∣∣ ≤ ∫

E

|f |dµ.

Proof. First observe that |f | = f+ +f−. Therefore, both assertions are equivalent to I+, I− <∞,where I± =

∫Sf± dµ and hence the result follows. To prove the required estimate note that by

the triangle inequality and linearity of the integral for positive functions:∣∣∣ ∫E

f dµ∣∣∣ = |I+ − I−| ≤ I+ + I− =

∫E

f+ + f− dµ =

∫E

|f |dµ.

,

By considering the positive and negative part separately one can check Proposition 5.8 (i),(ii),(v) hold for all integrable functions f, g : S → R again (see Exercise 5.4). The following exampleshows that Parts (iii) and (vi) do not extend to this setting.

Example 5.13. Let S = R and λ the Lebesgue measure. Let f = 1(0,1] − 1(1,2]. Then f+ = 1(0,1]

and f− = 1(1,2] and ∫(0,1]

f dλ =

∫(0,1]

f+ dλ = λ((0, 1]) = 1,∫(0,2]

f dλ =

∫(0,2]

f+ dλ−∫(0,2]

f− dλ = λ((0, 1])− λ((1, 2]) = 1− 1 = 0.

The extension of the linearity (iv) is more difficult and proved below.

Proposition 5.14. Let f, g : S → R be integrable functions and α, β ∈ R. Then39 αf + βg isintegrable, and for all E ∈ A

(linearity)

∫E

αf + βg dµ = α

∫E

f dµ+ β

∫E

g dµ.

Proof. Note that by Proposition 5.12 each of the following functions is integrable f±, g±, |f |, |g|.Therefore, |α| |f |+ |β| |g| is integrable by Proposition 5.8 (iv). Since |αf + βg| ≤ |α| |f |+ |β| |g|,the function αf + βg is integrable as well (see Exercise 5.7).

39Of course we only consider those functions for which αf + β + g is well-defined.

Page 22: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

22 MEASURE AND INTEGRATION

It remains to prove the identity for each E ∈ A. To do this we first consider the case α = β = 1.Write f + g = φ−ψ, where φ = f+ + g+ and ψ = (f− + g−). Then by Proposition 5.8 (iv) φ andψ are integrable and Exercise 5.3 yields

(5.6)

∫E

f + g dµ =

∫E

φ dµ−∫E

ψ dµ

=

∫E

f+ dµ+

∫E

g+ dµ−∫E

f− dµ−∫E

g− dµ =

∫E

f dµ+

∫E

g dµ.

It remains to show that for all α ∈ R,∫E

αf dµ = α

∫E

f dµ.

If α ≥ 0, then αf = (αf)+ − (αf)− = αf+ − αf−. Proposition 5.8 (iv) yields∫E

αf dµ =

∫E

αf+ dµ−∫E

αf− dµ = α

∫E

f+ dµ− α∫E

f− dµ.

In case α < 0, then by (5.6) and the previous case, we find

0 =

∫E

αf + (−α)f dµ =

∫E

αf dµ+

∫E

(−α)f dµ =

∫E

αf dµ+ (−α)

∫E

f dµ

and the result follows by subtracting the second term on both sides. ,

Example 5.15. Every simple function f : S → R given by f =∑nj=1 xj1Aj with A1, . . . , An ∈ A

and µ(Aj) <∞ for all j, is integrable and by Proposition 5.14,∫E

f dµ =

n∑j=1

xj

∫E

1Aj dµ =

n∑j=1

xjµ(E ∩Aj).

In the next example we show that for continuous functions the integral with respect to theLebesgue measure coincides with the Riemann integral.40

Example 5.16. Let a < b be real numbers. Let λ be the restriction of the Lebesgue measure toB([a, b]) and let f : [a, b] → R be continuous. Note that f is measurable by Proposition 4.4.Moreover, since there is a constant M ≥ 0 such that |f | ≤ M , we can use Exercise 5.7 to deducethat f is (Lebesgue) integrable. Since f is continuous, it is Riemann integrable. Below we showthat L(f) = R(f), where we used the abbreviations

L(f) :=

∫[a,b]

f dλ (Lebesgue integral) and R(f) :=

∫ b

a

f(x) dx (Riemann integral).

To prove this identity let ε > 0. Since f is uniformly continuous we can find a δ > 0 such thatfor all x, y ∈ [a, b], |x− y| < δ implies |f(x)− f(y)| < ε

b−a . Let n ∈ N be such that b−an < δ. Let

xj = a + j b−an for j = 0, . . . , n. Let mj = min{f(x) : x ∈ [xj−1, xj ]} and Mj = max{f(x) : x ∈[xj−1, xj ]}. Let g(x) = 1{a}f(a) +

∑nj=1 1(xj−1,xj ]mj and G(x) = 1{a}f(a) +

∑nj=1 1(xj−1,xj ]Mj .

Then g and G are integrable in the sense of Riemann and Lebesgue and one can check

L(g) = R(g) =

n∑j=1

(xj − xj−1)mj =: α and L(G) = R(G) =

n∑j=1

(xj − xj−1)Mj =: β.

Since g ≤ f ≤ G, by monotonicity we find that L(f),R(f) ∈ [α, β]. Therefore,

|R(f)− L(f)| ≤ β − α =

n∑j=1

(xj − xj−1)(Mj −mj) ≤n∑j=1

b− an

ε

b− a = ε,

Since ε > 0 is arbitrary, we find that R(f) = L(f).

40It is even known that every Riemann integrable function f is Lebesgue integrable and the integrals coincide

(see [1, Theorem 23.6]). The proof uses a similar method as ours. However, one should be aware that not every

improper Riemann integral is Lebesgue integrable. See Theorem 6.12 and Exercise 6.10 for more details on improperRiemann integrals

Page 23: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 23

Example 5.17. Let S = R, A = P(R). Let x ∈ R and let δx be the Dirac measure from Example2.5. Then for any f : R→ R. ∫

Rf dδx = f(x).

Indeed, for simple functions f this is obvious. For f : S → [0,∞] this follows by approximationfrom below by simple functions. For general f : S → [0,∞] this follows by writing f = f+ − f−.

Example 5.18. 41 Consider the setting of the counting measure of Example 5.9. A function f :N→ R is integrable if and only if

∑∞j=1 |f(j)| <∞. In that case∫

Nf dτ =

∞∑j=1

f(j).

Finally we briefly indicate how one can extend the integral to complex functions f : S → C.On the complex numbers C we consider its Borel σ-algebra. If f : S → C is measurable, then wewrite f = u+ iv with u, v : S → R and u and v are measurable (see Exercise 5.11).

Definition 5.19. A measurable function f : S → C given by f = u + iv with u, v : S → R iscalled integrable if both u and v are integrable. In this case let∫

E

f dµ =

∫E

udµ+ i

∫E

v dµ, E ∈ A.

Exercise 5.11 yields that Propositions 5.12 and 5.14 extend to the complex setting.

Exercises about general integralsIn the exercises below (S,A, µ) denotes a measure space.

Exercise 5.1. Let f : S → [0,∞] be an integrable function. Show that f <∞ a.e.

Exercise 5.2.

(a) Prove Proposition 5.8 (iv).Hint: Approximate by simple functions as in the definition of the integral.

(b) Prove Proposition 5.8 (ii),Hint: Write 1Eg = 1E(g − f) + 1Ef and use Proposition 5.8 (iv).

(c) Prove Proposition 5.8 (iii).(d) Prove Proposition 5.8 (v).

The following was used in the proof of Proposition 5.14.

Exercise 5.3. Assume g, h : S → [0,∞] and f : S → R are all integrable functions and f = g−h.Show that ∫

E

f dµ =

∫E

g dµ−∫E

hdµ, E ∈ A.

Hint: Note that f+ + h = f− + g and use Proposition 5.8 (iv).

Exercise 5.4. Extend the following results to integrable functions f, g : S → R:

(a) Proposition 5.8 (i). First show that 1Ef is integrable for E ∈ A.(b) Proposition 5.8 (ii).(c) Proposition 5.8 (v)

Exercise 5.5. Let f, g : S → R both be measurable functions. Assume f is integrable and f = ga.e. Show that g is integrable and∫

E

f dµ =

∫E

g dµ, E ∈ A.

Hint: First try to prove this in the case f and g take values in R. In this case apply Proposition5.8 (vi) to h = |f − g|. For the case where f and g are R-valued one needs a careful analysis ofpositive and negative parts.

41For details see Exercise 5.8

Page 24: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

24 MEASURE AND INTEGRATION

Exercise 5.6. Let λ denote the Lebesgue measure on (R,B(R)). Let f : R→ R be an integrablefunction. Show that for all −∞ < a < b <∞∫

[a,b]

f dλ =

∫(a,b)

f dλ.

Is this true if we replace λ by an arbitrary measure on B(R)?

Exercise∗ 5.7 (Domination). Let f, g : S → R be measurable functions.

(a) Assume |f | ≤ g and g is integrable. Prove that f is integrable.(b) Show that f is integrable if and only if

∑∞n=−∞ 2nµ({s ∈ S : 2n < |f(s)| ≤ 2n+1}) <∞.

Hint: Use (a) for a suitable function g : S → [0,∞].

In the next exercise you are asked to give the details of Example 5.18

Exercise∗ 5.8. Consider the setting of the counting measure of Example 5.9.

(a) Show that a function f : N→ R is integrable if and only if∑∞n=1 |f(n)| <∞.

(b) Assume∑∞n=1 |f(n)| <∞. Show that∫

Nf dτ =

∞∑j=1

f(j).

Hint: Write f = f+ − f− and approximate f+ and f− in a similar way as in Example 5.9.

Exercise∗ 5.9 (Chebyshev’s inequality). Let f : S → [0,∞] be a measurable function. Prove thatfor every t > 0,

µ({s ∈ S : f(s) ≥ t}) ≤ 1

t

∫S

f dµ.

Hint: Let At = {s ∈ S : f(s) ≥ t} and write µ(At) =∫S

1At dµ.

Exercise∗ 5.10. Let λ be the Lebesgue measure on (Rd,B(Rd)). Let f : Rd → R be a measurablefunction. For h ∈ Rd define the translation fh : Rd → R by fh(x) = f(x− h).

(a) Show that fh is measurable.(b) Assume that f is integrable. Show that fh is integrable and∫

Rdfh dλ =

∫Rdf dλ.

Exercise∗∗ 5.11 (Complex functions).

(a) Let f : S → C be a measurable function and write f = u+ iv with u, v : S → R. Show that uand v are measurable.

(b) Prove Proposition 5.14 for integrable functions f, g : S → C and α, β ∈ C.(c) Prove Proposition 5.12 for integrable functions f : S → C.(d) Which parts of Proposition 5.8 remain true for integrable functions f, g : S → C?

Page 25: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 25

6. Convergence theorems and applications

In this section (S,A, µ) is a measure space.One of the problems with the Riemann integral is that the cases where limit and integral can

be interchanged are rather limited. In modern mathematics it is crucial to have better “tools”for this. We use the integration theory developed so far in order to obtain these tools. In Section6.1 we prove three famous convergence results. In Section 6.2 we give several consequences andapplications.

We start with several simple examples which illustrate some difficulties.

Example 6.1. For instance if (xj)j≥1 is an enumerate of Q, and An = {x1, . . . , xn}, then 1An → 1Qpointwise. Each 1An is Riemann integrable, but 1Q is not.

Example 6.2. Let fn : R → R be defined by one of the following n1(0, 1n ],1n1(n,2n) or 1(n,∞) for

n ≥ 1. Then fn → 0, but in each case∫R fndλ 6→ 0.

6.1. The three main convergence results. In this section we prove the three most famousconvergence results of integration theory:

• Monotone Convergence Theorem (MCT);42

• Fatou’s lemma;43

• Dominated Convergence Theorem (DCT).44

Theorem 6.3 (Monotone Convergence Theorem (MCT)). Let (fn)n≥1 be measurable functionssuch that 0 ≤ fn ↑ f . Then f is measurable and

limn→∞

∫S

fn dµ =

∫S

f dµ.

Proof. By Theorem 4.9, f : S → [0,∞] is measurable. By monotonicity of the integral

α := limn→∞

∫S

fn dµ ≤∫S

f dµ.

