APPUNTI
LECTURE NOTES
Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Scuola Normale Superiore
Piazza dei Cavalieri, 7
56126 Pisa, Italy
Introduction
to Measure Theory
and Integration
c 2011 Scuola Normale Superiore Pisa
ISBN: 978-88-7642-385-7
e-ISBN: 978-88-7642-386-4
Contents
Preface ix
Introduction xi
1 Measure spaces 1
1.1 Notation and preliminaries . . . . . . . . . . . . . . . . 1
1.2 Rings, algebras and σ –algebras . . . . . . . . . . . . . . 2
1.3 Additive and σ –additive functions . . . . . . . . . . . . 4
1.4 Measurable spaces and measure spaces . . . . . . . . . . 7
1.5 The basic extension theorem . . . . . . . . . . . . . . . 8
1.5.1 Dynkin systems . . . . . . . . . . . . . . . . . . 9
1.5.2 The outer measure . . . . . . . . . . . . . . . . 11
1.6 The Lebesgue measure in R . . . . . . . . . . . . . . . 14
1.7 Inner and outer regularity of measures on metric spaces . 18
2 Integration 23
2.1 Inverse image of a function . . . . . . . . . . . . . . . . 23
2.2 Measurable and Borel functions . . . . . . . . . . . . . 24
2.3 Partitions and simple functions . . . . . . . . . . . . . . 25
2.4 Integral of a nonnegative E –measurable function . . . . 27
2.4.1 Integral of simple functions . . . . . . . . . . . 27
2.4.2 The repartition function . . . . . . . . . . . . . 28
2.4.3 The archimedean integral . . . . . . . . . . . . . 31
2.4.4 Integral of a nonnegative measurable function . . 32
2.5 Integral of functions with a variable sign . . . . . . . . . 35
2.6 Convergence of integrals . . . . . . . . . . . . . . . . . 36
2.6.1 Uniform integrability and Vitali convergence
theorem . . . . . . . . . . . . . . . . . . . . . . 38
2.7 A characterization of Riemann integrable functions . . . 39
vi Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
4 Hilbert spaces 61
4.1 Scalar products, pre-Hilbert and Hilbert spaces . . . . . 61
4.2 The projection theorem . . . . . . . . . . . . . . . . . . 63
4.3 Linear continuous functionals . . . . . . . . . . . . . . . 66
4.4 Bessel inequality, Parseval identity and orthonormal
systems . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.5 Hilbert spaces on C . . . . . . . . . . . . . . . . . . . . 70
5 Fourier series 73
5.1 Pointwise convergence of the Fourier series . . . . . . . 75
5.2 Completeness of the trigonometric system . . . . . . . . 79
5.3 Uniform convergence of the Fourier series . . . . . . . . 80
6 Operations on measures 83
6.1 The product measure and Fubini–Tonelli theorem . . . . 83
6.2 The Lebesgue measure on Rn . . . . . . . . . . . . . . . 87
6.3 Countable products . . . . . . . . . . . . . . . . . . . . 90
6.4 Comparison of measures . . . . . . . . . . . . . . . . . 94
6.5 Signed measures . . . . . . . . . . . . . . . . . . . . . 101
6.6 Measures in R . . . . . . . . . . . . . . . . . . . . . . . 105
6.7 Convergence of measures on R . . . . . . . . . . . . . . 107
6.8 Fourier transform . . . . . . . . . . . . . . . . . . . . . 112
6.8.1 Fourier transform of a measure . . . . . . . . . . 113
A 137
A.1 Continuity and differentiability of functions depending
on a parameter . . . . . . . . . . . . . . . . . . . . . . . 137
vii Introduction to Measure Theory and Integration
References 183
Preface
Even though this statement makes perfectly sense within Riemann’s the-
ory, any attempt to prove this result within the theory (try, if you don’t
xii Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
believe!) seems to fail, and leads (more or less explicitely, see [2]) to
a larger theory. In addition to that, the continuity assumption on the
limit function f is not natural, because a pointwise limit of continuous
functions need not be continuous, and we would like to give a sense to
1
−1 f (x) dx even without this assumption. This necessity emerges for
instance in the study of the convergence of Fourier series
∞
f (x) = ai cos(iπ x) + bi sin(iπ x) x ∈ [−1, 1].
i=0
Under this condition the limit function f need not be continuous: for
instance, if f (x) = 1 for x ∈ [−1/2, 1/2] and f (x) = 0 otherwise, then
we will see that the coefficients of the Fourier series are given by bi = 0
for all i (because f is even) and by
⎧
⎪ 1
⎪
⎪ if i = 0;
⎨2
ai =
⎪
⎪
⎪
⎩ sin(πi/2) if i > 0.
iπ
(4) The spaces of integrable functions, as for instance
1
H := f : [−1, 1] → R : f 2 (x) dx < ∞
−1
(1) Notice the analogy with liminf and limsup limits for a sequence (a ) of real numbers. We have
n
lim sup an = inf sup am and lim inf an = sup inf am . This is something more than an analogy,
n→∞ n∈N m≥n n→∞ n∈N m≥n
see Exercise 1.1.
∞
(ii) if (An ) is a sequence of elements of E then An ∈ E .
n=0
If E is a σ –algebra and (An ) ⊂ E we have n An ∈ E by the De
Morgan identity. Moreover, both sets
belong to E .
Obviously, {∅, X} and P (X) are σ –algebras, respectively the smal-
lest and the largest ones. Let K be a subset of P (X). As the intersection
of any family of σ –algebras is still a σ -algebra, the minimal σ –algebra
including K (that is the intersection of all σ –algebras including K ) is
well defined, and called the σ –algebra generated by K . It is denoted by
σ (K ).
In contrast with the case of generated algebras, it is quite hard to give
a constructive characterization of the generated σ -algebras: this requires
the transfinite induction and it is illustrated in Exercise 1.18.
Definition 1.3 (Borel σ -algebra). If (E, d) is a metric space, the σ –
algebra generated by all open subsets of E is called the Borel σ –algebra
of E and it is denoted by B (E).
In the case when E = R the Borel σ -algebra has a particularly simple
class of generators.
Example 1.4 (B (R)). Let I be the set of all semi–closed intervals [a,b)
with a ≤ b. Then σ (I ) coincides with B (R). In fact σ (I ) contains all
open intervals (a, b) since
∞
1 1
(a, b) = a + ,b with n 0 > .
n=n 0 n b−a
(2) Indeed, let (a ) be a sequence including all rational numbers of A and denote by I the largest
k k
open interval contained in A and containing ak . We clearly have A ⊃ ∞k=0 Ik , but also the opposite
inclusion holds: it suffices to consider, for any x ∈ A, r > 0 such that (x − r, x + r) ⊂ A, and k
such that ak ∈ (x − r, x + r) to obtain (x − r, x + r) ⊂ Ik , by the maximality of Ik , and then x ∈ Ik .
4 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
μ(L) = lim μ(Bn ) = inf μ(Bn ) ≥ inf sup μ(Am ) = lim sup μ(An ).
n→∞ n∈N n∈N m≥n n→∞
Thus, we have proved (1.3). The inequality (1.2) can be proved similarly
using Proposition 1.6, thus without using the assumption μ(X) < ∞.
The following result is very useful to estimate the measure of a lim sup
of sets.
Lemma 1.9. Let μ be σ –additive on a σ –algebra E and let (An ) ⊂ E .
∞
Assume that μ(An ) < ∞. Then
n=0
μ lim sup An = 0.
n→∞
Proof. Set L = lim sup An and define Bn as in (1.4). Then the inclusion
n→∞
L ⊂ Bn gives
∞
μ(L) ≤ μ(Bn ) ≤ μ(Am ) for all n ∈ N.
m=n
As n → ∞ we find μ(L) = 0.
7 Introduction to Measure Theory and Integration
This example can be generalized as follows, see also Exercise 1.5 and
Exercise 1.23.
Example 1.10 (Discrete measures). Assume that Y ⊂ X is a finite or
countable set. Given c : Y → [0, +∞] we can define a measure on
(X, P (X)) as follows:
μ(B) := c(x) ∀B ⊂ X.
x∈B∩Y
Clearly μ = x∈Y c(x)δx is a finite measure if and only if x∈Y c(x) <
∞, and it is σ –finite if and only if c(x) ∈ [0, +∞) for all x ∈ Y .
More generally, the construction above works even when Y is uncount-
able, by replacing the sum with
sup c(x),
c∈B∩Y
where the supremum is made among the finite subsets Y of Y . The meas-
ures arising in the previous example are called atomic, and clearly if X is
either finite or countable then any measure μ in (X, P (X)) is atomic: it
suffices to notice that
μ= c(x)δx with c(x) := μ({x}).
x∈X
In the next section we will introduce a fundamental tool for the construc-
tion of non-atomic measures.
8 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
{x ∈ X : P(x) is false}
A, B ∈ K ⇒ A ∩ B ∈ K .
(i) X, ∅ ∈ D ;
(ii) A ∈ D ⇒ Ac ∈ D ;
(iii) (Ai ) ⊂ D mutually disjoint ⇒ i Ai ∈ D .
H (B) = {F ∈ D 0 : B ∩ F ∈ D 0 }.
We claim that H (B) is a Dynkin system. In fact properties (i) and (iii)
are clear. It remains to show that if F ∩ B ∈ D 0 then F c ∩ B ∈ D 0 or,
equivalently, F ∪ B c ∈ D 0 . In fact, since F ∪ B c = (F \ B c ) ∪ B c =
(F ∩ B) ∪ B c and F ∩ B and B c are disjoint, we have that F ∪ B c ∈ D 0
as required.
Notice first that if K ∈ K we have K ⊂ H (K ) since K is a
π–system. Therefore H (K ) = D 0 , by the minimality of D 0 . Con-
sequently, the following implication holds
K ∈ K , B ∈ D 0 ⇒ K ∩ B ∈ D 0 ,
10 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
which implies K ⊂ H (B) for all B ∈ D 0 . Again, the fact that H (B) is
a Dynkin system and the minimality of D 0 give that H (B) = D 0 for all
B ∈ D 0 . By the definition of H (B), this proves (1.5).
The uniqueness part in Caratheodory’s theorem is a direct consequence
of the following coincidence criterion for measures; in turn, the proof of
the criterion relies on Theorem 1.14.
Proposition 1.15 (Coincidence criterion). Let μ1 , μ2 be measures in
(X, E ) and assume that:
(i) the coincidence set
D := {A ∈ E : μ1 (A) = μ2 (A)}
contains a π–system K with σ (K ) = E ;
(ii) there exists a nondecreasing sequence (X i ) ⊂ K with μ1 (X i ) =
μ2 (X i ) < ∞ and X i ↑ X.
Then μ1 = μ2 .
Proof. We first assume that μ1 (X) = μ2 (X) is finite. Under this as-
sumption D is a Dynkin system including the π–system K (stability of
D under complement is ensured precisely by the finiteness assumption).
Thus, by the Dynkin theorem, D = E , which implies that μ1 = μ2 .
Assume now that we are in the general case and let X i be given by
assumption (ii). Fix i ∈ N and define the σ –algebra E i of subsets of X i
by
E i := {A ⊂ X i : A ∈ E } .
We may obviously consider μ1 and μ2 as finite measures in the measur-
able space (X i , E i ). Since these measures coincide on the π–system
K i := {A ⊂ X i : A ∈ K }
we obtain, by the previous step, that μ1 and μ2 coincide on σ (K i ) ⊂
P (X i ).
Now, let us prove the inclusion
{B ∈ E : B ⊂ X i } ⊂ σ (K i ). (1.6)
Indeed
B ⊂ X : B ∩ X i ∈ σ (K i )
is a σ –algebra containing K (here we use the fact that X i ∈ K ), and
therefore contains E . Hence any element of E contained in X i belongs to
σ (K i ).
By (1.6) we obtain μ1 (B ∩ X i ) = μ2 (B ∩ X i ) for all B ∈ E and all
i ∈ N. Passing to the limit as i → ∞, since B is arbitrary we obtain that
μ1 = μ2 .
11 Introduction to Measure Theory and Integration
Consequently
∞
∞
μ(Ai, j ) ≤ μ∗ (E i ) + ε.
i, j=0 i=0
∞
Since E ⊂ Ai, j we have
i, j=0
∞
∞
μ∗ (E) ≤ μ(Ai, j ) ≤ μ∗ (E i ) + ε
i, j=0 i=0
and the first part of the statement follows from the arbitrariness of ε.
Now, let us assume that μ is σ -subadditive on A and choose E ∈ A ;
since E ⊂ i Ai then μ(E) ≤ i μ(Ai ), so we deduce μ∗ (E) ≥ μ(E);
but, by choosing A0 = E and An = ∅ for n ≥ 1, we obtain that μ∗ (E) =
μ(E). This proves that μ∗ extends μ.
12 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
= μ∗ (E ∩ A) + μ∗ (E ∩ Ac ∩ B) + μ∗ (E ∩ Ac ∩ B c ) (1.10)
Since
(E ∩ A) ∪ (E ∩ Ac ∩ B) = E ∩ (A ∪ B),
we have by the subadditivity of μ∗ ,
μ∗ (E ∩ A) + μ∗ (E ∩ Ac ∩ B) ≥ μ∗ (E ∩ (A ∪ B)).
So, by (1.10) it follows that
μ∗ (E) ≥ μ∗ (E ∩ (A ∪ B)) + μ∗ (E ∩ (A ∪ B)c ),
and A ∪ B ∈ G as required. The additivity of μ∗ on G follows directly
from (1.9).
Step 3. G is a σ –algebra. Let (An ) ⊂ G . We are going to show that
S := n An ∈ G . Since we know that G is an algebra, it is not
restrictive
to assume that all sets An are mutually disjoint. Set Sn := n0 Ai , for
n ∈ N.
For any n ∈ N, by using the σ –subadditivity of μ∗ and by applying
(1.7) repeatedly, we get
∞
∗ ∗ ∗
μ (E ∩ S ) + μ (E ∩ S) ≤ μ (E ∩ S ) +
c c
μ∗ (E ∩ Ai )
i=0
n
= lim μ∗ (E ∩ S c ) + μ∗ (E ∩ Ai )
n→∞
i=0
= lim μ∗ (E ∩ S c ) + μ∗ (E ∩ Sn ) .
n→∞
also belongs to A .
We prove the additivity property first in the case when F = (x, y] ∈
I . It is also not restrictive to assume that the series n λ(Fn ) is con-
vergent. As any Fn is a finite union of intervals, say Nn , we can find,
given any ε > 0, a finite union Fn ⊃ Fn of intervals in I such that
λ(Fn ) ≤ λ(Fn ) + ε/2n and the internal part of Fn contains Fn (just shift
the endpoints of each interval in Fn by a small amount, to obtain a lar-
ger interval in I , increasing the length at most by ε/(Nn 2n )). Let also
Fn be the internal part
of Fn , that still includes Fn , and let x ∈ (x, y].
Then, since [x , y] ⊂ n Fn , Lemma 1.20 provides an integer k such
16 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
k
that [x , y] ⊂ 0 Fn . Hence, the additivity of λ in A gives
k k
y − x ≤ λ Fn ≤ λ(Fn )
n=0 n=0
k
ε ∞
≤ λ(Fn ) + ≤ 2ε + λ(Fn ).
n=0
2n n=0
k
∞
k
∞
λ(F) = λ(Ii ∩ F) = λ(Ii ∩ Fn ) = λ(Fn ).
i=1 n=0 i=1 n=0
p 1 p
ν([0, )) = pν [0, ) = c .
q q q
By approximation we eventually obtain that ν([0, t)) = ct for all
t ≥ 0.
The completion of the Borel σ –algebra with respect to λ is the so-
called σ -algebra of Lebesgue measurable sets. It coincides with the
class C of additive sets with respect to λ∗ considered in the proof of
Carathéodory theorem (see Exercise 1.12).
Remark 1.23 (Outer Lebesgue measure and non-measurable sets).
The measure λ∗ used in the proof of Carathéodory’s theorem is also called
outer Lebesgue measure, and it is defined on all parts of R. The termin-
ology is slightly misleading here, since λ∗ , though σ –subadditive, fails
to be σ –additive. In particular, there exist subsets of R that are not Le-
besgue measurable. To see this, let us consider the equivalence relation
in R defined by x ∼ y if x − y ∈ Q and let us pick a single element
x ∈ [0, 1] in any equivalence class induced by this relation, thus forming
a set A ⊂ [0, 1]. Were this set Lebesgue measurable, all the sets A + h
18 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
where d(x, Ac ) := inf y∈Ac d(x, y) is the distance function from Ac . Then
Cn are closed subsets of A, and moreover Cn ↑ A, which implies μ(A \
Cn ) ↓ 0. Thus the conclusion follows.
Notice that inner and outer approximation hold for μ–measurable sets
B as well: one has just to notice that there exist Borel sets B1 , B2 such
that B1 ⊂ B ⊂ B2 with μ(B2 \ B1 ) = 0, and apply inner approximation
to B1 and outer approximation to B2 .
Remark 1.25 (Inner and outer approximation for σ-finite measures).
