Pmath 450
Pmath 450
Laurent W. Marcoux
We are now at the third instance of the notes. As always, there may be typos,
and it is crucial that you analyse what is written and not just accept things without
thinking.
Thanks to M. Esipova, M.E. Gusak, J. Vendryes, Y. Wu and D. Yang for catching
some of the remaining typos/errors.
These notes are a work in progress, and - this being the “first edition” - they
are replete with typos. As of April 10, 2018, I have had the opportunity to look
over Chapters 1 - 4. That is not to say that they are mistake-free. It is, instead, an
admission that I simply haven’t had the chance to look over the remaining Chapters.
A student should approach these notes with the same caution he or she would
approach buzz saws; they can be very useful, but you should be thinking the whole
time you have them in your hands. Enjoy.
ii
From the moment I picked your book up until I laid it down I was
convulsed with laughter. Someday I intend reading it.
Groucho Marx
Reading this book is like waiting for the first shoe to drop.
Ralph Novak
Thank you for sending me a copy of your book. I’ll waste no time
reading it.
Moses Hadas
Sometimes you just have to stop writing. Even before you begin.
Stanislaw J. Lec
That’s not writing, that’s typing.
Truman Capote
Contents
1. Riemann integration 1
2. Lebesgue outer measure 16
3. Lebesgue measure 27
4. Lebesgue measurable functions 42
5. Lebesgue integration 57
6. Lp Spaces 83
7. Hilbert spaces 109
8. Fourier analysis - an introduction 123
9. Convolution 134
10. The Dirichlet kernel 159
11. The Féjer kernel 171
12. Which sequences are sequences of Fourier coefficients? 191
Bibliography 197
Index 199
iii
1. RIEMANN INTEGRATION 1
1. Riemann integration
What the world needs is more geniuses with humility, there are so
few of us left.
Oscar Levant
1.1. In first year, we study the Riemann integral for functions defined on inter-
vals, taking values in R. The same circle of ideas can be greatly extended, as we
shall now see.
Note. Much of the theory contained in these notes applies equally well to
the setting of real- or complex-valued functions and vector spaces. When writing a
statement which remains valid in either context, we shall simply write K to denote
the base field. In other words, we shall always have K ∈ {R, C}.
1.3. Remarks.
(a) Typically we shall denote seminorms by ν or µ. When the seminorm is
known to be a norm, we shall more often write ∥ ⋅ ∥.
(b) Let ν be a seminorm on a vector space X. Let z ∈ X denote the zero vector.
It follows from condition (ii) above that
1.5. The norm ∥ ⋅ ∥ on a nls X gives rise to a metric via the formula:
d∶ X×X → R
(x, y) ↦ ∥x − y∥.
We refer to this as the metric induced by the norm. (That this is indeed a
metric is left as an exercise to the reader.) When referring to metric properties of
X, it is understood that we are referring to the metric induced by the norm, unless
it is explicitly stated otherwise.
1.6. Definition. A normed linear space (X, ∥ ⋅ ∥) is said to be a Banach space
if (X, d) is a complete metric space, where d is the metric on X induced by the norm.
1.7. Examples.
(a) The motivating example is X = K itself, where the norm is given by the
absolute value function. Since (K, ∣ ⋅ ∣) is complete, it is a Banach space.
Of course, C is a one-dimensional Banach space over C, and a two-
dimensional Banach space over R.
(b) Let N ≥ 1 be an integer. For x = (xn )N N
n=1 ∈ K , we define three functions:
● ∥x∥1 ∶= ∣x1 ∣ + ∣x2 ∣ + ⋯ + ∣xN ∣;
● ∥x∥∞ ∶= max(∣x1 ∣, ∣x2 ∣, . . . , ∣xN ∣); and
1
● ∥x∥2 ∶= (∑N 2 2
n=1 ∣xn ∣ ) .
It is a routine exercise that ∥ ⋅ ∥1 and ∥ ⋅ ∥∞ define norms on KN , and that
KN becomes a Banach space when equipped with either of these norms.
It is a slightly more interesting exercise (left to the reader) to show that
N
(K , ∥ ⋅ ∥2 ) is also a Banach space. The standard proof that ∥ ⋅ ∥2 satisfies
1. RIEMANN INTEGRATION 3
(d) We saw in Example 1.4 that C([0, 1], K) is a nls when equipped with the
norm
∥f ∥sup = sup ∣f (x)∣ = max ∣f (x)∣.
x∈[0,1] x∈[0,1]
Then (C([0, 1], K), ∥ ⋅ ∥1 ) is a nls, but it is not a Banach space. The details
are yet again left to the reader.
(f) Let (X, ∥ ⋅ ∥X ) and (Y, ∥ ⋅ ∥Y ) be normed linear spaces over the field K. Let
T ∶ X → Y be a K-linear map. Consider
Set
B(X, Y) ∶= {T ∶ X → Y ∣ T is linear and ∥T ∥ < ∞}.
As we shall see in the assignments, B(X, Y) is a vector space over K, and
∥ ⋅ ∥ ∶ B(X, Y) → R defines a norm on B(X, Y), referred to as the operator
norm on B(X, Y). One can show that (B(X, Y), ∥ ⋅ ∥) is complete if and
only if (Y, ∥ ⋅ ∥Y ) is complete.
4 L.W. Marcoux Introduction to Lebesgue measure
1.9. Remarks.
(a) When (X, ∥ ⋅ ∥) = (R, ∣ ⋅ ∣), then this is the usual Riemann sum that one
studies in first-year calculus. In particular, in the case where 0 ≤ f (x) for
all x ∈ [a, b], we see that S(f, P, P ∗ ) estimates the area under the curve
y = f (x), x ∈ [a, b].
(b) Observe that in general,
N
1
S(f, P, P ∗ ) = ∑ λk f (p∗k ),
b−a k=1
pk − pk−1
where λk ∶= , 1 ≤ k ≤ N . Clearly λk ≥ 0 for all k, while ∑N
k=1 λk = 1.
b−a
1
Thus S(f, P, P ∗ ) is a convex combination of the f (p∗k )’s, and as
b−a
such, “averages” f over [a, b].
1.10. Example. Let X = C([−π, π], C), equipped with the supremum norm
∥ ⋅ ∥sup . Let 1 ≤ n ∈ N be a fixed integer and consider the function f ∶ [0, 1] → X given
by
[f (x)](θ) = e2πx sin (nθ) + cos x cos (nθ), θ ∈ [−π, π].
1 1
If P = {0, 10 , 2 , 1}, and if P ∗ = { 50
1 1 4
, 3 , 5 }, then
1
S(f, P, P ∗ ) = (eπ/25 sin (nθ) + cos(1/50) cos (nθ))( − 0)
10
1 1
+ (e2π/3 sin (nθ) + cos(1/3) cos (nθ))( − )
2 10
1
+ (e8π/5 sin (nθ) + cos(4/5) cos (nθ))(1 − ).