It remains to prove∫Sf dµ ≤ α. For this choose a sequence of simple functions (gm)m≥1 such

that 0 ≤ gm ↑ f . We claim that∫Sgm dµ ≤ α for each m ≥ 1. The required estimate follows from

the claim by letting m → ∞ and using the definition of∫Sf dµ. To prove the claim we repeat

part of the argument of Lemma 5.6. Fix m ∈ N and write g = gm.Let ε ∈ (0, 1) and for each n ≥ 1, set En = {s ∈ S : (1 − ε)g(s) ≤ fn(s)}. Then using the

indicated parts of Proposition 5.8 in the estimates below, we find

(1− ε)∫En

g dµ(iv)=

∫En

(1− ε)g dµ(ii)

≤∫En

fn dµ(iii)

≤∫S

fn dµ(ii)

≤ α.

Finally,∫Eng dµ →

∫Sg dµ follows from (5.5) with E = S in Lemma 5.6. Therefore, we can

conclude (1− ε)∫Sg dµ ≤ α and the claim follows since ε ∈ (0, 1) was arbitrary. ,

Lemma 6.4 (Fatou’s Lemma). Let (fn)n≥1 be a sequence of measurable functions with values in[0,∞]. Then ∫

S

lim infn→∞

fn dµ ≤ lim infn→∞

∫S

fn dµ.

Proof. Note that f = lim infn→∞ fn is a measurable function by Theorem 4.9. Fix n ∈ N and letgn = inf

k≥nfk. For all m ≥ n, we have gn ≤ fm and by the monotonicity of the integral this gives∫

Sgn dµ ≤

∫Sfm dµ. Therefore, taking the infimum over all m ≥ n, we find

(6.1)

∫S

gn dµ ≤ infm≥n

∫S

fm dµ.

42This result is due to Beppo Levi 1875–1961 who was an Italian mathematician.43This result is named after the French mathematician Pierre Fatou 1878–1929.44This is due to Lebesgue (see footnote 18)

Page 26: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

26 MEASURE AND INTEGRATION

Note that 0 ≤ gn ↑ f and thus∫S

f dµ =

∫S

limn→∞

gn dµ(MCT)

= limn→∞

∫S

gn dµ(6.1)

≤ limn→∞

infm≥n

∫S

fm dµ = lim infn→∞

∫S

fn dµ.

,

Theorem 6.5 (Dominated Convergence Theorem (DCT)). Let (fn)n≥1 be a sequence of integrablefunctions such that lim

n→∞fn = f pointwise. If there exists an integrable function g : S → [0,∞]

such that |fn| ≤ g for all n ≥ 1, then f is integrable and

(6.2) limn→∞

∫S

fn dµ =

∫S

f dµ.

Proof. The function f is measurable by Theorem 4.9. Moreover, since also |f | ≤ g, we see thatf is integrable. Let xn :=

∫Sfn dµ and x :=

∫Sf dµ. It suffices to show that lim sup

n→∞xn ≤ x ≤

lim infn→∞

xn. It follows that∫S

g dµ± x =

∫S

g ± f dµ =

∫S

limn→∞

(g ± fn) dµ (linearity)

≤ lim infn→∞

∫S

g ± fn dµ (Fatou’s lemma with g ± fn ≥ 0)

= lim infn→∞

(∫S

g dµ± xn)

(linearity).

=

∫S

g dµ+ lim infn→∞

(±xn).

Since∫Sg dµ <∞, it follows that ±x ≤ lim inf

n→∞(±xn). This implies the estimates

x ≤ lim infn→∞

xn and lim supn→∞

xn ≤ x,

where for the second part we used lim infn→∞

(−xn) = − lim supn→∞

xn. ,

For functions f, f1, f2 : S → R we say that fn → f a.e. if there exists a set A ∈ A such thatµ(A) = 0 and for all s ∈ S \ A, fn(s)→ f(s). For the details of the following remark we refer toExercise 6.5.

Remark 6.6. Let f, f1, f2, . . . : S → R be measurable functions.

(1) Theorem 6.3 also holds under the weaker assumption that 0 ≤ fn ↑ f a.e.(2) Theorem 6.5 also holds under the weaker assumptions that fn → f a.e. and |fn| ≤ g a.e.

Example 6.7. Let f : R→ R be integrable with respect to the Lebesgue measure λ. Let F : R→ Rbe given by F (x) =

∫(−∞,x] f dλ. Then F is continuous on R. Indeed, if xn < x and xn → x, then

F (xn) =

∫R

1(−∞,xn]f dλ(DCT)−→

∫R

1(−∞,x)f dλ = F (x),

where the last step follows from Exercise 5.5 and the fact that 1(−∞,x)f = 1(−∞,x]f almosteverywhere. Similarly, one checks that F (xn)→ F (x) if xn ≥ x and xn → x.

6.2. Consequences and applications. We continue with several consequences and applications.We start with a result on integration of series of positive functions.

Corollary 6.8 (Series and integrals). Let f1, f2, . . . : S → [0,∞] be measurable functions. Then

(6.3)

∫S

∞∑n=1

fn dµ =

∞∑n=1

∫S

fn dµ.

Proof. This follows from the MCT (See Exercise 6.4). ,

From a measure µ one can build many other measures in the following way:

Page 27: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 27

Theorem 6.9 (Density). Let f : S → [0,∞] be a measurable function. Define ν : A → [0,∞] by

ν(A) =

∫A

f dµ.

Then ν is a measure.45 Moreover, g : S → R is integrable with respect to ν if and only if fg isintegrable with respect to µ. In this case

(6.4)

∫S

g dν =

∫S

fg dµ

Proof. Step 1: Clearly, ν(∅) = 0. For (An)n≥1 a disjoint sequence in A, and A :=∞⋃n=1

An,

ν(A) =

∫S

1Af dµ =

∫S

∞∑n=1

1Anf dµ(6.3)=

∞∑n=1

∫S

1Anf dµ =

∞∑n=1

ν(An).

Step 2: First we show that (6.4) holds for all measurable g : S → [0,∞]. For g = 1A this isimmediate from

∫S

1A dν = ν(A) =∫S

1Af dµ. For simple functions g : S → [0,∞) this followsby linearity. For a measurable function g : S → [0,∞], by Theorem 4.12 we can find a sequenceof simple functions (gn)n≥1 such that 0 ≤ gn ↑ g. By the previous case, we obtain∫

S

g dν(MCT)

= limn→∞

∫S

gn dν = limn→∞

∫S

fgn dµ(MCT)

=

∫S

fg dµ.

Step 3: To prove the “if and only if” assertion and (6.4), let g : S → R be a measurablefunction. Since step 2 yields that ∫

S

g± dν =

∫S

fg± dµ,

both the equivalence and (6.4) follows by writing g = g+ − g−. ,

Example 6.10. Let f : R→ [0,∞] be measurable. Define ν : B(Rd)→ [0,∞) by

ν(A) =

∫A

f dλ, A ∈ B(R).

Then ν is a measure. For example one could take f(x) = 1√2πe−

x2

2 . Then ν is the standard

Gaussian measure on R.

In Example 5.16 we have discussed a connection between the Riemann and Lebesgue integral.A similar connection holds for improper Riemann integrals.

Definition 6.11. Let f : R→ R be a continuous function. We say that∫∞−∞ f(x) dx exists as an

improper Riemann integral if the limits L1 = limt→∞∫ t0f(x) dx and L2 = limt→−∞

∫ 0

tf(x) dx

exist in R. In that case define∫∞−∞ f(x) dx = L1 + L2.

We show that there is a connection with the Lebesgue integral with respect λ.

Theorem 6.12. Let f : R→ R be a continuous function. Then the following are equivalent:

(i) f is integrable;(ii)

∫∞−∞ |f(x)|dx exists as an improper Riemann integral.

Moreover, in this case∫∞−∞ f dx exists as am improper Riemann integral and

(6.5)

∫Rf dλ =

∫ ∞−∞

f dx.

It may happen that∫∞−∞ f dx exists as an improper Riemann integral without f being integrable

(see Exercise 6.10).

45The function f is usually called the density of ν.

Page 28: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

28 MEASURE AND INTEGRATION

Proof. (i)⇒(ii): First we make a general observation. Let g : R→ R be a continuous and integrable

function. In Example 5.16 we have seen that∫ t0g(x) dx =

∫R 1[0,t]g dλ. Setting L1 =

∫R 1[0,∞]g dλ,

we find ∣∣∣L1 −∫ t

0

g(x) dx∣∣∣ =

∣∣∣L1 −∫R

1[0,t]g dλ∣∣∣ =

∣∣∣ ∫R

1[t,∞)g dλ∣∣∣ ≤ ∫

R1[t,∞)|g|dλ.

Now the DCT yields that∫R 1[t,∞)|g|dλ→ 0 as t→∞, and we may conclude that

∫∞0g(x) dx =

L1 exists. The part on (−∞, 0] goes similarly and we find (6.5) with f replaced by g.Now (i) ⇒ (ii) follows by letting g = |f |. Moreover, (6.5) for f follows by taking g = f .

(ii)⇒ (i): By Example 5.16, monotonicity of improper Riemann integrals,∫R

1[−n,n]|f |dλ =

∫ n

−n|f(x)|dx ≤

∫ ∞−∞|f(x)|dx =: M <∞.

Therefore, by the MCT,∫R |f |dλ = limn→∞

∫R 1[−n,n]|f |dλ ≤M , and hence f is integrable. ,

Example 6.13. Let fn : [0,∞) → R be defined by fn(x) =(1 + x

n

)ne−2x. Below we show that

limn→∞∫ n0fn(x) dx = 1. Recall the standard limit 0 ≤

(1 + x

n

)n ↑ ex for every x ∈ [0,∞) and

thus 0 ≤ 1[0,n]fn ↑ f , where f(x) = e−x. Therefore, by Example 5.16∫ n

0

fn(x) dx =

∫[0,∞)

1[0,n]fn dλ(MCT)−→

∫[0,∞)

f dλ(6.5)=

∫ ∞0

f(x) dx = limt→∞

(1− e−t) = 1.

We end this section with a standard application of the DCT to calculus.

Theorem 6.14 (Differentiating under the integral sign). Suppose f : R× S → R is such that thefollowing hold:

(i) f is continuous and differentiable with respect to its first coordinate and for each y0 ∈ (a, b)there exists a δ > 0 and integrable function g : S → [0,∞] such that

|∂f∂y (y, s)| ≤ g(s), y ∈ (y0 − δ, y0 + δ), s ∈ S.(ii) s 7→ f(y, s) is integrable with respect to µ.

Then for all y ∈ R,d

dy

∫S

f(y, s) dµ(s) =

∫S

∂f

∂y(y, s) dµ(s).

Proof. Fix y0 ∈ R and let δ and g be as in (i). Let hn ∈ (0, δ) for n ≥ 1 be such that hn → 0. Letφn : (y0 − δ, y0 + δ)× S → R and F : (y0 − δ, y0 + δ)→ R be given by

φn(y, s) =f(y + hn, s)− f(y, s)

hnand F (y) =

∫S

f(y, s) dµ(s).

Then φn(y, s) → ∂f∂y (y, s) for each y ∈ R and s ∈ S. Therefore, Theorem 4.12 yields that

s 7→ φn(y, s) is measurable. From the mean value theorem46 we obtain φn(y0, s) = ∂f∂y (yn, s) for

some yn ∈ (y0, y0 + δ), and hence |φn(y0, s)| ≤ g(s) for all s ∈ S. It follows that

F ′(y0) = limn→∞

F (y0 + hn)− F (y0)

hn= limn→∞

∫S

φn(y0, s) dµ(s)(DCT)

=

∫S

∂f

∂y(y0, s) dµ(s).

,

Exercises

Exercise 6.1. Use convergence theorems to find limn→∞

∫Rfn dλ in each of the following cases:

(a) fn(x) = 1[4,32](x)(1 + log(4x)

n

)n(use MCT; answer is 2016).

46See [11, Theorem 6.2.3]: if g : [a, b]→ R is continuous on [a, b] and differentiable on (a, b), then there exists a

point c ∈ (a, b) such that g′(c) =g(b)−g(a)b−a .

Page 29: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 29

(b) fn(x) = 1(1,∞)(x)sin(nx)

nx2(use DCT; answer is 0).