It is possible to extend the inner approximation property to σ -finite meas-
ures: suffices to assume the existence of a sequence of closed sets Cn with
finite measure such that μ(X \∪n Cn ) = 0. Indeed, assuming with no loss
of generality that Cn ⊂ Cn+1 , we know that for any Borel set B and any
n ∈ N it holds
so that
μ(B ∩ Cn ) ≤ sup {μ(C) : C closed, C ⊂ B} .
Letting n ↑ ∞ we recover the inner approximation property.
Analogously, if we assume the existence of a sequence of open sets An
with finite measure satisfying X = ∪n An , we have the outer approxim-
ation property: indeed, for any n and any > 0 we can find (assuming
with no loss of generality μ(B) < +∞) open sets Bn ⊂ An containing
20 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
B ∩ An and such that μ Bn \ (B ∩ An ) < 2−n . It follows that ∪n Bn
contains B and
μ Bn \ B < 2 .
n∈N
Since Bn are also open in X, the set ∪n Bn is open and since is arbitrary
we get the outer approximation property.
We conclude this chapter with the following result, whose proof is a
straightforward consequence of Proposition 1.24 (alternatively, one can
use Dynkin’s argument, since the class of closed sets is a π-system and
generates the Borel σ -algebra).
Corollary 1.26. Let μ, ν be finite measures in (E, B (E)), such that
μ(C) = ν(C) for any closed subset C of E. Then μ = ν.
Exercises
1.1 Given A ⊂ X, denote by 1 A : X → {0, 1} its characteristic function, equal
to 1 on A and equal to 0 on Ac . Show that
and that
lim sup An = A ⇐⇒ lim sup 1 An = 1 A ,
n→∞ n→∞
lim inf An = A ⇐⇒ lim inf 1 An = 1 A .
n→∞ n→∞
1.2 Let A ⊂ Rn be a Borel set. Show that for h ∈ Rn and t ∈ R the sets
A + h := {a + h : a ∈ A} , t A := {ta : a ∈ A}
1.8 Let λ be the Lebesgue measure in [0, 1]. Show the existence of a λ–negli-
gible set having the cardinality of the continuum. Hint: consider the classical
Cantor’s middle third set, obtained by removing the interval (1/3, 2/3) from
[0, 1], then by removing the intervals (1/9, 2/9) and (7/9, 8/9), and so on.
1.9 Let λ be the Lebesgue measure in [0, 1]. Show the existence, for any ε > 0,
of a closed set C ⊂ [0, 1] containing no interval and such that λ(C) > 1 − ε.
Hint: remove from [0, 1] a sequence of open intervals, centered on the rational
points of [0, 1].
1.10 Using the previous exercise, write [0, 1] = A ∪ B where A is negligible
in the measure-theoretic sense (i.e. λ(A) = 0) and B is negligible in the Baire
category sense (i.e. it is the union of countably many closed sets with empty
interior). So, the two concepts of negligible should be never used at the same
time.
1.11
Let λ be the Lebesgue measure in [0, 1]. Construct a Borel set E ⊂ (0, 1)
such that
0 < λ(E ∩ I ) < λ(I )
for any open interval I ⊂ (0, 1).
1.12 Let (X, E , μ) be a measure space and let μ∗ : P (X) → [0, +∞] be
the outer measure induced by μ. Show that the completed σ –algebra E μ is
contained in the class C of additive sets with respect to μ∗ .
1.13 Let (X, E , μ) be a measure space and let μ∗ : P (X) → [0, +∞] be the
outer measure induced by μ. Show that for all A ⊂ X there exists B ∈ E
containing A with μ(B) = μ∗ (A).
1.14 Let (X, E , μ) be a measure space. Check the following statements, made
in Definition 1.12:
(i) E μ is a σ –algebra;
(ii) the extension μ(A) := μ(B), where B ∈ E is any set such that AB is
contained in a μ–negligible set of E , is well defined and σ –additive on
E μ;
(iii) μ–negligible sets of E μ are characterized by the property of being coin-
tained in a μ–negligible set of E .
1.15
Let (X, E , μ) be a measure space and let μ∗ : P (X) → [0, +∞] be
the outer measure induced by μ. Show that if μ(X) is finite, the class C of
additive sets with respect to μ∗ coincides with the class of E μ –measurable sets.
Hint: one inclusion is provided by Exercise 1.12. For the other one, given an
additive set A, by applying Exercise 1.13 twice, find first a set B ∈ E with
μ∗ (B \ A) = 0, and then a set C ∈ E with μ(C) = 0 and B \ A ⊂ C.
1.16 Find a σ –algebra E ⊂ P (N) containing infinitely many sets and such that
any B ∈ E different from ∅ has an infinite cardinality.
1.17 Find μ : P(N) → {0, +∞} that is additive, but not σ –additive.
1.18
Let ω be the first uncountable ordinal and, for K ⊂ P (X), define by
transfinite induction a family F (i) , i ∈ ω, as follows: F (0) := K ∪ {∅},
∞
(i) c ( j) ( j)
F := Ak , B : (Ak ) ⊂ F , B ∈ F ,
k=0
22 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
if i is the successor of j, and F (i) := j∈i F ( j) otherwise.
Show that i∈ω F (i) = σ (K ).
1.19
Show that B (R) has the cardinality of the continuum. Hint: use the con-
struction of the previous exercise, and the fact that ω has at most the cardinality
of continuum.
1.20
Show that the σ –algebra L of Lebesgue measurable sets has the same
cardinality of P(R), thus strictly greater than the continuum. Hint: consider all
subsets of Cantor’s middle third set.
1.21
Show that the cardinality of any σ –algebra is either finite or uncount-
able.
1.22 Let X be a set and let A ⊂ P (X) be an algebra with finite cardinality.
Show that its cardinality is equal to 2n for some integer n ≥ 1.
1.23
Let (X, E , μ) be a a measure space and suppose that X is finite or count-
able. Show the existence of a measure μ̃ on P (X) that extends μ, that is,
μ(A) = μ̃(A) for all A ∈ E .
1.24
Find an example of an additive set function μ : P (N) → {0, 1}, with
μ(N) = 1 and μ({n}) = 0 for all n ∈ N (in particular μ is not σ –additive, the
construction of this example requires Zorn’s lemma).
1.25
Let C ∈ B ([0, 1]) with λ(C) > 0. Without using the continuum hypo-
thesis, show that C has the cardinality of continuum.
1.26
Let (K , d) be a compact metric space and let μ be as in Exercise 1.24.
Let’s say that a sequence (xn ) ⊂ K μ-converges to x ∈ K if
μ {n ∈ N : d(xn , x) > ε} = 0 ∀ε > 0.
Show that any sequence (xn ) ⊂ K is μ-convergent and that the μ-limit is
unique.
Chapter 2
Integration
ϕ −1 (I ) := {x ∈ X : ϕ(x) ∈ I } = {ϕ ∈ I }.
Let us recall some elementary properties of ϕ −1 (the easy proofs are left
to the reader as an exercise):
(i) ϕ −1 (I c ) = (ϕ −1 (I ))c for all I ∈ P (Y );
(ii) if {Ji }i∈I ⊂ P (Y ) we have
ϕ −1 (Ji ) = ϕ −1 Ji , ϕ −1 (Ji ) = ϕ −1 Ji .
i∈I i∈I i∈I i∈I
Proof. Let us prove that ϕ(x) := supn ϕn (x) is E –measurable (all other
cases can be deduced from this one, or directly proved by similar argu-
ments). For any a ∈ R we have
∞
−1
ϕ ([−∞, a]) = ϕn−1 ([−∞, a]) ∈ E .
n=0
where A1 , . . . , AN need not be mutually disjoint and ak need not be in
the range of ϕ. For instance
It is easy to check that a simple function is E –measurable if, and only if,
all level sets Ak in (2.3) are E –measurable; in this case we shall also say
that {Ak } is a finite E –measurable partition of X.
Now we show that any nonnegative E –measurable function can be ap-
proximated by simple functions; a variant of this result, with a different
construction, is proposed in Exercise 2.8.
Proposition 2.4. Let ϕ be a nonnegative extended E –measurable func-
tion. For any n ∈ N∗ , define
⎧
⎪ i −1 i −1 i
⎨ n if n
≤ ϕ(x) < n , i = 1, 2, . . . , n2n ;
ϕn (x) = 2 2 2 (2.4)
⎪
⎩
n if ϕ(x) ≥ n.
Then ϕn are simple and E –measurable, (ϕn ) is nondecreasing and con-
vergent to ϕ. If in addition ϕ is bounded the convergence is uniform.
Proof. It is not difficult to check that (ϕn ) is nondecreasing. Moreover,
we have
1
0 ≤ ϕ(x) − ϕn (x) ≤ if ϕ(x) < n, x ∈ X,
2n
and
0 ≤ ϕ(x) − ϕn (x) = ϕ(x) − n if ϕ(x) ≥ n, x ∈ X.
So, the conclusion easily follows.
27 Introduction to Measure Theory and Integration
It is easy to see that the definition does not depend on the choice of the
formula for ϕ. Indeed, let {b1 , . . . , b M } be the range of ϕ
representation
and let ϕ = 1M b j 1 B j , with B j := ϕ −1 (b j ), be the canonical representa-
tion of ϕ. We have to prove that
N
M
ak μ(Ak ) = b j μ(B j ). (2.5)
k=1 j=1
In order to show (2.6) we fix j and consider, for I ⊂ {1, . . . , N }, the sets
A I := x ∈ B j : x ∈ Ai iff i ∈ I ,
so that {A I } are a E –measurable partition of B j and x ∈ A I iff the set
of i’s for which x ∈ Ai coincides with I . Then, using first the fact that
A I ⊂ Ai if i ∈ I , and Ai ∩ A I = ∅ otherwise,
N and then the fact that
ak = b j whenever A I = ∅ (because 1 ak 1 Ak coincides with b j , the
k∈I
constant value of ϕ on B j ), we have
N
N
N
ak μ(Ak ∩ B j ) = ak μ(Ak ∩ A I ) = ak μ(Ak ∩ A I )
k=1 k=1 I I k=1
= ak μ(A I ) = b j μ(A I ) = b j μ(B j ).
I k∈I I
28 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Proof. Let
n
m
ϕ= ak 1 A k , ψ= bh 1 Bh
k=1 h=1
lim F(t) = lim F(n) = lim μ({ϕ > −n}) = μ({ϕ > −∞}),
t→−∞ n→−∞ n→∞
and, if μ is finite,
lim F(t) = lim F(n) = lim μ({ϕ > n}) = μ({ϕ = +∞}),
t→+∞ n→∞ n→∞
since
∞
∞
{ϕ > −∞} = {ϕ > −n}, {ϕ = +∞} = {ϕ > n}.
n=1 n=1
ϕ F
10
9
8
7
6
5
4
3
2
1
1 1 2 3 4 5 6 7 8 9 10
The color scheme used for the areas below the two graphs in 2.1 proves
graphically that the areas are identical.
Now, we want to define the integral of any nonnegative extended E –
measurable function by generalizing formula (2.7). For this, we need
first to define the integral of any nonnegative nonincreasing function in
(0, +∞).
We define ∞
f (t) dt := sup{I f (σ ) : σ ∈ }. (2.9)
0
∞
The integral 0 f (t) dt is called the archimedean integral of f . It enjoys
the usual properties of the Riemann integral (see Exercise 2.5) but, among
these, we will need only the monotonicity with respect to f in the sequel.
For our purposes the most relevant property of the Archimedean integral
is instead the continuity under monotonically nondecreasing sequences.
Proposition 2.8. Let f n ↑ f , with f n : [0, +∞) → [0, +∞] nonin-
creasing. Then ∞ ∞
f n (t) dt ↑ f (t) dt.
0 0
This implies
∞ ∞
sup f n (t) dt ≥ f (t) dt
n∈N 0 0
Notice that the function t → μ({ϕ > t}) ∈ [0, +∞] is nonnegative and
nonincreasing in [0, +∞), so that its archimedean integral is well defined
and (2.10) extends, by the remarks made at the end of Section 2.4.2, the
integral elementarily defined on simple functions. If the integral is finite
we say that ϕ is μ–integrable.
It follows directly from the analogous properties of the archimedean
integral that the integral so defined is monotone, i.e.
ϕ ≥ ψ ⇒ ϕ dμ ≥ ψ dμ.
X X
Indeed, ϕ ≥ ψ implies {ϕ > t} ⊃ {ψ > t} and μ({ϕ > t}) ≥ μ({ψ >
t}) for all t > 0. Furthermore, the integral is invariant under modifica-
tions of ϕ in μ–negligible sets, that is
ϕ = ψ μ–a.e. in X ⇒ ϕ dμ = ψ dμ.
X X
≥ aμ({ϕ ≥ a}).
Since
∞
{ϕ = ∞} = {ϕ > n},
n=1
by applying the continuity along decreasing sequences in the space ({ϕ >
1} (with finite μ measure) we obtain
Proof. Setting ϕ(x) := lim infn ϕn (x), and ψn (x) = infm≥n ϕm (x), we
have that ψn (x) ↑ ϕ(x). Consequently, by the monotone convergence
theorem,
ϕ(x) dμ(x) = lim ψn (x) dμ(x).
X n→∞ X
On the other hand
ψn (x) dμ(x) ≤ ϕn (x) dμ(x),
X X
so that
ϕ(x) dμ(x) ≤ lim inf ϕn (x) dμ(x).
X n→∞ X
Consequently,
ϕ dμ ≤ lim inf ϕn dμ. (2.13)
X n→∞ X
Consequently,
ϕ dμ ≥ lim sup ϕn dμ. (2.14)
X n→∞ X
This means that for any ε > 0 there exists δ > 0 such that
μ(A) < δ ⇒ |ϕi (x)| dμ(x) ≤ ε ∀ i ∈ I.
A
Now, given any ε > 0, we can choose m > 1/ε to obtain that
1
|ϕn (x) − ϕ(x)| ≤ < ε for all x ∈ Bn(m),m , n ≥ n(m).
m
As X \ Aδ ⊂ Bn(m),m , this proves the uniform convergence of ϕn to ϕ on
X \ Aδ .
Proof
of the Vitali Theorem. Fix ε > 0 and find δ > 0 such that
A |ϕn | dμ < ε whenever μ(A) < δ. Again, Fatou’s Lemma yields that
A |ϕ| dμ ≤ ε whenever μ(A) < δ.
Assume now that A is given by Egorov Lemma, so that ϕn → ϕ uni-
formly on X \ A. Then, writing
(ϕ − ϕn ) dμ = (ϕ − ϕn ) dμ + (ϕ − ϕn ) dμ
X X\A A
and the latter considering the infimum in correspondence of all step func-
tions h ≥ f . We denote by I ( f ) the Riemann integral, equal to the upper
and lower integral whenever the two integrals coincide.
Asn−1
the Lebesgue integral of the function h in (2.19) coincides with
i ai (ti+1 − ti ), we have
g dλ = I (g) for any step function g : J → R.
J
40 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Proof. Let
⎧
⎪
⎪ f ∗ (x) := inf lim inf f (xh ) : x h → x
⎪
⎪
⎨ h→∞
(2.18)
⎪
⎪
⎪
⎪ ∗
⎩ f (x) := sup lim sup f (x h ) : x h → x .
h→∞
It is not hard to show (see Exercise 2.6 and Exercise 2.7) that f ∗ is lower
semicontinuous and f ∗ is upper semicontinuous, therefore both f ∗ and
f ∗ are Borel functions.
We are going to show that I∗ ( f ) = J f ∗ dλ and I ∗ ( f ) = J f ∗ dλ.
These two equalities yield the conclusion, as f is continuous at λ–a.e.
point in J iff f ∗ − f ∗ = 0 λ–a.e. in J , and this holds iff (because f ∗ − f ∗ ≥
0)
( f ∗ − f ∗ ) dλ = 0.
J
Exercises
2.1 Show that any of the conditions listed below is equivalent to the E –mea-
surability of ϕ : X → R.
(i) ϕ −1 ((−∞, t]) ⊂ E for all t ∈ R;
(ii) ϕ −1 ((−∞, t)) ⊂ E for all t ∈ R;
(iii) ϕ −1 ([a, b]) ⊂ E for all a, b ∈ R;
(iv) ϕ −1 ([a, b)) ⊂ E for all a, b ∈ R;
(v) ϕ −1 ((a, b)) ⊂ E for all a, b ∈ R.
2.2 Let ϕ, ψ : X → R be E –measurable. Show that ϕ + ψ and ϕψ are E –
measurable. Hint: prove that
{ϕ + ψ < t} = [{ϕ < r} ∩ {ψ < t − r}]
r∈Q
and √ √
{ϕ 2 > a} = {ϕ > a} ∪ {ϕ < − a}, a ≥ 0.
(i) Show that (R, d) is a compact metric space (the so-called compactification
of R) and that A ⊂ R is open relative to the Euclidean distance if, and only
if, it is open relative to d;
(ii) use (i) to show that, given a measurable space (X, E ), f : X → R is E –
measurable according to (2.2) if and only if it is measurable between E and
the Borel σ –algebra of (R, d).