2
1. RIEMANN INTEGRATION 5
1.11. Definition. Let a < b be real numbers and (X, ∥ ⋅ ∥) be a Banach space.
We say that a function f ∶ [a, b] → X is Riemann integrable if there exists a vector
x0 ∈ X such that for any ε > 0, there exists a partition P ∈ P[a, b] with the property
that if Q is any refinement of P and Q∗ is any choice of test values for Q, then
∥x0 − S(f, Q, Q∗ )∥ < ε.
As is the case with the usual version of Riemann integration, the usefulness of
the Cauchy Criterion below is that it allows us to verify that a given function is
Riemann integrable without first having to know what its integral is.
Since f is continuous on the compact set [a, b], and since X is a metric space,
we see that f is uniformly continuous on [a, b]. Let ε > 0 and choose δ > 0 such that
ε
x, y ∈ [a, b] and ∣x − y∣ < δ implies that ∥f (x) − f (y)∥ < .
2(b − a)
Let R = {a = r0 < r1 < r2 < ⋯ < rN = b} ∈ P[a, b] be a partition of [a, b] such that
rj − rj−1 < δ for all 1 ≤ j ≤ N , and let R∗ = {rj∗ }N
j=1 be a set of test values for R.
Suppose that P = {pi }i=0 is any refinement of R, and that P ∗ = {p∗i }M
M
i=1 is a set
of test values for P . Then we can find a sequence 0 = k0 < k1 < k2 < ⋯ < kN ∶= M
such that
pkj = rj , 1 ≤ j ≤ N.
In other words,
P = {a = r0 = p0 = pk0 < p1 < ⋯ < pk1 = r1 < pk1 +1 < ⋯ < pk2 = r2 < ⋯ < pkN = rN = b}.
Now S(f, R, R∗ ) = ∑N ∗ ∗ M ∗
j=1 f (rj )(rj − rj−1 ). while S(f, P, P ) = ∑i=1 f (pi )(pi − pi−1 ).
But then
N
S(f, R, R∗ ) = ∑ f (rj∗ )(rj − rj−1 )
j=1
N kj
∗
= ∑ f (rj ) ∑ (pi − pi−1 )
j=1 i=kj−1 +1
N kj
=∑ ∑ f (rj∗ )(pi − pi−1 ),
j=1 i=kj−1 +1
while
N kj
S(f, P, P ∗ ) = ∑ ∑ f (p∗i )(pi − pi−1 ).
j=1 i=kj−1 +1
But for kj−1 + 1 ≤ i ≤ kj , we have that rj−1 ≤ rj∗ , p∗i ≤ rj , and therefore ∣rj∗ − p∗i ∣ ≤
rj − rj−1 < δ. From our estimate above, we see that
N kj
∥S(f, R, R∗ ) − S(f, P, P ∗ )∥ = ∥ ∑ ∗ ∗
∑ (f (rj ) − f (pi ))(pi − pi−1 )∥
j=1 i=kj−1 +1
N kj
≤∑ ∑ ∥f (rj∗ ) − f (p∗i )∥(pi − pi−1 )
j=1 i=kj−1 +1
N kj
ε
<∑ ∑ (pi − pi−1 )
j=1 i=kj−1 +1 2(b − a)
N j k
ε
= ∑ ∑ (pi − pi−1 )
2(b − a) j=1 i=kj−1 +1
8 L.W. Marcoux Introduction to Lebesgue measure
M
ε
∑(pi − pi−1 )
=
2(b − a) i=1
ε
= (pM − p0 )
2(b − a)
ε
= .
2
As such, if Q is any other refinement of R and if Q∗ is any set of test values for
Q, then
ε
∥S(f, Q, Q∗ ) − S(f, R, R∗ )∥ < ,
2
whence
∥S(f, Q, Q∗ ) − S(f, P, P ∗ )∥ ≤ ∥S(f, Q, Q∗ ) − S(f, R, R∗ )∥+
∥S(f, R, R∗ ) − S(f, P, P ∗ )∥
ε ε
< +
2 2
= ε.
By the Cauchy Criterion 1.13, f is Riemann integrable.
◻
We provide a second proof of the above result in the Appendix to this section.
We now turn our attention to real-valued functions, with the intention of show-
ing what one of the failings of the Riemann integral is, and thereby (hopefully)
motivating the study of the Lebesgue integral in the forthcoming chapters.
Indeed, observe that E and [0, 1] ∖ E are each dense in [0, 1]. Let
P = {0 = p0 < p1 < p2 < ⋯ < pN = 1}
be any partition of [0, 1]. For 1 ≤ n ≤ N , choose p∗n ∈ Q ∩ [pn−1 , pn ], and choose
qn∗ ∈ Qc ∩ [pn−1 , pn ], so that P ∗ ∶= {p∗1 , p∗2 , . . . , p∗N } and Q∗ ∶= {q1∗ , q2∗ , . . . , qN
∗
} are
1. RIEMANN INTEGRATION 9
Appendix to Section 1.
1.18. There are more notions of integration in a Banach space than just Rie-
mann integration. Indeed, in general, there exist more notions of integration than
one can shake a stick at, even if one has strong arms, a very light stick, ample
dexterity and a solid and enviable history of stick-shaking.
This notion, however, will prove sufficient for our purposes. In Theorem 1.14,
we showed that every continuous function from a closed, bounded interval in R to a
Banach space is Riemann integrable over that closed interval. A minor modification
will show that every piecewise continuous, bounded function from a closed interval
to a Banach space is Riemann integrable in the sense of Definition 1.11.
1.19. Culture. The notion of integrating in a Banach space is not simply some
arcane and useless generalization of integration of real- or complex-valued functions.
Let 1 ∈ n be an integer, and let H ∶= Cn , equipped with the Euclidean norm ∥ ⋅ ∥2 .
Consider (B(Cn ), ∥ ⋅ ∥), where ∥ ⋅ ∥ denotes the operator norm on B(H) = B(Cn ).
We leave it as an exercise for the reader to show that every linear map T ∶ Cn →
C satisfies ∥T ∥ < ∞, and thus we may identify B(Cn ) with Mn (C), the space of
n
n × n matrices with entries in C with respect to a fixed, orthonormal basis {ek }nk=1
for Cn .
Let d1 , d2 , . . . , dn ∈ C be n distinct points, and let D ∈ B(Cn ) be the unique
linear operator whose corresponding matrix is the diagonal matrix
⎡d1 0 ... 0 ⎤⎥
⎢
⎢ d 0 . . . 0 ⎥⎥
⎢ 2
⎢ ⎥
⎢ ⋱ ⎥.