(c) fn(x) = 1[0,1](x)nxn sin(nx)− n√

x+ 2n2(use DCT; answer is − 1

2

√2).

Exercise 6.2. Assume f, f1, f2, . . . : S → R are measurable functions such that

(i) There is a constant M ≥ 0 such that for all n ∈ N,∫S|fn|dµ ≤M

(ii) fn → f pointwise.

Use Fatou’s lemma to show that∫S|f |dµ ≤M .

Exercise 6.3. Find limn→∞

∞∑m=1

1

m2 + n2.

Hint: Use the counting measure and the DCT.

Exercise 6.4. Deduce Corollary 6.8 from the MCT applied to sn =∑nj=1 fj for n ≥ 1.

Exercise∗ 6.5. Give the details of Remark 6.6 (use Exercise 5.5).

Exercise 6.6. Let (fn)n≥1 be a sequence of measurable functions with values in R or C andfn → f pointwise. Assume there exists an integrable function g : S → [0,∞) such that |fn| ≤ g.

(a) Show that∫S|fn − f |dµ→ 0 as n→∞.

Hint: Apply the DCT in a suitable way.(b) In the real case we know

∫Sfn dµ→

∫Sf dµ. Derive this in the complex case as well.

Hint: Use (a) and Exercise 5.11 (c).

Exercise∗ 6.7. Let f : S → R be a measurable function. Define ν : B(R) → [0,∞] by ν(B) =µ(f−1(B)). Prove the following:

(a) ν is a measure.(b) For every measurable g : R→ [0,∞] one has∫

Rg(t) dν(t) =

∫S

g(f(s)) dµ(s).

Hint: Argue as in Theorem 6.9 step 2.(c) A function g : R → R is integrable with respect to ν if and only if g ◦ f is integrable with

respect to µ. Moreover, in this case∫Rg(t) dν(t) =

∫S

g(f(s)) dµ(s).

Hint: Argue as in Theorem 6.9 step 3.

Exercise∗ 6.8. Assume f : R→ R is integrable with respect to λ.

(a) Show that for each y ∈ R the function x 7→ sin(xy)f(x) is integrable with respect to λ.

Define g : R→ R by

g(y) =

∫R

sin(xy)f(x) dλ(x)

(b) Show that g is continuous.Hint: Use the sequential characterization of continuity and one of the convergence theorems.

Exercise∗∗ 6.9. Use induction and Theorem 6.14 to show that for each integer n ≥ 0,∫ ∞0

xne−yx dx =n!

yn+1, y > 0.

In particular, setting y = 1 one obtains:∫∞0xne−x dx = n!.

Exercise∗∗ 6.10. Let f : [0,∞)→ R be given by f(x) = sin(x)x if x 6= 0 and f(0) = 1.

(a) Show that f is not integrable with respect to λ.Hint: Use the estimate | sin(x)| ≥ 1

2 on [πn+ 13π, πn+ 2

3π] for all integers n ≥ 0.

(b) Show that∫∞0f(x) dx exists as an improper Riemann integral.47

47Using some smart tricks one could actually show that the integral equals π2

.

Page 30: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

30 MEASURE AND INTEGRATION

7. Lp-spaces

In this section (S,A, µ) is measure space. In this section we want to allow the scalar field to becomplex as well and we use the notation K for this. So K = R or K = C.

Definition 7.1. For p ∈ [1,∞) let

Lp(S) ={f : S → K : f is measurable and

∫S

|f |p dµ <∞}.

For f ∈ Lp(S) let

‖f‖p :=(∫

S

|f |p dµ) 1p

.

Note that L1(S) coincides with the set of integrable functions f : S → K.If ‖f − g‖p = 0, then from Proposition 5.8 (vi) we can conclude f = g a.e. However, we would

like to have f = g. Therefore, we will identify f and g whenever f = g a.e.48 So one has to berather careful if one talks about f(s) for a certain fixed s ∈ S. In integration theory this usuallydoes not lead to any problems since f = g a.e. implies that

∫Ef dµ =

∫Eg dµ for any E ∈ A.

Example 7.2. Let S = R with the Lebesgue measure λ. Then 1{0} = 1Q = 0 and 1[0,1]\Q = 1[0,1]

in Lp(R).

7.1. Minkowski and Holder’s inequalities. Note that for scalars α ∈ K, ‖αf‖p = |α|‖f‖p.Therefore, the next result can be used to show that Lp(S) is a normed vector space.

Proposition 7.3 (Minkowski’s inequality 49). For all f, g ∈ Lp(S) we have f + g ∈ Lp(S) and

‖f + g‖p ≤ ‖f‖p + ‖g‖p.Proof. Observe that for all a, b ∈ [0,∞) one has (See Exercise 7.1)

(7.1) (a+ b)p = infθ∈(0,1)

θ1−pap + (1− θ)1−pbp.

It follows from (7.1) that for all θ ∈ (0, 1) and all s ∈ S,

|f(s) + g(s)|p ≤ (|f(s)|+ |g(s)|)p ≤ θ1−p|f(s)|p + (1− θ)1−p|g(s)|p.Therefore, by monotonicity and linearity of the integral, we obtain that for all θ ∈ (0, 1),∫

S

|f + g|p dµ ≤ θ1−p∫S

|f |p dµ+ (1− θ)1−p∫S

|g|p dµ.

Stated differently, this says that for all θ ∈ (0, 1),

‖f + g‖pp ≤ θ1−p‖f‖pp + (1− θ)1−p‖g‖pp.Now the result follows by taking the infimum over all θ ∈ (0, 1) and applying (7.1). ,

Another famous inequality which can be proved with the same method is the following.

Proposition 7.4 (Holder’s inequality50). Let p, q ∈ (1,∞) satisfy51 1p + 1

q = 1. If f ∈ Lp(S) and

g ∈ Lq(S), then fg ∈ L1(S) and

‖fg‖1 ≤ ‖f‖p‖g‖q.Proof. See Exercise 7.2. ,

48More precisely, one can build an equivalent relation f ∼ g if f = g almost everywhere and then consider a

quotient space. We will use the above imprecise but more intuitive definition.49Hermann Minkowski (1864-1909) was a German mathematician who worked in geometry. He also was Albert

Einstein’s teacher and provided the 4-dimensional mathematical framework for part of Einstein’s relativity theory.50Otto Holder (1859-1937) is most famous for this result and for the notion of Holder continuity of a function.51These exponents are called conjugate exponents. If p = 2, then q = 2 and in this case the inequality is known

as the Cauchy-Schwarz inequality.

Page 31: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 31

7.2. Completeness of Lp.

Theorem 7.5 (Riesz–Fischer52). Let p ∈ [1,∞). Then (Lp(S), ‖ · ‖p) is a Banach space.53

Proof. Let (fk)∞k=1 be a Cauchy sequence with respect to the norm ‖ · ‖p of Lp(S). By a standardargument it suffices to show that (fk)∞k=1 has a subsequence which is convergent in Lp(S).

Recursively, we can find a subsequence (fkn)n≥1 such that

(7.2) ‖fkn+1− fkn‖p ≤

1

2n, n = 1, 2, . . .

For notational convenience let φn = fkn for n ∈ N. Also let φ0 = 0. We will show that there existsan f ∈ Lp(S) such that ‖φn − f‖p → 0.

Define g, g1, g2, . . . : S → [0,∞] by

g :=

∞∑n=0

|φn+1 − φn| and gm :=

m−1∑n=0

|φn+1 − φn|, ,

By Minkowski’s inequality we obtain for each m ≥ 1,

‖gm‖p ≤m−1∑n=0

‖φn+1 − φn‖p =

∞∑n=0

‖φn+1 − φn‖p(7.2)

≤ ‖φ1‖p +∑n≥1

2−n ≤ ‖φ1‖p + 1.

Since 0 ≤ gm ↑ g, the MCT yields,∫S

|g|p dµ = limm→∞

∫S

|gm|p dµ = limm→∞

‖gm‖pp ≤ (‖φ1‖p + 1)p.

Letting A = {g <∞}, Exercise 5.1 yields µ(Ac) = 0. Therefore, we can define f : S → R by

f =

∞∑n=0

1A(φn+1 − φn),

where the series is absolutely convergent and f is measurable by Theorem 4.9. By a telescopingargument it follows that pointwise on S

f = limm→∞

m−1∑n=0

1A(φn+1 − φn) = limm→∞

1Aφm.

Clearly,

|φm| =∣∣∣m−1∑n=0

(φn+1 − φn)∣∣∣ ≤ m−1∑

n=0

|φn+1 − φn| ≤ |g|

and by letting m→∞ we see that also |f | ≤ |g| and in particular f ∈ Lp(S). It follows that

|f − φm|p ≤ (|f |+ |φm|)p ≤ (2|g|)p = 2p|g|p.Since |f − φm|p → 0 a.e. and 2p|g|p is integrable, it follows from the DCT (see the a.e. version ofRemark 6.6) that

limm→∞

‖f − φm‖pp = limm→∞

∫S

|f − φm|p dµ = 0.

,

In the sequel we will say that fn → f in Lp(S) if ‖fn − f‖p → 0.From the proof of Theorem 7.5 we deduce the following result.

Corollary 7.6. Let p ∈ [1,∞). Suppose f, f1, f2, . . . ∈ Lp(S). If fk → f in Lp(S), then thereexists a subsequence (fkn)n≥1 such that fkn → f a.e.

52Frigyes Riesz (1880–1956) was a Hungarian mathematician who worked in functional analysis. Ernst Fischer

(1875–1954) was a Austrian mathematician who worked in analysis.53Recall that a Banach space is a complete normed vector space. Stefan Banach (1892–1945) was a Polish

mathematician who is one of the world’s most important 20th-century mathematicians. He is most famous for hisbook on functional analysis [2].

Page 32: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

32 MEASURE AND INTEGRATION

In general one does not have fn → f a.e. (see Exercise 7.7).

Proof. Indeed, since (fn)n≥1 is convergent it is a Cauchy sequence in Lp(S). In the proof of

Theorem 7.5 there exists a subsequence (fkn)n≥1 and a function f such that fkn → f in Lp(S)

and fkn → f a.e. Now Minkowski’s inequality yields:

‖f − f‖p ≤ ‖f − fkn‖p + ‖fkn − f‖p → 0.

Thus f = f a.e. and thus fkn → f a.e. ,

For f, g ∈ L2(S) we define

〈f, g〉 =

∫S

f(s)g(s) dµ(s),

where g(s) stands for the complex conjugate of the number g(s) ∈ K. Then 〈·, ·〉 is an innerproduct and 〈f, f〉 = ‖f‖22.54

The next theorem is immediate from Theorem 7.5.

Theorem 7.7 (Riesz–Fischer). L2(S) is a Hilbert space.55

Example 7.8. Let p ∈ [1,∞). If µ = τ is the counting measure on N, then56 `p := Lp(N) coincideswith a space of sequences and

‖(an)n≥1‖p =(∑n≥1

|an|p) 1p

.

If p ≤ q <∞, then `p ⊆ `q (see Exercise 7.8).

Example 7.9. Assume µ(S) <∞. If 1 ≤ p ≤ q <∞, then Lq(S) ⊆ Lp(S) (see Exercise 7.3).

We end this section with a simple density result. Clearly, the definition of a simple functionextends to the complex setting. The following result shows that any Lp-function can be approxi-mated by simple functions.

Proposition 7.10 (Density of simple functions). Let p ∈ [1,∞). The set of simple functions isdense in Lp(S).

Proof. Write f = u + iv, where u and v are both real-valued. Write u = g − h, where g = u+

and h = u−. By Theorem 4.12 we can find simple functions gn and hn such that 0 ≤ gn ↑ gand 0 ≤ hn ↑ h. Then un := gn − hn is a simple function, un → u pointwise, and |un| ≤gn + hn ≤ g + h ≤ |u|. Similarly, one constructs simple functions vn such that vn → v pointwise,|vn| ≤ |v|. We can conclude that fn := un + ivn is a simple function, fn → f pointwise and

|fn| ≤ (|un|2 + |vn|2)12 ≤ (|u|2 + |v|2)

12 ≤ |f |. Now since |fn − f |p ≤ (|fn|+ |f |)p ≤ 2p|f |p and the

latter is integrable, it follows from the DCT that ‖fn − f‖p → 0 as n→∞. ,

7.3. Lp-spaces on intervals. In the remaining part of this section we discuss Lp(I), where I isan interval and µ = λ is the Lebesgue measure on I. These results will not be used in Section 8or in the exercises.