2.4 Let (X, E , μ) be a measure space and let E μ be the completion of E induced
by μ. Show that f : X → R is E μ –measurable iff there exists a E –measurable
function g such that { f = g} is contained in a μ–negligible set of E .
2.5 Let us define I f as in (2.8) and let us endow with the usual partial ordering
σ = {t1 , . . . , t N } ≤ ζ = {s1 , . . . , s M } if and only if σ ⊂ ζ . Show that σ →
∞
I f (σ ) is nondecreasing. Use this fact to show that f → 0 f (t) dt is additive.
2.6 Let f : R → R be a function. Show that the functions f ∗ , f ∗ defined in
(2.18) are respectively lower semicontinuous and upper semicontinuous.
2.7 Let f : R → R be a bounded function. Using Exercise 2.6 show that
{ f ∗ ≤ t} and { f ∗ ≥ t} are closed for all t ∈ R. In particular deduce that
= {x ∈ R : f is continuous at x}
belongs to B (R).
2.8 Let (an ) ⊂ (0, ∞) with
∞
ai = ∞, lim ai = 0.
i→∞
i=0
Show that for any ϕ : X → [0, +∞] E –measurable there exist Ai ∈ E such that
ϕ = i ai 1 Ai . Hint: set ϕ0 := ϕ, A0 := {ϕ ≥ a0 } and ϕ1 := ϕ0 − a0 1 A0 ≥ 0.
Then, set A1 := {ϕ1 ≥ a1 } and ϕ2 := ϕ0 − a1 1 A1 and so on.
2.9 Let ϕ : X → R be μ–integrable. Show that the property (2.15) holds. Hint:
−i
by contradiction its failure for some ε > 0 and find Ai with μ(Ai ) < 2
assume
and Ai |ϕ| dμ ≥ ε. Then, notice that B := lim supi Ai is μ–negligible, consider
Bn := Ai \ B ↓ ∅
i≥n
Check that:
(a) |gλ (x) − gλ (x )| ≤ λd(x, x ) for all x, x ∈ X;
(b) gλ ↑ g as λ ↑ ∞.
2.12 Let f : R2 → R be satisfying the following two properties:
(i) x → f (x, y) is continuous in R for all y ∈ R;
(ii) y → f (x, y) is continuous in R for all x ∈ R.
Show that f is a Borel function. Hint: first reduce to the case when f is
bounded. Then, for ε > 0 consider the functions
x+ε
1
f ε (x, y) := f (x , y) dx ,
2ε x−ε
We have clearly
and
so that conditions (ii) and (iii) in the definition of the norm are fulfilled.
However, · 1 is not a norm in general, since ϕ1 = 0 if and only if
ϕ = 0 μ–a.e. in X, so (i) fails.
ϕ̃ + ψ̃ := ϕ
+ ψ, α ϕ̃ := α"
ϕ. (3.2)
it is also easy to see that this definition does not depend on the particular
element ϕ chosen in ϕ̃, and that (ii), (iii) still hold. Now, if ϕ̃1 = 0
we have that the integral of |ϕ| is zero, and therefore ϕ̃ = 0. Therefore
L 1 (X, E , μ), endowed with the norm · 1 , is a normed space.
To simplify the notation typically ϕ̃ is identified with ϕ whenever the
formula does not depend on the choice of the function in the equival-
ence class: for instance, quantities as μ({ϕ > t}) or X ϕ dμ have this
independence, as well as most statements and results in Measure Theory
and Probability, so this slight abuse of notation is justified. It should be
noted, however, that formulas like ϕ(x̄) = 0, for some fixed x̄ ∈ X, do
not make sense in L 1 (X, E , μ), since they depend on the representative
chosen (unless μ({x̄}) > 0).
More generally, if an exponent p ∈ (0, ∞) is given, we can apply a
similar construction to the space
L (X, E , μ) := ϕ : ϕ is E –measurable and
p
|ϕ| dμ < ∞ .
p
X
We are going to show that · p is a norm for any p ∈ [1, +∞). Notice
that we already checked this fact when p = 1, and that the homogen-
eity condition (ii) trivially holds, whatever the value of p is. Further-
more, condition (i) holds precisely because L p (X, E , μ) consists, strictly
speaking, of equivalence classes induced by (3.1). So, the only condi-
tion that needs to be checked is the subadditivity condition (ii), and in the
sequel we can assume p > 1.
The concept of Legendre transform will be useful. Let f : R → R be
a function; we define its Legendre transform f ∗ : R → R ∪ {+∞} by
f ∗ (y) = sup{x y − f (x)}, y ∈ R.
x∈R
Proposition 3.6. Let p ∈ [1, +∞) and let (ϕn ) be a Cauchy sequence in
L p (X, E , μ). Then:
Next, set
∞
g(x) := |ϕn(k+1) (x) − ϕn(k) (x)|, x ∈ X.
k=0
Therefore, g is finite μ–a.e., that is, there exists B ∈ E such that μ(B) =
0 and g(x) < ∞ for all x ∈ B c . Set now
∞
ϕ(x) := ϕn(0) (x) + (ϕn(k+1) (x) − ϕn(k) (x)), x ∈ Bc.
k=0
∞ 1/ p
X
∞
≤ |ϕn(k+1) (x) − ϕn(k) (x)| dμ(x)
p
≤ 2−k ,
k=h X k=h
Since (ϕn ) is Cauchy, for any ε > 0 there exists n ε ∈ N such that
Now choose k ∈ N such that n(k) > n ε and ϕ − ϕn(k) p < ε. For any
n > n ε we have
n if x ∈ [0, 1/n],
ϕn (x) =
0 if x ∈ [1/n, 1].
This easily follows from the fact that the function t → μ({|ϕ| > t}) is
right continuous (Proposition 2.6), so the infimum is attained.
Notice also that ϕ∞ is characterized by the property
Indeed, we have just to notice that |ϕ(x)ψ(x)| ≤ ϕ∞ |ψ(x)| for μ–a.e.
x ∈ X, and then integrate with respect to μ. This inequality can be still
written as (3.6), provided we agree that q = 1 is the dual exponent of
p = ∞ (and conversely).
For finite measures we can apply Hölder’s inequality to obtain that the
L spaces are nested; in particular L ∞ is the smaller one and L 1 is the
p
larger one.
Remark 3.10 (Inclusions between L p spaces). Assume that μ is finite.
Then, if 1 ≤ r ≤ s ≤ ∞ we have
and so
ϕr ≤ (μ(X))(s−r)/rs ϕs . (3.12)
−1/ p
By (3.12) we obtain that p → μ(X) ϕ p is nondecreasing for ϕ in
the intersection of the spaces L (X, E , μ), so that it has a limit as p →
p
Then ϕ ∈ L ∞ (X, E , μ) if and only if the limit lim p→∞ ϕ p is finite. If
this is the case, we have that ϕ∞ coincides with the value of the limit.
Proof. If p ≥ 1 we have by the Markov inequality
By several approximations (see Exercise 3.7) one can prove that a convex
function f satisfies g(t x +(1−t)y) ≤ tg(x)+(1−t)g(y) for all x, y ∈ J
and t ∈ [0, 1], and even that
n n
n
g ti xi ≤ ti g(xi ) whenever ti ≥ 0, xi ∈ J and ti = 1.
i=1 i=1 i=1
(3.14)
In the proof we use an elementary property of convex functions g : R →
R satisfying g(t) → +∞ as |t| → +∞, namely the existence of a
minimum point t0 ; moreover, the function g is nondecreasing in [t0 , +∞)
and nonincreasing in (−∞, t0 ] (see Exercise 3.8).
55 Introduction to Measure Theory and Integration
In the general case, let us first assume that g(t) → +∞ as |t| → +∞.
Then, by Exercise 3.8 we know that g has a minimum point t0 , and that
g is nondecreasing in [t0 , +∞), and nonincreasing in (−∞, t0 ]. We can
assume with no loss of generality (possibly replacing g(t) by g(t − t0 )
ϕ by ϕ + t0 ) that g attains its minimum value at t0 = 0, and that
and
X g(ϕ) dμ is finite. Furthermore, replacing g by g−g(0), we can assume
that the minimum value of g is 0.
Let ϕn± be nonnegative simple functions satisfying ϕn± ↑ ϕ ± ; the simple
functions ϕn+ − ϕn− converge to ϕ + − ϕ − = ϕ in L 1 (X, E , μ). In addition,
since g is monotone in (−∞, 0] and [0, +∞), the monotone convergence
theorem gives
g(ϕn+ ) dμ ↑ g(ϕ + ) dμ, g(−ϕn− ) dμ ↑ g(−ϕ − ) dμ,
X X X
+ − + − + −
= 0, ϕ−n ϕn = 0 and ϕ ϕ = 0)
so that +(since g(0) +
X g(ϕn − ϕn ) dμ
−
=
X g(ϕn ) dμ + X g(−ϕn ) converges to X g(ϕ ) dμ + X g(−ϕ ) =
X g(ϕ) dμ. Passing to the limit as n → ∞ in Jensen’s inequality for the
simple functions ϕn+ − ϕn−
+ −
g (ϕn − ϕn ) dμ ≤ g(ϕn+ − ϕn− ) dμ
X X
we get (3.15).
56 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Taking the supremum in the right hand side we obtain Jensen’s inequality.
Both viewpoints are important in the theory of convex functions.
To be more precise, Jensen’s inequality holds provided g is convex on
an interval containing the image of ϕ. The next example is very important
in Probability and Information theory.
Example 3.14 (Entropy functional). By applying Jensen’s inequality
with the convex function g(z) = z ln z in [0, +∞) we obtain
ϕ ln ϕ dμ ≥ ϕ dμ ln ϕ dμ (3.16)
X X X
all ϕ ∈ L (X, E , μ) nonnegative. If X ϕ dμ = 1 we obtain that
1
for
X ϕ ln ϕ dμ ≥ 0 even though the function g has a variable sign (it attains
the minimum value −1/e at z = 1/e).
Proposition 3.16. For any p ∈ [1, +∞) and any finite measure μ, Cb (X)
is dense in L p (X, E , μ).
where
d(x, I ) := inf{|x − y| : y ∈ I }.
It is easy to see that ϕn are continuous, that 0 ≤ ϕn ≤ 1 and that ϕn (x) →
1 I (x), hence the dominated convergence theorem implies that ϕn → 1 I in
L p (X, E , μ).
Now, let
G := {I ∈ B (X) : 1 I ∈ C }.
It is easy to see that G is a Dynkin system (which includes the π–system
of closed sets), so that by the Dynkin theorem we have G = B (X).
Remark 3.17. Cb (X) (or more precisely, the equivalence classes of con-
tinuous bounded functions) is a closed subspace of L ∞ (X, E , μ), and
therefore it is not dense in general. Indeed, if (ϕn ) ⊂ Cb (X) is Cauchy
in L ∞ (X, E , μ), then it uniformly converges, up to a μ-negligible set
B (just take in Remark 3.12 as B the union of the μ–negligible sets
{|ϕn −ϕm | > ϕn −ϕm }). Therefore (ϕn ) uniformly converges on B c and
on its closure K . Denoting by ϕ ∈ Cb (K ) its uniform limit, by Tietze’s
exension theorem we may extend ϕ to a function, that we still denote by
ϕ, in Cb (X). As X \ K ⊂ B is μ–negligible, it follows that ϕn → ϕ in
L ∞ (X, E , μ).
Exercises
3.1 Assume that μ is σ –finite, but not finite. Provide examples showing that no
inclusion holds between the spaces L p (X, E , μ) in general. Nevertheless, show
that for any E –measurable function ϕ : X → R the set
p ∈ [1, ∞] : ϕ ∈ L p (X, E , μ)
3.2 Let 1 ≤ p ≤ q < ∞ and f ∈ L q (X, E , μ). Show that for any δ ∈ (0, 1)
we can write f = g + f˜, with g ∈ L q (X, E , μ), f˜ ∈ L p (X, E , μ) and gq ≤
δ f q (notice that if μ is finite we can take g = 0).
3.3 Let p ∈ (1, ∞), ϕ ∈ L p and ψ ∈ L q , with q = p , be such that ϕψ1 =
ϕ p ψq . Show that either ψ = 0 or there exists a constant λ ∈ [0, +∞) such
that |ϕ| = λ|ψ|q−1 μ–a.e. in X. Hint: first investigate the case of equality in
Young’s inequality.
3.4 Prove the following variant of Hölder’s inequality, known as Young’s in-
equality: if ϕ ∈ L p , ψ ∈ L q and 1p + q1 = r1 , with r ≥ 1, we have that ϕψ ∈ L r
and ϕψr ≤ ϕ p ψq .
3.5 Let (ϕn ) ⊂ L 1 (X, E , μ) be nonnegative and satisfying lim infn ϕn ≥ ϕ μ–
a.e. in X. Show that
ϕn dμ = ϕ dμ = 1 ⇒ |ϕ − ϕn | dμ → 0.
X X X
Hint: notice that the positive part and the negative part of ϕ − ϕn have the same
integral to obtain
|ϕ − ϕn | dμ = 2 (ϕ − ϕn )+ dμ.
X X
Hint: prove first the statement under the additional assumption that ψn → ψ
μ–a.e. in X.
3.7 Show that (3.13) implies g(t x + (1 − t)y) ≤ g(x) + (1 − t)g(y) for all
x, y ∈ J and t ∈ [0, 1]. Then, deduce from this property (3.14). Hint: it is
useful to consider dyadic numbers t = k/2m , with k ≤ 2m integer.
3.8 Let g : R → R be a convex function such that g(z) → +∞ as |z| → +∞.
Show the existence of z 0 ∈ R where g attains its minimum value. Then, show
that g is nondecreasing in [z 0 , +∞) and nonincreasing in (−∞, z 0 ].
3.9 Let (ϕn ) ⊂ L 1 (X, E , μ) be nonnegative functions. Show that the conditions
lim inf ϕn ≥ ϕ μ–a.e. in X, lim sup ϕn dμ ≤ ϕ dμ < ∞
n→∞ n→∞ X X
with (c) := (c)/c, and then choose c sufficiently large, such that M/(c) <
ε/2.
3.11
Assuming that (X, d) is a metric space, E = B (X) and μ is finite, prove
Lusin’s theorem: for any ε > 0 and any f ∈ L 1 (X, E , μ), there exists a closed
set C ⊂ X such that μ(X \ C) < ε and f |C is continuous and bounded. Hint:
use the density of Cb (X) in L 1 and Egorov’s theorem.
Chapter 4
Hilbert spaces
In this chapter we recall the basic facts regarding real vector spaces en-
dowed with a scalar product. We introduce the concept of Hilbert space
and show that, even for the infinite-dimensional ones, continuous linear
functionals are induced by the scalar product. Moreover, we see that even
in some classes of infinite dimensional spaces (the so-called separable
ones) there exists a well-defined notion of basis (the so-called complete
orthonormal systems), obtained replacing finite sums with converging
series. Even though the presentation will be self-contained, we assume
that the reader has already some familiarity with these concepts (basis,
scalar product, representation of linear functionals) in finite-dimensional
spaces.
(iii). Let 2 be the space of all sequences of real numbers x = (xk ) such
∞
that xk2 < ∞. 2 is a vector space with the usual operations,
k=0
(x −yn )+(x −ym )2 +(x −yn )−(x −ym )2 = 2x −yn 2 +2x −ym 2 .
Consequently
$ $2
$ yn + ym $
2 2 $
yn − ym = 2x − yn + 2x − ym − 4 $x −
2 $ .
2 $
Taking into account that (yn + ym )/2 ∈ Y we find
n
n
= x2 − xk2 + (xk − yk )2 .
k=1 k=1
eh , ek = δh,k , h, k ∈ N.
∞
(ii) For any x ∈ H the series x, ek ek is convergent in H (1) .
k=0
(iii) Equality holds in (4.7) holds if and only if
∞
x= x, ek ek . (4.8)
k=0
Inequality (4.7) is called Bessel inequality and when the equality holds,
Parseval identity.
Proof. (i) Let n ∈ N. Then by (4.5) we have
$ $2
$ n $ n
$ $
$x − x, ek ek $ = x −
2
|x, ek |2 , (4.9)
$ k=0
$ k=0
∞
(1) A series x of vectors in a Banach space E is said to be convergent if the sequence of the
i
k=0
n
finite sums xi is convergent in E
k=0
68 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Then
$ $2
$n+ p $
n+ p
$ $
sn+ p − sn 2 = $ x, ek ek $ = |x, ek |2 .
$k=n+1 $ k=n+1
∞
Since the series |x, ek |2 is convergent by (i), the sequence (sn ) is
k=0
Cauchy and the conclusion follows.
Passing to the limit as n → ∞ in (4.9) we find
$ $2
$ ∞ $ ∞
$ $
$x − x, ek ek $ = x −
2
|x, ek |2 .
