Kesavan
Kesavan
Kesavan
S. Kesavan
Measure and
Integration
Texts and Readings in Mathematics
Volume 77
Advisory Editor
C. S. Seshadri, Chennai Mathematical Institute, Chennai
Managing Editor
Rajendra Bhatia, Ashoka University, Sonepat
Editors
Manindra Agrawal, Indian Institute of Technology, Kanpur
V. Balaji, Chennai Mathematical Institute, Chennai
R. B. Bapat, Indian Statistical Institute, New Delhi
V. S. Borkar, Indian Institute of Technology, Mumbai
Apoorva Khare, Indian Institute of Sciences, Bangalore
T. R. Ramadas, Chennai Mathematical Institute, Chennai
V. Srinivas, Tata Institute of Fundamental Research, Mumbai
Technical Editor
P. Vanchinathan, Vellore Institute of Technology, Chennai
The Texts and Readings in Mathematics series publishes high-quality textbooks,
research-level monographs, lecture notes and contributed volumes. Undergraduate
and graduate students of mathematics, research scholars, and teachers would find
this book series useful. The volumes are carefully written as teaching aids and
highlight characteristic features of the theory. The books in this series are
co-published with Hindustan Book Agency, New Delhi, India.
123
S. Kesavan (emeritus)
Institute of Mathematical Sciences
Chennai, Tamil Nadu, India
This work is a co-publication with Hindustan Book Agency, New Delhi, licensed for sale in all countries
in electronic form only. Sold and distributed in print across the world by Hindustan Book Agency, P-19
Green Park Extension, New Delhi 110016, India. ISBN: 978-93-86279-77-4 © Hindustan Book Agency
2019.
© Hindustan Book Agency 2019 and Springer Nature Singapore Pte Ltd. 2019
This work is subject to copyright. All rights are reserved by the Publishers, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publishers, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publishers nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made. The publishers remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Dedicated to
Professor Philippe G. Ciarlet,
to whom I owe more than I can
possibly express,
on the occasion of his eightieth
birthday.
Preface
A course on the theory of measure and the Lebesgue integral is now
an essential component in any masters or graduate programme in math-
ematics in Indian universities. It is part of the training a student receives
in analysis. The most interesting examples of Banach spaces are func-
tion spaces of various kinds and the Lebesgue spaces, also known as Lp
spaces, are amongst the most important of these. A knowledge of the
theory of measure and integration is essential for the study of several
advanced topics in functional analysis like the theory of distributions
and Sobolev spaces, which constitute the functional analytic framework
for the modern study of partial differential equations. Of course, the
theory of measure and integration is vital to the study of probability
and stochastic processes.
This book grew out of the notes I prepared for lectures on measure
theory and the theory of integration. These lectures were delivered,
over the past four decades, to masters and graduate students in sev-
eral leading institutions like the Centre for Applicable Mathematics,
Tata Institute of Fundamental Research, Bangalore, The Institute of
Mathematical Sciences, Chennai, the Chennai Mathematical Institute,
Siruseri, and the Indian Institute of Technology, Madras. Portions of
the book were also taught at numerous refresher or summer courses.
In particular, it was taught by me at several refresher courses at the
Ramanujan Institute for Advanced Study in Mathematics, of the Uni-
versity of Madras, Chennai. I am indeed thankful to these institutions
and organizers of refresher courses for having given me the opportunity
to deliver these lectures.
Chapter 9 concerns signed measures and the main result of this chap-
ter is the Radon-Nikodym theorem.
Chennai S. Kesavan
November, 2018
Notations
Certain general conventions followed throughout the text regarding
notations are described below. All other specific notations are explained
as and when they appear in the text.
• The set of natural numbers {1, 2, 3, · · ·}, is denoted by the symbol
N, the integers by Z, the rationals by Q, the reals by R and the
complex numbers by C.
• The union and intersection of sets are denoted using the usual
symbols ∪ and ∩ respectively.
• If a, b ∈ R ∪ {±∞}, then
[a, b] = {x ∈ R | a ≤ x ≤ b},
[a, b) = {x ∈ R | a ≤ x < b},
(a, b] = {x ∈ R | a < x ≤ b}.
N
! 12
X
|x| = |xi |2 .
i=1
x
Contents
Preamble 1
1 Measure 9
1.1 Algebras of sets . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2 Measures on rings . . . . . . . . . . . . . . . . . . . . . . 11
1.3 Outer-measure and measurable sets . . . . . . . . . . . . . 16
1.4 Completion of a measure . . . . . . . . . . . . . . . . . . . 24
1.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3 Measurable functions 54
3.1 Basic properties . . . . . . . . . . . . . . . . . . . . . . . . 54
3.2 The Cantor function . . . . . . . . . . . . . . . . . . . . . 61
3.3 Almost everywhere . . . . . . . . . . . . . . . . . . . . . . 65
3.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4 Convergence 68
4.1 Egorov’s theorem . . . . . . . . . . . . . . . . . . . . . . . 68
4.2 Convergence in measure . . . . . . . . . . . . . . . . . . . 70
4.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5 Integration 81
5.1 Non-negative simple functions . . . . . . . . . . . . . . . . 81
5.2 Non-negative functions . . . . . . . . . . . . . . . . . . . . 85
5.3 Integrable functions . . . . . . . . . . . . . . . . . . . . . 94
xi
xii Contents
6 Differentiation 118
6.1 Monotonic functions . . . . . . . . . . . . . . . . . . . . . 118
6.2 Functions of bounded variation . . . . . . . . . . . . . . . 124
6.3 Differentiation of an indefinite integral . . . . . . . . . . . 131
6.4 Absolute Continuity . . . . . . . . . . . . . . . . . . . . . 136
6.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
10 Lp spaces 196
10.1 Basic properties . . . . . . . . . . . . . . . . . . . . . . . . 196
10.2 Approximation . . . . . . . . . . . . . . . . . . . . . . . . 205
10.3 Some applications . . . . . . . . . . . . . . . . . . . . . . 208
10.4 Duality . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
10.5 Convolutions . . . . . . . . . . . . . . . . . . . . . . . . . 221
10.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
Bibliography 235
About the Author
S. KESAVAN is former professor at the Institute of Mathematical
Sciences, Chennai, and adjunct faculty at the Indian Institute of
Technology Madras, Chennai, India. He started his career at the Tata
Institute of Fundamental Research Centre for Applicable Mathematics
(TIFR-CAM), Bangalore, India, in 1973. He has also been associated
with the Chennai Mathematical Institute, where he was deputy director
during 2007–2010 and he is currently adjunct professor at the Indian
Institute of Technology, Madras. He earned his PhD from Université
Pierre-et-Marie-Curie, Paris, in 1979. He is a fellow of the National
Academy of Sciences, India, at Allahabad and the Indian Academy of
Sciences, Bangalore, India. He is a life member of the National Board
for Higher Mathematics, since 2000, Indian Mathematical Society,
International Society for the Interaction of Mechanics and Mathematics
(ISIMM), Indian Society of Industrial and Applicable Mathematics
(ISIAM), Ramanujan Mathematical Society, American Mathematical
Society and an elected fellow of the Forum d’Analystes, Chennai.
He was also the Secretary (Grant Selection) of the Commission for
Developing Countries of the International Mathematical Union, during
2011–2014 and 2015–2018. He has published four books and authored
over 50 research articles, apart from several contributions to conference
proceedings and popular articles. His research interests are in partial
differential equations, homogenization, control theory, and isoperimetric
inequalities.
xiii
Preamble
From the time of the Greeks, the problem of computing the area
enclosed by a curve had been exercising the minds of scientific thinkers.
This crucial question, at the base of the theory of integral calculus, was
treated as early as the third century B.C. by Archimedes, who calcu-
lated the area of a circular disc, the area of a segment of a parabola
and other such figures. He used the ‘method of exhaustion’. The basic
idea was to exhaust the given area by a sequence of polygonal domains
and calculate the area as the limit of the area of the inscribed polygons.
During the seventeenth century, many such areas were calculated and in
each case the problem was solved by an ingenious device specially suited
for the case in hand. One of the achievements of calculus was to develop
a general and powerful method to replace these special restricted proce-
dures.
From the time of Archimedes until the time of Gauss, the attitude
was that the area was an intuitively obvious entity which need not be
defined, but which had to be computed. Before Cauchy, there was no
definition of the integral in the precise sense of the term. One was often
limited to saying which areas one had to add, or subtract, to get the
integral.
where a = x0 < x1 < ... < xi < xi+1 < ... < xN = b is a partition of
[a, b] and ξi ∈ [xi , xi+1 ]. He then deduced the value of the integral
Z b
f (x)dx
a
In what follows, we will briefly recall the salient features of the Rie-
mann integral and see what are its principal drawbacks which will mo-
tivate the study of Lebesgue’s theory of measure and integration.
Then, we define the lower and upper (Darboux) sums associated to the
function f and the partition P by
PN
L(P, f ) = i=1 mi (xi − xi−1 )
PN
U (P, f ) = i=1 Mi (xi − xi−1 ).
Rb
a f (x)dx = inf P U (P, f )
where the supremum and infimum are taken over all possible partititions
of [a, b]. The function f is said to be Riemann integrable over [a, b]
if its lower and upper integrals are equal and the common value, called
the Riemann integral of f over [a, b], is denoted by the symbol
Z b
f (x)dx.
a
for all partitions P. Thus, the lower and upper integrals of f always
exist but the question of their being equal is a delicate one.
N
X
S(P, f ) = f (ti )(xi − xi−1 ).
i=1
lim S(P, f ) = A
µ(P)→0
if, for every ε > 0, there exists a δ > 0 such that, for all partitions P
such that µ(P) < δ, and for all choices of points ti compatible with the
partition, we have
|S(P, f ) − A| < ε.
Example Let us consider the unit interval [0, 1]. Let us choose some
numbering of all the rational numbers in this interval and write them as
r1 , r2 , .... Define
1, if x = r1 , r2 , ..., rn ,
fn (x) =
0, otherwise.
Thus,
1
n2
Z
fn (x)dx = → ∞
0 2n + 2
Preamble 5
R1
while, since f ≡ 0, we have 0 f (x)dx = 0. Similarly, if we define
fn (x) = nx(1 − x2 )n ,
R1
again fn → f ≡ 0 pointwise but 0 fn (x)dx → 1/2 6= 0.
So, when do the two limit processes - the pointwise limit of functions
and Riemann integration (which has been defined as a limit of sums as
shown in Theorem 1) - commute?
The idea of Lebesgue is to work, not from the domain, but from the
range of a function. We take a particular value and consider the set of
all points where this value is assumed when we define the integral. Let
us illustrate this via an example.
Measure
Throughout this section X will stand for a non-empty set. We will define
various classes of subsets of X. The power set of X, i.e. the collection
of all subsets of X, will be denoted by P(X).
Definition 1.1.1 A non-empty collection R of subsets of X is called a
ring if it is closed under the formation of unions and differences, i.e. if
E and F are members of R, then so are E ∪ F and E\F . A ring is said
to be an algebra if, in addition, X itself is a member of R.
Remark 1.1.1 By induction, it is clear that every ring is closed under
the formation of finite unions.
Remark 1.1.2 The empty set always belongs to any ring since, if E is
a member, so is ∅ = E\E.
E\F = E ∩ F c = (E c ∪ F )c ∈ R.
The next example is one which we will deal with in detail in this book
since it will be the starting point of the construction of the Lebesgue
measure.
P = {[a, b) | a, b ∈ R, a ≤ b}
From this it easily follows that R is closed under the formation of dif-
ferences as well. It is also clear that each member of R can, in fact, be
written as a finite disjoint union of members of P.
Remark 1.1.6 Thus, a σ-ring is a ring which is closed under the forma-
tion of countable unions. If {Ei }∞
i=1 is a countable collection in a σ-ring
S, then
∩∞ ∞
i=1 Ei = E\ ∪i=1 (E\Ei ) ∈ S,
E = E ∪ ∅ ∪ ∅ ∪ ∅···
Remark 1.2.2 Since µ(Ei ) ≥ 0 for all i, the order of the summands in
(1.2.1) is unimportant.
E = ∪N
i=1 Ei ∪ ∅ ∪ ∅ ∪ ∅ · · · ,
we have
N
X
µ(E) = µ(Ei ).
i=1
We need to check only the countable additivity. Let {Ei }∞ i=1 be a col-
lection of mutually disjoint subsets whose union is E. If E is a finite
set, then only at most finitely many of the Ei will be non-empty and
they will also be finite sets. Then (1.2.1) is obviously true. If E is an
infinite set, then either at least one of the Ei is an infinite set, or there
1.2 Measures on rings 13
are infinitely many Ei which are non-empty finite sets. In either case,
both sides of (1.2.1) take the value +∞ and so countable additivity is
established.
This measure is called the counting measure on the set X.
∪i Gi = ∪i Fi = E.
µ(∪∞
i=1 Ei ) = lim µ(En ). (1.2.2)
n→∞
1.2 Measures on rings 15
since the sets Ei \Ei−1 are all mutually disjoint. This completes the
proof.
µ(∩∞
i=1 Ei ) = lim µ(En ). (1.2.3)
n→∞
Proof: Since the sequence of sets is decreasing, we have that µ(En ) <
+∞ for all n ≥ m. Further {Em \En }n≥m is an increasing sequence of
sets. Hence by the preceding proposition and by the subtractive property
of the measure, we have
µ(Em ) − µ(∩∞ ∞ ∞
i=1 Ei ) = µ(Em ) − µ(∩i=m Ei ) = µ(Em \(∩i=m Ei ))
= µ(∪∞
i=m (Em \Ei )) = limn→∞ µ(Em \En )
The result now follows on subtracting µ(Em ), which is finite, from both
sides of the above relation.
Example 1.2.4 The preceding proposition is not valid without the as-
sumption that µ(Em ) is finite for some m. Consider the set of natural
numbers, N, equipped with the counting measure (cf. Example 1.2.1).
Let En = {m ∈ N | m ≥ n}. Then µ(En ) = +∞ for all n while
∩∞
n=1 En = ∅.
Thus, the Dirac measure (cf. Example 1.2.2) and the measure de-
fined in Example 1.2.3 are both finite measures. The counting measure
on N is a σ-finite measure, since any subset of N can be covered by a
countable number of singleton sets, and each singleton set has measure
unity.
But the sum on the extreme right is the tail of a convergent series and
hence can be made arbitrarily small for large n. Thus it follows that
µ(E) = 0.
∞
X
∗
µ (∪∞
n=1 En ) ≤ µ∗ (En ). (1.3.1)
n=1
P∞
we have, by subadditivity of the measure, µ(E) ≤ n=1 µ(En ) and
∗ ∗
so, by definition of µ , we have µ(E) ≤ µ (E). Thus, for E ∈ R, we
have µ(E) = µ∗ (E) and so µ∗ extends µ. In particular, we have that
µ∗ (∅) = 0.
∞
X ε
µ(Eij ) < µ∗ (Ei ) + .