For an interval I ⊆ R let the set of step functions Step(I) be defined by

Step(I) = span{1J : J ⊆ I is an interval with finite length}.One can check that each φ ∈ Step(I) is a simple function. The converse does not hold. For

instance 1Q∩(0,1) is not a step function. Step function have a lot of structure and often questionson Lp can be reduced to this special class by the following density result.

54Note that the conjugation is need otherwise the square of a complex number can be negative. Physicists often

put the conjugation on f(s) instead of g(s). If K = R, then the conjugation does not play a role.55Recall that a Hilbert space is a Banach space where ‖x‖ = 〈x, x〉

12 . David Hilbert (1862–1943) is sometimes

said to be the last universal mathematician (which means he knew “all” mathematics of his time). He was one of

the most influential mathematicians of the 19th and early 20th centuries.56Sometimes we wish to use the counting measure on Z instead of N. In this case we write `p(Z) := Lp(Z).

Page 33: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 33

1 2 3 4 5

1

2

Figure 7.1. Example step function 141

(− 12 ,1]

+ 321

(1,52 ]− 1

31(52 ,4)

+ 1[4,5]

Theorem 7.11. Let λ be the Lebesgue measure on I = (a, b) with −∞ ≤ a < b ≤ ∞ and letp ∈ [1,∞). Then Step(I) is dense in Lp(I).

Proof. Let f ∈ Lp(I) and let ε > 0. We will construct a step function φ such that ‖f − φ‖p < 3ε.Step 1: Reduction to bounded I. Let fn = 1(−n,n)f for n ∈ N. Then fn → f pointwise.

Moreover, |fn − f |p ≤ |f |p. Therefore, the DCT yields ‖fn − f‖p → 0. Therefore, for n ∈ N largeenough ‖f − g‖p < ε, where g = fn ∈ Lp(I).

Step 2: By Step 1 we can assume I is bounded, and moreover we can assume I = (a, b] with−∞ < a < b < ∞. By Proposition 7.10 there exists a simple function h : I → K such that‖g − h‖p < ε. We can write h =

∑nj=1 1Ajxj with (Aj)

nj=1 in B(I) disjoint.

Step 3: By Exercise 7.5 there exist Fj ∈ F(a,b] such that λ(Aj4Fj) < εn(|xj |+1) . Now let

φ =∑nj=1 1Fjxj . Observe that |1Aj − 1Fj | = 1Aj4Fj . Therefore, Minkowski’s inequality yields

that

‖h− φ‖p ≤n∑j=1

|xj |‖1Aj4Fj‖p =

n∑j=1

|xj |µ(Aj4Fj) ≤n∑j=1

ε|xj |n(|xj |+ 1)

< ε.

Conclusion: Clearly, φ is a step function. Moreover, by Minkowski’s inequality we find

‖f − φ‖p ≤ ‖f − g‖p + ‖g − h‖p + ‖h− φ‖p ≤ 3ε.

,

Corollary 7.12. Let λ be the Lebesgue measure on I = [a, b] with −∞ < a < b < ∞ and letp ∈ [1,∞). Then C([a, b]) is dense in Lp(I).

Proof. Let f ∈ Lp(I) and let ε > 0. By Theorem 7.11 there exists a step function φ such that‖f − φ‖p < ε. It remains find ψ ∈ C([a, b]) such that ‖φ − ψ‖p < ε. For this it suffices toapproximate an arbitrary 1J . Using Figure 7.2 and the DCT, the reader can easily convince himor herself that this can indeed be done. ,

1 2 3 4

1

2

1 2 3 4

1

2

1 2 3 4

1

2

Figure 7.2. Approximation of 1(1, 52 ]in Lp by continuous functions

Exercises

Page 34: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

34 MEASURE AND INTEGRATION

Exercise 7.1. Prove the identity (7.1).Hint: First consider the case where a, b > 0 and minimize the right-hand side of (7.1).

Exercise∗ 7.2. Let p, q ∈ (1,∞) be such that 1p + 1

q = 1.

(a) Show that for all a, b ≥ 0,

ab = inft>0

( tpapp

+bq

qtq

).

(b) Use the above identity to prove Proposition 7.4.Hint: Argue as in Proposition 7.3, but use the identity from (a) instead.

Exercise 7.3. Assume µ(S) < ∞ and 1 ≤ q ≤ p < ∞. Show that Lp(S) ⊆ Lq(S) and for allf ∈ Lp(S),

‖f‖q ≤ µ(S)1q−

1p ‖f‖p.

Hint: Apply Holder’s inequality to |f |q · 1.

Exercise 7.4. Let p ∈ [1,∞) and λ be the Lebesgue measure on R. Determine for which α ∈ Rone has f ∈ Lp(R) in each of the following cases:

(a) f(x) = 1(0,1)(x)xα.(b) f(x) = 1(1,∞)(x)xα.

Explain why Lp(R) * Lq(R) for all p, q ∈ [1,∞) with p 6= q.

For sets A,B ⊆ R, A4B = (A \B)∪ (B \A) denotes the symmetric difference of A and B.

Exercise∗ 7.5. Let λ be the Lebesgue measure restricted to ((a, b],B((a, b])). Let F(a,b] be the

finite unions of half-open intervals in (a, b].57 Let

A = {A ∈ B((a, b]) : ∀ε > 0 ∃F ∈ F(a,b] such that λ(A4F ) < ε}.(a) For A,B ∈ B((a, b]) show that A4B = Ac4Bc.(b) Let I be an index set and Ai, Bi ⊆ B((a, b]) for all i ∈ I. Let A =

⋃i∈I

Ai and B =⋃i∈I

Bi.

Show that A4B ⊆ ⋃i∈I

Ai4Bi.(c) Deduce from (a) that A ∈ A implies Ac ∈ A.

(d) Assume (An)n≥1 is a disjoint sequence in A. Deduce from (b) that∞⋃n=1

An ∈ A.

Hint: Use that limn→∞ λ(∞⋃k=n

Ak) = 0 in order to reduce to finitely many sets.

(e) Show that A = B((a, b]).Hint: Use Exercises 3.1 and 3.8.

Exercise∗ 7.6. Let −∞ < a < b < ∞ and f ∈ L1((a, b]) and assume∫(a,t]

f dλ = 0 for all

t ∈ (a, b]. We will derive that f = 0 in L1(R). By considering real and imaginary part separately,one can reduce to the case where f is real valued.

(a) Show that for all A ∈ F1 with A ⊆ (a, b],∫Af dλ = 0.

(b) Let A ∈ B(R) be such that A ⊆ (a, b]. Construct sets A1, A2, . . . ∈ F1 such that Ak ⊆ (−a, b]and 1An → 1A a.e.Hint: Use ‖1A−1B‖1 = λ(A4B) for all A,B ∈ B(R) and Exercise 7.5. Now apply Corollary7.6 and the DCT.

(c) Show that for all A ∈ B(R) with A ⊆ (a, b],∫Af dλ = 0.

Hint: Use (a) and (b).(d) Derive that f = 0 in L1(R).

Hint: Consider A = {x ∈ R : f(x) ≥ 0} and A = {x ∈ R : f(x) ≤ 0}.

57By Exercise 1.7 of the lecture notes we can take them disjoint

Page 35: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 35

Exercise∗∗ 7.7. Let λ be the Lebesgue measure on R. Observe that each n ∈ N can be uniquelywritten as n = 2k + j with k ∈ N0 and j ∈ {0, . . . , 2k − 1}. Now for such n define fn =1(j2−k,(j+1)2−k]. Show that fn → 0 in L1(R), but for all x ∈ (0, 1], there exists infinitely manyn ∈ N such that fn(x) = 1.Hint: First make a picture for k = 1 and j = 0, 1, and k = 2 and j = 0, 1, 2, 3.

Exercise∗∗ 7.8. Let 1 ≤ p ≤ q <∞).

(a) Prove `p ⊆ `q and that for all (an)n≥1 ∈ `p one has ‖(an)n≥1‖q ≤ ‖(an)n≥1‖p.Hint: By homogeneity one can assume ‖(an)n≥1‖`p = 1, and therefore |an| ≤ 1 for all n ∈ N.

(b) Let an = nα for n ∈ N. For which α ∈ R does one have (an)n≥1 ∈ `p?There is a natural limiting space of Lp(S) for p→∞:

Exercise∗∗ 7.9. A measurable function f : S → K is said to be in L∞(S) if there exists an M ≥ 0such that µ({|f | > M}) = 0.58 Define

‖f‖∞ = inf{M ≥ 0 : µ({|f | > M}) = 0}.As usual we identify functions f and g in L∞(S) if f = g a.e.

(a) Show that (L∞(S), ‖ · ‖∞) is a Banach space.(b) Show that for all f ∈ L∞(S) and g ∈ L1(S), fg ∈ L1(S) and

‖fg‖1 ≤ ‖f‖∞‖g‖1.(c) Assume µ(S) <∞ and p ∈ [1,∞). Show that L∞(S) ⊆ Lp(S) and ‖f‖p ≤ µ(S)

1p ‖f‖∞.

(d) Assume S = N with the counting measure and p ∈ [1,∞). Let `∞ := L∞(N). Show that`p ⊆ `∞ and ‖(an)n≥1‖∞ ≤ ‖(an)n≥1‖p.

(e) Assume I is a finite interval and µ = λ is the Lebesgue measure. Show that the simplefunctions are dense in L∞(I), but the step functions are not.

58In order words |f | ≤M a.e.

Page 36: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

36 MEASURE AND INTEGRATION

8. Applications to Fourier series

Fourier59 analysis plays a role in a large part of mathematics. In particular, it is one of thecentral mathematical tools in Physics and Electrical Engineering. A Fourier series is of the form60∑

n∈Zcne

inx, or equivalentlya02

+∑n≥1

an cos(nx) + bn sin(nx),

where x ∈ [0, 2π]. Of course one can also use x ∈ R here. Clearly, the above functions will beperiodic whenever they are well-defined.

One of the reasons that Fourier series naturally arise in mathematics is that each of the functions

e±inx, cos(nx), sin(nx) as an eigenfunctions of d2

dx2 with eigenvalue −n2. Indeed, for instance

cos(nx)′′ = −n2 cos(nx).We have seen that Taylor series can be used to represent functions which are smooth enough.61

Fourier series provides another tool to represent functions. The class of functions which can berepresented as a Fourier series will turn out to be enormous.

In this section we will prove a couple of central results in the theory of Fourier series. Theinterested reader can read more on the subject in [9], [10], [13], [14] and [19]. In particular, veryinteresting but mostly elementary applications to geometry, ergodicity, number theory and PDEscan be found in [13].

8.1. Fourier coefficients. In this section S = [0, 2π] and λ is the Lebesgue measure on (0, 2π).For notational convenience we will write∫ b

a

f(x) dx :=

∫[a,b]

f dλ.

for f ∈ L1(0, 2π) and [a, b] ⊆ [0, 2π].

Definition 8.1. Let ek : [0, 2π]→ C be given by ek(x) = eikx for k ∈ Z.62 Let f ∈ L1(0, 2π).

(i) For k ∈ Z the k-th order Fourier coefficient is defined by63

f(k) =1

∫ 2π

0

f(x)ek(x) dx.

(ii) for n ∈ N0, the n-th partial sum of the Fourier series sn(f) : [0, 2π]→ C is defined by

(8.1) sn(f) =∑|k|≤n

f(k)ek.

(iii) A function of the form∑|k≤n ckek with n ∈ N0 and (ck)|k|≤n in C is called a trigonometric

polynomial.

Remark 8.2.

(1) The reason to use the notion “trigonometric polynomial” is that ek(x) = (eix)k. Also notethat ek(x) = cos(kx) + i sin(kx) by Euler’s formula.

(2) Observe that by (the complex version of) Proposition 5.12 for each k ∈ Z,

(8.2) |f(k)| ≤ 1

∫ 2π

0

|f(x)ek(x)|dx =1

2π‖f‖1.