$ k=0
$ k=0
∞
a → ai ei
i=0
m
n
n
x − x, ei ei ≤ x − x, ei ei ≤ x − ai ei < ε
i=1 i=1 i=1
for m ≥ n. Since ε is arbitrary this proves that the sum of the series is
equal to x.
generating the same vector space generated by (vn ). Let S be this vec-
tor space, and let us represent it as ∪n Sn , where Sn is the vector space
generated by {e0 , . . . , en }. Notice that S is dense, as all vn belong to S.
By applying the Gram-Schmidt process to ei , an operation that does not
change the vector spaces Sn generated by the vectors e0 , . . . , en , we can
also assume that (ei ) is an orthonormal system. Then, Proposition 4.13
gives that (ei ) is complete.
and prove that it induces an Hilbert space structure. The space 2 (C) of
complex-values sequences (z n ) with (|z n |) ∈ 2 (R) is a particular case.
The norm still satisfies the parallelogram identity, so that we can still
prove the existence of orthogonal projections on closed subspaces and its
characterization in terms of
Re x − πY (x), z = 0 ∀z ∈ Y.
Analogously, in Remark 4.4, one has to replace the scalar product by its
real part.
71 Introduction to Measure Theory and Integration
Riesz representation theorem still holds (now for continuous and C-linear
functionals) and the concepts of orthonormal system and complete or-
thonormal system make sense. We have Bessel’s inequality for orthonor-
mal systems and Parseval’s identity for complete orthonormal systems.
Finally, 2 (C) is the canonical model of all separable Hilbert spaces; as
in the real case the correspondence is induced by the choice of a complete
orthonormal system, which provides coordinates of a vector.
We conclude this chapter providing a natural example, considered in
the literature, of non-separable Hilbert space.
Example 4.15 (Quasi-periodic functions). We define the space A P(R)
of almost periodic functions as the closure, with respect to uniform con-
vergence in R, of the vector space generated by complex-valued periodic
functions (of arbitrary period). This space has been extensively studied
by Bochner and Bohr. It is easy to show that the space of almost periodic
functions is not only a vector space (it is a subspace of C(R, C)), but also
an algebra, i.e. f g ∈ A P(R) whenever f, g ∈ A P(R).
If f is almost periodic one can also show (by approximation, taking
into account that this property is linear with respect to f and holds for
periodic functions) that there exists the limit
T
1
M( f ) := lim f (x + t) dt.
T →+∞ 2T −T
f, g A P := M( f ḡ) f, g ∈ A P(R).
eλ (t) = eiλt , t ∈ R.
Exercises
4.1 Let (X, · ) be a normed space, and assume that the norm satisfies the
parallelogram identity (4.2). Set
1 1
x, y := x + y2 − x − y2 , x, y ∈ X.
4 4
Show that ·, · is a scalar product whose induced norm is · . Use this identity
to show that any linear isometry between pre-Hilbert spaces preserves also the
scalar product.
4.2 Show that, in the situation considered in Remark 4.4, π K (x) is characterized
by the property
x − π K (x), z − π K (x) ≤ 0 ∀z ∈ K .
4.3 Let H be a finite dimensional pre-Hilbert space and let {v1 , . . . , vn }, with
n = dim H , be a basis of it. Define
v2 , f 1 v3 , f 1 v3 , f 2
f 1 = v1 , f 2 = v2 − f 1 , f 3 = v3 − f1 − f 2 , ......
f1, f1 f1, f1 f2 , f2
Show that ei = f i / f i is an orthonormal system in H (notice that vk − f k is
the projection of vk on the vector space generated by {v1 , . . . , vk−1 }).
4.4 Let H be a Hilbert space, and let X be an infinite-dimensional separable
subspace. Show that
∞
π X (x) = x, ek ek ∀x ∈ H,
k=0
where(ek ) is any complete orthonormal system of X. Hint: show that the vector
x − k x, ek ek is orthogonal to all vectors of X.
4.5 Let X be the space of functions
f : [0, 1] → R such that f (x) = 0 for at
most countably many x, and x f 2 (x) < +∞. Show that X, endowed with the
scalar product
f, g := f (x)g(x),
x∈[0,1]
is a non-separable Hilbert space.
4.6 Let (ek )k∈N be a complete orthonormal system of H . Show that, for any
x, y ∈ H we have
∞
x, ek y, ek = x, y. (4.10)
k=0
4.7
Show that for any Hilbert space H there exists a family (not necessarily
finite or countable) of vectors {ei }i∈I such that:
(i) ei , e j is equal to 1 if i = j, and to 0 otherwise;
(ii) for any vector x ∈ H there exists a countable set J ⊂ I with
x= x, ei ei .
i∈J
Hint: use Zorn’s lemma.
Chapter 5
Fourier series
and π
1
bk := f (y) sin kydy, k ∈ N, k ≥ 1.
π −π
√
Indeed, it is easily seen that a02 π/2 = ( f, 1/ 2π)2 and, for k ≥ 1,
% &2 % &2
1 1
ak2 π = f, √ cos kx , bk2 π = f, √ sin kx .
π π
π
1
f (x) = an e inx
where an := f (x)e−inx dx.
n∈Z
2π −π
75 Introduction to Measure Theory and Integration
with
⎧ T
⎪ 1
⎪
⎪ f (x) dx if k = 0;
⎪
⎨ T −T
ak :=
⎪
⎪
⎪
⎪ 1 T π
⎩ f (x) cos kx d x if k > 0,
T −T T
1 T
π
bk := f (x) sin kx dx.
T −T T
1 N
SN (x) := a0 + (ak cos kx + bk sin kx), x ∈ [−π, π).
2 k=1
'
f (x + 2πn) = f (x), x ∈ [−π, π), n = ±1, ±2, . . . . (5.4)
Proof. Write
1 N
S N (x) = a0 + (ak cos kx + bk sin kx)
2 k=1
1 π
1 N
= f (y) + (cos kx cos ky + sin kx sin ky) dy
π −π 2 k=1
1 π
1 N
= f (y) + cos k(x − y) dy.
π −π 2 k=1
1 1 N
= sin 2 z + sin k + 1
2
z − sin k − 1
2
z
2 k=1
1
= sin N + 12 z .
2
Therefore
1 N
1 sin N + 12 z
+ cos kz = (5.6)
2 k=1 2 sin 12 z
and so,
π
1 sin N + 12 (x − y)
S N (x) = f (y) dy. (5.7)
2π −π sin 12 (x − y)
Now, setting τ = y − x we get
π−x
1 sin N + 12 τ
SN (x) = '
f (x + τ ) dτ
2π −π−x sin 12 τ
π
1 sin N + 12 τ
= '
f (x + τ ) dτ
2π −π sin 12 τ
since the function under the integral is 2π–periodic. Now, integrating
(5.6) over [−π, π] yields
π
1 sin N + 12 τ
1= dτ,
2π −π sin 12 τ
77 Introduction to Measure Theory and Integration
so that
π
1 sin N + 12 τ 1 0 sin N + 12 τ
dτ = 1 = dτ.
π 0 sin 12 τ π −π sin 12 τ
If we multiply both sides by l and r, and subtract the resulting identities
from (5.7), (5.5) follows.
Proposition 5.3 (Dini’s test). Let x, l, r ∈ R be such that
π '
| f (x + τ ) − Hl,r (τ )|
dτ < ∞. (5.8)
−π | sin(τ/2)|
Then the Fourier series of f converges to (l + r)/2 at x.
Dini’s test shows a remarkable property of the Fourier series: while the
specific value of the coefficients ak and bk depends on the behaviour of f
on the whole interval (−π, π), and the same holds for the Fourier series,
the character of the series (convergent or not) at a given point x depends
only on the behaviour of f in the neighbourhood of x: indeed, it is this be-
haviour that influences the integrability of ( '
f (x +τ )− Hl,r (τ ))/ sin(τ/2)
(the only singularity being at τ = 0).
In the next example we provide sufficient conditions for the conver-
gence of the Fourier series.
Example 5.4. Assume that f : [−π, π] → R is L-Lipschitz continuous,
i.e.
| f (x) − f (y)| ≤ L|x − y| ∀ x, y ∈ [−π, π]
for some L ≥ 0. Then Dini’s test is fulfilled at any x ∈ R \ Zπ choosing
l =r = ' f (x), and at any x ∈ Zπ choosing l = ' f (x− ) and r = '
f (x+ ) (1) .
Indeed, with these choices of l and r, the quotient
'
f (x + τ ) − Hl,r (τ )
sin(τ/2)
is bounded in a neighbourhood of 0.
The same conclusions hold when f is α–Hölder continuous for some
α ∈ (0, 1], i.e.
| f (x) − f (y)| ≤ L|x − y|α , ∀ x, y ∈ [−π, π]
for some L ≥ 0: in this case the quotient is bounded from above, near 0,
by the function L|τ |α /| sin(τ/2)| ∼ 2L|τ |α−1 which is integrable.
(1) here we denote by g(x ), g(x ) the left and right limits of g at x
− +
78 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
we have limk f, ek = 0.
Let us now consider the general case. We know that bounded continu-
ous functions are dense in L 1 (−π, π), hence for any ε > 0 we can find
g ∈ Cb (−π, π) such that f − g1 < ε. As a consequence
| f, ek | = | f − g, ek | + |g, ek | ≤ Mε + |g, ek |
and letting k → ∞ we obtain lim supk | f, ek | ≤ Mε. Since ε is arbit-
rary the proof is achieved.
Proof of Proposition 5.3. Set
'
f (x + τ ) − Hl,r (τ )
g(τ ) := ∈ L 1 (−π, π). (5.10)
sin(τ/2)
Then, writing
1 1 1
sin[(N + )t] = sin N t cos t + cos N t sin t
2 2 2
and applying the Riemann–Lebesgue lemma to g cos t/2 (with e N =
sin N t) and to g sin(t/2) (with e N = cos N t) we obtain from (5.5) that
SN (x) converge to (l + r)/2.
79 Introduction to Measure Theory and Integration
For instance, taking f (x) = x one finds the following nice relation
between π and the harmonic series with exponent 2:
∞
1 π2
= .
k=1
k2 6
Notice that (5.11) provides, for any f ∈ L 2 (−π, π), the existence of
a subsequence N (k) such that SN (k) f (x) → f (x) for L 1 –a.e. x ∈
80 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
1 ∞
f (x) = a0 + (ak cos kx + bk sin kx) ∀x ∈ [−π, π].
2 k=1
Notice that a0 = 0 because f (−π) = f (π) implies that the mean value
of f on (−π, π) is 0. As easily checked through an integration by parts
(using again the fact that f (−π) = f (π)), we have ak = kbk and bk =
−kak . Then, by the Bessel inequality it follows that
∞ ∞ π
2 2 1
k 2
(ak2 + bk2 ) = (ak ) + (bk ) ≤ | f (x)|2 dx < ∞. (5.14)
k=1 k=1
π −π
Exercises
5.1 Check that the trigonometric system (5.1) is orthogonal.
5.2 Let E be a Banach space. Show that any totally convergent series n xn ,
with (xn ) ⊂ E, is convergent. Moreover,
$∞ $ ∞
$ $
$ xn $ ≤ xn . (5.15)
n=0 n=0
N M
Hint: estimate 0 xn − 0 xn with the triangle inequality.
5.3 Prove that the following systems on L 2 (0, π) are orthonormal and complete
(
2
sin kx, k ≥ 1,
π
and (
1 2
√ ; cos kx, k ≥ 1.
π π
5.4 Show that
1
ek (x) := √ eikx , k∈Z
2π
is a complete orthonormal system in L 2 ((−π, π); C). Hint: in order to show
completeness, consider first the cases where f is real-valued or i f is real-valued.
82 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
5.5 Let (ek ) be as in Exercise 5.4. Using the Parseval identity show that
π 2
π 1
| f (x)|2 dx = f (x)e−ikx dx ∀ f ∈ L 2 ((−π, π); C) .
−π 2π k∈Z −π
N
5.6 Let f ∈ L 2 ((−π, π); C) and let S N f = −N f, ek ek , with N ≥ 1, be the
Fourier sums corresponding to the complete orthonormal system in Exercise 5.4.
Show that
π
f (x) − S N f (x) = G N (x − y)( f (x) − f (y)) dy
−π
with
sin((N + 1/2)z)
G N (z) := .
sin(z/2)
Hint: use the identities 0N eiky = 0N (eiy )k = (ei(N +1)z − 1)/(eiy − 1).
−4
5.7 Arguing as in Remark 5.7, show that ∞ 1 k = π 4 /90. Hint: consider the
function f (x) = x . 2
5.8 Chebyschev polynomials Cn in L 2 (a, b), with (a, b) bounded interval, are
the ones obtained by applying the Gram-Schmidt procedure to the vectors 1, x,
x 2 , x 3 , . . .. They are also called Legendre polynomials when (a, b) = (−1, 1).
(a) Compute explicitly the first three Legendre polynomials.
(b) Show that {Cn }n∈N is a complete orthonormal system. Hint: use the density
of polynomials in C([a, b]).
(c)
Show that the n-th Legendre polynomial Pn is given by
(
2n + 1 1 d n 2
Pn (x) = (x − 1)n .
2 2n n! d n x
5.9 Let f ∈ C m [−π, π]; C with f ( j) (−π) = f ( j) (π) for all j = 0, . . . , m −
1. Show that ck(m) , the k-th Fourier coefficient of f (m) is linked to ck , the k-th
Fourier coefficient of f , by ck(m) = (ik)m ck .
Chapter 6
Operations on measures
x → ν(E x ), y → μ(E y ),
B if x ∈ A A if y ∈ B
Ex = E =
y
∅ if x ∈ / A, ∅ if y ∈ / B.
Consequently,
so (i), (ii) and (iii) follow from Proposition 6.1. Consequently, by lin-
earity, (i)–(iii) hold when F is a simple function. If F is general, it
86 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
= f (x) dμ(x).
X
Notice that, strictly speaking, the function in (ii) is defined only out of
a μ–negligible set; by μ–integrability of it we mean μ–integrability of
any F –measurable
extension of it (for instance we may set it equal to 0
wherever Y |F(x, y)| dν(y) is not finite).
Remark 6.7 (Finite products). The previous constructions extend with-
out any difficulty to finite products of measurable spaces (X i ,F i ). Name-
ly, the product σ -algebra F := × n
i F i in the cartesian product X :=
× n
1 X i is generated by the rectangles
{A1 × · · · × An : Ai ∈ F i , 1 ≤ i ≤ n} .
87 Introduction to Measure Theory and Integration
and any permutation in the order of the integrals would produce the same
result. Finally, the product measure is uniquely determined, in the σ –
finite case, by the product rule
n
μ (A1 × · · · × An ) = μi (Ai ) Ai ∈ F i , 1 ≤ i ≤ n.
i=1
It is also not hard to show that the product is associative, both at the level
of σ –algebras and measures, see Exercise 6.1.
× n
1 L 1 . We say that L n is the Lebesgue measure on Rn .
Since (see Exercise 6.2)
n
B (Rn ) = × B (R),
i=1
Indeed, we have
π n/2
ωn = . (6.5)
( n2 + 1)
A proof of this formula, based on the identity (z + 1) = z(z) (which
gives also (n) = (n − 1)! for n ≥ 1 integer) is proposed in Exercise 6.7.
We are going to show that L n is invariant under translations and rota-
tions. For this we need some notation. For any a ∈ Rn and any δ > 0 we
set
Q(a, δ) : = x ∈ Rn : ai ≤ xi < ai + δ, ∀ i = 1, . . . , n
n
= ×[a , a + δ).
i=1
i i
It is also clear that each box in Q N is Borel and that its Lebesgue measure
is 2−n N . Now we set
∞
Q = Q N.
N =0
It is clear that all boxes in Q N are mutually disjoint and that their union
is Rn . Furthermore, if N < M, Q ∈ Q N and Q ∈ Q M , then either
Q ⊂ Q or Q ∩ Q = ∅. If follows that if Q, Q ∈ Q intersect, then
one of the two sets is contained in the other one.
Lemma 6.9. Let U be a non empty open set in Rn . Then U is the disjoint
union of boxes in Q .
Proof. For any x ∈ U , let Q x ∈ Q be the biggest box such that x ∈
Q x ⊂ U . This box is uniquely defined: indeed, fix an x; for any m there
is only one box Q x,m ∈ Q m such that x ∈ Q x,m ; moreover, since U is
open, for m large enough Q x,m ⊂ U ; we can then define Q x = Q x,m̃
where m̃ is the smallest integer m such that Q x,m ⊂ U .
89 Introduction to Measure Theory and Integration
E + x = {y + x : y ∈ Rn }.
μ(Q N ) = Cμ L n (Q N ).
So, Lemma 6.9 gives that μ(A) = Cμ L n (A) for any open set, and there-
fore for any Borel set.
90 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
This definition is well posed, again thanks to the fact that In,A = Im,B
with n < m when B = A × X n+1 × · · · × X m . It is easy to check that μ is
additive: indeed, if In,A and Im,B are disjoint, using the previous remark
we can assume with no loss of generality that n = m, and therefore the
equality μ(In,A ∪ In,B ) = μ(In,A ) + μ(In,B ) follows by
n n n
×μ
k=1
k (A ∪ B) = ×μ
k=1
k (A) + ×μ
k=1
k (B).