⎢ ⎥
⎢
⎢ d n−1 0 ⎥
⎥
⎢ d n⎦
⎥
⎣
Suppose that ∅ ≠ ∆ ⊆ {1, 2, . . . , n}.
Let Γ ⊆ C be a piecewise smooth curve in C for which
⎧
⎪1 if k ∈ ∆
⎪
ind(Γ, dk ) = ⎨
⎪
⎪ 0 otherwise.
⎩
It can be shown that
1 −1
P ∶= ∫ (sI − D) ds
2πi Γ
is the orthogonal projection P = diag(p1 , p2 , . . . , pn ), where pk = 1 if k ∈ ∆ and pk = 0
otherwise.
The astute reader will have observed the striking similarity of the integral above
to Cauchy’s Integral Formula. This is not a coincidence. Such integrals are studied
in much greater generality in the theory of Banach algebras.
12 L.W. Marcoux Introduction to Lebesgue measure
1.20. Theorem. (Theorem 1.14 revisited). Let (X, ∥ ⋅ ∥) be a Banach space and
a < b ∈ R. If f ∶ [a, b] → X is continuous, then f is Riemann integrable over [a, b].
Proof. We shall in fact show that if PN ∈ P([a, b]) is a regular partition of [a, b]
b−a
into 2N subintervals of equal length N , and if PN∗ = PN ∖ {a} is the set of test
2
values for PN consisting of “right-hand endpoints” of the subintervals of PN , then
the sequence (S(f, PN , PN∗ ))∞ N =1 converges in X to
b
∫ f (s) ds.
a
We begin by showing that the sequence (S(f, PN , PN∗ ))∞
N =1 is Cauchy, and there-
fore converges to something in X, which we temporarily designate by y.
Since f is continuous on the compact set [a, b], and since X is a metric space,
we see that f is uniformly continuous on [a, b]. Let ε > 0 and choose δ > 0 such that
ε
x, y ∈ [a, b] and ∣x − y∣ < δ implies that ∥f (x) − f (y)∥ < .
b−a
b−a
For each N ≥ 1, let PN be as above, and choose M ≥ 1 such that M < δ. If
2
K ≥ L ≥ M , then PM ⊆ PL ⊆ PK ; indeed, writing
PL = {a = p0 < p1 < ⋯ < p2L = b},
and
PK = {a = q0 < q1 < ⋯ < q2K = b},
and setting p∗j
= pj , 1 ≤ j ≤ 2L and qs∗ = qs , 1 ≤ s ≤ 2K , we see that pj = qj 2K−L for all
0 ≤ j ≤ 2L , and that for 1 ≤ j ≤ 2L ,
ε
∥f (p∗j ) − f (qs∗ )∥ < , (j − 1) 2K−L < s ≤ j 2K−L .
b−a
Thus
X
X
X 2L j 2K−L X
X
X
∥S(f, PL , PL∗ ) − S(f, PK , PK ∗ X X
)∥ = X
X
X
X ∑ ∑ (f (p j ) − f (q s ))(q s − q s−1 X
) X
X
X
X
X X
X
Xj=1 s=(j−1) 2K−L +1
X X
X
2L j 2K−L
≤∑ ∑ ∥f (pj ) − f (qs )∥(qs − qs−1 )
j=1 s=(j−1) 2K−L +1
2L j 2K−L
ε
≤∑ ∑ (qs − qs−1 )
j=1 s=(j−1) 2K−L +1
b−a
K
ε 2
= ∑ qs − qs−1
b − a s=1
ε
= (b − a) = ε.
b−a
1. RIEMANN INTEGRATION 13
So if we set yN = S(f, PN , PN∗ ), N ≥ 1, then the above argument shows that (yN )∞
N =1
is a Cauchy sequence in X, and as such, admits a limit y ∈ X.
b
There remains to show that y = ∫a f (s) ds. The proof is almost identical to that
above.
b−a
With ε > 0, choose T ≥ 1 such that T < δ and also such that
2
∥y − S(f, PT , PT∗ )∥ < ε.
Let R = {a = r0 < r1 < ⋯ < rJ = b} be any refinement of PT = {a = p0 < p1 < ⋯ < p2T =
b}. Thus there exists a sequence
0 = j0 < j1 < ⋯ < j2T = J
such that
rjk = pk , 0 ≤ k ≤ 2T .
Let R∗ be any set of test values for R. If jk−1 + 1 ≤ s ≤ jk , then ∣p∗k − rs∗ ∣ ≤ ∣pk − pk−1 ∣ =
b−a
< δ, 1 ≤ k ≤ 2T and so
2T
2T jk
∥S(f, PT , PT∗ ) − S(f, R, R∗ )∥ ≤ ∑ ∑ ∥f (p∗k ) − f (rs∗ )∥(rs − rs−1 )
k=1 s=jk−1 +1
T
jk
ε 2
< ∑ ∑ (rs − rs−1 )
b − a k=1 s=jk−1 +1
ε
= (b − a) = ε.
b−a
Thus
∥y − S(f, R, R∗ )∥ ≤ ∥y − S(f, PT , PT∗ )∥ + ∥S(f, PT , PT∗ ) − S(f, R, R∗ )∥ < ε + ε = 2ε.
This clearly shows that
b
y=∫ f (s) ds.
a
◻
14 L.W. Marcoux Introduction to Lebesgue measure
Exercise 1.1.
Let ∅ ≠ X be a compact, Hausdorff space. Prove that for each ∅ ≠ Ω ⊆ X, the
function
νΩ ∶ C(X, K) → R
f ↦ supx∈Ω ∣f (x)∣
defines a seminorm on C(X, K), and that it is a norm if and only if Ω is dense in X.
Exercise 1.2.
(a) Let ν be a seminorm on a vector space Y over K. Prove that
∣ν(x) − ν(y)∣ ≤ ν(x − y) for all x, y ∈ Y.
(b) Let (X, ∥ ⋅ ∥) be a nls. Prove that the map
d∶ X×X → R
(x, y) ↦ ∥x − y∥
defines a metric on X.
Exercise 1.3.
Let N ≥ 1 be an integer. Define three functions from KN to R as follows: for
x = (xn )N N
n=1 ∈ K , we set
(a) ∥x∥1 ∶= ∣x1 ∣ + ∣x2 ∣ + ⋯ + ∣xN ∣;
(b) ∥x∥∞ ∶= max(∣x1 ∣, ∣x2 ∣, . . . , ∣xN ∣); and
1
(c) ∥x∥2 ∶= (∑N 2 2
n=1 ∣xn ∣ ) .
Prove that each of these functions defines a norm on KN .
Exercise 1.4.
Let N ≥ 1 be an integer, and let x ∈ KN . Prove that
lim ∥x∥p = ∥x∥∞ .
p→∞
Exercise 1.5.