In Exercise 8.4 it will be shown that one even has limk→∞ f(k) = 0.

(3) If f, g ∈ L1(0, 2π), the following linearity property holds: (f + g)(k) = f(k) + g(k) for allk ∈ Z. This follows directly from the linearity of the integral.

59Fourier analysis is named after the French mathematician Jean-Baptiste Fourier (1768–1830), and were intro-duced in order to solve differential equations such as the heat equation

60Recall Euler’s formula: eix = cos(x) + i sin(x) for x ∈ R61In Complex Function Theory these functions will be characterized as the so-called analytic functions62Note that e0 = 1[0,2π].63To calculate Fourier transforms numerically one can use the so-called fast Fourier transform FFT (see [13,

Section 7.1.3])

Page 37: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 37

Example 8.3. Given f ∈ L1(0, 2π), the function sn(f) is a trigonometric polynomial for eachn ∈ N. Each of the functions

cos(nx) =einx + e−inx

2, sin(nx) =

einx − e−inx2i

is a trigonometric polynomial as well.Finally note that if P is a trigonometric polynomial, then for any j ∈ N0, P j is a trigonometric

polynomial as well.

The main questions in this section is whether we can reconstruct f from its Fourier coefficients.More precisely:

• (representation) Which functions f : [0, 2π] → C can we write as f =∑k∈Z ckek for

certain coefficients (ck)k∈Z?• (convergence) In what sense does the above series converge?

• (uniqueness) Does f(k) = g(k) for all k ∈ Z imply f = g ?

When considering convergence of Fourier series we will always consider the convergence of

∑|k|≤n

ckek =

n∑k=−n

ckek as n→∞.

8.2. Weierstrass’ approximation result and uniqueness. Before we consider convergence ofFourier series, we first we prove a fundamental result about the approximation by trigonometricpolynomials. It will be an essential ingredient in the uniqueness result in Theorem 8.5.

Theorem 8.4 (Weierstrass’ approximation theorem64 for periodic functions). The trigonometricpolynomials are dense in {f ∈ C([0, 2π]) : f(0) = f(2π)}.

Proof. Let f ∈ C([0, 2π]) be such that f(0) = f(2π) and let ε > 0 be arbitrary. It suffices to showthat there exists a trigonometric polynomial P such that ‖f −P‖∞ < ε. We extend f periodicallyto a function f : R→ C. Since f is also uniformly continuous we can choose δ ∈ (0, π) such that|x− y| < δ implies |f(x)− f(y)| < ε/2.

Define65 Fn = 1n

∑nk=1

∑|j|≤k−1 ej which is a periodic function as well. Define Pn : [0, 2π]→ C

by Pn(x) =∫ 2π

0Fn(x− y)f(y) dy. Then since ej(x− y) = e2πixe−2πiy the following identity holds

Pn(x) =1

∫ 2π

0

Fn(x−y)f(y) dz =1

2πn

n∑k=1

∑|j|≤k−1

ej(x)

∫ 2π

0

ej(−y)f(y) dy =1

n

n∑k=1

∑|j|≤k−1

ej(x)f(j).

This shows that Pn is a trigonometric polynomial.By Exercise 8.3 the following identity holds

Fn(z) =sin2(nz/2)

n sin2(z/2), z ∈ (0, 2π).(8.3)

Therefore, Fn(z) ≥ 0, and thus we can write

‖Fn‖1 =

∫ 2π

0

Fn(z) dz =1

n

n∑k=1

∑|j|≤k−1

∫ 2π

0

ej(z) dz =1

n

n∑k=1

2π = 2π.

64Originally Weierstrass proved that the polynomials are dense in C([a, b]). This can be derived from our version

of the theorem as indicated in Exercise 8.1165This is called Fejer’s kernel

Page 38: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

38 MEASURE AND INTEGRATION

Fix x ∈ [0, 2π]. It follows that

2π|f(x)− Pn(x)| =∣∣∣f(x)

∫ 2π

0

Fn(x− y) dy −∫ 2π

0

f(y)Fn(x− y) dy∣∣∣

=∣∣∣ ∫ 2π

0

Fn(x− y)(f(x)− f(y)) dy∣∣∣

≤∫ 2π

0

|f(x)− f(y)|Fn(x− y) dy∣∣∣ ≤ T1 + T2,

where T1 and T2 are the integrals over I := [x−δ, x+δ] and J := [0, x−δ]∪ [x+δ, 2π], respectively.On the interval I we can use the uniform continuity of f to estimate

T1 ≤ε

2

∫I

Fn(x− y) dy ≤ πε.

On J := [0, x− δ] ∪ [x+ δ, 2π] we can estimate

T2 ≤ 2‖f‖∞∫J

Fn(x− y) dy = 2‖f‖∞∫Bδ

Fn(z) dz,

where we substituted z := x− y and where Bδ = [δ, 2π − δ]. Therefore, using (8.3) again and thefact that | sin(z/2)| ≥ sin(δ/2) for z ∈ Bδ (recall that δ ≤ π) we obtain∫

Fn(z) dz ≤ |Bδ|n sin2(δ/2)

≤ 2π

n sin2(δ/2).

So choosing n ≥ 1 so large that 2‖f‖∞n sin2(δ/2)

< ε2 we obtain T2 < πε.

Therefore, combining the the estimates can conclude that |f(x) − Pn(x)| ≤ T1+T2

2π < ε. Sincex ∈ [0, 2π] was arbitrary it follows that ‖f − Pn‖∞ < ε as required. ,

Now we can deal with the uniqueness question for Fourier series. This is the most technicalpart of this section and could be skipped it at first reading.

Theorem 8.5 (Uniqueness). If f ∈ L1(0, 2π) satisfies f(n) = 0 for all n ∈ Z, then f = 0 inL1(0, 2π).

Proof. 66 Step 1: First assume f ∈ C([0, 2π]) and f(0) = f(2π). By linearity it follows that foreach trigonometric polynomial P we have∫ 2π

0

f(x)P (x) dx = 0.

By Theorem 8.4 we can find trigonometric polynomials (Pn) such that Pn → f uniformly. There-fore, it follows that ∫ 2π

0

|f(x)|2 dx = limn→∞

∫ 2π

0

f(x)Pn(x) dx = 0.

This implies f = 0.

Step 2: Next let f ∈ L1(0, 2π) and assume f(n) = 0 for all n ∈ Z. Let F : [0, 2π] → R bedefined by

F (t) =

∫ t

0

f(x) dx.

Then F ∈ C([0, 2π]) (see Example 6.7), and F (0) = F (2π) = 0.67 By Exercise 8.6 (b) F (k) =f(k)ik = 0 for all k 6= 0. Now let g = F − C, where C = F (0). Then g ∈ C([0, 2π]), g(0) = g(2π)

and g(k) = 0 for all k ∈ Z. Therefore, g = 0 by step 1 and hence F = C. Since F (0) = 0, thisyields F (x) = 0 for all x ∈ [0, 2π]. By Exercise 7.6 we find f = 0.

,

66There is a much better proof in the literature using the Fejer kernel as an approximate identity. Since

approximate identities are not part of these lecture notes, we proceed differently.67Note that F (2π) = (2π)f(0) = 0.

Page 39: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 39

8.3. Fourier series in L2(0, 2π). In this section we will consider Fourier series in the Hilbertspace L2(0, 2π). Here the inner product is given by

〈f, g〉 =

∫ 2π

0

f(x)g(x) dx.

Note that L2(0, 2π) ⊆ L1(0, 2π) (see Exercise 7.3). Therefore, if f ∈ L2(0, 2π) the Fourier coeffi-

cients are well-defined and f(k) = (2π)−1〈f, ek〉.Let us recall as special case of Proposition 7.3 for f, g ∈ L2(0, 2π),

(8.4) |〈f, g〉| ≤ ‖f‖2‖g‖2 (Cauchy–Schwarz inequality).

We will say that f, g ∈ L2(0, 2π) are orthogonal if 〈f, g〉 = 0. Note that in this case the followingform of Pythagoras theorem holds68

(8.5) ‖f + g‖22 = ‖f‖22 + ‖g‖22.Lemma 8.6 (Orthogonality). For j, k ∈ Z,

〈ej , ek〉 =

{2π, if j = k;0, if j 6= k.

Consequently, if finitely many (cj)j∈Z in C are nonzero, then

(8.6)∥∥∥∑j∈Z

cjej

∥∥∥2

= (2π)12

(∑j∈Z|cj |2

) 12

.

Proof. Indeed, if j 6= k, then using ek(x) = e−ikx we find that

〈ej , ek〉 =

∫ 2π

0

ei(j−k)x dx =

∫ 2π

0

cos((j − k)x) dx+ i

∫ 2π

0

sin((j − k)x) dx = 0

by periodicity of cos and sin. Similarly, one sees 〈ej , ej〉 = 2π.The final statement follows from∥∥∥∑

j∈Zcjej

∥∥∥22

=∑j∈Z

∑k∈Z

cjck〈ej , ek〉 = 2π∑j∈Z|cj |2.

,

We extend this result to series using the completeness of L2(0, 2π).

Theorem 8.7 (Riesz–Fischer, Convergence of Fourier series in L2).

(i) If (cn)n∈Z ∈ `2, then g :=∑n∈Z

cnen converges in L2(0, 2π), and g(n) = cn for all n ∈ Z, and

(8.7) ‖g‖2 = (2π)12 ‖(cn)n∈Z‖`2 (Parseval’s identity)

(ii) If f ∈ L2(0, 2π), then (f(n))n∈Z in `2 and f =∑n∈Z

f(n)en in L2(0, 2π) and (8.7) holds with

g = f and cn = f(n) for n ∈ Z.

Part (ii) shows that every L2-function can be represented as a Fourier series. A similar resultholds for series of sine and cosine functions and can be derived as a consequence of the aboveresult (see Exercise 8.8).

Proof. (i): Let gn =∑|k|≤n ckek for n ∈ N. We show that (gn)n≥1 is a Cauchy sequence. Let

ε > 0 and choose N ∈ N such that ( ∑|k|≥N

|ck|2) 1

2

(2π)12

.

68This follows by writing out ‖f + g‖22 = 〈f + g, f + g〉.

Page 40: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

40 MEASURE AND INTEGRATION

Then for all integers n > m ≥ N by Lemma 8.6,

‖gn − gm‖2 =∥∥∥ ∑m<|k|≤n

ckek

∥∥∥2

= (2π)12

( ∑m<|k|≤n

|ck|2) 1

2 ≤ (2π)12

( ∑|k|≥N

|ck|2) 1

2

< ε.

This proves that (gn)n≥1 is a Cauchy sequence. By completeness (see Theorem 7.7), g :=limn→∞ gn exists in L2(0, 2π).

To check (8.7) note that by the continuity of ‖ · ‖2 and Lemma 8.6,

‖g‖2 = limn→∞

‖gn‖2 = limn→∞

(2π)12

( ∑|k|≤n

|ck|2) 1

2

= (2π)12 ‖(cn)‖`2 .

Finally, note that g(k) = (2π)−1〈g, ek〉 = limn→∞(2π)−1〈gn, ek〉 = ck.69

(ii): Fix n ∈ N. Since 〈f − sn(f), ek〉 = 0 for each |k| ≤ n, also 〈f − sn(f), sn(f)〉 = 0 andhence (8.5) yields

(8.8) ‖f‖22 = ‖f − sn(f) + sn(f)‖22 = ‖f − sn(f)‖22 + ‖sn(f)‖22 ≥ ‖sn(f)‖2.Let ck = f(k) for k ∈ Z. Then (8.8) yields:

‖f‖22 ≥ ‖sn(f)‖2(8.6)=

∑|k|≤n

2π|ck|2.

Letting n→∞, we find ‖(ck)k∈Z‖`2 ≤ (2π)−1‖f‖2 <∞.By (i) we can define g =

∑n∈Z cnen where the series converges in L2(0, 2π). We claim that

f = g in L2(0, 2π). To see this note that by (i), g(n) = cn = f(n). Therefore, the claim followsfrom the uniqueness Theorem 8.5 applied to f − g. ,

For f ∈ L2(0, 2π) Theorem 8.7 yields that f − sn(f) =∑|k|>n f(k)ek and by (8.7)

(8.9) (L2-error estimate) ‖f − sn(f)‖22 = 2π∑|k|>n

|f(k)|2.