2i
j=1
Then E ⊂ ∪∞ ∞
i=1 ∪j=1 Eij and
∞ X
X ∞ ∞
X
µ∗ (E) ≤ µ(Eij ) ≤ µ∗ (Ei ) + ε.
i=1 j=1 i=1
Step 4: Let E ∈ H(R). Then, there exists a countable cover {Ei }∞ i=1 of
E such that Ei ∈ R for each i. Since µ is σ-finite, we can find {Eij }∞
i,j=1
in R such that Ei ⊂ ∪∞j=1 Eij and µ(Eij ) < +∞ for each 1 ≤ i, j < ∞.
Now, we have E ⊂ ∪∞ ∞ ∗
i=1 ∪j=1 Eij and µ (Eij ) = µ(Eij ) < +∞. This
completes the proof.
Example 1.3.1 Let X = N and consider the ring R of all finite subsets,
with the counting measure. Then µ is finite. Since any countable union
of singletons has to be in H(R), it follows that N is in H(R) and by
heredity, it follows that H(R) = P(N), the power set of N. It is now
immediate to see that if E is any infinite subset of N, then µ∗ (E) = +∞.
1.3 Outer-measure and measurable sets 19
µ∗ (A) = µ∗ (A ∩ E) + µ∗ (A ∩ E c ). (1.3.2)
µ∗ (A) ≥ µ∗ (A ∩ E) + µ∗ (A ∩ E c ).
µ∗ (A) = µ∗ (A ∩ E) + µ∗ (A ∩ E c ),
µ∗ (A ∩ E) = µ∗ (A ∩ E ∩ F ) + µ∗ (A ∩ E ∩ F c ),
µ∗ (A ∩ E c ) = µ∗ (A ∩ E c ∩ F ) + µ∗ (A ∩ E c ∩ F c ).
Thus,
µ∗ (A) = µ∗ (A ∩ (E ∪ F )) + µ∗ (A ∩ (E ∪ F )c ).
µ∗ (A ∩ (E\F )c )) = µ∗ (A ∩ E ∩ F ) + µ∗ (A ∩ E c ∩ F ) + µ∗ (A ∩ E c ∩ F c ).
(1.3.5)
20 1 Measure
µ∗ (A) = µ∗ (A ∩ (E\F )c ) + µ∗ (A ∩ E ∩ F c )
= µ∗ (A ∩ (E\F )c ) + µ∗ (A ∩ (E\F )).
µ∗ (A∩(E1 ∪E2 )) = µ∗ (A∩E1 ∩E2 )+µ∗ (A∩E1 ∩E2c )+µ∗ (A∩E1c ∩E2 ).
µ∗ (A ∩ (E1 ∪ E2 )) = µ∗ (A ∩ E1 ) + µ∗ (A ∩ E2 ).
By induction, it follows that for any n mutually disjoint sets {Ei }ni=1 ,
we have
Xn
∗ n
µ (A ∩ (∪i=1 Ei )) = µ∗ (A ∩ Ei ). (1.3.7)
i=1
Set Fn = ∪ni=1 Ei .
Since S is a ring, we have that Fn ∈ S. Further, since
Fn ⊂ E, we have that E c ⊂ Fnc . Consequently we have, by (1.3.7) and
the monotonicity of µ∗ ,
µ∗ (A) = µ∗ (A ∩ Fn ) + µ∗ (A ∩ Fnc )
Pn ∗ (A
≥ i=1 µ ∩ Ei ) + µ∗ (A ∩ E c ).
µ∗ (A) ≥ µ∗ (A ∩ E) + µ∗ (A ∩ E c )
µ(E) = µ∗ (E).
= inf{ ∞ ∞
P
i=1 µ(Ei ) | E ⊂ ∪i=1 Ei , Ei ∈ R}
≥ inf{ ∞ ∞
P
i=1 µ(Ei ) | E ⊂ ∪i=1 Ei , Ei ∈ S(R)}
≥ inf{µ(∪∞ ∞
i=1 Ei ) | E ⊂ ∪i=1 Ei , Ei ∈ S(R)}
≥ inf{µ(F ) | E ⊂ F, F ∈ S(R)}
≥ inf{µ(F ) | E ⊂ F, F ∈ S}
= inf{µ∗ (F ) | E ⊂ F, F ∈ S}
≥ µ∗ (E).
Remark 1.3.3 Let µ be a measure on a ring R of subsets of a non-
empty set X and let µ be its extension, as described in this section, to
S(R). Starting from this, one could again try to define an outer-measure
µ∗ on H(R) and try to extend it further. The proof of the preceding
proposition shows that µ∗ = µ∗ and so the σ-ring of measurable sets will
still be S and so the induced measure will also only be µ.
Definition 1.3.5 Let µ be a measure on a ring R of subsets of a non-
empty set X. Let E ∈ H(R) and let F ∈ S(R). We say that F is
a measurable cover of E if E ⊂ F and for all G ∈ S(R) such that
G ⊂ F \E, we have µ(G) = 0.
1
µ∗ (E) ≤ µ∗ (F ) ≤ µ∗ (Fn ) < µ∗ (E) + .
n
24 1 Measure
µ∗ (E) = µ∗ (F ) = µ(F ).
µ
e(E∆N ) = µ(E).
Thus,
Se = {E ∪ N | E ∈ S, N ⊂ A, A ∈ S, µ(A) = 0}.
It is now clear that Se is closed under the formation of countable unions.
Let E1 , E2 ∈ S and let N1 ⊂ A1 , N2 ⊂ A2 where Ai ∈ S, µ(Ai ) = 0
1.4 Completion of a measure 25
E1 ∆E2 = N1 ∆N2 .
N ⊂ A, we have, by definition,
since
µ(E ∪ N ) = µ∗ (E ∪ N ) ≤ µ∗ (E) + µ∗ (N ) = µ∗ (E)
= µ(E) ≤ µ(E ∪ N ).
E = (F \G) ∪ (E ∩ G).
1.5 Exercises
P = ΠN
i=1 [ai , bi ) | ai , bi ∈ R, ai ≤ bi , 1 ≤ i ≤ N .
S = {F ⊂ X | F ∈ R or F c ∈ R}.
1.9 Let X = N and let R be the ring described in Exercise 1.6 above.
Define, for E ∈ R,
+∞, if E is an infinite set,
µ(E) =
0, otherwise.
µ∗ (E) + µ∗ (F ) = µ∗ (E ∪ F ) + µ∗ (E ∩ F ).
Show that
(i) µ∗ is a σ-finite outer-measure on H;
(ii) the only µ∗ -measurable sets are at most countable sets or their com-
plements;
(iii) the induced measure µ is not σ-finite.
µ(E∆E0 ) ≤ ε.
Chapter 2
Let P denote the class of all intervals of the form [a, b), where a ≤
b, a, b ∈ R. Let R be the ring of all finite unions of members of P (cf.
Example 1.1.3). As observed earlier, we can express each member of R
as a finite disjoint union of members of P. Let us define
µ([a, b)) = b − a.
However, we need to check that this is well-defined and also that it sat-
isfies the properties of a measure. In particular, we need to verify that
© Hindustan Book Agency 2019 and Springer Nature Singapore Pte Ltd. 2019 30
S. Kesavan, Measure and Integration, Texts and Readings in
Mathematics 77, https://doi.org/10.1007/978-981-13-6678-9_2
2.1 Construction of the Lebesgue measure 31
Lemma 2.1.1 (a) Let {Ei }ni=1 be a finite set of mutually disjoint inter-
vals in P, such that each of them is contained in E0 ∈ P. Then
n
X
µ(Ei ) ≤ µ(E0 ). (2.1.1)
i=1
Thus,
Pn Pn Pn Pn−1
i=1 µ(Ei ) = i=1 (bi − ai ) ≤ i=1 (bi − ai ) + i=1 (ai+1 − bi )
= bn − a1 ≤ b0 − a0 = µ(E0 ).
δ
Ui = ai − ,b
2i i
, 1 ≤ i < ∞.
Then, F0 ⊂ E0 and Ei ⊂ Ui , 1 ≤ i < ∞. Thus F0 ⊂ ∪∞ i=1 Ui . Since F0
is compact, there exists a positive integer n such that F0 ⊂ ∪ni=1 Ui . It
now follows from the preceding lemma (cf. (2.1.2)) that
n ∞
X δ X
b0 − a 0 − ε < bi − a i + i ≤ (bi − ai ) + δ.
2
i=1 i=1
In other words,
∞
X
µ(E0 ) − ε < µ(Ei ) + δ.
i=1
The result now follows since ε and δ are arbitrarily small quantities.
Proposition 2.1.2 The set function µ is countably additive on P.
Proof: Let E0 = ∪∞ ∞
i=1 Ei , where {Ei }i=1 is a sequence of mutually
disjoint members of P. Assume that E0 ∈ P as well. By the preceding
proposition, we have
X∞
µ(E0 ) ≤ µ(Ei ).
i=1
On the other hand, for any positive integer n, we have by Lemma 2.1.1
(cf. (2.1.1)),
Xn
µ(Ei ) ≤ µ(E0 )
i=1
which yields
∞
X
µ(Ei ) ≤ µ(E0 )
i=1
from which the result follows.
2.1 Construction of the Lebesgue measure 33
E = ∪ni=1 Ei = ∪m
j=1 Fj ,
Ei = ∪ m
j=1 Ei ∩ Fj .
Thus,
n
X n X
X m m X
X n m
X
µ(Ei ) = µ(Ei ∩ Fj ) = µ(Ei ∩ Fj ) = µ(Fj ).
i=1 i=1 j=1 j=1 i=1 j=1
Ei = ∪nk=1
i
Eik , 1 ≤ i < ∞,
Case 1: E ∈ P.
In this case, since
ni
E = ∪∞
i=1 ∪k=1 Eik
Case 2: E = ∪nj=1 Fj ,
where Fj ∈ P for each 1 ≤ j ≤ n, and the Fj ’s are
all mutually disjoint. Then, for each 1 ≤ j ≤ n, we have
Fj = ∪ ∞
i=1 Fj ∩ Ei .
Thus {Fj ∩ Ei }∞
i=1 is a collection of mutually disjoint sets in R whose
union is Fj ∈ P. Thus, by Case 1 above, we have
∞
X
µ(Fj ) = µ(Fj ∩ Ei ).
i=1
Ei = ∪nj=1 Ei ∩ Fj .
2.1 Construction of the Lebesgue measure 35
P = ΠN
i=1 [ai , bi ) | ai ≤ bi , ai , bi ∈ R, 1 ≤ i ≤ N ,
we see that B is, indeed, a σ-algebra. Thus, the hereditary σ-ring gen-
erated by R is clearly the power set of R. We can now define the
induced outer-measure µ∗ for all subsets of R. The collection L of all
µ∗ -measurable sets is thus a σ-algebra which is called the Lebesgue σ-
algebra and its members are called the Lebesgue measurable sets;
the induced measure on this σ-algebra is called the Lebesgue measure
on R. It is clear that the Lebesgue measure is σ-finite and complete.
Thus the Lebesgue measure is the completion of the measure induced
on the Borel σ-algebra (cf. Theorem 1.4.2) by µ.
Now consider any interval [a, b) as the set [a, b) × {0} ⊂ R2 . Then
∞ 1
[a, b) × {0} = ∩n=1 [a, b) × 0, .
n
Example 2.1.2 (i) Since any non-empty open set contains a non-empty
open interval, it follows that the Lebesgue measure of any non-empty
open set is strictly positive.
(ii) The set of rationals, Q, being countable, has Lebesgue measure zero.
Thus Q is a an example of a dense set which has measure zero. Its com-
plement, the set of irrationals, is a dense set of infinite measure.
(iii) If E ⊂ R is a measurable set of measure zero, then it cannot contain
any non-empty open set. Thus, every non-empty open set will intersect
E c and so E c will be dense.
(iv) If K ⊂ R is a compact set, then it is bounded and closed and so it
has finite measure.
Now let X3 be the union of the open middle-thirds of the four subinter-
vals of X\(X1 ∪ X2 ) and so on. Set
C = X\ ∪∞
n=1 Xn .
2n−1
m1 (Xn ) = .
3n
Since all these sets Xn are disjoint, we have
∞
X 2n−1
m1 (∪∞
n=1 Xn ) = = 1 = m1 ([0, 1]).
3n
n=1
Consequently, m1 (C) = 0.
(iii) It then follows that since it is closed and has measure zero, it cannot
contain a non-empty open set. Thus C is nowhere dense.
(iv) Let x ∈ C. If (a, b) is any interval containing x, then, for n suffi-
ciently large, it must also contain a sub-interval of Xn . End-points of all
such sub-intervals are in C. Thus no point of C is isolated. Since C is
closed as well, it follows that C is a perfect set and so it is uncountable
(cf. Rudin [7]). Thus C is an example of an uncountable set of measure
zero.
(iv) Another way of proving the uncountability of the Cantor set is as
follows. Consider the ternary expansion of real numbers in [0, 1]:
∞
X
x = an 3−n ,
n=1
It is now clear that the Cantor set contains the set of all points in
[0, 1] with infinite ternary expansion such that the digits in its ternary
expansion are either 0 or 2. Now a simple Cantor diagonalisation argu-
ment shows that C is uncountable.
BN ⊂ LN ⊂ P(RN ).
One would, naturally, like to know if these inclusions are strict. We will
show, in the sequel, that these inclusions are, indeed, strict. To start
with, since the Lebesgue measure is complete, we have that every subset
of C is Lebesgue measurable. Since C is uncountable, the cardinality of
L1 is 2c , where c is the cardinality of the continuum. It can be shown
that the cardinality of B1 is just c. Consequently the inclusion B1 ⊂ L1
is strict. We will later define the Cantor function and use it to prove the
existence of a Lebesgue measurable set which is not Borel measurable.
We will also prove the existence of sets in R which are not Lebesgue
measurable.
2.2 Approximation
B = ΠN
j=1 Ij ,
Proof: (i) ⇒ (ii): If µ∗ (E) < +∞, then, by the previous proposition,
there exists an open set U containing E such that µ∗ (U ) < µ∗ (E) + ε.
Since E is Lebesgue measurable, we have that µ∗ = mN and since a
measure is subtractive, we deduce that µ∗ (U \E) < ε. If µ∗ (E) = +∞,
then since µ∗ = mN is σ-finite, we can find disjoint measurable sets En
such that each of them has finite measure and such that E = ∪∞ n=1 En .
Then, we can find open sets Un such that En ⊂ Un and such that
µ∗ (Un \En ) < 2εn . Then U = ∪∞
n=1 Un is open, contains E and
∞
X ε
µ∗ (U \E) ≤ µ∗ (∪∞
n=1 (Un \En )) ≤ = ε.
2n
n=1
(ii) ⇒ (iv): For each positive integer n, choose Un open such that E ⊂ Un
and µ∗ (Un \E) < n1 . Set G = ∩∞ n=1 Un . Then G is a Gδ set containing E
and
1
µ∗ (G\E) ≤ µ∗ (Un \E) < ,
n
from which we deduce that µ∗ (G\E) = 0.