(a) Prove that the nls (C([0, 1], K), ∥⋅∥sup ) is complete, and that it is therefore
a Banach space.
1
(b) Recall that ∥f ∥1 ∶= ∫0 ∣f (x)∣dx defines a norm on C([0, 1], K). Prove that
(C([0, 1], K), ∥ ⋅ ∥1 ) is not complete.
Exercise 1.6.
Let X be a Banach space, a < b be real numbers and f ∶ [a, b] → X be a function.
Suppose that f is Riemann integrable over [a, b].
1. RIEMANN INTEGRATION 15
Prove that for all ε > 0 there exists a partition R ∈ P[a, b] with the property
that if P and Q are refinements of R, and if P ∗ and Q∗ are test values for P and Q
respectively, then
∥S(f, P, P ∗ ) − S(f, Q, Q∗ )∥ < ε.
Exercise 1.7.
Let X be a Banach space, a < b be real numbers, and g ∶ [a, b] → X be a bounded,
piecewise-continuous function. Show that g is Riemann integrable over [a, b].
Exercise 1.8.
1
Prove the claim from Remark 1.16, namely: ∫0 fn (s) ds = 0 for all n ≥ 1,
∞
where Q ∩ [0, 1] = {qn }n=1 , and where fn is the characteristic function of En ∶=
{q1 , q2 , . . . , qn }, n ≥ 1.
If you want to know what God thinks of money, just look at the people
he gave it to.
Dorothy Parker
2.1. Our goal in this section is to define a “measure of length” for as many
subsets of R as possible. We would like our new notion to agree with our intuition
in the cases we know; for example, it seems reasonable to ask that our generalized
notion of “length” of a finite interval (a, b) should be (b − a) when a < b in R. We
shall therefore use this as our starting point, and we shall use this intuition to extend
our notion of “length” to a greater variety of sets by approximation.
In order to help make the text more readable, and since these are the only covers
of sets we will consider in these notes, we shall abbreviate the expression “cover of E
by open intervals” to “cover of E”. For any set X, we denote by P(X) = {Y ∶ Y ⊆ X}
the power set of X.
(d) if E, F1 , F2 , F3 , . . . ⊆ X and if E ⊆ ∪∞
n=1 Fn , then
∞
µE ≤ ∑ µFn .
n=1
Condition (c) (or equivalently (d)) is generally referred to as the countable sub-
additivity or σ-subadditivity of µ.
2.4. Proposition. The function m∗ defined in Definition 2.2 is an outer mea-
sure on R.
Proof.
(a) Let E = ∅. With In = ∅, n ≥ 1, it is clear that {In }∞
n=1 is a cover of E, and
so
∞ ∞
0 ≤ m∗ ∅ ≤ ∑ ℓ(In ) = ∑ 0 = 0.
n=1 n=1
Thus m∗ ∅ = 0.
(b) Let E ⊆ F ⊆ R.
If {In }∞
n=1 is a cover of F , then it is also a cover of E. It follows
immediately from the definition that
m∗ E ≤ m∗ F.
(c) Suppose that {En }∞ n=1 is a countable collection of subsets of R. We wish
∞
to prove that m∗ (∪∞ ∗
n=1 En ) ≤ ∑n=1 m En .
∞
If ∑n=1 m∗ En = ∞, then we have nothing to prove. Thus we consider
the case where ∑∞ ∗ ∞
n=1 m En < ∞. Set E ∶= ∪n=1 En .
(n)
Let ε > 0 and for each n ≥ 1, choose a cover {Ik }∞ k=1 of En such that
∞
(n) ∗ ε
∑ ℓ(Ik ) < m En + .
k=1 2n
(n)
Then {Ik }∞k,n=1 is a cover of E, and so
∞ ∞
(n)
m∗ E ≤ ∑ ∑ ℓ(Ik )
n=1 k=1
∞
∗ ε
≤ ∑ (m En + )
n=1 2n
∞
= ( ∑ m∗ En ) + ε.
n=1
(Here we have used the fact that if 0 ≤ an ∈ R, n ≥ 1, and if σ ∶ N → N is
any permutation, then ∑∞ ∞
n=1 an = ∑n=1 aσ(n) .)
Since ε > 0 was arbitrary,
∞
m∗ E ≤ ∑ m∗ En .
n=1
18 L.W. Marcoux Introduction to Lebesgue measure
We have defined outer measure m∗ E for any subset E of R, and we have done this
based upon an intuitive notion of what the length of an open interval (a, b) should
be, namely b − a. At first glance, it seems obvious that m∗ (a, b) = ℓ(b − a) = b − a.
But upon reflection, we see that this is not how m∗ (a, b) is defined. This leaves
us with an interesting problem: how does our notion of measure m∗ (a, b) of an
interval compare with this notion of length? On the one hand, it is clear that
m∗ (a, b) ≤ ℓ(a, b) = b − a, since I1 ∶= (a, b) and In = ∅, n ≥ 2 yields a cover of (a, b).
On the other hand, the notion of outer measure of (a, b) requires us to consider all
covers of (a, b) by intervals, not only the obvious cover by the interval (a, b) itself.
We now turn to this problem. It will prove useful to first consider the outer measure
of closed, bounded intervals [a, b], as these are compact. Because of this, we will be
able to replace general covers of [a, b] (by open intervals) with finite covers of [a, b]
(by open intervals).
Now,
∞ N
∑ ℓ(In ) ≥ ∑ ℓ(In )
n=1 n=1
M
≥ ∑ ℓ((ank , bnk ))
k=1
= (bn1 − an1 ) + (bn2 − an2 ) + ⋯ + (bnM − anM )
= bnM + (bnM −1 − anM ) + (bnM −2 − anM −1 ) + ⋯ + (bn1 − an2 ) − an1
> bnM − an1
> b − a.
Since {In }∞
n=1 was an arbitrary cover of [a, b], it follows that
m∗ [a, b] ≥ b − a.
Combining this with the reverse inequality above, we conclude that
m∗ [a, b] = b − a.
20 L.W. Marcoux Introduction to Lebesgue measure
2.14. In light of the above Theorem, we are faced with a difficult choice. We
may either
(a) content ourselves with the notion of Lebesgue outer measure m∗ for all
subsets E of R, which agrees with our intuitive notion of length for in-
tervals, but in so doing we must sacrifice the highly-desirable property of
σ-additivity; or
(b) restrict the domain of our function m∗ to a more “tractable” family of
subsets of R, where we might in fact be able to prove that the restriction
of m∗ to this family is σ-additive.
The standard approach is the second, and it is the one we shall adopt. The
next section is devoted to describing our tractable family of sets, and to proving the
σ-additivity of the restriction of m∗ to this collection.