Set now
+ ε0 ,
F j,1 = x1 ∈ X 1 : μ(1) (E j (x1 )) ≥ , j ≥ 1.
2
Then F j,1 is not empty and by (6.8) we have
(1)
μ(E j ) = μ (E j (x1 )) dμ1 (x1 ) + μ(1) (E j (x1 )) dμ1 (x1 )
F j,1 c
F j,1
ε0
≤ μ1 (F j,1 ) + .
2
Therefore μ1 (F j,1 ) ≥ ε0 /2 for all j ≥ 1.
Obviously (F j,1 ) is a nonincreasing sequence of subsets of X 1 . Since
μ
1∞is σ –additive, it is continuous at 0. Therefore, there exists α1 ∈
1 F j,1 and so
ε0
μ(1) (E j (α1 )) ≥ , j ≥ 1. (6.9)
2
Consequently we have
E j (α1 ) = ∅, j ≥ 1. (6.10)
We set
+ ε0 ,
F j,2 = x2 ∈ X 2 : μ(2) (E j (α1 , x2 )) ≥ , j ≥ 1.
4
93 Introduction to Measure Theory and Integration
E j ((α1 , α2 )) = ∅. (6.12)
Exercises
6.1 Let (X 1 , F 1 ), (X 2 , F 2 ), (X 3 , F 3 ) be measurable spaces. Show that
(F 1 × F 2 ) × F 3 = F 1 × (F 2 × F 3 ).
L 1 × L 1 L 2.
94 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Hint: to show the strict inclusion, consider the set E = F × {0}, where F ⊂ R
is not Lebesgue measurable.
6.4 Show that the product σ –algebra is also generated by the family of products
× ∞
1 Ai where Ai ∈ F i and Ai = X i only for finitely many i.
6.5 Writing properly L 3 as a product measure, compute L 3 (T ), where
+ ,
T = (x, y, z) : x 2 + y 2 < r 2 and y 2 + z 2 < r 2 .
6.9 Using the notation of the Fubini-Tonelli theorem, let X = Y = [0, 1],
F = G = B ([0, 1]), let μ be the Lebesgue measure and let ν be the counting
measure.
Let D = {(x,
x) y: x ∈ [0, 1]} be the diagonal in X × Y ; check that
X ν(D x ) dμ(x) = Y μ(D ) dν(y).
6.10
Let ( f h ) be converging to f in L 1 (X × Y, μ × ν). Show the existence
of a subsequence h(k) such that f h(k) (x, ·) converge to f (x, ·) in L 1 (Y, ν) for
μ–a.e. x ∈ X. Show by an example that, in general, this property is not true for
the whole sequence.
and
μ(A) = 0 ⇒ ν(A) = 0.
For finite measures, the absolute continuity property can also be given in
a (seemingly) stronger way, see Exercise 6.14.
The following theorem shows that absolute continuity of ν with respect
to μ is not only necessary, but also sufficient to ensure the representation
ν = f μ.
Theorem 6.14 (Radon–Nikodým). Let μ and ν be finite measures on
(X, F ) such that ν # μ. Then there exists a unique nonnegative ρ ∈
L 1 (X, F , μ) such that
ν(E) = ρ(x) dμ(x) ∀E ∈ F . (6.16)
E
96 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
We are going to show a more general result, whose statement needs two
more definitions. We say that a measure μ is concentrated on a F –
measurable set A if μ(X \ A) = 0. For instance, the Dirac measure δa
is concentrated on {a}, and the Lebesgue measure in R is concentrated
on the irrational numbers, and f μ is concentrated (whatever μ is) on
{ f = 0}.
Definition 6.15 (Singular measures). Let μ, ν be measures in (X, F ).
We say that μ is singular with respect to ν, and write μ ⊥ ν, if there exist
disjoint F –measurable sets A, B such that μ is concentrated on A and
ν is concentrated on B.
The relation of singularity, as stated, is clearly symmetric. However, it
can also be stated in a (seemingly) asymmetric way, by saying that μ ⊥ ν
if μ is concentrated on a ν–negligible set A (just take B = Ac to see the
equivalence with the previous definition).
Example 6.16. Let X = R, F = B (R), μ the Lebesgue measure on
(X, F ) and ν = δx0 the Dirac measure at x0 ∈ R. Then μ is concentrated
on A := R \ {x0 }, whereas ν is concentrated on B := {x0 }. So, μ and ν
are singular.
Theorem 6.17 (Lebesgue). Let μ and ν be measures on (X, F ), with μ
σ –finite and ν finite. Then the following assertions hold.
(i) There exist two unique finite measures νa and νs on (X, F ) such that
ν = νa + νs , νa # μ, νs ⊥ μ. (6.17)
f (x)
ρ(x) := lim [ f (x) + f 2 (x) + · · · + f n+1 (x)] = , x ∈ A.
n→∞ 1 − f (x)
∞
X= Xn with μ(X n ) < ∞.
n=0
∞
∞
∞
νa := (νn )a , νs := (νn )s , ρ := ρn 1 X n .
n=0 n=0 n=0
Since
k
k
(νn )a + (νn )s = νn = 1∪k X n ν,
0
n=0 n=0
99 Introduction to Measure Theory and Integration
ν = νa + νs = νa + νs
and let B, B be μ–negligible sets where νs and νs are respectively con-
centrated. Then, as B ∪ B is μ–negligible and both νs and νs are con-
centrated on B ∪ B , for any set E ∈ F we have
Exercises
6.11 Show that a F –measurable function h is f μ–integrable if and only if f h
is μ–integrable.
6.12 Show that ( f μ)∨(gμ) = ( f ∨g)μ and ( f μ)∧(gμ) = ( f ∧g)μ whenever
f, g ∈ L 1 (X, F , μ) are nonnegative.
6.13 Let {μi }i∈I be a family of measures in (X, F ). Show that
∞
μ(B) := inf μi(k) (Bk ) : i : N → I,
k=0
(Bk ) countable F –measurable partition of B
is the greatest lower bound of the family {μi }i∈I , i.e. μ ≤ μi for all i ∈ I and it
is the largest measure with this property. Show also that
∞
μ(B) := sup μi(k) (Bk ) : i : N → I,
k=0
(Bk ) countable F –measurable partition of B
is the smallest upper bound of the family {μi }i∈I , i.e. μ ≥ μi for all i ∈ I and
it is the smallest measure with this property.
6.14 Let μ, ν be measures in (X, F ) with ν finite. Then ν # μ if and only if
for all ε > 0 there exists δ > 0 such that
Finally choose t = 2−h , with h sufficiently large so that ξh < μ(A) and B =
A \ Bh , to get a contradiction with (6.21).
Proposition 6.20. Let μ be a signed measure and let |μ| be its total vari-
ation. Then |μ| is a finite measure on (X, F ).
102 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Aj = A ∩ E j, Bj = B ∩ E j, j ∈ N.
∞
∞
∞
|μ(E j )| ≤ |μ(A j )| + |μ(B j )| ≤ |μ|(A) + |μ|(B),
j=0 j=1 j=0
∞
ε
∞
ε
|μ(Aεk )| ≥ |μ|(A) − , |μ(Bkε )| ≥ |μ|(B) − .
k=0
2 k=0
2
∞
|μ(X n )| > 2(1 + |μ(X)|).
n=0
Then either the sum of those μ(X n ) which are nonnegative or the absolute
value of the sum of those μ(X n ) which are nonpositive is greater than
1 + |μ(X)|. To fix the ideas, assume that for a subsequence (X n(k) ) we
have μ(X n(k) ) ≥ 0 and
∞
μ(X n(k) ) > 1 + |μ(X)|.
k=0
∞
Set A = 0 X n(k) and B = Ac . Then we have |μ(A)| > 1 + |μ(X)| and
Since
|μ|(X) = |μ|(A) + |μ|(B) = ∞,
either |μ|(B) = +∞ or |μ|(A) = +∞. In the first case we are done, in
the second one we exchange A and B. So, the claim is proved and the
proof is complete.
1 1
μ+ := (|μ| + μ), μ− := (|μ| − μ),
2 2
so that
μ = μ+ − μ− and |μ| = μ+ + μ− . (6.25)
The measure μ+ (respectively μ− ) is called the positive part (respectively
negative part) of μ and the first equation in (6.25) is called the Jordan
representation of μ.
104 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Remark 6.21. It is easy to check that Theorems 6.17 and 6.14 hold when
ν is a signed measure: it suffices to split it into its positive and negative
part, see also Exercise 6.19.
The following theorem proves also that μ+ and μ− are singular, and
provides a canonical representation of μ± as suitable restrictions of ±μ.
and the second inequality is strict if |μ|(E 1 ) > 0, we have that |μ|(E 1 ) =
0. In a similar way one can prove that |μ|(F1 ) = 0, so that |h| ≤ 1 |μ|–
a.e. in X. Now, let r ∈ (0, 1) and set
Therefore
∞
|μ(G r,k )| ≤ r|μ|(G r ),
k=0
and
− 1 1
μ (E) = (|μ|(E) − μ(E)) = (1 − h)d|μ|
2 2 E
=− hd|μ| = −μ(E ∩ B).
E∩B
Exercises
6.19 Using the decomposition of ν in positive and negative part, show that Le-
besgue decomposition is still possible when μ is σ –finite and ν is a signed meas-
ure. Using the Hahn decomposition extend this result to the case when even μ
is a signed measure. Are these decompositions unique?
6.20 Show that | f μ| = | f |μ for any f ∈ L 1 (X, E , μ).
6.6. Measures in R
In this section we estabilish a 1-1 correspondence between finite Borel
measures in R and a suitable class of nondecreasing functions. In one
direction this correspondence is elementary, and based on the concept of
repartition function.
Given a finite measure μ in (R, B (R)), we call repartition function of
μ the function F : R → [0, +∞) defined by
F(x) := μ ((−∞, x]) x ∈ R.
Notice that obviously (1) F is nondecreasing, right continuous, and satis-
fies
lim F(x) = 0, lim F(x) ∈ [0, +∞). (6.28)
x→−∞ x→+∞
Moreover, F is continuous at x if and only if x is not an atom of μ.
(1) The arguments are similar to those used in Section 2.4.2, in connection with the properties of the
function t → μ({ϕ > t})
106 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
The following result shows that this list of properties characterizes the
functions that are repartition functions of some finite measure μ; in addi-
tion the measure is uniquely determined by its repartition function.
Theorem 6.23. Let F : R → [0, +∞) be a nondecreasing and right
continuous function satisfying (6.28). Then there exists a unique finite
measure μ in (R, B (R)) such that F is the repartition function of μ.
Proof. The proof follows the same lines of the construction of the Le-
besgue measure in Section 1.6, with a simplification due to the fact that
we can also consider unbounded intervals (because we are dealing with
finite measures). We set
I := {(a, b] : a ∈ [−∞, +∞), b ∈ R, a < b}
and denote by A the ring generated by I : it consists, as it can be easily
checked, of all finite disjoint unions of intervals in I . We define, with
the convention F(−∞) = 0,
μ((a, b]) := F(b) − F(a) ∀(a, b] ∈ I . (6.29)
This definition is justified by the fact that, if μ were a measure and F
were its repartition function, (6.29) would be valid, because (a, b] =
(−∞, b] \ (−∞, a]. Then we extend μ to A with the same mechan-
ism used in the proof of Theorem 1.21, and check that μ is additive on
A . Also, the same argument used in that proof shows that μ is even σ –
additive: in order to prove that μ(F) = i μ(Fi ) whenever F and all Fi
belong to A one first reduces to the case when F = (a, b] belongs to I ;
then, one enlarges Fi to Fi ∈ A with μ(Fi ) < μ(Fi ) + δ2−i and, using
the fact that all intervals [a , b] with a > a are contained in a finite union
of the sets Fi , obtains
∞
∞
μ((a , b]) ≤ μ(Fi ) ≤ 2δ + μ(Fi ).
i=0 i=0
δ ↓ 0 and then a ↓ a we obtain the σ –subadditivity property
Letting first
μ(F) ≤ i μ(Fi ), and the opposite inequality follows by monotonicity.
By the Carathéodory theorem μ has a unique extension, that we still
denote by μ, to B (R) = σ (A ). Setting a = −∞ and letting b tend to
+∞ in the identity (6.29) we obtain that μ(R) = F(+∞) ∈ R. From
(6.29) with a = −∞ we obtain that the repartition function of μ is F.
Given a nondecreasing and right continuous function F satisfying (6.28),
the Stieltjes integral
f dF
R
107 Introduction to Measure Theory and Integration
is defined as f dμ F , where μ F is the finite measure built in the pre-
theorem. The notation d F is justified by the fact that, when f =
vious
i z i 1(ai ,bi ] , we have (by the very definition of μ F )
f dF = f dμ F = z i (F(bi ) − F(ai )).
R R i
This approximation of the Stieltjes integral will play a role in the proof
of Theorem 6.28.
lim μh (−∞, x]) = μ ((−∞, x]) with at most countably many exceptions.
h→∞
(6.30)
Since the repartition function is right continuous, it is uniquely determ-
ined by (6.30). Then, since the measure is uniquely determined by its
repartition function, we obtain that the weak limit, if exists, is unique.
The following fundamental example shows why we admit at most count-
ably many exceptions in the convergence of the repartition functions.
Example 6.25. [Convergence tothe Dirac mass] Let ρ ∈ C ∞ (R) be
a nonnegative function such that R ρ dx = 1 (an important example is
2
the Gauss function (2π)−1/2 e−x /2 ). We consider the rescaled functions
ρh (x) = hρ(hx) and the induced measures μh = ρh L 1 , all probability
measures. Then, it is immediate to check that μh weakly converge to δ0 :
for x > 0 we have indeed
x hx
μh ((−∞, x]) = ρh (y) dy = ρ(y) dy → 1
−∞ −∞
μi (R \ J ) ≤ ε ∀i ∈ I.
Clearly any finite family of probability measures is tight. One can also
check (see Exercise 6.24) that {μi }i∈I is tight if and only if
lim sup Fh(k) (x) ≤ inf lim sup Fh(k) (q) = inf G(q) = G(x),
k→∞ q∈Q, q>x k→∞ q∈Q, q>x
and analogously
lim inf Fh(k) (x) ≥ sup lim inf Fh(k) (q) = sup G(q) = G(x).
k→∞ q∈Q, q<x k→∞ q∈Q, q<x
109 Introduction to Measure Theory and Integration
and analogously
! !
! !
! g dμ −
! f dμ!! ≤ Mε + ε = (M + 1)ε. (6.35)
R (−t,t]
Since
n−1
f dμh = g(ti ) [Fh (ti+1 ) − Fh (ti )]
(−t,t] i=1
n−1
→ g(ti ) [F(ti+1 ) − F(ti )] = f dμ,
i=1 (−t,t]
adding and subtracting (−t,t] f dμh , and using (6.34) and (6.35), we con-
clude that
! !
! !
lim sup !! g dμh − g dμ!! ≤ (M + 1)ε.
h→∞ R R
Exercises
6.21 Show that the probability measures
1 h
μh := δi
h i=1 h
where h ∈ N and a−h , . . . , ah ≥ 0. Show that for any finite Borel measure μ
in R there exists a sequence of measures (μh ) of the previous form that weakly
converges to μ.
6.24 Show that a family {μi }i∈I of probability measures in R is tight if and only
if (6.31) holds.
6.25 Show that any sequence (μh ) of probability measures weakly convergent
to a probability measure is tight. Hint: if μ is the weak limit and ε > 0 is
given, choose an integer n ≥ 1 such that μ([1 − n, n − 1]) > 1 − ε and points
x ∈ (−n, 1 − n) and y ∈ (n − 1, n) where the repartition functions of μh are
converging to the repartition function of μ.
6.26 We want to extend what was shown in this section from the realm of prob-
ability measures to that of finite measures. Let (μh ), μ be finite positive Borel
measures on R, and let Fh , F be their repartition functions. Consider the fol-
lowing implications:
(a) limh R g dμh = R g dμ ∀g ∈ Cb (R) (that is, (6.32));
(b) limh R g dμh = R g dμ ∀g ∈ Cc (R);
(c) Fh converge to F at all points where F is continuous;
(d) Fh converge to F on a dense subset of R;
(e) limh μh (R) = μ(R);
(f) (μh ) is tight.
Find an example where (b) holds but (a), (c), (e) do not hold and prove the
following implications: a ⇒ b, e, a ⇒ c, d ⇔ c, b ∧ e ⇒ c, d ∧ e ⇒ f ,
d ∧ f ⇒ e, d ∧ f ⇒ a. As a corollary, if (e) holds (as it happens in the case when
all μh and μ are probability measures) we obtain that a ⇔ b ⇔ c ⇔ d ⇒ f .
112 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
with
1 T π π π π
an = f (x)e−in T x dx, ein T x = cos n x + i sin n x.
2T −T T T
(6.39)
Remark 6.32. (Inverse Fourier transform) For g ∈ L 1 (R, C) we de-
fine inverse Fourier transform of f the function
1
g̃(x) := g(ξ )ei xξ dξ x ∈ R.