µ∗ (E\F ) = µ∗ (E ∩ F c ) = µ∗ (E ∩ U ) = µ∗ (U \E c ) < ε.
(iii) ⇒ (v): For every positive integer n, choose Fn , a closed set contained
in E, such that µ∗ (E\Fn ) < n1 . Set F = ∪∞ n=1 Fn . Then F is an Fσ set
contained in E. Further,
1
µ∗ (E\F ) ≤ µ∗ (E\Fn ) <
n
E = F ∪ (E\F )
is measurable as well.
Proof: We can find a closed box J1 and an open box J2 , such that
J1 ⊂ J2 ⊂ J2 ⊂ I and such that mN (I\J1 ) < ε. By Urysohn’s lemma,
there exists a continuous function ϕ such that 0 ≤ ϕ(x) ≤ 1 for all x
and such that ϕ(x) = 1 for all x ∈ J1 and ϕ(x) = 0 for all x 6∈ J2 . Then,
the support of ϕ is contained in J2 ⊂ I, and so it is compact. Thus
ϕ ∈ Cc (RN ). Now,
{x ∈ RN | ϕ(x) 6= χI (x)} ⊂ I\J1
from which the result follows immediately.
Corollary 2.2.1 Let Ω ⊂ RN be an open set and let f : Ω → R be a
step function. Let ε > 0 be an arbitrary positive number. Then there
exists ϕ ∈ Cc (Ω) such that
mN ({x ∈ Ω | ϕ(x) 6= f (x)}) < ε,
and such that
max |ϕ(x)| ≤ max |f (x)|.
x∈Ω x∈Ω
Pk
Proof: Let f = j=1 αj χIj be a step function. Without loss of gen-
erality, we can assume that the boxes are disjoint. By the preceding
proposition, there exist functions ϕj ∈ Cc (RN ), 1 ≤ j ≤ k, such that, for
each such j, we have 0 ≤ ϕj ≤ 1, the support of ϕj is contained in the
box Ij , and
ε
mN ({x ∈ RN | ϕj (x) 6= χIj (x)}) < .
k
Pk
Set ϕ = j=1 αj ϕj . Then
and so
mN ({x ∈ Ω | ϕ(x) 6= f (x)}) < ε.
Since the supports of the ϕj , 1 ≤ j ≤ k, are all disjoint, it follows that
max |ϕ(x)| ≤ max |αj | = max |f (x)|.
x∈Ω 1≤j≤k x∈Ω
Proof: Set
S = {E ⊂ Ω | T (E) is a Borel set}.
Clearly Ω and ∅ are in S. Also, if E ⊂ Ω, T (E c ) = (T (E))c and if {Ei }∞ i=1
is a sequence of subsets of Ω, we have that T (∪∞ ∞
i=1 Ei ) = ∪i=1 T (Ei ). It
follows from these observations that S is closed under the formation of
countable unions and under complementation. Thus, S is a σ-algebra
on Ω. Since open sets get mapped onto open sets, we get that all open
sets are in S. Then it follows that all Borel sets are in S as well. The
converse follows by applying this reasoning to the map T −1 .
2N n ν(Q)
e = ν(Q) = c = cmN (Q) = c2N n mN (Q).
e
Thus, ν(Q) e = cmN (Q)e as well. Since any open set can be written as
the countable disjoint union of such boxes (cf. Lemma 2.2.1), it follows
that if V is any open set, then ν(V ) = cmN (V ). Then, for any Borel set
E, it follows, from condition (ii) in the statement of this theorem, that
ν(E) = cmN (E).
ν(E) = mN (A(E)).
Thus
mN (A(E)) = ΠN
i=1 λi .
Once again, in this case, we have cA = det(A) = |det(A)|.
Thus it follows that we must have that m1 (E) = 0. Thus the only mea-
surable subsets of P are those of measure zero. The same is true for any
Pi , 1 ≤ i < ∞.
Now let A ⊂ [0, 1) be a measurable set such that m1 (A) > 0. Set
Ei = A ∩ Pi . If Ei is measurable, then m1 (Ei ) = 0. Thus if all the Ei
are measurable, we have, since A = ∪∞ i=0 Ei ,
∞
X
0 < m1 (A) ≤ m1 (Ei ) = 0,
i=0
We can draw the same conclusion for any interval of the form [n, n +
1). Thus, if A ⊂ R is a measurable set with strictly positive measure,
then there exists a positive integer n such that A ∩ [n, n + 1) has strictly
positive measure and hence will contain a non-measurable subset. Thus,
we conclude that every measurable set in R with strictly positive measure
will contain a non-measurable subset.
52 2 The Lebesgue Measure
2.5 Exercises
2.2 Let S 1 denote the unit circle in the plane. The Borel sets in S 1 are
the members of the σ-algebra generated by all open arcs. Show that
there exists a Borel measure µ on S 1 such that µ(S 1 ) = 1 and such that
µ is invariant under all rotations of S 1 .
2.4 Show that the plane R2 cannot be expressed as the countable union
of straight lines.
2.7 (a) Let T be a triangle in the plane R2 with vertices at the points
(0, 0), (1, 0) and (0, 1). What is the value of m2 (T )?
(b) Let T be a triangle in the plane R2 with vertices at the points
(xi , yi ), i = 1, 2, 3. Show that m2 (T ) = |A|, where
1 1 1
1
A = x1 x2 x3 .
2
y1 y2 y 3
2.8 Let A : RN → RN be a non-singular linear transformation. Let
E ⊂ RN . With the notations of Section 2.2, show that
µ∗ (E) = |det(A)|µ∗ (E).
Deduce that E is Lebesgue measurable if, and only if A(E) is Lebesgue
measurable.
2.5 Exercises 53
µ(E) = m1 (T (E)),
Measurable functions
(ii) ⇒ (iii):
(iii) ⇒ (iv):
1
f −1 ([−∞, α]) = ∩∞
n=1 f
−1
−∞, α + ∈ S.
n
(iv) ⇒ (i):
f −1 ({α}) ∈ S.
The result now follows since every open set can be written as the count-
able union of open intervals.
Then
R\[0, 1), if α = −2,
{−α}, if − α ∈ [0, 1)\E,
f −1 ({α}) =
{α}, if α ∈ E,
∅, otherwise.
Since the rationals are countable, it follows that (f +g)−1 ([−∞, α)) ∈ S.
Thus f + g is measurable. Since f − g = f + (−1)g, it follows from (i)
above that f − g is also measurable.
(iii) Since constant functions are measurable, it follows from (ii) above
that f + c is measurable.
are measurable.
Proof: The result follows from the previous propositions and the fol-
lowing relations:
1
max{f, g} = 2 (f + g + |f − g|),
1
min{f, g} = 2 (f + g − |f − g|).
Remark 3.1.3 The functions f + and f − are called the positive and
negative parts of the function f . We have f = f + − f − and |f | =
f + + f − . Notice that both f + and f − are non-negative functions.
Lemma 3.1.1 Let (X, S) be a measurable space and let f be a real-
valued measurable function defined on X. Then f −1 (E) ∈ S whenever
E is a Borel set.
Proof: Consider
Se = {E ⊂ R | f −1 (E) ∈ S}.
k
X
f = αi χAi ,
i=1
n2 n
X i−1
fn = nχFn + χEn,i .
2n
i=1
We claim that fn (x) ≤ fn+1 (x) for each positive integer n and for
each x ∈ X. Indeed, if f (x) ≥ n+1, then fn+1 (x) = n+1 and fn (x) = n.
If n ≤ f (x) < n + 1, then fn+1 (x) = 2i−1
n+1 for some i such that
i−1 i
f (x) ∈ , ⊂ [n, n + 1).
2n+1 2n+1
In this case, fn (x) = n and so we still have fn (x) ≤ fn+1 (x). Finally, if
f (x) < n, then for some 1 ≤ i ≤ n2n , we have
i−1 i 2(i − 1) 2i
f (x) ∈ , = , .
2n 2n 2n+1 2n+1
i−1 2(i−1)
Consequently, fn (x) = 2n while fn+1 (x) = 2n+1
= i−1
2n = fn (x), if
f (x) ∈ [ 2(i−1) , 2i−1 ), or fn+1 (x) = 22i−1
2n+1 2n+1 n+1 > i−1
2n = fn (x), if f (x) ∈
[ 22i−1 , 2i
n+1 2n+1 ). This establishes the claim.
Proof: We can split f into its positive and negative parts. Thus f =
f + − f − , where f ± are non-negative measurable functions. We can find
sequences of non-negative simple functions {ϕn } and {ψn } such that
ϕn ↑ f + and ψn ↑ f − . Thus, fn = ϕn − ψn gives a sequence of simple
functions converging pointwise to f.
The Cantor function, like the Cantor set, provides a lot of interest-
ing examples, or counter-exmples, to illustrate fine points in the theory
of measure and integration. Several constructions are possible but the
essential properties are the same for all these functions and they serve
62 3 Measurable functions
the same purpose. In this section, we will present one such construction.
f (x), if x ∈ [a, c1 ],
g(x) = f (c1 ), if x ∈ [c1 , c2 ],
f (b)−f (c1 )
f (c1 ) + b−c2 (x − c2 ), if x ∈ [c2 , b].
In other words, we move along f in the first third of the interval, then
move horizontally along the second third, and finally climb up to f (b)
in a straight line on the third interval (cf. Figure 3.2.1 below).
f(b)
f(a)
a b
Figure 3.2.1
3.2 The Cantor function 63
Let us consider the interval [0, 1] and the function f0 (x) = x defined
on it. If we apply the procedure described above to this function, we
will get the function f1 given by
x, if x ∈ [0, 13 ],
1 1 2
f1 (x) = 3 , if x ∈ [ 3 , 3 ],
2x − 1, if x ∈ [ 23 , 1].
Figure 3.2.2
The maximum slope occurs in the last sub-interval and, as seen ear-
lier, each time we apply the iteration procedure, the slope doubles. Thus,
64 3 Measurable functions
Let us now define ψ(y) = y + f (y) for y ∈ [0, 1]. Then ψ is strictly
monotonic increasing and continuous. We have ψ(0) = 0 and ψ(1) = 2.
Thus ψ is a continuous bijection of [0, 1] onto [0, 2].
ϕ(x) − ϕ(y) ≤ x − y
{x ∈ X | |f (x)| > M }
3.4 Exercises
Convergence
Then, clearly,
Since, µ(Y ) = µ(X) < +∞, given ε > 0, there exists n0 (m) ∈ N such
that
ε
µ(Y \En0 (m),m ) = µ(Y ) − µ(En0 (m),m ) < m .
2
© Hindustan Book Agency 2019 and Springer Nature Singapore Pte Ltd. 2019 68
S. Kesavan, Measure and Integration, Texts and Readings in
Mathematics 77, https://doi.org/10.1007/978-981-13-6678-9_4
4.1 Egorov’s theorem 69
Set G = ∪∞
m=1 (Y \En0 (m),m ). Then G is measurable and
∞
X ε
µ(G) < = ε.
2m
m=1
F c = ∩∞
m=1 En0 (m),m .
1
Given any η > 0, choose m such that m < η. If x ∈ F c , then x ∈
En0 (m),m ⊂ En,m for all n ≥ n0 (m). Thus, for all x ∈ F c , and for all
n ≥ n0 (m), we have
1
|fn (x) − f (x)| < < η.
m
Since the choice of m depended only on η, this shows that we have uni-
form convergence of the sequence {fn } to f on F c . This completes the
proof.
Example 4.1.1 The result of the theorem does not hold, in general,
in infinite measure spaces. For instance, consider the set N of natural
numbers with the counting measure defined on the σ-algebra of all sub-
sets of N. If F ⊂ N is such that µ(F ) < ε < 1, then, clearly, F = ∅.
Thus uniform convergence on F c means uniform convergence on N. Now
consider the sequence {fn }∞ n=1 defined by fn = χ{1,2,···,n} . Then fn → f
on N, where f (i) = 1 for all i ∈ N, but this convergence is not uniform.
1
Proof: Let m ∈ N. Choose Fm ∈ S such that µ(Fm ) < m and such
c ∞
that fn → f uniformly on Fm . Set F = ∩m=1 Fm . Then µ(F ) = 0. Since
F c = ∪∞ c c
m=1 Fm , we have that fn (x) → f (x) for every x ∈ F .
Definition 4.1.2 Let (X, S, µ) be a measure space and let {fn }∞ n=1 be a
sequence of real-valued measurable functions defined on X. We say that
this sequence is almost uniformly Cauchy if, for every ε > 0, there
exists a set F ∈ S such that µ(F ) < ε and such that {fn } is a uniformly
Cauchy sequence on F c .
Proposition 4.1.2 Let (X, S, µ) be a measure space and let {fn }∞ n=1
be an almost uniformly Cauchy sequence of real-valued measuarble func-
tions defined on X. Then there exists a real-valued measurable function
f defined on X such that fn → f almost uniformly.
1
Proof: For each m ∈ N, choose Fm ∈ S such that µ(Fm ) < m and such
that the sequence {fn } is uniformly Cauchy on Fm . set F = ∩∞
c
m=1 Fm .
Then µ(F ) = 0. Since F c = ∪∞ F
m=1 m
c , we have that {f (x)} is a Cauchy
n
sequence for every x ∈ F c . Define
limn→∞ fn (x), if x ∈ F c ,
f (x) =
0, if x ∈ F.
Definition 4.2.1 Let (X, S, µ) be a measure space and let {fn }∞ n=1 be
a sequence of real-valued measurable functions defined on X. Let f be a
4.2 Convergence in measure 71
We say that the sequence {fn } is Cauchy in measure if for every ε > 0
and for every δ > 0, there exists N ∈ N such that for all n, m ≥ N , we
have
µ({x ∈ X | |fn (x) − fm (x)| ≥ ε}) < δ.
Proposition 4.2.1 Let (X, S, µ) be a finite measure space, i.e. µ(X) <
+∞. Let {fn }∞n=1 be a sequence of real-valued measurable functions,
defined on X, converging a.e. to a real-valued measurable function f .
µ
Then fn → f .
Proof: Let D denote the set of all points x ∈ X such that the sequence
{fn (x)} fails to converge to f (x). Thus µ(D) = 0. Let ε > 0. If we set
then,
D = ∪ε>0 ∩∞ ∞
n=1 ∪m=n Em (ε) = ∪ε>0 lim sup En (ε).
n→∞
Thus, µ(lim supn→∞ En (ε)) = 0. Since µ(X) < +∞, we have (cf. Exer-
cise 1.10),
Thus,
µ
from which we deduce that limn→∞ µ(En (ε)) = 0, i.e. fn → f.
72 4 Convergence
The converse is not true, even in finite measure spaces, i.e. conver-
gence in measure does not imply pointwise convergence as the following
example shows.
χin = χ , 1 ≤ i ≤ n.
[ i−1 i
n ,n)
Let x ∈ [0, 1). For each n ∈ N, there exists exactly one i such that
χin (x) = 1, 1 ≤ i ≤ n, while χjn (x) = 0 for all 1 ≤ j ≤ n, j 6= i. Thus, we
see that the above sequence fails to converge at every point x ∈ [0, 1).