24 L.W. Marcoux Introduction to Lebesgue measure
Appendix to Section 2.
2.15. For those with a historical bent (and there are exciting new treatments
for that now), the “construction” of the set V defined in Step 2 of Theorem 2.13 is
due to the Italian mathematician Giuseppe Vitali [6]. Strictly speaking, this isn’t
an explicit construction. The Axiom of Choice was invoked to prove the existence
of such a set V.
In the next Chapter, we shall give a name to the “tractable” family of sets
described in Paragraph 2.14. Indeed, we shall refer to elements of this family as
“Lebesgue measurable sets”. The set V under discussion is an example of a non-
measurable set – see Exercise 2. More generally, however, a Vitali set is a set
B ⊆ [0, 1] for which x ≠ y ∈ B implies that that x − y ∈ R ∖ Q; i.e. B contains at most
one representative from each coset of Q in R/Q.
Suppose that such a measure µ exists. Let {En }∞ n=1 be a countable family of
pairwise disjoint sets. Since µ is an outer measure on R, it is countably subadditive,
and so
∞
µ(⊍∞
n=1 En ) ≤ ∑ µEn .
n=1
By monotonicity and finite-additivity of µ, we find that for all N ≥ 1,
N ∞
N ∞
∑ µEn = µ(⊍n=1 En ) ≤ µ(⊍n=1 En ) ≤ ∑ µEn .
n=1 n=1
2. LEBESGUE OUTER MEASURE 25
Exercise 2.1.
Let E ⊆ R be a finite set. Prove that m∗ E = 0.
Exercise 2.2.
Prove the remaining cases from Proposition 2.7 and Corollary 2.8. That is, prove
that if a < b ∈ R, then
m∗ [a, b) = m∗ (a, b) = b − a,
while
m∗ (−∞, b] = m∗ (a, ∞) = m∗ [a, ∞) = m∗ R = ∞.
Exercise 2.3.
Prove that
m∗ ([0, 2] ∪ [7, 11]) = 6.
Exercise 2.4.
Let ∼ be the relation on R defined by x ∼ y if y − x ∈ Q. Prove that ∼ is an
equivalence relation on R.
Exercise 2.5.
Let m∗ denote Lebesgue outer measure on R, and let V ⊆ [0, 1] denote Vitali’s
set from the proof of Theorem 2.13.
Prove that
m∗ V > 0.
3. LEBESGUE MEASURE 27
3. Lebesgue measure
3.1. As mentioned at the end of the previous Chapter, our strategy will be to
restrict the domain of Lebesgue outer measure to a smaller collection of sets, where
Lebesgue outer measure will be σ-additive. We shall refer to this collection as the
collection of Lebesgue measurable sets.
We begin with Carathéodory’s definition of a Lebesgue measurable set, since it
is the most practical definition to use. Later we shall see that Lebesgue measurable
sets are “almost” countable intersections of open sets, in a way which we shall make
precise.
3.2. Definition. A set E is R is said to be Lebesgue measurable if, for all
X ⊆ R,
m∗ X = m∗ (X ∩ E) + m∗ (X ∖ E).
We denote by M(R) the collection of all Lebesgue measurable sets.
3.3. Remarks. Since our attention in this course is almost exclusively focused
upon Lebesgue measure, we shall allow ourselves to drop the adjective “Lebesgue”
and refer only to “measurable sets”.
Informally speaking, we see that a set E ⊆ R is measurable provided that it is
a “universal slicer ”, in the sense that it “slices” every other set X into two disjoint
sets, namely X ∩ E and X ∖ E, where Lebesgue outer measure is additive!
We also note that the inequality
m∗ X ≤ m∗ (X ∩ E) + m∗ (X ∖ E)
is free from the σ-subadditivity of Lebesgue outer measure. In checking to see
whether a given set is measurable or not, it therefore suffices to verify that the
reverse inequality holds for all sets X ⊆ R.
Before we proceed to the examples, which shall obtain a result which allows us
to show that the set M(R) of Lebesgue measurable sets itself has an interesting
structure.
3.4. Definition. Let Y be a non-empty set. A collection Ω ⊆ P(Y ) is said to
be an algebra of subsets of Y if
(a) Y ∈ Ω;
(b) E ∈ Ω implies that E c ∶= Y ∖ E ∈ Ω; and
(c) if N ≥ 1 and E1 , E2 , . . . , EN ∈ Ω, then
E ∶= ∪N
n=1 En ∈ Ω.
28 L.W. Marcoux Introduction to Lebesgue measure
m∗ (X ∩ (⊍M +1 ∗ M +1
n=1 Fn )) = m ((X ∩ (⊍n=1 Fn )) ∖ FM +1 )+
m∗ ((X ∩ (⊍M +1
n=1 Fn )) ∩ FM +1 )
= m∗ (X ∩ (⊍M ∗
n=1 Fn )) + m (X ∩ FM +1 )
M
= ∑ m∗ (X ∩ Fn ) + m∗ (X ∩ FM +1 )
n=1
by the induction hypothesis
M +1
= ∑ m∗ (X ∩ Fn ).
n=1
This completes the induction step and therefore proves our claim.
Step 4. Finally, observe that E = ∪∞ ∞ ∞
n=1 En = ∪n=1 Hn = ⊍n=1 Fn . We shall use
this to prove that E ∈ M(R).
Let X ⊆ R. For all N ≥ 1, HN ∈ M(R) and so
m∗ X = m∗ (X ∩ HN ) + m∗ (X ∖ HN )
= m∗ (X ∩ (⊍N ∗
n=1 Fn )) + m (X ∖ HN )
≥ m∗ (X ∩ (⊍N ∗
n=1 Fn )) + m (X ∖ E) as (X ∖ E) ⊆ (X ∖ HN )
N
= ∑ m∗ (X ∩ Fn ) + m∗ (X ∖ E) by Step 3.
n=1
It’s high time that we produce examples of Lebesgue measurable sets. Thanks
to the previous Theorem, given a subset S ⊆ M(R), the entire σ-algebra generated
by S (namely the smallest σ-algebra of subsets of R which contains S – why should
this exist?) is also contained in M(R).
3.6. Proposition.
(a) If E ⊆ R and m∗ E = 0, then E ∈ M(R).
(b) For all b ∈ R, E ∶= (−∞, b) ∈ M(R).
(c) Every open and every closed set is Lebesgue measurable.
Proof.
(a) Let E ⊆ R be a set with m∗ E = 0, and let X ⊆ R. By monotonicity of outer
measure,
m∗ (X ∩ E) ≤ m∗ E = 0,
and
m∗ (X ∩ E c ) ≤ m∗ X.
Thus
m∗ X = 0 + m∗ X ≥ m∗ (X ∩ E) + m∗ (X ∖ E).