2π R
It can be shown (see for instance Chapter VI.1 in [7]) that the maps f →
fˆ and g → g̃ are each the inverse of the other in the so-called Schwarz
space S(R, C) of smooth and rapidly decreasing functions at infinity:
∞
S(R, C) := f ∈ C (R, C) : lim |x| |D f |(x) = 0 ∀k, i ∈ N .
k i
|x|→∞
In particular we have
f
f (x) = 2π (x) = aξ ei xξ dξ with
2π R
1
aξ : = f (x)e−iξ x dx.
2π R
These formulas can be viewed as the continuous counterpart of the dis-
crete Fourier transform (6.38), (6.39). In this sense, aξ are generalized
Fourier coefficients, corresponding to the “frequency” ξ . The difference
with Fourier series is that any frequency is allowed, not only the integer
multiples nπ/T of a given one.
(2) This computation can be done using the residue theorem in complex analysis
115 Introduction to Measure Theory and Integration
1 2
ρσ (x − y) = ρσ (w)e−iw(x−y)/σ dw.
(2πσ )2 1/2
R
(6.41)
As a consequence, the integrals h σ (y) = R ρσ (y−x) dμ(x) are uniquely
determined by μ̂. But, still using the Fubini-Tonelli theorem, one can
check the identity
g(y)ρσ (x − y) dy dμ(x) = h σ (y)g(y) dy ∀g ∈ Cb (R).
R R R
(6.42)
Passing to the limit as σ ↓ 0 and noticing that (by Example 6.25, that
provides the weak convergence of ρσ λ to δ0 as σ ↓ 0, or a direct verific-
ation)
g(y)ρσ (x − y) dy = g(x − z)ρσ (z) dz → g(x) ∀x ∈ R
R R
from
the dominated convergence theorem we obtain that all integrals
R g dμ, for g ∈ Cb (R), are uniquely determined. Hence μ is uniquely
determined by its Fourier transform.
Remark 6.36. It is also possible to show an explicit inversion formula
for the Fourier transform. Indeed, (6.42) holds not only for continuous
functions, but also for bounded Borel functions;
choosing a < b that are
not atoms of μ and g = 1(a,b) , we have that R g(x)ρσ (x − y) dy → g(x)
for μ–a.e. x (precisely for x ∈/ {a, b}), so that (6.42) and (6.41) give
b b −w2 /2σ 2
e w 2
μ((a, b)) = lim h σ (y) dy = lim μ̂( 2 )eiyw/σ dwdy.
σ ↓0 a σ ↓0 a R 2πσ 2 σ
116 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
The following theorem, due to Lévy, gives essentially the converse im-
plication, allowing to deduce the weak convergence from the convergence
of the Fourier transforms.
Theorem 6.37 (Lévy). Let (μh ) be probability measures in R. If f h =
μ̂h pointwise converge in R to some function f , and if f is continuous at
0, then f = μ̂ for some probability measure μ in R and μh → μ weakly.
Proof. Let us show first that (μh ) is tight. Fixed a > 0, taking into
account that sin ξ is an odd function and using the Fubini theorem we get
a a a
−i xξ
σ̂ (ξ ) dξ = e dσ (x)dξ = cos(xξ ) dξ dσ (x)
−a −a R R −a
2
= sin(ax) dσ (x)
R x
for any probability measure σ . Hence, using the inequalities | sin t| ≤ |t|
for all t and | sin t| ≤ |t|/2 for |t| ≥ 2, we get
1 a sin(ax)
1 − σ̂ (ξ ) dξ = 2 − 2 dσ (x)
a −a R ax
sin(ax)
=2 1− dσ (x) (6.45)
R ax
) *
2 2
≥ σ R\ − , .
a a
For ε > 0 we can find, by the continuity of f at 0, a > 0 such that
a
(1 − f (ξ )) dξ < εa.
−a
a
As a −1 −a (1 − μ̂h (ξ )) dξ → 0 as a ↓ 0 for any fixed h, we infer that
we can find b ∈ (0, a] such that (6.46) holds with b replacing a for all
h ∈ N. From (6.45) we get μh (R \ [−n, n]) < ε for all h ∈ N, as soon
as n > 2/b.
Being the sequence tight, we can extract a subsequence (μh(k) ) weakly
converging to a probability measure μ and deduce from (6.44) that f =
μ̂. It remains to show that the whole sequence (μh ) weakly converges to
μ: if this is not the case there exist ε > 0, g ∈ Cb (R) and a subsequence
h (k) such that
! !
! !
! g dμh (k) − g dμ!! ≥ ε ∀k ∈ N.
!
R R
But, possibly extracting one more subsequence, we can assume that μh (k)
weakly converge to a probability measure σ ; in particular
! !
! !
! g dσ − g dμ! ≥ ε > 0. (6.47)
! !
R R
Exercises
6.27 Check the identity (6.37).
6.28
Show that μ̂ is uniformly continuous in R for any finite measure μ.
118 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
6.29 Let μ be a probability measure in R. Show that if |μ̂| attains its maximum
at ξ0 = 0, then there exist x0 ∈ R and cn ∈ [0, ∞) such that
2nπ
μ= cn δxn with xn = x0 + .
n∈Z
ξ0
Use this fact to show that |μ̂| ≡ 1 in R if and only if μ is a Dirac mass.
Chapter 7
The fundamental theorem of the integral
calculus
In this section we give a closer look at a classical theme, namely the fun-
damental theorem of the integral calculus, looking for optimal conditions
on f ensuring the validity of the formula
x
f (x) − f (y) = f (s) ds.
y
Notice indeed that in the classical theory of the Riemann integration there
is agap between the conditions imposed to give a meaning to the integ-
x
ral a g(s) ds (i.e. Riemann integrability of g) and those that ensure its
differentiability as a function of x (for instance, typically one requires
the continuity of g). We will see that this gap basically disappears in Le-
besgue’s theory, and that there is aprecise characterization of the class
x
of functions representable as c + a g(s) ds for a suitable (Lebesgue)
integrable function g and for some constant c.
The following definition is due to Vitali.
Definition 7.1 (Absolutely continuous functions). Let I ⊂ R be an in-
terval. We say that f : I → R is absolutely continuous if for any ε > 0
there exists δ > 0 for which the implication
n
n
(bi − ai ) < δ ⇒ | f (bi ) − f (ai )| < ε (7.1)
i=1 i=1
holds for any finite family {(ai , bi )}1≤i≤n of pairwise disjoint intervals
contained in I .
An absolutely continuous function is obviously uniformly continuous,
but the converse is not true, see Example 7.7.
Let f : [a, b] → R be absolutely continuous. For any x ∈ [a, b] define
n
F(x) = sup | f (xi ) − f (xi−1 )|,
σ ∈a,x i=1
2
F(x) ≤ (x − a) + 1.
δ
We set
1 1
f + (x) = (F(x) + f (x)), f − (x) = (F(x) − f (x)),
2 2
so that
n
F(y) ≥ | f (y) − f (x)| + | f (xi ) − f (xi−1 )|.
i=1
F(ai )| < 2ε. For any i = 1, . . . , n we can find σi = {ai = x0,i < x1,i <
· · · < xni ,i = bi } such that
ε ni
F(bi ) − F(ai ) < + | f (xk,i ) − f (xk−1,i )|, 1 ≤ i ≤ n. (7.2)
n k=1
ε mi
F(bi ) < + | f (yk ) − f (yk−1 )|
n k=1
n
(F(bi ) − F(ai )) < ε + ε = 2ε.
i=1
Theorem 7.4 (Vitali covering theorem). Let {Bri (xi )}i∈I be a finite
family of balls in a metric space (X, d). Then there exists J ⊂ I such
that the balls {Bri (xi )}i∈J are pairwise disjoint, and
Bri (xi ) ⊂ B3ri (xi ). (7.5)
i∈I i∈J
123 Introduction to Measure Theory and Integration
[min f, max f ].
B r (x) B r (x)
The previous theorem tells us that the same convergence occurs, for L n –
a.e. x ∈ Rn , for any integrable function f . This simply follows by the
124 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
inequality
! ! ! !
! 1 ! ! !
! f (y) dy − f (x)!= 1 ! f (y) − f (x) dy !
!ω rn ! ωn r n ! !
n Br (x) Br (x)
1
≤ | f (y) − f (x)| dy.
ωn r n Br (x)
By the local nature of this statement, the same property holds for locally
integrable functions.
Proof of Theorem 7.5. Given ε, δ > 0 and an open ball B = B R (0), it
suffices to check that the set
1
A := x ∈ B : lim sup | f (y) − f (x)| dy > 2ε
r↓0 ωn r n Br (x)
has Lebesgue measure less than (3n + 1)δ. To this aim, we write f as the
sum of a “good” part g and a “bad”, but small, part h, i.e. f = g + h
with g : B → R bounded and continuous, and h L 1 (B ) < εδ, with
B = B R+1 (0); this decomposition is possible, because Proposition 3.16
ensures the density of bounded continuous functions in L 1 (B).
The continuity of g gives
1
lim |g(y) − g(x)| dy = 0 ∀x ∈ B.
r↓0 ωn r n B (x)
r
is lower semi continuous, hence A3 is open. Notice also that for any
x ∈ A3 there exists r ∈ (0, 1), depending on x, such that
|h(y)| dy > εωn r n .
Br (x)
This is a closed subspace of the complete metric space C([0, 1]), hence
X is complete as well. For any f : [0, 1] → R we set
⎧
⎪
⎨ f (3x)/2 if 0 ≤ 3x ≤ 1,
T f (x) := 1/2 if 1 < 3x < 2, (7.8)
⎪
⎩
1/2 + f (3x − 2)/2 if 2 ≤ 3x ≤ 3.
of these intervals is Cantor’s middle third set (see also Exercise 1.8), and
since
∞ n
∞
2n−1 1 2
L 1 (A) = n
= =1
n=1
3 2 n=1
3
Exercises
7.1 Let H : R → R be satisfying the Lipschitz condition
Example 8.2. Let F : R → R be of class C 1 and such that F (t) > 0 for
all t ∈ R. Let A be the image of F (an open interval, by the assumptions
made on F) and let ψ : A → R be continuous. Then for any interval
[a, b] ⊂ A the following elementary formula of change of variables holds
(just put y = F(x) in the right integral):
F −1 (b) b
1
ψ(F(x)) dx = ψ(y) −1 dy.
F −1 (a) a F (F (y))
On the other hand, choosing ϕ = ψ 1 I with I = [a, b] in (8.2), we have
F −1 (b) b
ψ(F(x)) dx = ψ(y) d F# L 1 .
F −1 (a) a
(1) L(Rn ; Rm ) is the Banach space of all linear mappings T : Rn → Rm endowed with the sup norm
T = sup{|T x| : x ∈ Rn , |x| = 1}
131 Introduction to Measure Theory and Integration
Lemma 8.3. The image F(C F ) of the critical set is Lebesgue negligible.
Proof. Let K ⊂ C F be a compact set and ε > 0; for any x ∈ K the set
D F(x)(B 1 (0)) is Lebesgue negligible (because D F is singular at x, so
that D F(x)(Rn ) is contained in a (n − 1)-dimensional subspace of Rn ),
hence we can find δ = δ(ε, x) > 0 such that
L n {z ∈ Rn : dist (z − F(x), D F(x)(B 1 (0))) < δ} < ε.
On the other hand, since |F(y) − F(x) − D F(x)(y − x)| < δr in Br (x),
provided r is small enough, we get
F(Br (x)) ⊂ z ∈ Rn : dist (z − F(x), D F(x)(B r (0)) < δr .
It follows that Br (x) ⊂ U and L n (F(Br (x))) < εr n for r > 0 small
enough, depending on x.
132 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Since the family of balls {Br/3 (x)}x∈K covers the compact set K , we can
find a finite family {Bri /3 (xi )}i∈I whose union still covers K and extract
from it, thanks to Vitali’s covering theorem, a subfamily {Bri /3 (xi )}i∈J
made by pairwise disjoint balls such that the union of the enlarged balls
{Bri (xi )}i∈J covers K . In particular, covering F(K ) by the union of
F(Bri (xi )) for i ∈ J , we get
3n ε ri n 3n ε n
L n (F(K )) ≤ εrin = ωn ≤ L (U ).
i∈J
ωn i∈J 3 ωn
|F(x) − x| |G(y) − y|
lim = 0, lim =0
|y|→0 |x| |y|→0 |y|
So, for any ε ∈ (0, 1) there exists δε > 0 such that if |x| < δε we have
x ∈ U \C F and |F(x)−x| < ε|x| and if |y| < δε we have y ∈ F(U \C F )
and |G(y) − y| < ε|y|. It follows that r < δε implies
L n (G(Br (0)))
(1 − ε)n ≤ ≤ (1 + ε)n ,
ωn r n
provided r < δε , and this proves that (8.8) holds.
Step 2. Set T = DG(0) and H (x) = T −1 G(x), so that D H (0) = I .
Then we have G(Br (0)) = T (H (Br (0))) and so, thanks to (8.4),
which implies
we have U = (0, ∞) × (0, 2π) × (0, π) and the critical set is empty,
because the modulus of the Jacobian determinant is −ρ 2 sin φ.
135 Introduction to Measure Theory and Integration
Exercises
8.1 Let (X, F ), (Y, G ) and (Z , H ) be measurable spaces and let f : X → Y ,
g : Y → Z be measurable maps. Show that
g# ( f # μ) = (g ◦ f )# μ
8.3
Show the existence of a strictly increasing and C 1 function F : R → R
such that F# L 1 is not absolutely continuous with respect to L 1 .
8.4
Remove the injectivity assumption in Theorem 8.4, showing that
1
F# (1U L n ) = 1 L n.
|J F|(x) F(U \C F )
x∈F −1 (y)\C F
Then F(x) = 1 for all x = 0, while F(0) = 0. In this case the smallest
possible function satisfying (A.1) is |y|−1 which is not integrable.
Next, we assume that X is an open set of Rn endowed with the Euc-
lidean distance and we investigate the differentiability of F. Under suit-
able assumption, we can commute derivative and integral, namely
∂ ∂f
f (x, y) dμ(y) = (x, y) dμ(y) ∀x ∈ X, i = 1, . . . , n.
∂ xi Y Y ∂ xi
(A.2)
Theorem A.3 (Differentiability of F). Assume that for μ-almost all y ∈
Y the function f (·, y) is differentiable in X with a continuous gradient
∇x f (x, y) and that, for any ball Br (x0 ) ⊂ X, there exists m ∈ L 1 (Y, μ)
satisfying
For i large enough (as soon as |ti | < r) the functions of y inside the
integral are dominated by the function m in (A.3), hence we can pass to
the limit with the dominated convergence theorem to get (notice that the
measurability of ∂ f /∂ xi (x0 , ·) follows by the same limiting process)
∂F ∂f
(x0 ) = (x0 , y) dμ(y).
∂ xi Y ∂ xi
∞
α(B) ≤ α(Ai ).
i=1
Since B is arbitrary, (i) gives α(A) ≤ i α(Ai ).
Now, if we take (A.5) as the definition of α̃ for all subsets of X, Pro-
position 1.16 gives that α̃ extends α and is σ –subadditive. Then, The-
orem 1.17 gives that α is σ –additive on the Borel σ –algebra, provided
we are able to show that any Borel set is α̃–additive. Since the class of
additive sets is a σ –algebra, suffices to show that any closed set is α̃–
additive.
To this aim, we first show that α̃ is additive on distant sets, namely
(recall that dist(U, V ) is the infimum of the distances d(x, y) for x ∈ U
and y ∈ V )
the functional
L μ (g) := g dμ g ∈ C(X) (A.7)
X
Theorem A.6 (Riesz). Let (X, d) be a compact metric space. The space
(C(X))∗ is, via (A.7), isomorphic and isometric to M(K ). That is: all
functionals L μ belong to (C(X))∗ and, for any L ∈ (C(X))∗ , there exists
a unique μ ∈ M(K ) satisfying L = L μ . Finally, L μ = μ.
Proof. The proof will be achieved in three steps. In the first one we build
an auxiliary positive finite measure μ∗ and prove in the second one that
μ∗ provides the desired representation of L when L is nondecreasing.
In the last one we achieve the general case and provide equality of the
norms.
Step 1. Let α ∗ : A (X) → [0, +∞) be defined by
Notice that α ∗ (X) ≤ L and that α ∗ (∅) = 0. Notice also that we can
equivalently replace |L(g)| with L(g) inside the supremum and that a
simple approximation argument gives
But, since
1
N −1
1
N −1
1
+ χi ≥ g ≥ χi+1 (A.12)
N i=1
N i=1
N
we can let N → ∞ and use the continuity of L to get L μ∗ (g) ≥ |L(g)|.
If L is also monotone we can use the inequality (A.9) to get
1
N −1
1 ∗
N −1
1 1 N −1
L μ∗ (g) − ≤ μ (Ai ) ≤ L(χi ) = L χi .