On the other hand, if 0 < ε < 1, we have
i−1 i 1
m1 ({x ∈ [0, 1) | |χin (x)| ≥ ε}) = m1 , = ,
n n n
In the above example, notice that for any x ∈ [0, 1), we have χ1n (x) =
0, for every n ≥ x−1 . Thus, there exists a subsequence converging to
the zero function pointwise. This behaviour is generic, as the following
proposition shows.
Proposition 4.2.2 Let (X, S, µ) be a measure space and let {fn }∞ n=1
be a sequence of real-valued measurable functions defined on X which
converges in measure to a real-valued measurable function f defined on
X. Then, there exists a subsequence of {fn } which converges to f almost
everywhere.
4.2 Convergence in measure 73
Proof: Set
1
En,m = x ∈ X | |fn (x) − f (x)| ≥ .
m
Then, for every m ∈ N, we can find a positive integer n0 (m) such that
1
µ(En0 (m),m ) < .
2m
Thus,
∞
X
µ(En0 (m),m ) < +∞.
m=1
Hence, by the Borel-Cantelli lemma (cf. Proposition 1.2.6), there exists
a measurable set E, with measure zero, such that every point of E c
belongs to at most finitely many of the sets En0 (m),m . In other words,
for every x ∈ E c , there exist a positive integer N such that x 6∈ En0 (m),m ,
for all m ≥ N , i.e. for all m ≥ N , we have
1
|fn0 (m) (x) − f (x)| < .
m
This shows that fn0 (m) (x) → f (x) for all x ∈ E c . This completes the
proof.
The next result shows that the limit function, under convergence in
measure, is defined uniquely up to a set of measure zero.
Proposition 4.2.3 Let (X, S, µ) be a measure space and let {fn }∞n=1 be
a sequence of real-valued measurable functions defined on X. Let f and
µ
g be real-valued measurable functions defined on X such that fn → f
µ
and fn → g. Then f = g almost everywhere.
Proof: Let ε > 0. Then,
ε
{x ∈ X | |f (x) − g(x)| ≥ ε} ⊂ x ∈ X | |fn (x) − f (x)| ≥ 2
since
|f (x) − g(x)| ≤ |fn (x) − f (x)| + |fn (x) − g(x)|.
µ µ
Since fn → f and fn → g, we deduce that
Given δ > 0, we can then find a positive integer N such that for n and
m greater than, or equal to, N , we have that the measure of each of the
two sets on the right-hand side of the above relation will be less than 2δ .
Thus for m, n ≥ N , we have
Let δ > 0 be given. Then, there exists N ∈ N such that for all n ≥ N
and for all nk ≥ N , we have
ε δ
µ x ∈ X | |fnk (x) − f (x)| ≥ 2 < 2.
4.2 Convergence in measure 75
Set
1
Ek = x ∈ X | |fnk (x) − fnk+1 (x)| ≥ k .
2
Then µ(Ek ) < 2−k . Given δ > 0, choose k such that 2−(k−1) < δ. Set
F = ∪∞
i=k Ei . Then
∞
X 1
µ(F ) ≤ µ(Ei ) < < δ.
2k−1
i=k
76 4 Convergence
Pm 1
< j=` 2j
1 1
= 2`−1
< 2N −1
< ε.
Proposition 4.2.8 Let (X, S, µ) be a measure space and let {fn }∞n=1 be
a sequence of real-valued measurable functions, defined on X, which is
Cauchy in measure. Then, there exists a real-valued measurable function
µ
f defined on X such that fn → f .
and
µ
it follows that fn2 → 0 as well.
µ µ
Step 2: If fn → f , then fn − f → 0. Now, let
En = {x ∈ X | |f (x)| > n},
so that En ↓ ∅. Since µ(X) < +∞, it follows that (cf. Proposition 1.2.5)
µ(En ) ↓ 0. Given δ > 0, choose m ∈ N such that µ(Em ) < δ.
Now,
The measure of the first set on the right-hand side of the above relation
is, evidently, less than δ. Since |f (x)| ≤ m for x ∈ Em c , it follows that
for all points x in the second set on the right-hand side of the above
relation, we have
ε
ε ≤ m|fn (x) − f (x)|, i.e. |fn (x) − f (x)| ≥ .
m
Thus, we can find N ∈ N such that, for all n ≥ N , we have
µ({x ∈ X | |fn f (x) − f 2 (x)| ≥ ε} ∩ Em
c
) < δ.
Thus, for all n ≥ N , we have
µ({x ∈ X | |fn f (x) − f 2 (x)| ≥ ε}) < 2δ,
µ
which shows that fn f → f 2 .
Step 3: Now
fn2 − f 2 = (fn − f )2 + 2(fn f − f 2 ).
µ µ
Since fn −f → 0, we have (fn −f )2 → 0. Combining this with the result
µ
of Step 2 above, we deduce that fn2 → f 2 , which completes the proof.
4.3 Exercises
Integration
where {Ii }ni=1 is a finite collection of disjoint intervals and the αi are all
non-negative, the Riemann integral of ϕ is nothing but the area under
the graph of ϕ, i.e.
Z Xn
ϕ(x) dx = αi m1 (Ii ).
R i=1
© Hindustan Book Agency 2019 and Springer Nature Singapore Pte Ltd. 2019 81
S. Kesavan, Measure and Integration, Texts and Readings in
Mathematics 77, https://doi.org/10.1007/978-981-13-6678-9_5
82 5 Integration
Ai = ∪{j | βj =αi } Bj .
Set
E σ = ∩ki=1 Eiσi .
Thus, if σ0 = (−1, · · · , −1), then
c
E σ0 = ∩ki=1 Eic = ∪ki=1 Ei .
Given two such k-tuples σ and σ 0 which are not equal, there must
exist i with 1 ≤ i ≤ k such that σi 6= σi0 . Without loss of generality,
assume that σi = +1 and σi0 = −1. In that case, by definition, E σ ⊂ Ei
0 0
while E σ ⊂ Eic . Thus, if σ 6= σ 0 , we have that E σ and E σ are disjoint.
Now let us assume that ϕ is given in the form (5.1.3). For any
1 ≤ i ≤ k, we have
X
χ Ei = χ ∪ E σ = χE σ .
{σ | σi =+1}
{σ | σi =+1}
Consequently, we have
k
X X X X
ϕ = γi χE σ = γi χE σ .
i=1 {σ | σi =+1} σ6=σ0 {i | σi =+1}
Pk P σ)
= i=1 γi {σ | σi =+1} µ(E
Pk
= i=1 γi µ(Ei ).
84 5 Integration
The last equality comes from the finite additivity of the measure and
from the result of Lemma 5.1.1 above.
Remark 5.1.1 Notice that the measure of some (or all) of the sets Ei
could be +∞. Thus the integral of ϕ could be +∞ as well. This is
the reason why we only consider non-negative functions. If the γi were
of different signs and if the corresponding sets were of infinite measure,
then we cannot add them meaningfully. For consistency, if γi = 0 and
the set Ei has infinite measure, we adopt the convention that 0.∞ = 0.
Remark 5.2.1 In view of Remarks 5.1.2 and 5.1.3, it is clear that the
above definition is consistent with the definitions made in the previous
section in case f is itself a non-negative simple function. Again, if ϕ is a
simple function such that 0 ≤ ϕ ≤ f , then ϕχE is a non-negative simple
function defined on E which is bounded above by f |E . Conversely, if ϕ
is a non-negative simple function defined on E bounded above by f |E ,
86 5 Integration
Proof: Set
1
Fn = x ∈ X | f (x) > , n ∈ N.
n
Then
{x ∈ X | f (x) 6= 0} = ∪∞
n=1 Fn .
Pk P∞ P∞ Pk
= j=1 αj i=1 µ(Aj ∩ Ei ) = i=1 j=1 αj µ(Aj ∩ Ei )
P∞ R P∞
= i=1 Ei ϕ dµ = i=1 ν(Ei ).
Then,
Z Z Z
(ϕ + ψ) dµ = (αi + βj )µ(Eij ) = ϕ dµ + ψ dµ.
Eij Eij Eij
The result now follows immediately from the preceding proposition since
the Eij are all disjoint.
We now need to prove the reverse inequality, which will complete the
proof.
Let 0 < c < 1 be any fixed constant. Let ϕ be a simple function such
that 0 ≤ ϕ ≤ f . For n ∈ N, define
En = {x ∈ X | fn (x) ≥ cϕ(x)}.
Then each set En is measurable. Since the sequence {fn }∞ n=1 is increas-
ing, we have that E1 ⊂ E2 ⊂ · · · ⊂ En ⊂ · · ·. Further, given any x ∈ X,
we have two possibilities.
5.2 Non-negative functions 89
• Either, f (x) = 0 which imples that fn (x) = 0 for all n and also
that ϕ(x) = 0. In this case x ∈ E1 .
• Or, f (x) > 0 which implies that f (x) > cϕ(x), since 0 < c < 1. In
this case, there exists n ∈ N such that cϕ(x) < fn (x) ≤ f (x) and
so x ∈ En .
Thus, X = ∪∞
n=1 En . Now,
Z Z Z
def
fn dµ ≥ fn dµ ≥ c ϕ dµ = cν(En ).
X En En
Since 0 < c < 1 was arbitrarily chosen, it now follows, on letting c tend
to unity, that Z
α ≥ f dµ,
X
which completes the proof.
R
Remark 5.2.3 It is possible that X f dµ is infinite. In R that case,
we conclude from the preceding theorem that the limit of X fn dµ, as
n → ∞, is also infinite.
Proposition 5.2.5 Let (X, S, µ) be a measure space and let {fn }∞ n=1
be a sequence of non-negative extended real-valued measurable functions
defined on X. Set
∞
X
f (x) = fn (x), x ∈ X.
n=1
Let {ϕn }∞ ∞
n=1 and {ψn }n=1 be increasing sequences of non-negative
simple functions increasing to f1 and f2 respectively (cf. Theorem 3.1.1).
By Proposition 5.2.4,
Z Z Z
(ϕn + ψn ) dµ = ϕn dµ + ψn dµ.
X X X
We also have that ϕn + ψn increases to f1 + f2 . Hence, by the monotone
convergence theorem, we get
Z Z
lim (ϕn + ψn ) dµ = (f1 + f2 ) dµ,
n→∞ X X
and Z Z
lim ϕn dµ = f1 dµ,
n→∞ X
Z Z
lim ψn dµ = f2 dµ.
n→∞ X
We thus conclude that
Z Z Z
(f1 + f2 ) dµ = f1 dµ + f2 dµ.
X X X
It now follows by induction that, for any n ∈ N,
Z n Z
X
(f1 + · · · + fn ) dµ = fk dµ.
X k=1 X
Using the result of Example 5.2.1, this translates into the following re-
lation:
∞
X X∞ X ∞
f (j) = fi (j).
j=1 i=1 j=1
Proof: Set gn (x) = inf i≥n fi (x) for x ∈ X. Then {gn }∞n=1 is an in-
creasing sequence of non-negative measurable functions whose limit is
lim inf n→∞ fn . Thus, by the monotone convergence theorem, we get
R R
X (lim inf n→∞ fn )
dµ = limn→∞ X gn dµ
R R
≤ limn→∞ inf i≥n X fi dµ = lim inf n→∞ X fn dµ.
5.2 Non-negative functions 93
fn = χ[n,n+1) .
R
Then fn (x) → 0 for each x ∈ R and so RR(lim inf n→∞ fn ) dm1 = 0. On
the other hand, for every n ∈ N, we have R fn dm1 = m1 ([n, n + 1)) =
1.
Proposition 5.2.6 Let (X, S, µ) be a measure space and let {fn }∞ n=1
and f be non-negative, extended real-valued measurable functions defined
on X. Assume that, for all n ∈ N, and for all x ∈ X, we have 0 ≤
fn (x) ≤ f (x), and that fn (x) → f (x) as n → ∞. Then
Z Z
lim fn dµ = f dµ.
n→∞ X X
Proof: By Fatou’s lemma and the the fact that the fn are all bounded
above by f , we get
Z Z Z Z
f dµ ≤ lim inf fn dµ ≤ lim sup fn dµ ≤ f dµ,
X n→∞ X n→∞ X X
Proof: Clearly ν(∅) = 0 and ν(E) ≥ 0 for all E ∈ S. Let {Ei }∞i=1 be a
sequence of mutually disjoint sets in S whose union is E. Then
∞
X
χE = χ Ei .
i=1
in view of the linearity of the integral with respect to the integrand, try
to define the interal of f as the difference between the integrals of f +
and f − , which are well-defined, since these functions are non-negative.
However, if both these integrals turn out to be infinite, we cannot de-
fine
R their difference. So we need that at least one of the two quantities,
+ dµ or − dµ, be finite. In view of this requirement, we make
R
X f X f
the following definition.
Notation
R Let (X, S, µ) be a measure space. We generally use the sym-
bol X f dµ to denote the (Lebesgue) integral of an integrable function
96 5 Integration
defined over X, with respect to the measure µ. However, there may arise
situations when f depends on more than one variable. For instance, if
f is a function of two variables x and y varying over different measure
spaces, say, (X, S, µ) and (Y, S 0 , ν), and if we wish to integrate f as a
function of x with y fixed in Y , we will write the integral as
Z
f (x, y) dµ(x).
X
We now prove the full linearity of the Lebesgue integral with respect to
the integrand.
Theorem 5.3.1 Let (X, S, µ) be a measure space and let f and g be in-
tegrable, complex-valued functions defined on X. Let α and β be complex
constants. Then
Z Z Z
(αf + βg) dµ = α f dµ + β g dµ. (5.3.3)
X X X
h+ − h− = f + g = f + − f − + g + − g − ,
h+ + f − + g − = f + + g + + h− .
Since all the functions involved in the above relation are non-negative,
we deduce that (cf. the proof of Proposition 5.2.5)
Z Z Z Z Z Z
− −
+
h dµ + f dµ + g dµ = +
f dµ + +
g dµ + h− dµ.
X X X X X X
5.3 Integrable functions 97
Since all the quantities in the above relation are finite, we can rearrange
the terms to get
Z Z Z Z Z Z
− −
+
h dµ − h dµ = +
f dµ − f dµ + +
g dµ − g − dµ,
X X X X X X
which will complete the proof. If c ≥ 0, then (10.3.4) follows from the
definition of the integral and from (5.2.3). If c = −1, then the result is
again true since, for a real-valued function f , we have (−f )+ = f − and
(−f )− = f + , and we can again use the definition of the integral. Thus,
(10.3.4) is true for all real constants c. The proof will be complete if we
can prove the relation when c = i. Let f = u + iv be the decomposition
of f into its real and imaginary parts. Then, by definition,
Z Z Z Z Z
if dµ = (−v + iu) dµ = − v dµ + i u dµ = i f dµ.
X X X X X
where |α| = 1. Let u denote the real part of the function αf . Then
u ≤ |αf | ≤ |f |. Then, by the preceding theorem, we have
Z Z Z Z Z
f dµ = α
f dµ = αf dµ = u dµ ≤ |f | dµ.