As we have seen, this is the statement that E ∈ M(R).
(b) Fix b ∈ R and set E = (−∞, b). Let X ⊆ R be arbitrary. We must show that
m∗ X ≥ m∗ (X ∩ E) + m∗ (X ∖ E).
If m∗ X = ∞, then there is nothing to prove, and so we assume that
m X < ∞. Let ε > 0 and let {In }∞
∗
n=1 be a cover of X by open intervals such
that
∞
∗
∑ ℓ(In ) < m X + ε < ∞.
n=1
It follows that each interval In has finite length, and so we may write
In = (an , bn ), n ≥ 1. (As always, there is no harm in assuming that each
In ≠ ∅, otherwise we simply remove In from the cover.)
Set Jn = In ∩ E = (an , bn ) ∩ (−∞, b). Clearly each Jn , n ≥ 1 is an open
interval, possibly empty.
3. LEBESGUE MEASURE 31
depending upon the values of an and bn . But then we can find cn < dn in
R such that
Kn ⊆ Ln ∶= (cn , dn )
and ℓ(Ln ) − m∗ Kn < ε
2n ,
n ≥ 1.
In particular, for each n ≥ 1, In ⊆ Jn ∪ Ln and
ε
(ℓ(Jn ) + ℓ(Ln )) − ℓ(In ) < .
2n
Now
∞
m∗ X > ( ∑ ℓ(In )) − ε
n=1
∞
ε
> ( ∑ (ℓ(Jn ) + ℓ(Ln ) − )) − ε
n=1 2n
∞ ∞ ∞
ε
= ∑ ℓ(Jn ) + ∑ ℓ(Ln ) − ∑ n
− ε.
n=1 n=1 n=1 2
But X ∩ E ⊆ ∪∞
n=1 Jn and X ∖ E ⊆ ∪∞
n=1 Ln , and so
m∗ X > m∗ (X ∩ E) + m∗ (X ∖ E) − 2ε.
m∗ X ≥ m∗ (X ∩ E) + m∗ (X ∖ E),
(b, ∞) = ∪∞
n=1 En ∈ M(R) for all b ∈ R.
Now that we know that M(R) ≠ ∅, the following definition makes sense.
32 L.W. Marcoux Introduction to Lebesgue measure
Let us recall that our strategy was to try to restrict the domain of m∗ sufficiently
to allow the restriction of m∗ to this smaller collection of sets to be σ-additive. The
next result shows that M(R) may serve as such a domain.
3.12. Remarks. Note that since Bor (R) is a σ-algebra which contains all
open subsets of R, it also contains all closed subsets of R. In fact, we could have
defined Bor (R) to be the σ-algebra of subsets of R generated by the collection
F ∶= {F ⊆ R ∶ F is closed} of closed subsets of R, and concluded that it would have
contained G.
Given a family A ⊆ P(R) with ∅, R ∈ A, set
Aσ ∶= {∪∞
n=1 An ∶ An ∈ A, n ≥ 1}
Aδ ∶= {∩∞
n=1 An ∶ An ∈ A, n ≥ 1}.
We refer to elements of Aσ as A-sigma sets and elements of Aδ as A-delta sets.
Observe that Gδ and Fσ are both subsets of Bor (R).
3.13. Admittedly, our definition of a Lebesgue measurable set is not the most
intuitive definition, and Carathéodory’s definition is quite different from Lebesgue’s
original definition, which we shall now investigate.
Proof.
(a) implies (b).
Case 1. Suppose that mE < ∞.
Let ε > 0 and choose a cover {In }∞
n=1 of E by open intervals such that
∞
∑ ℓ(In ) < mE + ε.
n=1
Set G = ∪∞
n=1 In so that G is open and E ⊆ G. Then
∞ ∞
mG ≤ ∑ mIn = ∑ ℓ(In ) < mE + ε.
n=1 n=1
Also,
G = (G ∩ E) ∪ (G ∖ E) = E ⊍ (G ∖ E),
and so
mE + m(G ∖ E) = mG < mE + ε,
whence
m∗ (G ∖ E) = m(G ∖ E) < ε.
34 L.W. Marcoux Introduction to Lebesgue measure
3.15. Example. The Cantor middle thirds set. Recall from Corollary 2.5
that if E ⊆ R is countable, then m∗ E = 0. By Proposition 3.6, it follows that
E ∈ M(R), and thus mE = m∗ E = 0. In other words, every countable set is Lebesgue
measurable with Lebesgue measure zero.
We shall now construct an uncountable set C – in fact one whose cardinality is
c, the cardinality of the real line R – whose measure mC is equal to zero. Since
mR = ∞, we see that the Lebesgue measure of a set is not so much a reflection of
its cardinality, as much as a question of how the points in the set are distributed.
Having said that, when the set in question is countable, the above argument shows
that it is always “thinly distributed”, in this analogy.
The Cantor set is typically obtained as the intersection of a countable family
of sets, each iteratively constructed from the previous as follows:
We set C0 = [0, 1], and for n ≥ 1, we set Cn = 13 Cn−1 ∪ ( 32 + 13 Cn−1 ). Thus
3. LEBESGUE MEASURE 35
● C1 = [0, 13 ] ∪ [ 32 , 1];
● C2 = [0, 91 ] ∪ [ 92 , 13 ] ∪ [ 96 , 79 ] ∪ [ 89 , 1],
1 2 3 6 7 8 9
● C3 = [0, 27 ]∪[ 27 , 27 ]∪[ 27 , 27 ]∪[ 27 , 27 ]∪[ 18 19 20 21 24 25 26
27 , 27 ]∪[ 27 , 27 ]∪[ 27 , 27 ]∪[ 27 , 1];
● ⋮
The figure below shows each of the sets Cn , 0 ≤ n ≤ 7.
The Cantor middle thirds set is defined as the intersection of all of these
sets, i.e.
C ∶= ∩∞
n=0 Cn .
Alternatively, beginning with C0 = [0, 1], one can think of obtaining C1 from C0
by removing the (open) “middle third” interval ( 31 , 23 ), resulting in the two intervals
which comprise C1 above. To obtain C2 from C1 , one removes the (open) “middle
third” of each of the two intervals in C1 , and so on. This motivates the term middle
thirds in the above nomenclature.
It should be clear from the construction above that
(a) C0 ⊇ C1 ⊇ C2 ⊇ C3 ⊇ ⋯ ⊇ C. Furthermore, each set Cn is closed, n ≥ 0 and
mC0 = 1 < ∞. From our work in the assignments,
2n
m∗ C = lim mCn = lim = 0.
n→∞ n→∞ 3n
Now let us approach things from a different angle. Given x ∈ [0, 1], consider the
ternary expansion of x, namely
x = 0.x1 x2 x3 x4 . . .
where xk ∈ {0, 1, 2} for all k ≥ 1. As with decimal expansions, the expression above
is meant to express the fact that
∞
xk
x= ∑ k
.
k=1 3
36 L.W. Marcoux Introduction to Lebesgue measure
Appendix to Section 3.