Moreover, since sn(f) is a trigonometric polynomial, we also find the following:70

Corollary 8.8. The trigonometric polynomials are dense in L2(0, 2π).

Example 8.9 (Sawtooth function). Let f : [0, 2π) → R be defined by f(x) = x − π and extendedperiodically on R. For k ∈ Z \ {0}, by the Fundamental Theorem of Calculus (see [11, Theorem7.3.1]) and integration by parts,

(8.10) f(k) =1

∫ 2π

0

(x− π)e−ikx dx =1

[ (x− π)e−ikx

−ik]2π0− 1

∫ 2π

0

e−ikx dx = − 1

ik.

Clearly, f(0) = 0. Therefore, Theorem 8.7 yield that f = −∑k∈Z\{0}ekik with convergence in

L2(0, 2π). Moreover, using that 2 sin(kx) = ek−e−ki we also find that f = −2

∑∞k=1

sin(k·)k with

convergence in L2(0, 2π) (see Figure 8.1) for plots of the partial sums).The L2-error can be estimated using (8.9):

‖f − sn(f)‖22 = 2π∑|k|>n

|f(k)|2 ≤ 4π∑

k≥n+1

1

k2≤ 4π

∫ ∞n

1

x2dx =

n.

One can show that sn(f) will not converge to f uniformly (also see Figure 8.1). This is in particularclear for x = 0 and x = 2π, because f(0) = π and f(2π) = −π, but sn(f)(0) = sn(f)(2π) = 0.

By applying (8.7) one can obtains a remarkable identity:

‖f‖22 = 2π∑

k∈Z\{0}

1

k2.

69Here we used |〈g − gn, ek〉| ≤ ‖g − gn‖2‖ek‖2 as follows from (8.5).70With more advanced techniques one can show that the trigonometric polynomials are dense in any Lp(0, 2π)

with p ∈ [1,∞).

Page 41: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 41

On the other hand, if we calculate ‖f‖22 with the fundamental theorem of calculus, we obtain

‖f‖22 =

∫ 2π

0

(x− π)2 dx =[13 (x− π)3

]2π0

=2

3π3.

Combining both identities gives∑k∈Z\{0}

1k2 = 1

3π2, and so we find

∑∞k=1

1k2 = 1

6π2.

−10 −5 5 10

−4

−2

2

4

−10 −5 5 10

−4

−2

2

4

−10 −5 5 10

−4

−2

2

4

Figure 8.1. The Fourier series of the sawtooth function with n = 2, n = 5 andn = 10.

8.4. Fourier series in C([0, 2π]). In this section we will give some sufficient condition on f whichimply the Fourier series is uniformly convergent (or equivalently convergent in C([0, 2π]) with thesupremum norm ‖ · ‖∞). Note that fn → f uniformly implies that fn → f in L2(0, 2π). Indeed,this follows from

(8.11) ‖fn − f‖22 ≤∫ 2π

0

|fn(x)− f(x)|2 dx ≤ ‖fn − f‖2∞∫ 2π

0

1 dx = 2π‖fn − f‖2∞.

From the above we see that convergence of Fouier series in C([0, 2π]) is stronger than conver-gence in L2(0, 2π). However, there are example of functions f ∈ C([0, 2π]) with f(0) = f(2π) forwhich the uniform convergence (and even the pointwise convergence) fails (see [1, Example 35.11]and [10, Example 2.5.1]). So apparently more restrictive conditions are needed.

All the different types of convergence can be confusing. Let us summarize some convergenceresults for a sequence (fn)n≥1 in L2(0, 2π).71

uniform conv.

L2-conv. =⇒ L1-conv. =⇒ a.e.-conv. subsequence.

=⇒=⇒

pointwise conv. =⇒ a.e.-conv.

L1 and L2-convergence are only implied by a.e. conv. under additional assumptions on the (fn)n≥1(as given for instance in the DCT).

The following result provides sufficient conditions for uniform convergence.

Theorem 8.10 (Absolute and uniform convergence of Fourier series). Let f ∈ C([0, 2π]). If

(f(k))k∈Z ∈ `1, then f(x) =∑k∈Z f(k)ek(x), x ∈ [0, 2π] where the series is absolutely and

uniformly convergent.

As a consequence we see that f(0) = f(2π) holds in this situation, because ek(0) = ek(2π).

Proof. For all x ∈ [0, 2π], ∑k∈Z|f(k)ek(x)| =

∑k∈Z|f(k)| <∞.

71For completeness we note that a.e.-conv. =⇒ conv. in measure =⇒ conv. in distribution.

Page 42: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

42 MEASURE AND INTEGRATION

Therefore, we can let g =∑k∈Z f(k)ek, where the series is absolutely convergent. Moreover,

‖g − sn(f)‖∞ ≤∑|k|>n

|f(k)| → 0,

and hence g ∈ C([0, 2π]) and sn(f) → g uniformly. By (8.11) the convergence holds in L2(0, 2π)as well, and hence

g(k) = 〈g, ek〉 = limn→∞

〈sn(f), ek〉 = f(k)

and therefore, g = f a.e. by Theorem 8.5. Let A = {s ∈ [0, 2π] : f(s) = g(s)}. Then A is closedand λ(A) = 2π. We claim that A is dense. Indeed, if not then there exists an nonempty openinterval I ⊆ [0, 2π] \A. It follows that 0 < λ(I) ≤ λ([0, 2π] \A) = λ([0, 2π])− λ(A) = 0. This is acontradiction and thus the claim follows. Since A is also closed in [0, 2π], the claim implies thatA = [0, 2π]. ,

From the proof we see that the following error estimate holds:

(8.12) (uniform error estimate) ‖f − sn(f)‖∞≤∑|k|>n

|f(k)|.

The condition of Theorem 8.10 holds in the following situation:

Corollary 8.11. Assume f ∈ L2(0, 2π) satisfies f(0) = 0. Suppose c0 ∈ C and F : [0, 2π]→ K is

given by F (t) = c0 +∫ t0f(x) dx. Then (F (k))k∈Z ∈ `1 and F =

∑k∈Z F (k)ek where the series is

absolutely and uniformly convergent.

Proof. Since f ∈ L2(0, 2π) ⊆ L1(0, 2π), it follows from Example 6.7 that F is continuous on [0, 2π].Moreover, for every t ∈ [0, 2π],

|F (t)| ≤ |c0|+∫ 2π

0

|f(x)|dx ≤ |c0|+ ‖f‖1.

In particular, by monotonicity we see that ‖F‖2 ≤ (2π)12 (|c0|+ ‖f‖1).

By Exercise 8.6, F (k) = f(k)ik for k 6= 0. By Theorem 8.7, ‖f(k))k∈Z‖`2 = ‖f‖2. Therefore, by

the Cauchy–Schwarz inequality (8.4)∑k∈Z|F (k)| = |F (0)|+

∑k∈Z\{0}

|f(k)||k| ≤ |F (0)|+ C‖(f(k))k∈Z‖`2 ≤ |F (0)|+ C

(2π)12

‖f‖2,

where used (8.7) in the last step. Therefore, the absolute and uniform convergence follows fromTheorem 8.10. ,

Example 8.12. If g ∈ C([0, 2π]) satisfies g(0) = g(2π), g is piecewise continuously differentiableon (0, 2π) and g′ ∈ L2(0, 2π), then g =

∑k∈Z g(k)ek where the series is absolutely and uniformly

convergent. Indeed, let F = g− g(0). Then F (t) = −g(0) +∫ t0g′(x) dx, where f := g′ satisfies the

assumptions of Corollary 8.11.

Exercises

Exercise 8.1. Let f : [0, 2π)→ R be given by f = 1[0,π].

(a) Show that f(k) = 0 for even k 6= 0, and f(k) = 1πik for odd k, and f(0) = 1

2 .

(b) Write f as a series of sines as in Example 8.9 and give an estimate of L2-error given by (8.9).(c) Evaluate

∑j∈Z

1(2j+1)2 .

Hint: Argue as in Example 8.9.

Exercise 8.2. Let f : [0, 2π)→ R be given by f(x) = |x−π|. Example 8.12 yields that the Fourierseries of f is absolutely and uniformly convergent. Calculate the Fourier series explicitly and giveestimates for the L2-error (8.9) and uniform error (8.12). One can also obtain the exact form of

another famous series∑k∈Z |f(k)|2 using Parseval and the fundamental theorem of calculus.

Page 43: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 43

Exercise 8.3 (Special Fourier series and kernels). The following kernel’s play a central role inmore advanced theory of Fourier series. Prove the identities below for x ∈ (0, 2π). For each

exercise you should use the geometric sum∑nk=0 a

k = 1−an+1

1−a for a ∈ C \ {1}.

(a) (Dirichlet kernel) Show that Dn(x) :=∑

|k|≤n−1

ek(x) =sin((n− 1

2 )x)

sin( 12x)

for n ≥ 1.

(b) (Fejer kernel) Show that Fn(x) :=1

n

n∑j=1

Dj(x) =1

n

sin2(nx2 )

sin2( 12x)

for n ≥ 1.

Exercise∗ 8.4 (Riemann-Lebesgue lemma).

(a) Show that for any step function f : [0, 2π]→ C (see Section 7.3) one has lim|k|→∞ f(k) = 0.Hint: By linearity it suffices to consider f = 1(a,b), where (a, b) ⊆ [0, 2π].

(b) Show that for any f ∈ L1(0, 2π) one has lim|k|→∞ f(k) = 0.Hint: Use Theorem 7.11 and (a).

Exercise∗ 8.5.

(a) Let (〈H, 〈·, ·〉) be a Hilbert space (over the complex scalars). Prove that for all u, v ∈ H,

(polarization) 4〈u, v〉 = ‖u+ v‖2 − ‖u− v‖2 + i‖u+ iv‖2 − i‖u− iv‖2.(b) Use (a) and (8.7) to prove that for all f, g ∈ L2(0, 2π):∫ 2π

0

f(x)g(x) dx = 2π∑k∈Z

f(k)g(k).

Exercise∗ 8.6. Let g ∈ C1([0, 2π])72 and f ∈ L1(0, 2π). Define F : [0, 2π] → C by F (t) =∫ t0f(x) dx. By Example 6.7, F is continuous.

(a) Prove the following integration by parts formula:∫ 2π

0

f(x)g(x) dx = F (2π)g(2π)− F (0)g(0)−∫ 2π

0

F (x)g′(x) dx.

Hint: For continuous f this is just the standard integration by parts formula. Use Corollary7.12 and approximation to deduce the general case.

(b) Show that f(k) = f(0) + ikF (k) for all k ∈ Z \ {0}.Hint: Apply (a) with g = e−k.

Exercise∗ 8.7. Assume F : [0, 2π] → C is continuously differentiable and satisfies F (0) = F (2π)

and F (0) = 0. Let f = F ′.

(a) Show that f(k) = ikF (k) for all k ∈ Z.Hint: Apply Exercise 8.6 (b).

(b) Show that ‖F‖2 ≤ ‖f‖2 and that equality holds if and only if F = c1e1+c−1e−1 for c1, c−1 ∈ C.Hint: Apply (8.7).

Exercise∗ 8.8. Consider Γ = { 121[0,2π]} ∪ {cos(n·) : n ∈ N} ∪ {sin(n·) : n ∈ N} ⊆ L2(0, 2π).

(a) Show that φ, ψ ∈ Γ with φ 6= ψ are orthogonal and ‖φ‖2 = π.(b) Let f ∈ L1(0, 2π) be such that 〈f, φ〉 = 0 for all φ ∈ Γ. Show that f = 0.

Hint: Use Theorem 8.5(c) Show that for every (an)n≥0, (bn)n≥1 ∈ `2 the following series converges in L2(0, 2π).

g :=a02

+∑n≥1

an cos(n·) +∑n≥1

bn sin(n·)

Hint: Argue as in Theorem 8.7.(d) Show that ‖g‖2L2(0,2π) = π

∑n≥0 |an|2 + π

∑n≥1 |bn|2.

Hint: Argue as in Theorem 8.7.

72That means g is differentiable and its derivative is continuous on [0, 2π]

Page 44: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

44 MEASURE AND INTEGRATION

(e) Show that (an)n≥0 and (bn)n≥1 satisfy

an =1

π

∫ 2π

0

cos(nx)g(x) dx, bn =1

π

∫ 2π

0

sin(nx)g(x) dx.