X X X X X
98 5 Integration
In particular, we have
Z Z
lim fn dµ = f dµ. (5.3.8)
n→∞ X X
Theorem 5.3.4 Let (X, S, µ) be a measure space and let {fn }∞ n=1 be
a sequence of integrable functions defined on X converging pointwise,
almost everywhere, to an integrable function f defined on X. Assume
that for all x ∈ X, and for all n ∈ N, we have that |fn (x)| ≤ gn (x), where
{gn }∞
n=1 is a sequence of non-negative integrable functions defined on X
converging to a non-negative integrable function g almost everywhere in
X. Finally, assume that
Z Z
lim gn dµ = g dµ.
n→∞ X X
and R R
X (g + f ) dµ ≤ lim inf n→∞ X (gn + fn ) dµ
R R
= X g dµ + lim inf n→∞ X fn dµ.
R
Since X g dµ < +∞, we deduce that
Z Z Z Z
f dµ ≤ lim inf fn dµ ≤ lim sup fn dµ ≤ f dµ,
X n→∞ X n→∞ X X
Theorem 5.3.5 Let (X, S, µ) be a measure space and let {fn }∞ n=1 be
a sequence of integrable functions defined on X converging pointwise
5.3 Integrable functions 101
| |fn | − |f | | ≤ |fn − f |,
Thus,
∞
X
f (j) = aij , j ∈ N.
i=1
Define, for j ∈ N, g(j) = bj . Then g is a non-negative integrable func-
tion. By hypothesis, for each n ∈ N,
∞ ∞
Xn X X
fi (j) ≤ |fi (j)| = |aij | ≤ bj = g(j).
i=1 i=1 i=1
Pn
Since i=1 fi → f , we have, by the dominated convergence theorem,
∞
n X
X ∞ X
X n ∞
X ∞ X
X ∞
lim aij = lim aij = f (j) = aij .
n→∞ n→∞
i=1 j=1 j=1 i=1 j=1 j=1 i=1
5.3 Integrable functions 103
Thus, the given conditions are sufficient to ensure that the order of
summation can be reversed when the double sequence is not necessarily
non-negative. (cf. Example 5.2.3).
Proposition 5.3.2 Let (X, S, µ) be a measure space and let f be an
integrable function defined on X. Then, given ε > 0, there exists δ > 0
such that, whenever µ(E) < δ, E ∈ S, we have
Z
|f | dµ < ε. (5.3.10)
E
Thus, for every set E ∈ S such that µ(E) < δ, (5.3.10) holds.
On the real line, R, we have two notions of integrals. The first is that of
the Riemann integral, defined primarily for bounded functions defined
over bounded intervals. If f : [a, b] → R is a bounded function, then the
Riemann integral, if it exists, is denoted by the symbol
Z b
f (x) dx.
a
The Lebesgue integral, on the other hand, is defined for all non-
negative measurable functions defined on any (Lebesgue) measurable
subset of R. The value of the integral may be finite or infinite. If
it is finite, we say the function is Lebesgue integrable. The Lebesgue
integral is then defined for any measurable function such that |f | is
Lebesgue integrable. In this case, of course, the integral will be again a
real number. The Lebesgue integral of a non-negative or an integrable
function, f , over a (Lebesgue) measurable set E, is denoted by the
symbol Z
f dm1 .
E
5.4 The Riemann and Lebesgue integrals 105
be a partition of the interval [a, b]. The points {xi }ni=0 are called the
nodes of the partition. The mesh size of the partition, denoted ∆(P), is
defined as follows:
where, for 1 ≤ i ≤ n,
Then the upper and lower (Darboux) sums associated to f and this
partition (cf. Preamble) are given by
Z Z
U (Pk , f ) = Uk dm1 , and L(Pk , f ) = Lk dm1 . (5.4.1)
[a,b] [a,b]
Proof: With the notations established above, the sequence {Uk (x)}∞ k=1
is monotonic decreasing and bounded below and the sequence {Lk (x)}∞ k=1
is monotonic increasing and bounded above, for each x ∈ [a, b]. Thus
both sequences are convergent. Let their respective limits be U (x) and
L(x). Then
L(x) ≤ f (x) ≤ U (x), x ∈ [a, b].
Since f is bounded, assume that |f (x)| ≤ M for all x ∈ [a, b]. Then,
for all x ∈ [a, b], and for all k ∈ N, we also have |Lk (x)| ≤ M and
|Uk (x)| ≤ M . Consequently, by the dominated convergence theorem, we
have R R
limk→∞ [a,b] Uk dm1 = [a,b] U dm1 , and
R R (5.4.3)
limk→∞ [a,b] Lk dm1 = [a,b] L dm1 .
Since f is Riemann integrable, the upper and lower Darboux sums con-
verge to the Riemann integral of f . Thus, in view of (5.4.1), we get
Z Z Z b
U dm1 = L dm1 = f (x) dx.
[a,b] [a,b] a
i.e.
|U (Pk , f ) − L(Pk , f )| < ε,
which proves that f is Riemann integrable.
Example 5.4.1 The result of Theorem 5.4.1 is not true for unbounded
intervals. Consider the interval (0, ∞) ⊂ R. Let
sin x
f (x) = .
x
It is well known (a
R ∞standard exercise on contour integration) that the
Riemann integral 0 f (x) dx is well-defined and that its value is, infact,
π
2 (cf., for example, Ahlfors [1]). However, f is not Lebesgue integrable.
To see this, consider the intervals
h π πi
In = nπ + , nπ + , n ∈ N,
4 2
which are all disjoint. On In , we have
1 π
| sin x| ≥ √ and x = |x| ≤ (2n + 1) .
2 2
108 5 Integration
and the sum on the right becomes arbitrarily large, for large N , since it
is a partial sum of a divergent series.
Example 5.4.2 Let f (x) = √1x on the interval (0, 1). This is a non-
negative function and its Lebesgue integral is well-defined. Define
(
0, if x ∈ (0, n1 ),
fn (x) = √ , if x ∈ [ 1 , 1).
1
x n
Now Z Z
fn dm1 = f dm1 .
1
(0,1) [n ,1)
On the interval [0, 1], we have that f is a bounded and continuous func-
tion and so it is Riemann and hence Lebesgue integrable. Now, set
1
fn (x) = χ (x).
x2 (1,n)
Then for each x ∈ (1, ∞), fn (x) increases to x12 . Since fn is a bounded
and continuous function on the interval (1, n), it is Riemann integrable
there. Consequently,
Z Z n
1 1 1
2
dm1 (x) = lim dx = lim 1 − = 1.
(1,∞) x n→∞ 1 x2 n→∞ n
1
Thus, the function x 7→ x2
is integrable on the interval (1, ∞). Since
2
sin x 1
≤ ,
x x2
where
n n!
= ,
k k!(n − k)!
is the usual binomial coefficient.
Let fi (x) = xi , i = 0, 1, 2. Now,
Z n
X n
µtn (X) = f0 dµtn = tk (1 − t)n−k = 1.
X k
k=0
Next,
R t
Pn n k n−k k
X f1 dµn = k=0 k t (1 − t) n
Pn n − 1 k−1
= t k=1 k − 1 t (1 − t)(n−1)−(k−1)
= t.
Lemma 5.5.1 Let t ∈ [0, 1] and n ∈ N be fixed. Let ε > 0 be given. Set
Aε = {x ∈ X | |x − t| ≥ ε}.
Aδ = {x ∈ X | |x − t| ≥ δ}.
We can write Z
|f (x) − f (t)| dµtn (x) = I1 + I2 ,
X
where
|f (x) − f (t)| dµtn (x), and
R
I1 = Aδ
Now, given any η > 0, choose ε < η2 and choose N ∈ N such that, for
all n ≥ N , we have
2M η
< .
4n 2
112 5 Integration
5.6 Probability
BB = {A ∩ B | A ∈ B}.
5.6 Probability 113
p(A ∩ B)
pB (A ∩ B) = , A ∈ B.
p(B)
p(A ∩ B) = p(A)p(B).
5.7 Exercises
5.2 Let (X, S, µ) be a measure space such that µ(X) < +∞. Let
{fn }∞
n=1 be a sequence of integrable functions defined on X converg-
ing uniformly to a function f , on X. Show that
Z Z
lim fn dµ = f dµ.
n→∞ X X
5.6 In each of the following cases, check the function f for (Lebesgue)
integrability over the indicated domain.
1
(a) f (x) = 1+x 2 on R.
1
(b) f (x) = x on (0, 1).
1
(c) f (x) = x sin x1 on (0, 1).
xn−1
f (x) =
(1 + x2 )k
5.8 Let (X, S, µ) be a measure space. Show that the dominated con-
vergence theorem is true if we replace ‘fn → f almost everywhere’ by
µ
‘fn → f ’.
Show that f ≡ 0.
uniformly on X.
(g ◦ h)(x) = g(h(x))
.
5.15 Let (X, S, µ) be a measure space and let M denote the collection
of all equivalence classes of real-valued measurable functions defined
on X, modulo equality almost everywhere. If f : X → R is a mea-
surable function, denote the equivalence class containing f by f . Let
ϕ : [0, +∞) → [0, 1] be a strictly monotonic increasing continuous func-
tion such that ϕ(0) = 0. Asume, further, that, for all x, y ∈ [0, +∞), we
have
ϕ(x + y) ≤ ϕ(x) + ϕ(y).
Let µ(X) = 1. For f , g ∈ M, define
Z
d(f , g) = ϕ(|f − g|) dµ.
X
5.7 Exercises 117
Differentiation
We would like to investigate how far such a result is true when we deal
with functions which are only differentable almost everywhere and with
the Lebesgue integral of the derivative of that function.
We will first study, briefly, various classes of functions which are dif-
ferentiable almost everywhere. Following the treatment as in Royden [6],
© Hindustan Book Agency 2019 and Springer Nature Singapore Pte Ltd. 2019 118
S. Kesavan, Measure and Integration, Texts and Readings in
Mathematics 77, https://doi.org/10.1007/978-981-13-6678-9_6
6.1 Monotonic functions 119
Proof: We observe, first of all, that the intervals can be of any type:
open, closed, or half-open. The addition, or removal, of end-points do
not change the results since these points constitute a set of measure
zero. Consequently, without loss of generality, we will assume that the
intervals are all closed.
length less than, say, d2 . Then, it is clear that I will not intersect
any of the intervals Ik , 1 ≤ k ≤ n. Set
kn = sup m1 (I).
I∈I
I ∩ Ik = ∅
1≤k≤n
Set
R = E\ ∪N
k=1 Ik .
∞
X ∞
X
µ∗ (R) ≤ m1 (Jk ) ≤ 5 m1 (Ik ) < ε.
k=N +1 k=N +1
Remark 6.1.1 There are several similar results in the literature, each of
them being called a Vitali covering lemma. The spirit of these are all the
same: a set of finite outer-measure in RN is covered by basic open sets
of arbitrarily small size and we can find a finite disjoint sub-collection
which almost completely covers the given set, i.e. the outer-measure of
the uncovered portion can be made as small as we wish.
f (x+h)−f (x)
D+ f (x) = lim suph↓0 h ,
f (x)−f (x−h)
D− f (x) = lim suph↓0 h ,
f (x+h)−f (x)
D+ f (x) = lim inf h↓0 h ,
f (x)−f (x−h)
D− f (x) = lim inf h↓0 h .
We have that f is differentiable at x if, and only if, these four values co-
incide, and, in that case, we denote the common value as f 0 (x), which is
the derivative of f at x. Notice that we always have D+ f (x) ≥ D+ f (x)
and D− f (x) ≥ D− f (x) at any point x ∈ (a, b).
We can write
E = ∪r,s∈Q Ers ,
r>s
where
Ers = {x ∈ (a, b) | D+ f (x) > r > s > D− f (x)}.
Since Q is countable, again, it suffices to show that m1 (Ers ) = 0.
The collection of all such closed intervals, as x varies over Ers , forms a
Vitali covering of Ers and so by the Vitali covering lemma, we can find
a disjoint collection of such intervals {I1 , · · · , IN } such that the union of
their interiors covers a set A ⊂ Ers such that
µ∗ (A) > m − ε.
N
X N
X
(f (xk ) − f (xk − hk )) < s hk < sm1 (U ) < s(m + ε). (6.1.3)
k=1 k=1
Now, let y ∈ A. Then, for h0 sufficiently small, we have that (y, y+h0 )
is contained in an interval Ik , where 1 ≤ k ≤ N and
R R
= n [b,b+ 1 ) f dm1 −
n
(a,a+ 1 ] f dm1
n
R
= f (b) − n 1 f
(a,a+ n ] dm1 .
124 6 Differentiation
Thus,
Z Z
0
f dm1 ≤ f (b) − lim sup n f dm1 ≤ f (b) − f (a),
n→∞ 1
(a,b) (a,a+ n ]
since f (x) ≥ f (a) for all x ∈ (a, a + n1 ]. This completes the proof.
Define
n
X
t(P, f ) = |f (xi ) − f (xi−1 )|.
i=1
where the supremum is taken over all possible partitions of the interval
[a, b]. The function f is said to be of bounded variation over the
interval [a, b] if Tab (f ) < +∞.
Example 6.2.1 Let f : [a, b] → R be Lipschitz continuous, i.e. there
exists L > 0 such that, for all x, y ∈ [a, b], we have
[a, b].
Thus,
t(P, f ) = p(P, f ) + n(P, f ),
(6.2.2)
f (b) − f (a) = p(P, f ) − n(P, f ).
Define
Thus,
f (x) = g(x) − (h(x) − f (a)),
which completes the proof.
128 6 Differentiation
|f 0 (x)| ≤ |(Pax (f ))0 | + |(Nax (f ))0 | = (Pax (f ))0 + (Nax (f ))0 = (Tax (f ))0 ,
R PN R
= [a,b] i=1 yi fi dm1 ≤ [a,b] |y||f | dm1
R
= |y| [a,b] |f | dm1 ,
from which (6.2.5) follows on dividing throughout by |y|.
130 6 Differentiation
Therefore,
R xi 0 0
xi−1 |f (t)| dt − ε(xi − xi−1 ) ≤ |f (xi )|(xi − xi−1 )
R
xi
= xi−1 (f 0 (t) + f 0 (xi ) − f 0 (t)) dt
R
xi 0
≤ xi−1 f (t) dt
R
xi
+ xi−1 (f 0 (xi ) − f 0 (t)) dt
Thus,
Z b
|f 0 (t)| dt ≤ Tab (f ) + 2ε(b − a).
a
Since ε > 0 was arbitrarily chosen, we get
Z Z b
|f 0 | dm1 = |f 0 (t)| dt ≤ Tab (f ),
[a,b] a
which completes the proof.
In this section, we will show that the derivative of the indefinite intergal
of an integrable function is equal to the integrand, almost everywhere.