3.16. With regards to Theorem 3.14 above, first note that the definition of
Lebesgue outer measure says that for all H ⊆ R,
∞
m∗ H = inf{ ∑ ℓ(In ) ∶ {In }∞
n=1 is a cover of H}.
n=1
In other words, instead asking that m∗ be additive with respect to the decom-
position X = E ⊍ (X ∖ E) for every set X that contains E, it suffices to ask that
this condition holds in the single, solitary case where X = A!
In particular, this shows that if E is bounded (i.e. there exists M > 0 such that
E ⊆ [−M, M ]), then E is Lebesgue measurable if and only if
m∗ (E) + m∗ ([−M, M ] ∖ E) = 2M.
Suddenly Carathéodory’s definition of a Lebesgue measurable set doesn’t seem
so bad!
Define
f∶ A → B
⎧
⎪
⎪κ(a) if a ∈ T ,
a ↦ ⎨ −1
⎪
⎪λ (a) if a ∈ A / T .
⎩
Observe that λ is a bijection between B / κ(T ) and A / T , and that κ is a
bijection between T and κ(T ), so that f is a bijection between A and B.
◻
40 L.W. Marcoux Introduction to Lebesgue measure
Exercise 3.1.
Let E ∈ M(R), so that E is a Lebesgue measurable set. Let κ ∈ R, and set
E + κ ∶= {x + κ ∶ x ∈ E}. Prove that E + κ ∈ M(R), and that mE = m(E + κ).
Thus Lebesgue measure m is translation-invariant.
Exercise 3.2.
Let V ⊆ [0, 1] be Vitali’s set from Theorem 2.13. Prove that V is not Lebesgue
measurable.
Exercise 3.3.
Let S ⊆ M(R).
Prove that there exists a σ-algebra N ⊆ M(R) of subsets which contains S and
with the property that if K is any σ-algebra of measurable sets which contains S,
then N ⊆ K. We say that N is the σ-algebra generated by S.
Exercise 3.4.
Let x ∈ [0, 1) and consider the ternary expansion of x given by
x = 0.x1 x2 x3 x4 . . . .
Show that the expansion is unique except when there exists N ≥ 1 such that
r
x= N for some 0 < r < 3N , where 3 ∤ r.
3
Exercise 3.5.
Prove that the cardinality ∣M(R)∣ of the collection of Lebesgue measurable sets
is equal to that of the collection P(R) ∖ M(R) of non-measurable sets.
Don’t ever wrestle with a pig. You’ll both get dirty, but the pig will
enjoy it.
Cale Yarborough
4.5. Proposition. Let E ∈ M(R), and suppose that (X, dX ) and (Y, dY ) are
metric spaces. Suppose furthermore that f ∶ E → X is measurable and that g ∶ X → Y
is continuous. Then g ○ f ∶ E → Y is measurable.
Proof. Let G ⊆ Y be open. Then g −1 (G) ⊆ X is open, as g is continuous. But f is
measurable, so
(g ○ f )−1 (G) = f −1 (g −1 (G)) ∈ M(E),
proving that g ○ f is measurable.
◻
h−1 (G) = ∪∞ −1 ∞ −1 −1
n=1 h (An × Bn ) = ∪n=1 f (An ) ∩ g (Bn ).
Thus the set L(E, K) forms a subalgebra of the algebra KE of all functions from E
into K, and the constant function φ(x) = 1 for all x ∈ E clearly serves as the identity
of this algebra.
◻
f = u ⋅ ∣f ∣.
By definition, f is measurable.
◻
4. LEBESGUE MEASURABLE FUNCTIONS 47
4.14. We are now going to examine a number of results that deal with pointwise
limits of sequences of measurable, real-valued functions. It will prove useful to
include the case where the limit at a given point exists as an extended real number;
that is, when the sequence diverges to ∞, or to −∞. While this is useful for treating
measure-theoretic and analytic properties of sequences of functions, there is a price
to pay: the extended real numbers have poor algebraic properties. In particular, we
can not add ∞ to −∞, and so the class of functions we shall examine will not form
a vector space.
Note. We remark that condition (a) above can be replaced with the condition that
f −1 (F ) ∈ M(E) for all closed sets F ⊆ R.
As was the case with real-valued measurable functions, in testing whether or not
a given extended real-valued function is measurable or not, it suffices to check that
the inverse images of certain intervals are measurable. In what follows, we write
(a, ∞] to mean (a, ∞) ∪ {∞} and [−∞, b) = (−∞, b) ∪ {−∞} for all a, b ∈ R.
4.17. Proposition. Let E ∈ M(R) and suppose that f ∶ E → R is a function.
The following statements are equivalent.
(a) f if Lebesgue measurable.
(b) For all a ∈ R, f −1 ((a, ∞]) ∈ M(E).
(c) For all b ∈ R, f −1 ([−∞, b)) ∈ M(E).
Proof. The proof of this Proposition is left as an exercise for the reader.
◻
4.18. Proposition. Let E ∈ M(R) and suppose that (fn )∞ n=1 is a sequence
of extended real-valued, measurable functions on E. The following (extended real-
valued) functions are also measurable.
(a) g1 ∶= supn≥1 fn ;
(b) g2 ∶= inf n≥1 fn ;
(c) g3 ∶= lim supn≥1 fn ; and
(d) g4 ∶= lim inf n≥1 fn .
Proof.
(a) By Proposition 4.17 above, it suffices to prove that g1−1 ((a, ∞]) ∈ M(E) for
all a ∈ R.
But
g1−1 ((a, ∞]) = ∪∞ −1
n=1 fn ((a, ∞]).
Since each fn−1 ((a, ∞]) ∈ M(E) (as each fn is measurable), we have that
g1−1 ((a, ∞]) ∈ M(E).
4. LEBESGUE MEASURABLE FUNCTIONS 49
Hence g1 is measurable.
(b) The proof of (b) is similar to that of (a). For each b ∈ R,
g2−1 ([−∞, b)) = ∪n≥1 fn−1 ([−∞, b)) ∈ M(E).
Thus g2 is measurable.
(c) For each N ≥ 1, hN ∶= supn≥N fn is measurable by (a) above. Clearly
h1 ≥ h2 ≥ h3 ≥ ⋯.
But then lim supn≥1 fn = limn→∞ hn = inf n≥1 hn , and this is measurable by
(b) above.
(d) The proof of this is similar to that of (c), and is left as an exercise.
◻
The next result is an immediate corollary of the above Proposition.