Hint: Argue as in Theorem 8.7.(f) Show that every f ∈ L2(0, 2π) can be written as

f =a02

+∑n≥1

an cos(n·) + bn sin(n·)

with converges in L2(0, 2π).Hint: Apply Theorem 8.7 or argue as in Theorem 8.7.

It follows from the previous exercise and (8.2) that for C1-functions F one has |F (k)| ≤ ‖f‖12π|k|

for all k ∈ Z \ {0}. Moreover, in Exercise 8.4 we have seen that for general F ∈ L1(0, 2π) one has

F (k) → 0 as |k| → ∞. In the next exercise we show that the convergence can be arbitrary sloweven for periodic functions F ∈ C([0, 2π]).

Exercise∗∗ 8.9. Show that for any sequence (ck)k≥1 with ck 6= 0 and ck → 0 there exists a

function F ∈ C([0, 2π]) with F (0) = F (2π) such that |F (k)| ≥ |ck| for infinitely many k ∈ N.Hint: Choose a subsequence such that

∑∞n=1 |ckn | <∞.

Finally we deduce Weierstrass’ classical approximation result. We first need an elementaryresult about even trigonometric polynomials.

Exercise∗∗ 8.10. Let P =∑nk=−n ckek be a trigonometric polynomial.

(a) Show that there exists a polynomial qn of degree n such that cos(nx) = qn(cos(x)).Hint: Use induction and the recursion formula cos(kx)+cos((k−2)x) = 2 cos((k−1)x) cos(x).

(b) Show that there exists a polynomial qn of degree n such that sin(nx)sin(x) = rn(cos(x)).

Hint: Use induction and the recursion formula sin((k+1)x)+sin((k−1)x) = 2 cos(kx) sin(x).(c) From (a) and (b) derive that there there exit polynomials q, r : [−π, π]→ R of degree n such

that P (x) = q(cos(x)) + r(cos(x)) sin(x).(d) If additionally P is even (i.e. P (−x) = P (x) for x ∈ [−π, π]). Show that there exists a

polynomial q of degree n such that P (x) = q(cos(x)).Hint: Write 2P (x) = P (x) + P (−x) and use (c).

Exercise∗∗ 8.11 (Weierstrass’ approximation theorem for continuous functions). Let f : [−1, 1]→R be continuous and let ε > 0. Let g : [−π, π]→ R given by g(x) = f(cos(x)).

(a) Show that there is an even trigonometric polynomial P such that ‖g − P‖∞ < ε.Hint: First apply Theorem 8.4 on [−π, π] to obtain a trigonometric polynomial P such that

‖g − P‖∞ < ε. Now consider P (x) = P (−x)+P (x)2 .

(b) Use Exercise 8.10 (d) to find a a polynomial q such that ‖f − q‖∞ < ε.(c) The above shows that the polynomials are dense in C([−1, 1]). Use a scaling argument to

show that the polynomials are dense in C([a, b]).

Page 45: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 45

Final remarks:

• One can show that ‖f − Fn ∗ f‖p → 0 for all f ∈ Lp(0, 2π) with p ∈ [1,∞). In Theo-rem 8.4 The convergence holds uniformly if f ∈ C([0, 2π]) satisfies f(0) = f(2π). HereFn ∗ f is the so-called convolution product of Fn and f and is defined by Fn ∗ f(t) =∫ 2π

0Fn(t − x)f(x) dx. Using the definition of Fn one can check that Fn ∗ f is a trigono-

metric polynomial.• Similar results for Dn are true as well as long as p ∈ (1,∞), but this a much deeper

result. Since Dn ∗ f = sn(f), this implies that ‖f − sn(f)‖p → 0 for all f ∈ Lp(0, 2π) withp ∈ (1,∞).

• Finally, we note that sn(f)→ f a.e. for any f ∈ Lp(0, 2π) with p > 1. This is one of thedeepest result in the theory of Fourier series and was proved by Carleson for p = 2 andextended to p > 1 by Hunt in 1968. It was proved a long time before that the result failsfor p = 1 by Kolmogorov in 1923.

For details we refer to the elective Bachelor course on Fourier analysis !

Page 46: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

46 MEASURE AND INTEGRATION

Appendix A. Dynkin’s lemma

The results of this section are not part of the exam material. We will proof the uniquenessresult of Proposition 3.5. In this section S denotes a set.

Definition A.1 (π-system). A collection E ⊆ P(S) is called a π-system if for all A,B ∈ E onehas A ∩B ∈ E.

Example A.2. Every ring is a π-system.

Definition A.3. A collection D ⊆ P(S) is called a Dynkin-system73 if the following conditionshold:

(D1) S ∈ D;(D2) A,B ∈ D and A ⊆ B implies B \A ∈ D;(D3) If (An)n≥1 in D and 74 An ↑ A, then A ∈ D.

Example A.4. Let S = {1, 2, 3, 4} and D = {∅, S, {1, 2}, {3, 4}, {1, 3}, {2, 4}}. Then D is a Dynkinsystem, but it is not a π-system.

Proposition A.5. For a collection F ⊆ P(S) the following are equivalent:

(i) F is a σ-algebra;(ii) F is a Dynkin system and a π-system.

Proof. (i)⇒(ii): This is Exercise A.1.(ii)⇒(i): ∅, S ∈ F follows from (D1) and (D2) of the definition of a Dynkin system. Let

(An)n≥1 be a sequence in F . Let A =∞⋃j=1

Aj and Bn =n⋃j=1

Aj for n ≥ 1. Since F is a π-system it

is closed under finite intersections. Therefore, using (D2) we obtain Bn =( n⋂j=1

Acj

)c∈ F . Since

Bn ↑ A, it follows from (D3) that A ∈ F . ,

Lemma A.6 (Dynkin). Let E ⊆ P(S) be a π-system and D ⊆ P(S) be a Dynkin system. IfE ⊆ D, then σ(E) ⊆ D.

Proof. Let D0 denote the intersection of all Dynkin system which contain E . Observation: E ⊆D0 ⊆ D and D0 is a Dynkin system (see Exercise A.2). We claim that D0 is a π-system as well.As soon as we have proved this claim, Proposition A.5 yields that D0 is a σ-algebra. Therefore,from the observation it follows that σ(E) ⊆ D0 ⊆ D. To prove the claim we need two steps.

Step 1: Define a new collection by

D1 = {D ∈ D0 : D ∩ E ∈ D0 for each E ∈ E}.Since E is a π-system also E ⊆ D1. The collection D1 is a Dynkin-system again. Indeed, S ∈ D1 isclear. If A,B ∈ D1 and B ⊆ A, then for each E ∈ E we find (B \A)∩E = (B∩E)\ (A∩E) ∈ D0,because A ∩ E,B ∩ E ∈ D0 and D0 is a π-system, and thus B \ A ∈ D1. Next let (An)n≥1 in D1

with An ↑ A. Then for each E ∈ E , A∩E =∞⋃n=1

(An ∩E) ∈ D0 since An ∩E ∈ D0. Since D0 ⊆ D1

and D1 is a Dynkin system which contains E , we find D1 = D0.

Step 2: Define a new collection by

D2 = {D ∈ D0 : D ∩ C ∈ D0 for each C ∈ D0}.Since D1 = D0, we find that E ⊆ D2. As before one checks that D2 is a Dynkin system. Moreover,as before this yields D2 = D0. This proves the claim. -

Proposition A.7 (Uniqueness). Let µ1 and µ2 both be measures on measurable space (S,A).Assume the following conditions:

(i) E ⊆ A is a π-system with σ(E) = A;

73Eugene Dynkin 1924–2014 was a Russian mathematician who worked on Algebra and Probability theory.74See Definition 2.9 for the meaning of An ↑ A

Page 47: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 47

(ii) µ1(S) = µ2(S) <∞ and µ1(E) = µ2(E) for all E ∈ E.

Then µ1 = µ2 on A.

Proof. Let D = {A ∈ A : µ1(A) = µ2(A)}. Then E ⊆ D. We claim that D is a Dynkin system.From the claim and Lemma A.6 it follows that A = σ(E) ⊆ D ⊆ A. This implies D = A and therequired result follows from the Definition of D.

To prove the claim note that S ∈ D by assumption. If A,B ∈ D with A ⊆ B, then µ1(B \A) =µ1(B)− µ1(A) = µ2(B)− µ2(A) = µ2(B \A) and hence B \A ∈ D. Finally, if (An)n≥1 in D withAn ↑ A, then by Theorem 2.10, µj(An) ↑ µj(A) for j = 0, 1. Since µ1(An) = µ2(An), this yieldsµ1(A) = µ2(A) and thus A ∈ D. ,

Exercises

Exercise A.1. Prove Proposition A.5 (i) ⇒ (ii).

Exercise∗ A.2. Prove that the intersection of Dynkin systems is again a Dynkin system.

Exercise∗ A.3. Find a version of Proposition A.7 which for measures with µ1(S) = µ2(S) =∞.Hint: See the proof of Theorem 3.10.

Page 48: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

48 MEASURE AND INTEGRATION

Appendix B. Caratheodory’s extension theorem

The results of this section are not part of the exam material (although the statement of theCaratheodory’s extension in Theorem 3.1 is part of the exam).

In order to state and prove Caratheodory extension Theorem 3.1 we first a new concept ofmeasurability associated to a mapping α : P → [0,∞] which satisfies α(∅) = 0.

Definition B.1. Let α : P(S) → [0,∞] be a mapping which satisfies α(∅) = 0. We say thatA ⊆ S is α-measurable if

α(Q) = α(Q ∩A) + α(Q ∩Ac), for all Q ∈ P(S).

The collection of all α-measurable sets is denoted by Mα.

The mapping α could be rather general and it will not be additive in general. However, thecleverly defined collection Mα turns out to be a ring on which α is additive.

Lemma B.2. Let α : P(S)→ [0,∞] be a mapping which satisfies α(∅) = 0. Then Mα is a ringand α is additive on Mα.

Proof. In order to check that Mα is a ring we check the following:

(i) ∅ ∈Mα;(ii) A ∈Mα =⇒ Ac ∈Mα;(iii) A,B ∈Mα =⇒ A ∩B ∈Mα.

Given these properties it is straightforward to check that Mα is a ring. Indeed, this follows fromthe formulas B \A = B ∩Ac and A ∪B = (Ac ∩Bc)c.

Properties (i) and (ii) are clear. It remains to check (iii). Let A,B ∈Mα and write C = A∩B.Let Q ∈ P(S) be arbitrary. Observation: A ∩Bc = Cc ∩A and Ac = Cc ∩Ac. One has

α(Q) = α(Q ∩A) + α(Q ∩Ac) since A ∈Mα

= α(Q ∩A ∩B) + α(Q ∩A ∩Bc) + α(Q ∩Ac) since B ∈Mα

= α(Q ∩ C) + α(Q ∩ Cc ∩A) + α(Q ∩ Cc ∩Ac) by the observation

= α(Q ∩ C) + α(Q ∩ Cc) since A ∈Mα.

Therefore, A ∩B = C ∈Mα which yields (iii).To check that α is additive fix two disjoint sets A,B ∈Mα and let Q = A∪B. Then Q∩A = A

and Q ∩Ac = B. Therefore, since A ∈Mα we find

α(A ∪B) = α(Q) = α(Q ∩A) + α(Q ∩Ac) = α(A) + α(B).

,

In Lemma B.2 we have seen that α is additive. In order to obtain a measure we need a furthercondition on α.

Definition B.3. Let S be a set. A function α : P(S)→ [0,∞] is called an outer measure if

(i) α(∅) = 0;(ii) (monotonicity) A ⊆ B =⇒ α(A) ≤ α(B).

(iii) (σ-subadditivity) For each sequence (An)n≥1 in P(S) one has α( ∞⋃n=1

An

)≤∞∑n=1

α(An).

An outer measure is not necessarily a measure. For instance α(∅) = 0 and α(A) = 1 if A ⊆ Sis non-empty is an example of an outer measure which is not a measure (if S contains at least twoelements).

Next, we show that being an outer measure is the right additional ingredient to prove that αis a measure on Mα.