Proposition 6.3.1 Let f : [a, b] → R be an integrable function. The
indefinite integral of f , defined by
Z
F (x) = f dm1 , x ∈ [a, b],
[a,x]
132 6 Differentiation
Given ε > 0, we know that (cf. Proposition 5.3.2) there exists δ > 0
such that, whenever |x − y| < δ, we have
Z
|f | dm1 < ε,
[x,y]
Remark 6.3.1 The above result shows that in order that a function
be the indefinite integral of an integrable function, it must be at least a
function of bounded variation. In fact, we will see in the next section that
even more will be required. Thus, in general, a differentiable function
may not be the indefinite integral of its derivative, even if the derivative
is integrable. It needs to be at least a uniformly continuous function of
bounded variation, with additional properties.
for all x ∈ [a, b]. Then f (x) = 0 for almost every x ∈ [a, b].
6.3 Differentiation of an indefinite integral 133
Proof: Set
Assume that m1 (E+ ) > 0. Then (cf. Proposition 2.2.2), there exists
a closed set F ⊂ E+ such that m1 (F ) > 0. Set U = (a, b)\F , which
is open. Then, U can be written as the disjoint union of a countable
number of half-open intervals (cf. Lemma 2.2.1):
U = ∪∞
n=1 [an , bn ).
Set
f n = f χ ∪n [ak ,bk )
.
k=1
Since f > 0 onR F and since F has positive measure, we have (cf. Propo-
sition 5.2.2), F f dm1 > 0 and so since
Z Z Z
0 = f dm1 = f dm1 + f dm1 ,
[a,b] U F
R
we deduce
R that U f dm1 6= 0. Consequently, there exists n ∈ N such
that [an ,bn ) f 6= 0. But this is a contradiction, since
Z Z Z
f dm1 = f dm1 − f dm1 = 0,
[an ,bn ) [a,bn ) [a,an )
R R
= limn→∞ n [c,c+ 1 ] F dm1 − [a,a+ 1 ] F dm1 .
n n
(We have used the result of Exercise 5.3.) Now, since F is uniformly
continuous, given ε > 0, there exists N ∈ N such that for all n ≥ N ,
and for all x ∈ [a, b), we have |F (x + n1 ) − F (x)| < ε. Consequently, for
any x ∈ [a, b) and for any n ≥ N , we have
Z Z 1
x+ n
n F dm1 − F (x) = n (F (t) − F (x)) dt ≤ ε.
[x,x+ 1 ] x
n
It now follows, from the preceding arguments, that for almost every
x ∈ [a, b], we have
From the previous section, we see that in order that a given function f :
[a, b] → R be written as an indefinite integral of an integrable function, it
must be at least uniformly continuous and of bounded variation. We will
now introduce a new concept which will provide both a necessary and
sufficient condition for a function to be written as an indefinite integral.
we have
n
X
|f (yk ) − f (xk )| < ε. (6.4.2)
k=1
The existence of δ > 0, given ε > 0, such that (6.4.2) is true follows
from Proposition 5.3.2. (See also Remark 5.3.2).
Our aim now is to prove the converse of the result in the above
example, thereby establishing absolute continuity as the necessary and
sufficient condition for a function to be written as an indefinite integral
(of its derivative).
Proposition 6.4.1 Let f : [a, b] → R be absolutely continuous. Then,
f is of bounded variation over [a, b]. In particular, f is differentiable
almost everywhere in [a, b].
Proof: Let δ > 0 correspond to ε = 1 in the definition of absolute
continuity of f . Let K be the integral part of 1 + (b−a) δ . Then, given
any partition P of [a, b], we can refine it to a partition P 0 such that the
constituent intervals of P 0 can be grouped into K sets of intervals, each
with total length less than δ. Then,
t(P, f ) ≤ t(P 0 , f ) ≤ K.
Thus Tab (f ) ≤ K < +∞.
Proposition 6.4.2 Let f : [a, b] → R be an absolutely continuous func-
tion such that f 0 = 0 almost everywhere in [a, b]. Then f is a constant
function.
Proof: Let c ∈ (a, b] be arbitrarily chosen. Set
E = {x ∈ (a, c) | f 0 (x) = 0}.
By hypothesis, we have that m1 (E) = c − a. Let ε and η, be arbitrarily
small positive numbers.
138 6 Differentiation
and
n
X
|xk+1 − yk | < δ.
k=0
Since ε and η were arbitrarily chosen, it follows that f (c) = f (a), and
this completes the proof, since c was fixed arbitrarily in (a, b].
0 0
R R
= [a,b] F1 dm1 + [a,b] F2 dm1
P2
≤ i=1 (Fi (b) − Fi (a)) < +∞.
Thus, F 0 is integrable. Let
Z
G(x) = F 0 dm1 .
[a,x]
or, equivalently, Z
F (x) = F (a) + F 0 dm1 .
[a,x]
6.5 Exercises
Show that f (x+) and f (x−) always exist. Deduce that the set of dis-
continuities of f is at most countable.
140 6 Differentiation
x sin x1 , if x 6= 0,
f (x) =
0, if x = 0.
x2 sin x1 , if x 6= 0,
f (x) =
0, if x = 0,
is of bounded variation.
Tab (f ) = 0|
R
[a,b] |f dm1 ,
Pab (f ) = 0 )+
R
[a,b] (f dm1 , and,
Nab (f ) = 0 )−
R
[a,b] (f dm1 .
Change of variable
|T (a + h) − T (a) − A(h)|
lim = 0, (7.1.1)
h→0 |h|
∂T
where { ∂x i
(a)}Ni=1 are the usual partial derivatives of T at the point
a. It can be shown that if T is differentiable at a ∈ U , then all partial
derivatives of T exist at that point. Thus, the action of T 0 (a) on a vector
h = (h1 , · · · , hN ) is given by
N
X ∂T
T 0 (a)(h) = (a)hi .
∂xi
i=1
Example 7.1.4 Let N > 1 and M > 1. Let T (x) = (T1 (x), · · · , TM (x)),
where Ti is a mapping of U ⊂ RN into R. If T is differentiable at a
point a ∈ U , then all the Ti , 1 ≤ i ≤ M are also differentiable at a. The
derivative T 0 (a) can now be represented by an M × N matrix. We have,
for h = (h1 , · · · , hN ) ∈ RN ,
∂T1 ∂T1
(a) . . . (a)
∂x1 ∂xN h1
··· ··· ··· ···
T 0 (a)(h) = .
··· ··· ··· ···
∂TM ∂TM
∂x1 (a) . . . ∂xN (a)
hN
We will now recall (without poof) some important results from the
differential calculus in RN , which are well known when N = 1.
Notation
Remark 7.1.3 The mean value theorem has numerous applications. For
instance, let T : U → R be a mapping such that its partial derivatives
∂T
exist at all points of U . Assume that the mappings x 7→ ∂x i
(x) are
continuous at a point a ∈ U for all 1 ≤ i ≤ N . Then, using the mean
value theorem, we can show that T is differentiable at a (cf. Example
7.1.3).
S = {x ∈ U | JT (x) = 0}.
for all x, y ∈ C, by virtue of the mean value thoerem (cf. Theorem 7.1.2).
7.3 Diffeomorphisms
for every y ∈ C.
e Thus, it follows that for any such sub-cube,
e ≤ LN mN (C).
mN (T (C)) e (7.3.2)
Let ν < δ so that the above relations are valid throughout C. Let C e
be any sub-cube of C, as described earlier. Then, by virtue of Theorem
2.3.3, we have
Thus,
e < (1 + ε)N |JT (x)|mN (C).
mN (T (C)) e (7.3.4)
On the other hand, we also have from (7.3.3), that,
Z
|JT (y)|dmN (y) > (1 − ε)|JT (x)|mN (C).
e (7.3.5)
C
e
K ⊂ W ⊂ W ⊂ U.
Let
L = max kT 0 (x)k.
x∈W
(1 + ε)N
Z
0
mN (T (C (x))) ≤ |JT | dmN . (7.3.9)
1−ε C 0 (x)
(We must remember that each C 0 (x) is a sub-cube of one of the original
cubes. The estimates in Sard’s theorem and Lemma 7.4.1 were observed
7.3 Diffeomorphisms 151
mN (T (C 0 (x)) + mN (T (C 0 (x))
P P
≤ x∈Js x∈Jns
(1+ε)N
mN (C 0 (x)) +
P P R
≤ C(K)ε x∈Js 1−ε x∈Jns C 0 (x) |JT | dmn
(1+ε)N R
≤ C(K)εmN (W ) + 1−ε W |JT | dmn ,
where
C(K) = 2ωN −1 LN −1 .
We have used here the disjointness of the cubes C 0 (x) when using mN (W )
as an upper bound for x∈Js mN (C 0 (x)) and when using the integral
P
over W as an upper bound for the sum of the integrals over the cubes
C 0 (x).
Since ε can be chosen arbitrarily small, we get
Z
mN (T (K)) ≤ |JT | dmN .
W
Consequently, we have, by Lemma 7.3.2,
Z Z
mN (T (K)) ≤ inf |JT | dmN = |JT | dmN .
W open W K
W compact
K⊂W ⊂W ⊂U
Let W ⊂ U be an open set. Then, W can be written as the countable
increasing union of compact sets, i.e.
W = ∪∞
n=1 Kn ,
where the sets Kn , n ∈ N are all compact and Kn ⊂ Kn+1 for all n ∈ N.
Then, since T is a homeomorphism,
T (W ) = ∪∞
n=1 T (Kn ),
and for all n ∈ N,we have that T (Kn ) ⊂ T (Kn+1 ) and T (Kn ) is compact.
Thus,
Z Z
mN (T (W )) = lim mN (T (Kn )) ≤ lim |JT | dmN ≤ |JT | dmN .
n→∞ n→∞ K W
n
But Z Z
mN (T (K)) ≤ |JT | dmN ≤ |JT | dmN ,
K E
from which (7.3.7) follows.
R
= V f (y) dmN (y),
since |JT (T −1 (y))|.|JT −1 (y)| = 1 for all y ∈ V . This gives the reverse
inequality of (7.3.10), thereby establishing (7.3.11).
Corollary 7.3.3 Let U and V be bounded open subsets of RN and let
T be a diffeomorphism of U onto V . If E is any Borel subset of U , then
Z
mN (T (E)) = |JT | dmN . (7.3.12)
E
to get
Z 1 Z 1
f (x) dx = f (−y) dy. (7.3.13)
−1 −1
Product spaces
Let (X, S, µ) and (Y, T , λ) be two measure spaces. We would like to de-
fine a σ-algebra and a measure on the product X ×Y which is compatible
with the structures given on X and Y and also relate the process of in-
tegration with respect to this measure with the processes of integration
on X and Y .
Ex = {y ∈ Y | (x, y) ∈ E}.
E y = {x ∈ X | (x, y) ∈ E}.
Thus, Ex ⊂ Y and E y ⊂ X.
and
(∪∞ ∞
i=1 Qi )\P = ∪i=1 (Qi \P ) ∈ M.
Finally {P \Qi }∞
i=1 is a decreasing sequence of sets in M. Hence
P \(∪∞ ∞
i=1 Qi ) = ∩i=1 (P \Qi ) ∈ M.
Thus, we see that ∪∞ i=1 Qi ∈ U (P ). In the same way, it is easy to see that
if {Qi }∞
i=1 is a decreasing sequence of sets in U (P ), then ∩∞i=1 Qi ∈ U (P )
as well. This completes the proof.
Lemma 8.1.2 Let X be a non-empty set and let R be an algebra of
subsets of X. Let M(R) denote the monotone class generated by R.
Then M(R) = S(R), the σ-algebra generated by R.
Proof: Let P ∈ R. If Q ∈ R, then, P ∪ Q, P \Q and Q\P belong to
R and hence to M(R) as well. Thus, if U (P ) is as defined in the pre-
ceding lemma, we have that Q ∈ U (P ). Since U (P ) is a monotone class
containing R, it follows that U (P ) ⊃ M(R).
∪∞ ∞
i=1 Ei = ∪n=1 Fn ∈ M(R).
{y ∈ Y | fx (y) > c} = Qx
Proof: Let U be the collection of all sets in S ×T such that (8.2.1) holds.
def
limi→∞ ϕi (x) = λ(Qx ) = ϕ(x), and
def
limi→∞ ψi (y) = µ(Qy ) = ψ(y).
R R
Since X ϕi dµ = Y ψi dλ, for each i, it follows from the monotone
convergence theorem that (8.2.1) holds. Thus Q ∈ U .
∪ni=1 Qi = ∪∞
n=1 Rn ∈ U .
Step 4. Let A ∈ S and B ∈ T be such that µ(A) < +∞ and λ(B) < +∞.
Let {Qi }∞
i=1 be a sequence of sets in U such that
A × B ⊃ Q1 ⊃ Q2 ⊃ · · · ⊃ Qi ⊃ Qi+1 ⊃ · · · .
162 8 Product spaces
Step 5. Since both the measure spaces are σ-finite, we can write
X = ∪∞ ∞
n=1 Xn and Y = ∪m=1 Ym ,
where {Xn }∞ ∞
n=1 and {Ym }m=1 are sequences of disjoint sets such that for
all n and m we have µ(Xn ) < +∞ and λ(Ym ) < +∞. Let Q ∈ S × T .
Set Qnm = Q∩(Xn ×Ym ). Let M be the collection of all sets Q in S ×T
such that Qnm ∈ U for all n and m. By Steps 2 and 4, it follows that
M is a monotone class. By Steps 1 and 3, it follows that all elementary
sets are in M. Thus, M is a monotone class containing all elementary
sets and is contained in S × T . It now follows, from Proposition 8.1.2,
that M = S × T .
Remark 8.2.1 In view of the above example, we can ask ourselves what
is the relationship between the product of Lebesgue measures and the
Lebesgue measure of the product space. We sketch the argument below.
Let ` = k + n and let us consider R` as the product space Rk × Rn .
We have the Borel sets B` , Bk and Bn in R` , Rk and Rn respectively.
Similarly we have the Lebesgue measurable sets L` , Lk and Ln as well.
Now, any open set in R` and can be expressed as the countable
disjoint union of (half-open) boxes (cf. Lemma 2.2.1). Thus, all open
sets are in Lk × Ln and so we deduce that
B` ⊂ L k × L n .
B` ⊂ L k × L n ⊂ L ` .
Cosequently,
This proves one part of (8.3.2). The proof of the other part is similar.
Then (8.3.2) holds for the triples (f ± , ϕ± , ψ± ) replacing the triple (f, ϕ, ψ).
All the integrals are now finite and so subtracting the relations for f −
from those of f + , we deduce (8.3.2) for the function f . This completes
the proof.
The first and the last term are referred to as iterated integrals.
Then
∞ X
X ∞ ∞
X
am,n = (am,1 + am,m+1 ) = 0,
m=1 n=1 m=1
while
∞
X
am,1 = +∞
m=1
and so
∞ X
X ∞
am,n = +∞.
n=1 m=1
0| 0|
R R
= [a,b] |f dm1 [a,b] |g dm1 < +∞.