N i=1
N i=1
N N i=1
L μ∗ = L + + L − = L μ+ + L μ− = L μ+ +μ−
Hence
(1 − |a + − a − |) dμ∗ = μ∗ (X) − L ≤ 0.
X
Chapter 1
Exercise 1.1. All verifications are very simple and we omit them.
Exercise 1.2. We prove the statement for the translations, the proof for
the dilations being similar. Fix h ∈ R and consider the class
F := {A ∈ B (R) : A + h ∈ B (R)} .
if this property fails for some τ and A, for all subsets B either μ(B) = 0
or μ(B) ≥ τ μ(A). Now, choose B1 ⊂ A with μ(B1 ) ∈ (0, μ(A)) (this is
possible by assumption), then B2 ⊂ A \ B1 with μ(B2 ) ∈ (0, μ(B1 )) and
so on. Since all these sets are contained in A, we have μ(Bi ) ≥ τ μ(A),
and this contradicts the fact that they are disjoint.
Now, given t ∈ (0, μ(X)) we define a sequence of pairwise disjoint
sets Bi and numbers si as follows: first set
and then choose B1 with t ≥ μ(B1 ) > s1 /2; then recursively set
sn+1 := sup μ(B) : B ⊂ Bnc , μ(B) ≤ t − μ(Bn )
and choose Bn+1 ⊂ Bnc with t − μ(Bn ) ≥ μ(Bn+1 ) > sn+1 /2. We now
claim that μ(∪i Bi ) = t. If this property fails, then i μ(Bi ) < t and the
convergence of the series implies that si → 0. On the other hand
si ≥ sup μ(B) : B ⊂ X \ Bi , μ(B) ≤ t − μ(Bi )
i i
The previous property with A = X \ ∪i Bi and τ = (t − i μ(Bi ))/μ(A)
shows that the supremum in the right hand side (independent of i) is
positive, contradicting the fact that si → 0.
Exercise 1.7. Let X be a separable metric space and let E = B (X). If
μ({x}) > 0 for some x ∈ X, obviously μ is not diffuse. Conversely,
if A ∈ B (X) is given, with μ(A) > 0 and μ(B) ∈ {0, μ(A)} for all
B ⊂ A, we can fix a countable dense set (xi ) ⊂ X and define
r0 := sup r ≥ 0 : μ(A ∩ B r (x0 )) = 0 .
By the density of the family (xi ), this intersection contains at most one
point (and at least one, because the measure is positive). It follows that
this point is an atom of μ.
Exercise 1.8. Cantor’s middle third set can be obtained as follows: let
C0 = [0, 1], let C1 the set obtained from C0 by removing the interval
(1/3, 2/3), let C2 be the set obtained from C1 by removing the intervals
(1/9, 2/9) and (7/9, 8/9), and so on. Each set Cn consists of 2n disjoint
closed intervals with length 3−n , so that λ(Cn ) = (2/3)n → 0. If follows
that the intersection C of all sets Cn is a closed and λ–negligible set.
In order to show that C has the cardinality of continuum (at this stage
it is not even obvious that C = ∅!) we recall that numbers x ∈ [0, 1]
can be represented with a ternary, instead of a decimal, expansion: this
means that we can write
x= ai 3−i = 0, a1 a2 a3 . . .
i≥1
with the ternary digits ai ∈ {0, 1, 2}. As for decimal expansions, this
representation is not unique; for instance 1/3 can be written either as 0.1
or as 0.0222 . . ., and 2/3 can be written either as 0.2 or as 0.1222 . . ..
It is easy to check that C1 corresponds to the set of numbers that can
be expressed by a ternary representation not having 1 as first digit, C2
corresponds to the set of numbers that admit a representation not having
1 as a first or second digit, and so on. It follows that C is the set of
numbers that admit a ternary representation not using the digit 1: since
the map
∗
∞
(a1 , a2 , . . .) ∈ {0, 2}N → x = ai 3−i
i=1
∗ ∗
provides a bijection of {0, 2}N with C, and the cardinality of {0, 2}N is
the continuum, this proves that C has the cardinality of continuum.
Exercise 1.9. Let {qn }n∈N be an enumeration of the rational numbers in
[0, 1], and set
∞
ε ε
A := (qn − 2−n , qn + 2−n ).
n=0
4 4
Then A ⊂ R is open and λ(A) < n ε2−n−1 = ε (why is the inequality
strict ?). Therefore [0, 1] \ A has Lebesgue measure strictly less than ε
and an empty interior, because [0, 1] \ A does not intersect Q.
Exercise 1.11 Let {In }n∈N be an enumeration of the open intervals with
rational endpoints of (0, 1). By the construction in Exercise (1.9), for any
150 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
interval I and any δ ∈ (0, λ(I )) we can find a compact set C ⊂ I with
an empty interior such that 0 < λ(C) < δ. We will define
∞
E := Ci
i=0
where Cn ⊂ In are compact sets with an empty interior, λ(Cn ) > 0 and
λ(Cn ) < δn . The choice of Cn and δn will be done recursively. Notice
first that
λ(E ∩ In ) ≥ λ(Cn ) > 0 ∀n ∈ N,
so we have only to take care of the condition λ(E ∩ In ) < λ(In ). Set
βn = λ(In \ ∪n0 Ci ) and notice that βn > 0 because all Ci have an empty
interior. Since
n
∞
∞
λ(In ∩ E) ≤ λ(In ∩ Ci ) + δi = λ(In ) − βn + δi
0 i=n+1 i=n+1
it suffices to choose δn (and Cn ) in such a way that ∞n+1 δi < βn . This
is possible, choosing for instance δn+1 > 0 satisfying
1 1 1
δn+1 < max βn , βn−1 , . . . , n+1 β0 ,
2 4 2
μ∗ (D ∩ A) + μ∗ (D \ A) ≤ μ∗ (D ∩ B) + μ∗ (D \ B) = μ∗ (D).
Exercise 1.20. Obviously L has a cardinality not greater than the car-
dinality of P (R); by Bernstein theorem (1) it suffices to show that the
cardinality of P (R) is not greater than the cardinality of L : if C is the
Cantor set of Exercise 1.8, we know that P (R) is in one-to-one corres-
pondence of P (C), because C has the cardinality of continuum; on the
other hand, any subset of C obviously belongs to L , because C has null
Lebesgue measure.
Exercise 1.21. Let E ⊂ P (X) be a σ –algebra. Assume by contradiction
that E is infinite and countable. We define the equivalence relation
(1) If A has cardinality not greater than B, and B has cardinality not greater than A, then there exists
a bijection between A and B
154 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
We will show that the maximality of (τ̄ , μ̄) forces τ̄ to coincide with
P (N), so that μ̄ will be the desired extension of μ0 .
Let us assume by contradiction that τ̄ P (N) and choose Z ⊂ N
with Z ∈
/ τ̄ . We notice that
(A1 ∩ Z ) ∪ (A2 ∩ Z c ) : A1 , A2 ∈ τ̄
1. μ̃ is well defined: if
This continuous function maps continuously [0, 1] onto [0, λ(C)], and
it is constant in any connected component of A, so that g(A) is at most
countable. Since g(C) contains [0, λ(C)] \ g(A) we obtain that C has
cardinality at least equal to the continuum (one can actually see that
g(C) = g([0, 1])).
Exercise 1.26 Since K is totally bounded, for all > 0 there exist fi-
nitely many balls B1 , . . . , B N with radius whose union covers K . The
properties of μ imply the existence of an index i such that μ({n : xn ∈
Bi }) = 1. Now we start with = 1 and find a closed ball B (1) with
radius 1 such that μ({n : xn ∈ B (1) }) = 1. Repeating this construction
in B (1) we find a closed ball B (2) with radius 1/2 contained in B (1) with
μ({n : xn ∈ B (2) }) = 1. Continuing in this way, if z is the common point
of the balls B (i) , we find xn μ-converges to z.
Chapter 2
Exercise 2.1 The verification is straightforward and is omitted.
156 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
C = (C ∩ R) ∪ (C ∩ {−∞}) ∪ (C ∩ {+∞})
δ0 = min{δ, mini δi }, we have A |ϕn | dμ < ε/2 whenever n ∈ N, A ∈ E
and μ(A) < δ.
A possible example for the second question is = [0, 1], μ = λ the
n
Lebesgue measure, and ϕn = 2n 1[2−n ,21−n ) . The uniform integrability is a
direct consequence of the convergence of ϕn to 0 in L 1 . If ϕn ≤ g, then
∞ ∞
ϕn = ϕn ≤ g
n=1 n=1
∞ ∞
but n=1 ϕn = 1 1/n = +∞.
Exercise 2.11. (a) For any y ∈ X we have
gλ (x) ≤ g(y) + λd(x, y) ≤ g(y) + λd(x , y) + λd(x, x ).
Since y is arbitrary we get gλ (x) ≤ gλ (x ) + λd(x, x ). Reversing the
roles of x and x the inequality is achieved.
(b) Clearly the family (gλ ) is monotone with respect to λ, and since we
can always choose y = x in the minimization problem we have gλ (x) ≤
g(x). Assume that supλ gλ (x) is finite (otherwise the statement is trivial)
and let xλ such that gλ (x) + λ−1 ≥ g(xλ ) + λd(x, xλ ). This inequality
implies that xλ → x as λ → ∞ and, now neglecting the term λd(x, xλ ),
that
1
gλ (x) + ≥ g(xλ ).
λ
Passing to the limit in this inequality as λ → ∞ and using the lower
semicontinuity of g we get supλ gλ (x) ≥ g(x).
Exercise 2.12. Let us first assume that f is bounded. For ε > 0 we
consider the functions
x+ε
1
f ε (x, y) := f (x , y) dx .
2ε x−ε
Since x → f (x, y) is continuous, we can apply the mean value theorem
to obtain that f ε (x, y) → f (x, y) as ε ↓ 0. So, in order to show that f
is a Borel function, we need only to show that f ε are Borel.
We will prove indeed that f ε are continuous: let xn → x and yn → y;
since f (x , yn ) → f (x , y) for all x ∈ R, we have
1[xn −ε,xn +ε] (x ) f (x , yn ) → 1[x−ε,x+ε] (x ) f (x , y)
for all x ∈ R \ {x − ε, x + ε}. Therefore, since f is bounded, the
dominated convergence theorem yields
1
f ε (x, y) = 1 (x ) f (x , y) dx
2ε R [x−ε,x+ε]
1
= lim 1 (x ) f (x , yn ) dx = lim f ε (xn , yn ).
2ε n→∞ R [xn −ε,xn +ε] n→∞
160 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Chapter 3
Exercise 3.1. On the real line, endowed with the Lebesgue measure,
the function (1 + |x|)−1 belongs to L 2 , but not to L 1 , and the function
|x|−1/2 1(0,1) (x) belongs to L 1 , but not to L 2 . Turning back to the general
case, if ϕ ∈ L p1 ∩ L p2 with p1 ≤ p2 , from the inequality
(that can be verified considering separately the cases |ϕ| ≤ 1 and |ϕ| > 1)
we get that ϕ ∈ L p for all p ∈ [ p1 , p2 ].
Exercise 3.2. The statement is trivial if f q = 0, so we assume that
f q > 0. For > 0 the set X := {| f | > } has finite μ–measure, by
the Markov inequality, hence the inclusion between L r spaces for finite
measures gives that | f |1 X ∈ L p (X, E , μ). Since the dominated conver-
gence theorem gives
lim | f − f 1 X |q dμ = lim | f |q dμ = 0
↓0 X ↓0 X\X
in [0, +∞). But this problem has a unique minimizer, given by |ψ(x)|q−1,
and we conclude.
Exercise 3.4. It suffices to apply Hölder’s inequality to the functions |ϕ|r
and |ψ|r , with the dual exponents p/r and q/r, to obtain
Exercise 3.5. The positive part and the negative part of ϕ − ϕn have the
same integral, hence
|ϕ − ϕn | dμ = 2 (ϕ − ϕn )+ dμ.
X X
Therefore
lim sup ψn dμ + lim inf ϕn dμ ≥ ϕ dμ + ψ dμ.
n→∞ X n→∞ X X X
Subtracting ψ dμ from both sides the statement is achieved. In the
general case, let n(k) be a subsequence such that limk ϕn(k) dμ =
lim infn X ϕn , and let n(k(s)) be a further subsequence converging to
ϕ μ–a.e. Then
lim inf ϕn dμ = lim ϕn(k(s)) dμ ≥ lim inf ϕn(k(s)) dμ
n→∞ s→∞ X n
X
X
for all x1 , . . . , x2m ∈ J . The case m = 1 is (3.13) and the induction step
can be achieved grouping the terms as follows:
m−1
2m
1 1 2 1 2m−1
1
xi = xi + x2m−1 +i .
i=1
2m 2 i=1 2m−1 i=1
2m−1
162 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Let us choose c sufficiently large, such that M/(c) < ε/2, and then
δ > 0 such that cδ < ε/2. The inequality above yields A |ϕi | dμ < ε
whenever μ(A) < δ.
Exercise 3.11. Let ( f n ) ⊂ Cb (X) be converging in L 1 to f , and let f n(k)
be a subsequence pointwise convergent μ–a.e. to f . Then, given any
ε > 0, by Egorov theorem we can find a Borel set B ⊂ X with μ(B) < ε
and f n(k) → f uniformly on B c . By the inner regularity of the measure
we can find a closed set C ⊂ B c such that μ(X \ C) < ε. The function f
restricted to C, being the uniform limit of bounded continuous functions,
is bounded and continuous.
163 Introduction to Measure Theory and Integration
Chapter 4
Exercise 4.1. Notice that ·, · is obviously symmetric, that x, −y =
−x, y = −x, y and that x, x = x2 ≥ 0, with equality only if
x = 0. Notice that the parallelogram identity gives
and
x + x − 2y2 + x − x 2 = 2x − y2 + 2x − y2
= 8x, −y + 8x , −y − 2x + y2 − 2x + y2 .
So, we proved that x + x , 2y = 2x, y + 2x , y. Using the relation
u, 2v = 4u/2, v (due to the definition of ·, · and the homogeneity
of · ), we get
% &
x + x 1 1
, y = x, y + x , y.
2 2 2
Setting x = t1 v, x = t2 v, and defining the continuous function φ(t) =
tv, y, we get
t1 + t2 1 1
φ = φ(t1 ) + φ(t2 ).
2 2 2
This means that φ and −φ are convex in R, so that φ is an affine function,
and since φ(0) = 0 we get φ(t) = tφ(0), i.e. tu, y = tu, y. Coming
back to the identity above, we get x + x , y = x, y + x , y.
Exercise 4.2. Assume that y = π K (x). For all z ∈ K and t ∈ [0, 1] we
have y + t (z − y) belongs to K , so that
y + t (z − y) − x2 ≥ y − z2 .
k−1
f k = vk − vk , ei ei .
i=1
Exercise 4.5 Since X and its scalar product coincide with L 2 ([0, 1],
P ([0, 1]), μ), where μ is the counting measure in [0, 1], we obtain
that X is an Hilbert space. Let us prove by contradiction that X is not
separable. If S = { f n }n≥1 were a dense subset, it could be possible to
find a countable set D ⊂ [0, 1] such that f n (x) = 0 for all n and all
x ∈ [0, 1] \ D. Since [0, 1] is not countable we can find x0 ∈ [0, 1] \ D
and define g0 (x) equal to 1 if x = x0 and equal to 0 if x = x0 . We
claim that g0 does not belong to the closure of S. If this property fails,
we can find a sequence ( f n(k) ) ⊂ S convergent to g0 μ–a.e. in [0, 1];
but, convergence μ–a.e. corresponds to pointwise convergence and since
g0 (x0 ) = 0, while f n(k) (x0 ) = 0 for all k, we obtain a contradiction.
Exercise 4.6. By Parseval identity we know that x → (x, ei )) is a
linear isometry from H to 2 . As a consequence, taking the parallelogram
identity into account, the scalar product is preserved.
Exercise 4.7. We consider the class of orthonormal systems {ei }i∈I of H ,
ordered by inclusion. Zorn’s lemma ensures the existence of a maximal
165 Introduction to Measure Theory and Integration
system {ei }i∈I . Let V be the subspace spanned by ei , let Y be its closure
(still a subspace) and let us prove that Y = H . Indeed, if Y were a proper
subspace of H , we would be able to find, thanks to Corollary 4.5, a unit
vector e orthogonal to all vectors in Y , and in particular to all vectors
ei . Adding e to the family {ei }i∈I the maximality of the family would be
violated. Now, by the just proved density of V in H , given any x ∈ H
we can find a sequence of vectors (vn ), finite combinations of vectors ei ,
such that x − vn → 0. If we denote by Jn ⊂ I the set of indexes used
to build the vectors {v1 , . . . , vn }, and by Hn the vector space spanned by
{ei }i∈Jn , we know by Proposition 4.6 that
x − x, ei ei ≤ x − vn → 0.
i∈Jn
As a consequence, setting J = ∪n Jn , we have x = i∈J x, ei ei .