Lemma B.4. Let α be an outer measure. ThenMα is a σ-algebra and α is a measure on (S,Mα).

Page 49: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 49

Proof. Lemma B.2 gives thatMα is a ring and α is additive onMα. It remains to check that forany disjoint sequence (An)n≥1,

(B.1) A :=

∞⋃n=1

An ∈Mα and α(A) =

∞∑n=1

α(An).

Let Bn =n⋃j=1

Aj for each n ≥ 1. Fix an arbitrary Q ⊆ S. For every n ∈ N the following holds:∑nj=1 α(Q ∩Aj) + α(Q ∩Ac) = α(Q ∩Bn) + α(Q ∩Ac) (by Lemma B.2)

≤ α(Q ∩Bn) + α(Q ∩Bcn) (since Ac ⊆ Bcn)

= α(Q) (since Bn ∈Mα)

Using first the σ-subadditivity of α and then the arbitrariness of n ∈ N in the above, we deduce

(B.2) α(Q ∩A) + α(Q ∩Ac) ≤∞∑j=1

α(Q ∩Aj) + α(Q ∩Ac) ≤ α(Q).

On the other hand by subadditivity also the converse estimate holds: α(Q) ≤ α(Q∩A)+α(Q∩Ac),and hence A ∈ Mα. Moreover, all inequalities in (B.2) have to be identities75, and hence (B.1)follows by taking Q = A in (B.2). -

In the above results we have seen that with outer measures one can construct measures oncertain σ-algebras. Our next aim is to show that there is a natural outer measure associated toan additive mapping µ on a ring R.

Lemma B.5. Let S be a set and R ⊆ P(S) be a ring. Suppose µ : R → [0,∞] is additive andsatisfies µ(∅) = 0. For A ⊆ S define

(B.3) µ∗(A) = inf{ ∞∑j=1

µ(Bj) : A ⊆∞⋃j=1

Bj , where Bj ∈ R for j ≥ 1},

where we let µ∗(A) =∞ if the above set is empty. Then µ∗ is an outer measure

Proof. The mapping µ∗ : P(S) → [0,∞] clearly satisfies (i) and (ii). In order to check (iii) let(An)n≥1 in P and let ε > 0. If µ∗(An) = ∞ for some n ≥ 1, then (iii) is trivial. Next assumeµ∗(An) < ∞ for all n ≥ 1. Then by definition of µ∗ for each fixed n ≥ 1 we can find Bn,j ∈ Rsuch that

An ⊆∞⋃j=1

Bn,j and µ∗(An) + 2−nε ≥∞∑j=1

µ(Bn,j).

Then∞⋃n=1

An ⊆∞⋃

n,j=1

Bn,j and again by the definition of µ∗ we find

µ∗( ∞⋃n=1

An

)≤

∞∑n,j=1

µ(Bn, j) ≤∞∑n=1

µ∗(An) + 2−nε = ε+

∞∑n=1

µ∗(An).

Since ε > 0 was arbitrary, this prove the required estimate. ,

Theorem B.6 (Caratheodory’s extension theorem). Let R ⊆ P(S) be a ring and µ : R → [0,∞]be σ-additive on R and µ(∅) = 0. Then µ∗ defined by (B.3) satisfies the following properties:

(i) µ∗ is a measure on the σ-algebra Mµ∗ ;(ii) µ∗(A) = µ(A) for all A ∈ R.

(iii) R ⊆Mµ∗ ;

In particular, µ : σ(R)→ [0,∞] defined by µ(A) = µ∗(A) defines a measure.

Clearly, Theorem 3.1 follows from the above statement and actually shows that there is a furtherextension to the possibly larger σ-algebra Mµ∗ .

75Clearly, x ≤ y ≤ z ≤ x enforces x = y = z

Page 50: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

50 MEASURE AND INTEGRATION

Proof. In Lemma B.5 we have proved that µ∗ is an outer measure. Therefore, assertion (i) followsfrom Lemma B.4. In remains to prove (ii) and (iii). The assertion concerning µ follows from(i)–(iii) as the restriction of a measure to a smaller σ-algebra is a measure again.

Step 1: Proof of (ii): Let A ∈ R. It is clear that µ∗(A) ≤ µ(A). Indeed, take B1 = A andBn = ∅ for n ≥ 2 in (B.3). For the converse estimate the case µ∗(A) =∞ is clear. Now suppose

µ∗(A) < ∞ and let B1, B2, . . . ,∈ R be such that A ⊆∞⋃n=1

Bn. Then by the σ-additivity of µ on

R and Theorem 2.8 (iii) applied to A =∞⋃n=1

A ∩Bn, we find

µ(A) ≤∞∑n=1

µ(A ∩Bn) ≤∞∑n=1

µ(Bn).

Taking the infimum over all (Bn)n≥1 as above yields µ(A) ≤ µ∗(A).

Step 2: Proof of (iii): Let A ∈ R and Q ⊆ S. Since µ∗ is subadditive we find µ∗(Q) ≤µ∗(Q ∩A) + µ∗(Q ∩Ac). For the converse estimate the case µ∗(Q) =∞ is trivial. In case µ∗(Q) <

∞, choose B1, B2, . . . ,∈ R such that Q ⊆∞⋃n=1

Bn. Then Bn ∩A,Bn ∩Ac ∈ R for all n ≥ 1 and

Q ∩A ⊆∞⋃n=1

Bn ∩A and Q ∩Ac ⊆∞⋃n=1

Bn ∩Ac.

Therefore, using first the definition of µ∗ and then the additivity of µ on R, we find

µ∗(Q ∩A) + µ∗(Q ∩Ac) ≤∞∑n=1

µ(Bn ∩A) +

∞∑n=1

µ(Bn ∩Ac) =

∞∑n=1

µ(Bn)

Taking the infimum over all (Bn)n≥1 as above gives µ∗(Q ∩A) + µ∗(Q ∩Ac) ≤ µ∗(Q). Combiningboth estimates we can conclude A ∈Mµ∗ . ,

ExercisesIn the following exercise we show that Mµ∗ 6= P(S) in general.76

Exercise B.1. Let S = {1, 2, 3} and define a σ-algebra by A = {∅, S, {1, 2}, {3}}. Assume µ is ameasure satisfying µ({1, 2}) = µ({3}) = 1

2 .

(a) Show that µ∗({1}) = µ∗({2}) = 12 .

(b) Show that {1}, {2} /∈Mµ∗ .

Exercise B.2. Let α : P(S) → [0,∞] be an outer measure and suppose that A ⊆ P(S) satisfiesα(A) = 0. Show that A ∈Mα.

Exercise∗ B.3. Assume the conditions of Theorem B.6 and assume µ is σ-finite on R, that means

there exists a sequence (Sn)n≥1 in R such that µ(Sn) < ∞ for all n ≥ 1 and∞⋃n=1

Sn = S. Prove

that the following are equivalent:

(a) A ∈Mµ∗ ;(b) There exists a B ∈ σ(R) such that A ⊆ B and µ∗(B \A) = 0

Hint: First reduce to the case of finite measure by intersecting with Sn. Use the definition of µ∗

given in (B.3).

76For the Lebesgue measure one also has Mµ∗ 6= P(R), but this is much harder to prove. See Appendix C

Page 51: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

MEASURE AND INTEGRATION 51

Appendix C. Non-measurable sets

Let λ be the Lebesgue measure on B(Rd). Let λ∗ : P(S) → [0,∞] be the outer measureassociated with λ (see Lemma B.5). The σ-algebra M(Rd) :=Mλ∗ introduced in Definition B.1is usually called the Lebesgue σ-algebra. It follows from Theorem B.6 that Fd ⊆ M(Rd) andthus also B(Rd) ⊆M(Rd) and λ∗ is a measure onM(Rd). In the sequel we write λ again for thismeasure as it is just an extension of λ.

By Exercise B.3 for every A ∈M(Rd) there exists a B ∈ B(Rd) such that A ⊆ B and λ(B\A) =0. This shows that the Lebesgue σ-algebra is almost the same as the Borel σ-algebra up to setsof measure zero. Some strange things can happen with nonmeasurable sets.

The balls of Banach and TarskiOne can cut a ball of radius one in Rd in such a way that it can be used to form two balls of

radius one. Of course something has to be nonmeasurable there. See:https://en.wikipedia.org/wiki/Banach%E2%80%93Tarski_paradox

A set which Lebesgue measurable but not Borel measurableThere exist a set A ∈ M(Rd) with λ(A) = 0, but A /∈ B(Rd). See [3, Appendix C] and [12,

page 53]http://onlinelibrary.wiley.com/doi/10.1002/9781118032732.app3/pdf

http://www.math3ma.com/mathema/2015/8/9/lebesgue-but-not-borel

You can also read about this in the Bachelor thesis of Gerrit Vos:http://resolver.tudelft.nl/uuid:30d69b56-b846-435e-9d44-6a31b840a836

There exist sets which are not Lebesgue measurable A subset of R which is not inM(Rd)is given by Vitali’s example (see for example [4, Theorem 16.31]):

https://en.wikipedia.org/wiki/Vitali_set

Page 52: MEASURE AND INTEGRATION Contents · 2019-01-18 · theory is almost not measurable. The lecture notes are based on [1], [8], [16] and [17]. A very complete treatment of measure theory

52 MEASURE AND INTEGRATION

References

[1] C. D. Aliprantis and O. Burkinshaw. Principles of real analysis. Academic Press, Inc., San Diego, CA, thirdedition, 1998.

[2] S. Banach. Theorie des operations lineaires. Editions Jacques Gabay, Sceaux, 1993. Reprint of the 1932 original.[3] F. Burk. Lebesgue measure and integration. Pure and Applied Mathematics (New York). John Wiley & Sons,

Inc., New York, 1998. An introduction, A Wiley-Interscience Publication.

[4] N. L. Carothers. Real analysis. Cambridge University Press, Cambridge, 2000.[5] D. H. Fremlin. Measure theory. Vol. 1, 2, 3, 4. Torres Fremlin, Colchester, 2004-2006.

[6] D.J.H. Garling. Inequalities: a journey into linear analysis. Cambridge University Press, Cambridge, 2007.

[7] R. Haberman. Applied Partial Differential Equations with Fourier Series and Boundary Valve Problems.Pearson Higher Ed, 2012.

[8] O. Kallenberg. Foundations of modern probability. Probability and its Applications (New York). Springer-

Verlag, New York, second edition, 2002.[9] Y. Katznelson. An introduction to harmonic analysis. Cambridge Mathematical Library. Cambridge University

Press, Cambridge, third edition, 2004.

[10] J Korevaar. Fourier analysis and related topics. https://staff.fnwi.uva.nl/j.korevaar/Foubook.pdf.[11] S.R. Lay. Analysis. With an introduction to proof. 5th ed. Pearson, 2015.

[12] H. Royden and P. Fitzpatrick. Real analysis. 4th ed. New York, NY: Prentice Hall, 4th ed. edition, 2010.[13] E.M. Stein and R. Shakarchi. Fourier analysis, volume 1 of Princeton Lectures in Analysis. Princeton University

Press, Princeton, NJ, 2003. An introduction.

[14] E.M. Stein and R. Shakarchi. Real analysis. Princeton Lectures in Analysis, III. Princeton University Press,Princeton, NJ, 2005. Measure theory, integration, and Hilbert spaces.

[15] E.M. Stein and R. Shakarchi. Functional analysis, volume 4 of Princeton Lectures in Analysis. Princeton

University Press, Princeton, NJ, 2011. Introduction to further topics in analysis.[16] D. Werner. Einfuhrung in die hohere Analysis. Springer-Lehrbuch. [Springer Textbook]. Springer-Verlag,

Berlin, corrected edition, 2009.

[17] D. Williams. Probability with martingales. Cambridge Mathematical Textbooks. Cambridge University Press,Cambridge, 1991.

[18] A. Zygmund. The role of Fourier series in the development of analysis. In Proceedings of the American Academy

Workshop on the Evolution of Modern Mathematics (Boston, Mass., 1974), volume 2, pages 591–594, 1975.[19] A. Zygmund. Trigonometric series. Vol. I, II. Cambridge Mathematical Library. Cambridge University Press,

Cambridge, third edition, 2002. With a foreword by Robert A. Fefferman.