0 (x)g 0 (y)χ
R R
= [a,b] [a,b] f E (x, y) dm1 (x) dm1 (y). (8.3.3)
The left-hand side of (8.3.3) is equal to
Z Z !
g (y) dm1 (y) f 0 (x) dm1 (x),
0
[a,b] [x,b]
168 8 Product spaces
which yields
Z Z
0
(g(b) − g(x))f (x) dm1 (x) = g(b)f (b) − g(b)f (a) − gf 0 dm1 .
[a,b] [a,b]
We have used here the fact that both f and g are absolutely continuous
and so the integral of the derivative is the difference of the values of the
function at the end points of the interval (cf. Theorem 6.4.1). Similarly,
the right-hand side of (8.3.3) is equal to
Z Z !
f 0 (x) dm1 (x) g 0 (y) dm1 (y),
[a,b] [a,y]
which yields
Z
−g(b)f (a) + g(a)f (a) + f g 0 dm1 .
[a,b]
Now, on one hand I2 = I12 , by the above reasoning. On the other hand,
by the previous example,
Z ∞ Z 2π Z ∞
−r 2 2
I2 = e r dθ dr = 2π e−r r dr.
0 0 0
F (t) = µ(E(t)),
Thus, Z Z
|f | dµ = F dm1 .
X [0,+∞)
ϕ(x, y) = x − y, ψ(x, y) = y.
These are continuous and hence Borel measurable. Thus (cf. Proposition
3.1.4), we have that the mappings
Assume that f and g are integrable as well. Since the Lebesgue mea-
sure agrees with the product measure on Borel measurable sets, we can
apply Fubini’s theorem. We have
R R
RN
RN f (x − y)g(y) dmN (y) dmN (x)
R R
≤ RN RN |f (x − y)|.|g(y)| dmN (y) dmN (x)
R R
= RN |g(y)| RN |f (x − y)| dmN (x) dmN (y).
x = r cos θ, y = r sin θ,
f (x) = fe(|x|).
172 8 Product spaces
ωN = mN (B).
If R > 0, then the linear map T (x) = Rx maps the open unit ball
diffeomorphically onto B(0; R), the open ball of radius R, and so we
have (cf. Theorem 2.3.)
mN (B(0; R)) = ωN RN .
Let us denote the closed ball centred at the origin and of radius
R by B(0; R). Let us assume that we have a continuous function f :
B(0; R) → R which is radial. Thus, f (x) = fe(|x|), where fe : [0, R] → R
is continuous. Consider a partition of the interval [0, R]:
Let
∆(P) = max (ri − ri−1 ).
1≤i≤n
Pn N N )
= i=1 f (ξi )ωN (ri
e − ri−1
Pn N −1
= i=1 f (ξi )N ωN ξi
e (ri − ri−1 ),
in view of (8.4.1). Since f is continuous, we have that
Xn Z R
N −1
lim f (ξi )N ωN ξi
e (ri − ri−1 ) = N ωN fe(r)rN −1 dr.
∆(P)→0 0
i=1
Thus we have
Z Z R
f dmN = N ωN fe(r)rN −1 dr. (8.4.2)
B(0;R) 0
The interested reader is referred to the books of Evans and Gariepy [2] or
Folland [3]. Essentially, the coarea formula says that when we integrate
over RN , we first integrate over the surface of sphere of radius r, centred
at the origin, and then integrate over r. If R > 0, we have
Z Z Z
f dmN = f (rx0 )rN −1 dσN −1 (x0 ) dm1 (r).
B(0;R) [0,R) S N −1
Example 8.4.1 (Volume of the unit ball) We now compute the value of
2
ωN = mN (B(0; 1)). We start with the function f (x) = e−|x| . We saw
earlier (cf. Example 8.3.6) that
Z
2 N
e−|x| dmN (x) = π 2 .
RN
Since f is a radial function, we can also compute it using polar coordi-
nates. By (8.4.3), we get
Z Z ∞
−|x|2 2
e dmN (x) = N ωN e−r rN −1 dr.
RN 0
Setting s = r2 , we get
Z Z ∞
−|x|2 N −s N −1 N N
e dmN (x) = ωN e s2 = ωN Γ ,
RN 2 0 2 2
8.5 Exercises 175
where Γ(s) is the familiar gamma function. Thus, equating the two
expressions we got for the integral, we obtain,
N N
π2 π2
ωN = N N
= N
,
2Γ 2 Γ 2 +1
Using the last mentioned property of the gamma function and the
√
fact that Γ 12 = π, we can easily verify that
4
ω2 = π and ω3 = π.
3
We can also see that
1 2 8 2
ω4 = π and that ω5 = π ,
2 15
and so on.
8.5 Exercises
where
E(t) = {x ∈ X | |f (x)| > t}.
8.3 (a) For x > 0, show that
Z
1
e−xt dm1 (t) = .
[0,+∞) x
(b) Use the above relation and Fubini’s theorem to show that
Z R
sin x π
lim dx = .
R→+∞ 0 x 2
176 8 Product spaces
∗ g = fb · gb.
f[
respectively.
(a) If f is a non-negative simple function, show that V ∗ (f ) and V∗ (f ) are
measurable in X × R (where R is equipped with the Lebesgue measure).
(b) If f and g are non-negative functions such that f (x) ≤ g(x) for all
x ∈ X, show that V ∗ (f ) ⊂ V ∗ (g) and that V∗ (f ) ⊂ V∗ (g).
(c) Let {fn }∞ n=1 be a sequence of non-negative measurable functions
defined on X. If fn ↑ f , show that {V∗ (fn )}∞
n=1 is an increasing sequence
of sets whose union is V∗ (f ). If fn ↓ f , show that {V ∗ (fn )}∞ n=1 is a
decreasing sequence of sets whose intersection is V ∗ (f ).
(d) If f is a non-negative measurable function defined on X, show that
V ∗ (f ) and V∗ (f ) are measurable subsets of X × R.
(e) If f is any measurable real-valued function defined on X, show that
its graph, G(f ), is a measurable subset of X × R, where
Signed measures
If{En }∞
n=1 is a decreasing sequence of measurable sets such that µ(Em )
is finite for some m ∈ N, then
µ(∩∞
n=1 En ) = lim µ(En ). (9.1.3)
n→∞
Remark 9.1.2 The empty set is both a positive and a negative set.
Let {Bi }∞
i=1 be a sequence of sets in N such that µ(Bi ) ↓ β. If B =
∪∞ B
i=1 i , we have seen that B ∈ N and so β ≤ µ(B). On the other hand,
if we set Cn = ∪ni=1 Bi , then µ(B) = limn→∞ µ(Cn ), by Proposition
9.1.3. But, by Proposition 9.1.4, we have that µ(Cn ) ≤ µ(Bn ) and so
µ(B) ≤ β. Thus, µ(B) = β. In particular, by our assumption on µ, we
have that β is finite.
1
µ(E0 \E1 ) = µ(E0 ) − µ(E1 ) ≤ µ(E0 ) − < 0.
k1
Step 3. We can now apply the procedure of Step 2 to the set E0 \E1 .
Then there exist measurable subsets of E0 \E1 with positive measure,
182 9 Signed measures
and let k2 be the smallest positive integer with the property that there
exists such a set E2 with µ(E2 ) ≥ k12 . (In other words, at each stage,
we choose a set with positive measure with the measure being as large
as possible.)
Proceeding in this way, for each positive integer i, there exists a
measurable set of positive measure contained in E0 \ ∪i−1
k=1 Ek and let ki
be the smallest positive integer with the property that there exists such
a measurable subset Ei with µ(Ei ) ≥ k1i . Then, since Ei ⊂ E0 \ ∪i−1
`=1 E` ,
the sets {Ei }∞
i=1 are clearly disjoint, and so
∞
X
µ(Ei ) = µ(∪∞
i=1 Ei ) < +∞,
i=1
since ∪∞
i=1 Ei ⊂ E0 , which has finite measure. In particular, it follows
that µ(Ei ) → 0 as i → ∞ and so ki → ∞.
Remark 9.1.4 The decomposition of X into two disjoint sets, one posi-
tive and the other negative, is called a Hahn decompositon of X. Such a
decomposition is not unique. Let X = A ∪ B be a Hahn decomposition
and assume that there exists N ⊂ B such that µ(N ) = 0. Let F ⊂ N .
Then µ(F ) ≤ 0. If µ(F ) < 0, then 0 = µ(N ) = µ(F ) + µ(N \F ), which
implies that µ(N \F ) > 0, which is not possible since B and, hence, N ,
is a negative set. Thus, µ(F ) = 0, for all F ⊂ N . Then it is clear that
X = (A ∪ N ) ∪ (B\N ) gives another Hahn decomposition of X.
The situation described in the preceding remark is, in fact, the only
way non-uniqueness can occur for the Hahn decomposition. More pre-
cisely, we have the following result.
9.1 Hahn and Jordan decompositions 183
and so µ(E ∩ (A1 \A2 )) ≤ 0. Thus µ(E ∩ (A1 \A2 )) = 0 and, similarly,
µ(E ∩ (B1 \B2 )) = 0 as well. Consequently,
for E ∈ S.
Then ν << µ.
Example 9.2.2 Let (X, S) be a measure space and let µ and ν be mea-
sures defined on it. Then µ << µ + ν and ν << µ + ν.
Let, for E ∈ L1 ,
Z
µi (E) = fi dm1 , i = 1, 2.
E
while
Z 1
1
µ2 (X) = x dx = 6= 0.
0 2
Thus, if µ2 << µ1 and if µ1 (E) = 0, it does not imply that µ2 (E) = 0.
0 ≤ |µ|(E ∩ A) ≤ |µ|(E) = 0,
0 ≤ |µ|(E ∩ B) ≤ |µ|(E) = 0.
Proof: Assume the contrary. Then, there exists ε > 0 such that, for
every n ∈ N, there exists a set En ∈ S with |µ|(En ) < 21n and |ν|(En ) ≥
ε. Set E = lim supn→∞ En . Then, for every n ∈ N,
∞
X 1
|µ|(E) ≤ |µ|(Em ) < ,
m=n
2n−1
188 9 Signed measures
and so |µ|(E) = 0. Since |ν| is finite (cf. Exercise 1.10 (b)), we have
|ν|(E) ≥ ε, which contradicts the absolute continuity of ν with respect
to µ.
Example 9.2.6 The above result is not true, in general, if ν is not finite.
Let X = N and let S = P(N). Let µ({n}) = 2−n and let ν({n}) = 2n .
Then µ and ν define measures and since the only set, in either case, with
measure zero is the empty set, we have that ν << µ. However, for any
δ > 0, we can find n0 ∈ N such that for all n ≥ n0 , we have µ({n}) < δ,
while {ν({n})}n≥n0 is unbounded.
1
0 ≤ ν(B0 ) ≤ µ(B0 ).
n
Thus, ν(B0 ) = 0. Since A0 = B0c , and since ν is not identically zero, it
follows that ν(A0 ) > 0. By absolute continuity, it follows that µ(A0 ) > 0
as well. Then, there exists n ∈ N such that µ(An ) > 0. We can now set
A = An and ε = n1 .
Similarly, we have
Step 2. Let us denote by L(µ), the set of all measurable functions defined
on X which are integrable with respect to µ. Define
f ≥ 0 and
K = f ∈ L(µ) R .
E f dµ ≤ ν(E) for every E ∈ S
R
for every E ∈ S. Thus f ∈ K and X f dµ = εµ(A) > 0. Now set
Z
α = sup f dµ | f ∈ K .
X
Then,
0 < α ≤ ν(X) < +∞.
Step 3. Let {gn }∞
n=1 be a sequence of functions in K such that
Z
1
gn dµ > α − .
X n
≤ ν(E ∩ F c ) + ν(E ∩ F )
= ν(E).
Thus h ∈ K. But
Z Z
h dµ = f dµ + ηµ(F ) > α,
X X
Now, if E ∈ S, we have
dµ+ + fB dµ−
R R
= E∩A fA E∩B
R
= E (fA χA − fB χB ) dµ.
dν dν dµ
= ,
dλ dµ dλ
9.4 Singularity
A = {x ∈ X |f (x) ≥ 1} and B = Ac .
Then, Z
ν(A) ≥ d(µ + ν) = µ(A) + ν(A),
A
whence we deduce that µ(A) = 0. Define, for E ∈ S,
λ = ν0 − νe0 = νe1 − ν1 .
9.5 Exercises
Lp spaces
© Hindustan Book Agency 2019 and Springer Nature Singapore Pte Ltd. 2019 196
S. Kesavan, Measure and Integration, Texts and Readings in
Mathematics 77, https://doi.org/10.1007/978-981-13-6678-9_10
10.1 Basic properties 197
Lemma 10.1.1 Let 1 < p < ∞. Let p0 be its conjugate exponent. Then,
if a and b are non-negative real numbers, we have
1 1 a b
a p b p0 ≤ + 0. (10.1.3)
p p
f (t) = k(t − 1) − tk + 1,
where k ∈ (0, 1). Then f 0 (t) = k(1 − tk−1 ) ≥ 0 since 0 < k < 1. Thus,
f is an increasing function on [1, +∞) and, since f (1) = 0, we deduce
that
tk ≤ k(t − 1) + 1, (10.1.4)
for t ≥ 1 and for 0 < k < 1.
If a or b is zero, then (10.1.3) is obviously true. Let us assume,
without loss of generality (since p and p0 are conjugate exponents of
each other), that a ≥ b > 0. Then (10.1.3) follows from (10.1.4) on
setting t = ab , k = p1 and using the relation between p and p0 .
Let us now assume that 1 < p < ∞ so that 1 < p0 < ∞ as well. The
relation (10.1.5) is trivially true if kf kp (respectively, kgkp0 ) equals zero,
for then f (respectively, g) will be equal to zero almost everywhere. So
we assume further that kf kp 6= 0 and that kgkp0 6= 0. Then, by Lemma
10.1.1,
1 1 0
|f (x)g(x)| ≤ |f (x)|p + 0 |g(x)|p
p p
for all x ∈ X. Assume now that kf kp = kgkp0 = 1. Then, integrating
the above inequality over X, we get
Z
1 1
|f g| dµ ≤ + 0 = 1.
X p p
For the general case, apply the preceding result to the functions f /kf kp
and g/kgkp0 to get (10.1.5).
Thus,
p
0
kf + gkpp ≤ kf + gkpp (kf kp + kgkp ).
p
0
Dividing both sides by kf + gkpp and using, once again, the definition
of p0 , we get (10.1.6). The cases where p = 1 and p = ∞ follow trivially
from the inequality
N
! p1
X
kxkp = |xi |p ,
i=1
∞
!1
X p
p
kxkp = |xk | ,
k=1
Lp (µ) ⊂ Lq (µ)
R p
q 1− pq
= X |f | dµ (µ(X))
p
1− pq
= kfkqp .(µ(X))
which yields
kf kq ≤ C kf kp
where 1
− p1
C = (µ(X)) q .
This completes the proof.
Example 10.1.4 Nothing can be said about the reverse inclusions. For
√
example, if f (x) = 1/ x, then f ∈ L1 (0, 1) but f 6∈ L2 (0, 1).