Chapter 5
Exercise 5.1. The functions sin mx cos lx are odd, therefore their integral
on (−π, π) vanishes. To show that sin mx is orthogonal to sin lx when
l = m, we integrate twice by parts to get
π
m π
sin mx sin lx dx = cos mx cos lx dx
−π l −π
m2 π
= 2 sin mx sin lx dx.
l −π
we obtain that ( 0N xi ) is a Cauchy sequence in E. Therefore the com-
pleteness of E provides the convergence of the series. Passing to the
limit as N → ∞ in the inequality 0N xi ≤ 0N xi and using the
continuity of the norm we obtain (5.15).
√
Exercise 5.3. We consider only the first system gk = 2/π sin kx, the
proof for the second one being analogous. The fact that (gk ) is orthonor-
mal can be easily checked noticing that gk are restrictions to (0, π) of
odd functions, and using the orthogonality of sin kx in L 2 (−π, π). Ana-
logously, if f ∈ L 2 (0, π) let us consider its extension f˜ to (−π, π) as an
166 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
odd function and its Fourier series, which obviously contains no cosinus.
In (0, π) we have
N
N
bk sin kx = f, gk gk ,
k=1 k=1
where the scalar products are understood in L 2 (0, π). Therefore, from
the convergence of the Fourier series in L 2 (−π, π) to f˜, which implies
convergence in L 2 (0, π) to f , the completeness follows.
Exercise 5.4. Clearly ek , ek = 1, while
π ) *π
ikx −ilx 1
e e dx = e i(k−l)x
dx =0 whenever k = l.
−π i(k − l) −π
N
N
eikx 1
ck √ = (cos kx + i sin kx)(ak − ibk )
k=−N 2π 2 k=1
−1
+ (cos kx + i sin kx)(a−k − ib−k )
k=−N
N
a0
= + Re (cos kx + i sin kx)(ak − ibk )
2 k=1
a0 N
= + ak cos kx + bk sin kx,
2 k=1
where (ek ) is the orthonormal system of Exercise 5.4 and to use its com-
pleteness.
2N ikz
Exercise 5.6. From the identity i=0 e = (ei(2N +1)z − 1)/(ei z − 1),
we get
N
2N
ei(2N +1)z − 1
eikz = e−i N z eikz = e−i N z =
k=−N k=0
ei z − 1
i(N +1/2)z
e − e−i(N +1/2)z sin((N + 1/2)z)
= = (A.15)
e i z/2 −e −i z/2 sin(z/2)
Using the fact that sin((N + 1/2)z)/ sin(z/2) has, still because of (A.15),
mean value 1 on (−π, π), we get
π
1
f (x) − S N f (x) = ( f (x) − f (y))G N (x − y) dy.
2π −π
n−1
x n , Q k
n−1
Q n (x) := x n − Q k (x) = x n − x n , Pk Pk (x) ∀n ≥ 1.
k=0
Q k , Q k k=0
√
(a) Since Q 0 = 1, P0 = 1/ 2 and Q 1 = √ x − x, P0 P0 = x, because
x, P0 = 0. As a consequence P1 (x) = 3/2x. Since x 2 , P1 = 0, we
have also
1
Q 2 (x) = x 2 − x 2 , P0 P0 − x 2 , P1 P1 = x 2 −
3
√
and this leads, with simple calculations, to P2 (x) = 45/8(x 2 − 1/3).
(b) Let H be the closure of the vector space spanned by Cn . This space
contains all monomials x n , and therefore all polynomials. Since the poly-
nomials are dense in C([a, b]), for the sup norm, they are also dense in
L 2 (a, b). It follows that H = L 2 (a, b). By Proposition 4.13 we conclude
that (Cn ) is complete.
(c) Set
(
2n + 1 1 dn 2
z n := , P̃n (x) := z n n (x − 1)
n
2 2n n! d x
Clearly the polynomial P̃n has degree n. So, in order to show that P̃n =
Pn , we have to show that P̃n is orthogonal to all monomials x k , k =
0, . . . , n −1, and that P̃n 2 = 1. Since P̃n has zeros at ±1 with multipli-
city n, all its derivatives at ±1 with order less than n are zero. Therefore,
for k < n we have
) *1 1
d n−1 d n−1
P̃n , x k = z n x k n−1 (x 2 − 1)n − k x k−1 n−1 (x 2 − 1)n dx
d x −1 −1 d x
= ···
) n−k *1
d
= (−1) k!z n n−k (x − 1)
k 2 n
= 0.
d x −1
so that
1 1
2n
(1 − x ) dx =
2 n
(1 − x 2 )n−1 dx = · · ·
−1 2n + 1 −1
1
(2n)!! 2(2n)!!
= (1 − x 2 )0 dx = .
(2n + 1)!! −1 (2n + 1)!!
d 2n 2
(x − 1)n = (2n)! = (2n)!!(2n − 1)!! = 2n n!(2n − 1)!!
d 2n x
from (A.16) we get
2n + 1 1 2(2n)!! n
P̃n , P̃n = 2 n!(2n − 1)!! = 1.
2 22n (n!)2 (2n + 1)!!
Chapter 6
Exercise 6.1. Let us prove the inclusion
(F 1 × F 2 ) × F 3 ⊂ F 1 × (F 2 × F 3 ),
170 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
the proof of the converse one being analogous. We have to show that all
products A × B, with A ∈ F 1 × F 2 and B ∈ F 3 belong to F 1 ×
(F 2 × F 3 ). Keeping B fixed, the class of sets A for which this property
holds is a σ –algebra that contains the π–system of measurable rectangles
A = A1 × A2 (because A × B = A1 ×(A2 × B) and A2 × B ∈ F 2 ×F 3 ),
and therefore the whole product σ –algebra F 1 × F 2 .
For all A in the product σ –algebra we have
(μ1 × μ2 ) × μ3 (A) = μ3 (A x1 x2 ) dμ1 × μ2 (x1 , x2 )
X 1 ×X
2
= μ3 (A x1 x2 ) dμ2 (x2 ) dμ1 (x1 )
X1 X2
= μ2 × μ3 (A x1 ) dμ1 (x1 ) = μ1 × (μ2 × μ3 )(A).
X1
This class contains the open sets (because the product of open sets is
open) and it is a σ –algebra, so it contains B (R). We have thus proved
that all rectangles B1 × R × · · · × R, with B1 Borel belong to B (Rn ). By
a similar argument we can show that all rectangles
R × · · · × R × Bi × R × · · · × R
(A × B)(A × B ) ⊂ (N A × R) ∪ (R × N B )
It follows that k f h(k) (x, ·) − f (x, ·) L 1 (ν) is finite for μ–a.e. x ∈
X, and for any such x the functions f h(k) (x, ·) converge to f in L 1 (ν).
Choosing Y = { ȳ} and ν = δ ȳ , to provide a counterexample it is suffi-
cient to consider any example (see Remark 3.7) of a sequence converging
in L 1 but not μ–almost everywhere.
Exercise 6.11. It suffices
to apply (6.15) to |h| to show that |h| d f μ is
finite if and only if |h| f dμ is finite.
Exercise 6.12. We prove the property for the sup, the property for the inf
being analogous. If A = B1 ∪ B2 with B1 ∈ F and B2 ∈ F disjoint, we
have
f μ(B1 ) + gμ(B2 ) = f dμ + g dμ
B1 B2
≤ f ∨ g dμ + f ∨ g dμ
B1 B2
= f ∨ g dμ.
A
than μ): just write ν(B) =
greater k ν(Bk ) ≤ k μi(k) (Bk ) (resp.
≥ k μi(k) (Bk ). So, it remains to show that μ and μ are σ -additive.
For any map i : N → I , A1 , A2 ∈ F disjoint and any countable F –
measurable partition of A1 ∪ A2 we have
∞
∞
∞
μi(k) (Bk ) = μi(k) (Bk ∩ A1 ) + μi(k) (Bk ∩ A2 ).
k=0 k=0 k=0
Estimating the right hand side from below with μ(A1 ) + μ(A2 ) we get
(because (Bk ) is arbitrary) that μ is superadditive, i.e. μ(A1 ∪ A2 ) ≥
μ(A1 ) + μ(A2 ). With a similar argument one can prove not only that μ
is subadditive, but also that μ is σ –subadditive (it suffices to consider a
countable F –measurable family, instead of 2 sets).
Now, let us prove that μ is subadditive and μ is superadditive. Let
A1 , A2 ∈ F be disjoint and let Bk1 , Bk2 be countable F –measurable
partitions of A1 and A2 respectively. If i 1 , i 2 : N → I we define i(2k) =
i 1 (k), B2k = Bk1 and i(2n + 1) = i 2 (n), B2k+1 = Bk2 , so that
∞
∞
∞
μ(A1 ∪ A2 ) ≤ μi(k) (Bk ) = μi1 (k) (Bk1 ) + μi2 (k) (Bk2 ).
k=0 k=0 k=0
∞
ν Am ≥ ν(An ) ≥ ε0
m=n
Let us show now that the maximality of Bh implies that μ(C) ≤ 2h σ (C)
for any set C ⊂ A \ Bh , i.e. t 1 A\Bh μ ≤ σ . Indeed, if there is C0 ⊂ A \ Bh
with μ(C0 ) > 2h σ (C0 ), the maximality of Bh provides a minimal integer
h 1 ≥ 1 and C1 ⊂ C0 satisfying μ(C1 ) ≤ 2h σ (C1 ) − 1/ h 1 . Let us
consider C0 \ C1 ; we still have μ(C0 \ C1 ) > 2h σ (C0 \ C1 ) and the
maximality of Bh provide a minimal integer h 2 ≥ h 1 and C2 ⊂ C0 \
C1 satisfying μ(C2 ) ≤ 2h σ (C2 ) − 1/ h 2 . Continuing in this way we
have a nondecreasing sequence (h i ) of integers and (Ci ) ⊂ F such that
μ(Ci ) ≤ 2h σ (Ci ) − 1/ h i and Ci ⊂ C0 \ ∪i−1 j=1 C j for all i ≥ 2; moreover
h i is the least integer for which there is such Ci . Now limi h i = ∞, since
the Ci are pairwise disjoint. Setting C = C0 \ ∪∞ 1 Ci , for all F ∈ F
contained in C, since F ⊂ C0 \ ∪1 C j for all i ≥ 2, we have μ(F) ≥
i−1
hence ! !
! ! !!
! | f | dμ − ! !
f dμ !! ≤ 2εμ(Bi ).
!
Bi Bi
It follows that
!! !
!
| f μ(Bi )| = ! f dμ!≥ | f | dμ − 2εμ(Bi )
! !
i∈Z i∈Z B i i∈Z Bi
= | f | dμ − 2εμ(B).
B
Analogously
* * i
1 i
μ −∞, x + ≥μ −∞, = ai ≥ μh ((−∞, x]).
h h j=−h 2
In this case (c), (d), do not hold, because Fh (x) → 1 = 0 = F(x) for all
x ∈ R, (e) does not hold and (b) holds.
a ⇒ b, e. This is easy, because Cc (R) ⊂ Cb (R) and 1R ∈ Cb (R).
a ⇒ c. This follows by second part of the proof of Theorem 6.28.
d ⇔ c. This is Exercise 6.22.
b∧e ⇒ c. This follows by the same argument used in the proof of second
part of Theorem 6.28: the sequence (gk ) monotonically convergent to 1 A
can be chosen in Cc (R), and this shows that lim infh μh (A) ≥ μ(A) for
all A ⊂ R open. Using (e) and passing to the complementary sets, we
obtain lim suph μh (C) ≤ μ(C) for all C ⊂ R closed.
d ⇒ f . This follows by the same argument used in the solution of
Exercise 6.25.
d ∧ f ⇒ e. For all x ∈ D, with D dense, we have limh μh ((−∞, x]) =
μ((−∞, x]). Since μh ((−∞, x]) → μh (R) as x → +∞ uniformly in
h, we can pass to the limit as x ∈ D → +∞ to obtain limh μh (R) =
limx→+∞ μ((−∞, x]) = μ(R).
d ∧ f ⇒ a. This follows by the same argument used in the first part of
the proof of Theorem 6.28, choosing the points ti in the partitions to be
in the dense set where convergence occurs.
Exercise 6.27. Set
1 2 2
g(ξ ) := √ eiξ x e−x /(2σ ) dx.
2πσ R2
Notice that g(0) = 1, and that differentiation theorems under the integral
sign (2) and an integration by parts give
1 2 2
g (ξ ) = √ ieiξ x (xe−x /(2σ ) ) dx
2
2πσ R
σ2 d 2 2
= √ i eiξ x e−x /(2σ ) dx
2πσ R dx
2
ξσ2 2 2
= −√ eiξ x e−x /(2σ ) dx.
2πσ R2
(2) In this case, the application of the theorem is justified by the fact that sup d iξ x e−x 2 /(2σ 2 ) |
ξ ∈I | dξ e
is Lebesgue integrable for all bounded intervals I
179 Introduction to Measure Theory and Integration
Chapter 7
Exercise 7.1. Let C > 0 be such that |H (x) − H (y)| ≤ C|x − y| for all
x, y ∈ R. Let ε > 0 and let δ > 0 be such that i | f (bi )− f (ai )| < ε/C
whenever
i (bi − ai ) < δ. We
have i |H ( f (bi )) − H ( f (ai ))| ≤
C i | f (bi ) − f (ai )| whenever i (bi − ai ) < δ. In particular, choosing
f (t) = t, we see that Lipschitz functions are absolutely continuous.
Exercise 7.2. We assume that both L 1 (E) > 0 and L 1 (R \ E) > 0. Let
a ∈ R be such that L 1 ((a, ∞) ∩ E) > 0 and L 1 ((a, ∞) \ E) > 0, and
define F(t) = L 1 (E ∩(a, t)). By our choice of a, F(t) and (t −a)− F(t)
are not identically 0 in (a, +∞).
If t > a is a rarefaction point of E, we have
F(t + h) − F(t) L 1 ((t, t + h) ∩ E)
F+ (t) = lim = lim = 0.
h↓0 h h↓0 h
180 Luigi Ambrosio, Giuseppe Da Prato and Andrea Mennucci
Chapter 8
Exercise 8.1. Both are measures in (Z , H ). If B ∈ H then g◦ f # μ(B) =
μ( f −1 (g −1 (B))), because (g ◦ f )−1 = f −1 ◦ g −1 . On the other hand,
because their common value is 2−n . On the other hand, f −1 ({1}) consists
of a single point and therefore the identity above holds for I = {1}, the
common value being 0. By additivity the identity holds for finite unions
of sets of this type, a family stable under finite intersections. By the
coincidence criterion the two measures coincide.
Exercise 8.3. Let A ⊂ R be a dense open set whose complement C has
strictly positive Lebesgue measure (Exercise 1.9), and let
1
F# (1 Q̃ i L n ) = 1 Ln
|JF | ◦ F −1 F( Q̃ i )
and therefore we can pass to the limit to get
1
F# (1 Q i L n ) = 1 L n.
|JF | ◦ F −1 (y) F(Q i )
If we add both sides with respect to i ∈ I we get
1 1
F# (1U L n ) = 1 F(Q ) L n
= 1 F(U ) L n .
i∈I
|JF | ◦ F (y)
−1 i
x∈F −1 (y)
|J F |(x)
References
This series publishes polished notes dealing with topics of current re-
search and originating from lectures and seminars held at the Scuola Nor-
male Superiore in Pisa.
Published volumes
1. M. T OSI , P. V IGNOLO, Statistical Mechanics and the Physics of Flu-
ids, 2005 (second edition). ISBN 978-88-7642-144-0
2. M. G IAQUINTA , L. M ARTINAZZI , An Introduction to the Regularity
Theory for Elliptic Systems, Harmonic Maps and Minimal Graphs,
2005. ISBN 978-88-7642-168-8
3. G. D ELLA S ALA , A. S ARACCO , A. S IMIONIUC , G. T OMASSINI ,
Lectures on Complex Analysis and Analytic Geometry, 2006.
ISBN 978-88-7642-199-8
4. M. P OLINI , M. T OSI , Many-Body Physics in Condensed Matter Sys-
tems, 2006. ISBN 978-88-7642-192-0
P. A ZZURRI, Problemi di Meccanica, 2007. ISBN 978-88-7642-223-2
5. R. BARBIERI, Lectures on the ElectroWeak Interactions, 2007. ISBN
978-88-7642-311-6
6. G. DA P RATO, Introduction to Stochastic Analysis and Malliavin Cal-
culus, 2007. ISBN 978-88-7642-313-0
P. A ZZURRI, Problemi di meccanica, 2008 (second edition). ISBN 978-
88-7642-317-8
A. C. G. M ENNUCCI , S. K. M ITTER , Probabilità e informazione,
2008 (second edition). ISBN 978-88-7642-324-6
7. G. DA P RATO, Introduction to Stochastic Analysis and Malliavin Cal-
culus, 2008 (second edition). ISBN 978-88-7642-337-6
8. U. Z ANNIER, Lecture Notes on Diophantine Analysis, 2009.
ISBN 978-88-7642-341-3
9. A. L UNARDI, Interpolation Theory, 2009 (second edition).
ISBN 978-88-7642-342-0
186 Lecture notes