`p ,→ `q .
kxkq ≤ kxkp .
|xi | ≤ kxkp ,
202 10 Lp spaces
Let 1 ≤ p < q < ∞. Let x ∈ `p . Assume, for the moment, that kxkp = 1.
Then, if x = (x1 , · · · , xi , · · ·), we have seen that |xi | ≤ 1 for all i. Now
∞
X ∞
X ∞
X
|xi |q = |xi |p |xi |q−p ≤ |xi |p = 1.
i=1 i=1 i=1
x
Thus, x ∈ `q and kxkq ≤ 1 = kxkp . Now, if x ∈ `p , consider y = kxk p
.
Then kykp = 1. Consequently, y ∈ `q and kykq ≤ 1. Thus, it follows
that x ∈ `q as well (since it is a constant multiple of y) and
kxkq
≤ 1,
kxkp
Thus,
p−1 1
kf kp ≤ kf k∞p kf k1p .
Consequently,
lim sup kf kp ≤ kf k∞ .
p→∞
Then 1
kf kp ≥ (kf k∞ )(µ(E)) p ,
10.1 Basic properties 203
Thus, we have
lim kf kp = kf k∞ .
p→∞
This justifies, to some extent, the notation kf k∞ for the norm given by
the essential supremum.
Then
Z Z
p
|fn − fm | dµ ≥ |fn − fm |p dµ ≥ εp µ(An,m (ε)).
X An,m (ε)
Thus,
kfn − fm kpp
µ(An,m (ε)) ≤ ,
εp
and, since the right-hand side of the above inequality can be made arbi-
trarily small for large, n and m, the same is true for µ(An,m (ε)) as well
and that completes the proof.
Proof: Since the norm defines a continuous function on any normed lin-
ear space, it follows that if fn → f in Lp (µ), we have that kfn kp → kf kp .
10.2 Approximation
ϕx = χB(x;r) .
Set
∞ 1
Ux = f ∈ L (Ω) | kf − ϕx k∞ < .
2
208 10 Lp spaces
kϕk∞ ≤ kf k∞ .
En = {x ∈ E | |f (x)| ≤ n}.
Step 4. Now, by the Tietze extension theorem (cf., for instance, Sim-
mons [9]) we can find a continuous function g : RN → R such that
kgk∞ ≤ m and such that g = f on K.
Using this, and the fact that F ≥ 0, in (10.3.4), and applying Hölder’s
inequality, we get
p−1
Z
p
kF kp = F p−1 f dm1 ≤ kf kp kF p−1 kp0 ,
p (0,∞)
10.3 Some applications 211
Step 4. Now, for any x > 0, fn → f in Lp (0, x) and so, since (0, x) has
finite measure, fn → f in L1 (0, x). Consequently,
Z Z
fn dm1 → f dm1 ,
(0,x) (0,x)
212 10 Lp spaces
and so Fn (x) → F (x) for each x > 0. But, for a subsequence Fnk → G
pointwise, almost everywhere, and so F = G almost everywhere. Thus,
it follows that F ∈ Lp (0, ∞) and that Fn → F in Lp (0, ∞). Since
p
kFn kp ≤ kfn kp ,
p−1
1 − e−1
F (x) ≥ , x ≥ 1,
x
and so F is not integrable on (1, ∞) and so it cannot be integrable over
(0, ∞) either.
10.4 Duality
Proof: Let U = {t ∈ R | |t| > K}. This is an open set in R and hence
can be written as the countable union of open intervals. Let (a−r, a+r)
be one such interval. Let
E = {x ∈ X | g(x) ∈ (a − r, a + r)}.
Then Z
1
|AE (g) − a| =
(g − a) dµ < r.
µ(E) E
λ(E) = T (χE ).
χA∪B = χA + χB ,
By splitting any bounded function f into its positive and negative parts,
f ± , we deduce that (10.4.3) is true for any bounded measurable function.
Thus, Z
1
µ(E) g dµ ≤ kT k.
E
It then follows, from Lemma 10.4.1, that |g| ≤ kT k almost everywhere.
Thus, in this case g ∈ L∞ (µ) and kgk∞ ≤ kT k.
En = {x ∈ X | |g(x)| ≤ n}, n ∈ N.
0
Set f = χEn |g|p −1 ψ. Then
0 0
|f |p = χEn |g|p p−p = χEn |g|p ,
and so
Z Z 1
p
p0 p0
|g| dµ ≤ kT k · kf kp = kT k |g| dµ ,
En En
which yields
Z 10
p
p0
|g| dµ ≤ kT k.
En
But En ↑ X and so, by the monotone convergence theorem, we have
Z 10
p
p0
|g| dµ ≤ kT k.
X
218 10 Lp spaces
0
Thus, g ∈ Lp (µ) and kgkp0 ≤ kT k.
0
Step 5. Let 1 ≤ p < ∞. Then, we have seen that g ∈ Lp (µ) and that
kgkp0 ≤ kT k. Further, since µ(X) is finite, simple functions are dense
in Lp (µ) (cf. Lemma 10.2.1) and so L∞ (µ) is also dense in Lp (µ). Both
sides of (10.4.3) define continuous linear functionals on Lp (µ) and agree
on the dense subspace L∞ (µ) and so, they agree on all of Lp (µ). Thus,
we get that, in fact, T = Tg , in which case, we have that
kT k = kTg k ≤ kgkp0 ≤ kT k.
Thus T = Tg and kT k = kTg k = kgkp0 . This completes the proof.
where f ∈ Lp (ν) in the first line above and f ∈ Lp (µ) in the second.
This shows that S defines a continuous linear functional on Lp (ν) and
that kSk = kT k.
Theorem 10.4.2 (Riesz representation theorem) Let (X, S, µ) be a σ-
finite measure space. Let 1 ≤ p < ∞. Let T be a continuous linear
0
functional on Lp (µ). Then, there exists a unique g ∈ Lp (µ) such that
T = Tg and, further, kT k = kgkp0 .
Proof: We write X = ∪∞ n=1 Xn as a disjoint union of sets of finite
measure. We adopt the notation and definitions made in the preceding
paragraphs and define the function h and the measure ν as before. Given
a linear functional T on Lp (µ), we define, as before, the linear functional
0
S on Lp (ν). Then, by the preceding theorem, there exists ge in Lp (ν)
such that Z
S(f ) = f gedν, for all f ∈ Lp (ν),
X
and such that kSk = kgkLp0 (ν) .
1
Define g = h p0 ge if 1 < p < ∞ and g = ge, if p = 1.
0 0 0
g |p dν = kSkp = kT kp .
R
= X |e
kgkL∞ (µ) = ke
g kL∞ (ν) = kSk = kT k.
If f ∈ L1 (µ), then
Z Z
f g dµ = f geh−1 dν = S(h−1 f ) = T (f ).
X X
220 10 Lp spaces
This proves the result for p = 1 and the proof of the theorem is com-
plete.
Example 10.4.1 The above result is not true when p = ∞. For exam-
ple, consider the interval (0,1). Then C[0, 1] is a subspace of L∞ (0, 1).
If f ∈ C[0, 1], define
T (f ) = f (0).
This defines a linear functional on C[0, 1] and
|T (f )| = |f (0)| ≤ kf k∞ .
1 − nt, if 0 ≤ t ≤ n1 ,
fn (t) =
1
0, if ≤ t ≤ 1.
n
But |fn g| ≤ |g| for all n and g is integrable, while (fn g)(t) → 0 for all
0 < t ≤ 1. Thus, by the dominated convergence theorem, we have that
T (fn ) → 0, which is a contradiction.
Remark 10.4.1 The spaces Lp (µ) are reflexive (cf. for instance, Ke-
savan [5]) if 1 < p < ∞ since we have the canonical identifications
0 0
(Lp (µ))0 = Lp (µ) and (Lp (µ))0 = Lp (µ). The spaces L1 (µ) and L∞ (µ)
are not reflexive. We have that (L1 (µ))0 = L∞ (µ) but the reverse iden-
tity fails.
10.5 Convolutions 221
10.5 Convolutions
Theorem 10.5.1 Let 1 < p < ∞. Let f ∈ L1 (RN ) and let g ∈ Lp (RN ).
Then f ∗ g is well-defined. Further, f ∗ g ∈ Lp (RN ) and
0
Proof: Let p0 be the conjugate exponent of p. Let h ∈ Lp (RN ). Then
(x, y) 7→ f (x − y)g(y)h(x)
is measurable and using Hölder’s inequality and the translation invari-
ance of the Lebesgue measure, we have
R R
RN RN |f (x − y)g(y)h(x)| dmN (x)dmN (y)
R R
= RN |h(x)| RN |f (x − y)g(y)| dmN (y)dmN (x)
R R
= RN |h(x)| RN |f (w)g(x − w)| dmN (w)dmN (x)
R R
= RN |f (w)| RN |h(x)||g(x − w)| dmN (x)dmN (w)
exists for almost all x. We can choose h(x) 6= 0 for all x (for instance,
h(x) = exp(−|x|2 ), which belongs to all Lp spaces) and so we deduce
that f ∗ g defined via (10.5.1) is well-defined. Further, by the preceding
computation, it follows that
Z
h 7→ (f ∗ g)h dmN
RN
0
is a continuous linear functional on Lp (RN ) whose norm is bounded by
the quantity kgkp kf k1 which shows, by the Riesz representation theo-
rem, that f ∗ g ∈ Lp (RN ) and that (10.5.3) holds.
Remark 10.5.1 Notice that (10.5.2) is the same as (10.5.3) for the case
p = 1. The relation (10.5.3) is a particular case of Young’s inequality.
Let 1 ≤ p, q, r < ∞ be such that
1 1 1
+ = 1+ .
p q r
If f ∈ Lp (RN ) and g ∈ Lq (RN ), then f ∗ g is well-defined via (10.5.1),
f ∗ g ∈ Lr (RN ) and
kf ∗ gkr ≤ kf kp kgkq .
10.5 Convolutions 223
The actual pairing and order will be unimportant since we have commu-
tativity and associativity. The convolution product of functions occuring
within any pair of parantheses is well-defined since at least one of them
will have compact support as shown by the following result.
A + B = {x + y | x ∈ A, y ∈ B}.
The above integral needs to be taken only over a compact set K(x)
containing the supports of the functions y 7→ f (x−y) and y 7→ f (x+h−
y). For example, we can take K(x) = x+B(0; R)+B(0; 1) where B(0; r)
denotes the closed ball in RN with centre at the origin and radius r, and
R > 0 is such that supp(f ) ⊂ B(0; R). Since f has compact support, it
is uniformly continuous and so, given ε > 0, there exists η > 0 such that
|f (u) − f (v)| < ε whenever |u − v| < η. Thus, if |h| < η (we can assume
that 0 < η < 1), we get
Z
|(f ∗ g)(x + h) − (f ∗ g)(x)| ≤ ε |g| dmN
K(x)
R ∂f
= K(x) ∂xi ((x − y + θhei )g(y) dy
10.5 Convolutions 225
∂f
where θ ∈ (0, 1) (and depends on x, y and h). Since ∂x i
is assumed to be
continuous, it is bounded on the compact set K(x) and so the integrand
in the last integral above is bounded by M |g(y)| which is integrable on
the compact set K(x). Further, as h → 0, the integrand converges to
∂f
∂xi (x−y)g(y). Thus, by the dominated convergence theorem, we deduce
the validity of (10.5.4).
|α| = α1 + · · · + αN .
∂ |α|
Dα f = .
∂xα1 1 · · · ∂xαNN
∂4f
Dα f = .
∂x31 ∂x3
exp(−x−2 ) if x > 0,
f (x) =
0 if x ≤ 0.
Then f ∈ C ∞ (R).
Example 10.5.2 This is a slight, but very useful, variation of the pre-
ceding example. Let x = (x1 , · · · , xN ) ∈ RN . Let
N
! 12
X
|x| = |xi |2 . (10.5.5)
i=1
10.5 Convolutions 227
where Z
κ−1 = exp(−1/(1 − |x|2 )) dx.
|x|≤1
Definition 10.5.2 The family of functions {ρε }ε>0 is called the family
of mollifiers.
Proof: (i) Let x ∈ RN . Then, given η > 0, there exists δ > 0 such that
for all |y| < δ, we have |f (x − y) − f (x)| < η. Thus, if ε < δ, we have,
on observing that the integral of ρε is unity and that this function is
supported on B(0; ε),
Z
(ρε ∗ f )(x) − f (x) = (f (x − y) − f (x))ρε (y) dmN (y)
|y|≤ε
supp(ρε ∗ f ) ⊂ K + B(0; ε)
Then
But by (10.5.2)
η
kρε ∗ (g − f )kp ≤ kρε k1 kg − f kp <
3
since the integral of ρε is unity and the result now follows immediately.
Norosa, 1998.)
10.6 Exercises
kf gkr ≤ kf kp kgkq .
10.2 Let (X, S, µ) be a measure space. Let 1 ≤ p < ∞. Let {fn }∞ n=1
be a sequence in Lp (µ). Assume that there exists g ∈ Lp (µ) such that
|fn | ≤ g for each n ∈ N. If fn → f pointwise, show that f ∈ Lp (µ) and
that fn → f in Lp (µ).
Define, for x ∈ X, Z
g(x) = f (x, y) dµ(y).
X
10.6 Exercises 231
(This is called the amplitude-phase form of the series.) Write down the
relations between an , bn and dn , φn .
(b) Using the amplitude-phase form of a trigonometric series, show that
if the series converges pointwise over a set E whose (Lebesgue) measure
is strictly positive, then an → 0 and bn → 0 as n → ∞. (This is called
the Cantor-Lebesgue theorem.)
10.11 Let c0 denote the space of all real sequences which converge to
zero, equipped with the norm k · k∞ .
(a) Show that c0 is complete.
(b) Show that c00 is isometrically isomorphic to `1 .
we have
1
(f + g)
< 1 − δ.
2
p
10.6 Exercises 233
10.13 Let (X, S, µ) be a measure space and let 1 < p < ∞. Let p0
denote the conjugate exponent of p.
0
(a)If g ∈ Lp (µ), define
0
|g(x)|p −2 g(x), if g(x) 6= 0,
f (x) =
0, if g(x) = 0.
Show that f ∈ Lp (µ).
(b) Define the continuous linear functional (as in Section 10.5)
Z
Tg (f ) = f g dµ, for all f ∈ Lp (µ).
X
Show that kTg k = kgkp0 .
(c) Given that every uniformly convex Banach space is reflexive (cf. Ke-
savan [5], Theorem 5.5.1), deduce that Lp (µ) is reflexive for all p such
that 1 < p < ∞.
0
(d) Deduce that the dual of Lp (µ) is isometrically isomorphic to Lp (µ)
for every 1 < p < ∞.
© Hindustan Book Agency 2019 and Springer Nature Singapore Pte Ltd. 2019 235
S. Kesavan, Measure and Integration, Texts and Readings in
Mathematics 77, https://doi.org/10.1007/978-981-13-6678-9
Index