4.19. Corollary. Let E ∈ M(R) and suppose that (fn )∞ n=1 is a sequence of
real-valued, measurable functions on E such that f (x) = limn→∞ fn (x) exists as an
extended real number for each x ∈ E.
Then f ∈ L(E, R); i.e. f is measurable.
4.20. Definition. Let E ∈ M(R) and φ ∶ E → R be a function. We say that
φ is simple if ran φ is finite. Suppose that ran φ = {α1 < α2 < ⋯ < αN }, and set
En ∶= φ−1 ({αn }), 1 ≤ n ≤ N . We shall say that
N
φ = ∑ αn χEn
n=1
is the standard form of φ.
4.22. Example. The standard form is not the only way of expressing a simple
function as a linear combination of characteristic functions.
Consider φ ∶ R → R defined as φ = χQ + 9χ[2,6] . Then ran φ = {0, 1, 9, 10}.
Set
E1 = φ−1 ({0}) = R ∖ (Q ∪ [2, 6])
E2 = φ−1 ({1}) = Q ∖ [2, 6]
E3 = φ−1 ({9}) = [2, 6] ∖ Q
E4 = φ−1 ({10}) = Q ∩ [2, 6].
Then
φ = 0χE1 + 1χE2 + 9χE3 + 10χE4
is the standard from of φ.
4.24. Example.
(a) Let V = R3 , and let C = {(x, y, z) ∈ V ∶ 0 ≤ x, y, z}. Then C is a cone.
(b) Let V = C and let
π 2π
C = {w ∈ C ∶ w = reiθ , ≤ θ < , 0 ≤ r < ∞}.
6 6
Then C is a cone. We mention in passing that C is not closed in C.
(c) Let V = C([0, 1], C) ∶= {f ∶ [0, 1] → C ∶ f is continuous}. Let
C ∶= {f ∈ C([0, 1], C) ∶ f (x) ≥ 0 for all x ∈ [0, 1]}.
Then C is a cone.
We denote by
Simp(E, R)
the set of all simple, extended, measurable real-valued functions on E, and we set
Simp(E, [0, ∞]) ∶= {φ ∈ Simp(E, R) ∶ 0 ≤ φ(x) for all x ∈ E}.
Alas, Simp(E, R) is not a vector space over R, since if φ ∈ Simp(E, R) and
φ−1 ({−∞, ∞}) ≠ ∅, then φ does not admit an additive inverse (recall that we have
not defined −∞ + ∞ in R).
The next result will be the key to our definition of Lebesgue integrability of
functions in the next section.
4.26. Proposition. Let E ∈ M(R) and f ∶ E → [0, ∞] be a measurable function.
Then there exists an increasing sequence
φ1 ≤ φ2 ≤ φ3 ≤ ⋯ ≤ f
of simple, real-valued functions φn such that
f (x) = lim φn (x) for all x ∈ E.
n→∞
Proof. The proof of this Proposition is left as an Assignment question.
◻
52 L.W. Marcoux Introduction to Lebesgue measure
Appendix to Section 4.
4.27. Let us now examine the extended real numbers from a somewhat different
point of view. You may consider the discussion below “culture”.
The proof of the following result is left as yet another exercise for the reader.
We emphasize that the topology on [0, 1] ⊆ R is the usual (relative) topology that
it inherits as a subset of R, itself equipped with the usual topology.
4.29. Theorem. (R, τ2 ) is homeomorphic to the interval [0, 1], under a home-
omorphism that sends R ⊆ R to the dense subset (0, 1) ⊆ [0, 1].
Exercise 4.1.
Let E ∈ M(R). Prove that M(E) is a σ-algebra of sets.
Exercise 4.2.
Let (X, d) be a metric space.
(a) Let E ∈ M(R) be a measurable set. Verify that a function f ∶ E → X is
measurable if and only if f −1 (F ) ∈ M(R) for every closed subset F of X.
(b) Let H ⊆ R be a set. Verify that a constant function f ∶ H → X is measurable
if and only if H is measurable.
Exercise 4.3.
Let f ∶ R → K be a function and g ∶ K → R be the absolute value function
g(x) = ∣x∣, x ∈ K. Suppose that g ○ f = ∣f ∣ is measurable.
Give an example to show that f need not be measurable.
Exercise 4.4.
(a) Let G ⊆ K2 be an open set. Prove that there exists open sets An , Bn ⊆ K,
n ≥ 1, such that
G = ∪∞n=1 An × Bn .
(b) Suppose that G ⊆ R2 is a non-empty open set. Prove that there exist
countably many rectangles Rn = (an , bn ) × (cn , dn ) ⊆ R2 such that G =
∪∞
n=1 Rn .
Exercise 4.5.
Prove that the functions σ, µ and δ from Proposition 4.8 are all continuous.
Exercise 4.6.
Formulate and prove an analogue of Proposition 4.11 for complex-valued func-
tions.
Exercise 4.7.
Complete the proof of (d) implies (e) in Proposition 4.11.
Exercise 4.9.
Let E ∈ M(R). Show that an extended real-valued function f ∶ E → R is
measurable if and only the following two conditions hold.
(a) f −1 (F ) ∈ M(E) for all closed sets F ⊆ R, and
(b) f −1 ({−∞}) and f −1 ({∞}) ∈ M(E).
4. LEBESGUE MEASURABLE FUNCTIONS 55
Exercise 4.10.
Let E ∈ M(R) and suppose that f ∶ E → R is a function. Prove that the following
statements are equivalent.
(a) f if Lebesgue measurable.
(b) For all α ∈ R, f −1 ((α, ∞]) ∈ M(E).
(c) For all β ∈ R, f −1 ([−∞, β)) ∈ M(E).
Exercise 4.11.
Let E ∈ M(R) and suppose that (fn )∞n=1 is a sequence of extended real-valued,
measurable functions on E. Complete the proof of Proposition 4.18 by showing that
g ∶= lim inf fn
n≥1
is measurable.
Exercise 4.12.
Let E ∈ M(R). Prove that Simp(E, R) is an algebra over R.
Exercise 4.14.
Let E and F be measurable sets in R and suppose that f ∶ E → R and g ∶ F → R
are functions.
(a) Define the function
f̂ ∶ R → R
⎧
⎪f (x) if x ∈ E
⎪
x ↦ ⎨ .
⎪
⎪ 0 if x ∈/ E
⎩
Prove that f is measurable if and only if f̂ is measurable.
(b) Suppose that E ∩ F = ∅. Prove that the function h ∶ E ∪ F → R defined by
⎧
⎪f (x) if x ∈ E
⎪
h(x) = ⎨
⎪
⎪g(x) if x ∈ F
⎩
is measurable if and only if both f and g are measurable.
(c) Does the conclusion from (b) hold if E ∩ F ≠ ∅? Either prove that it does,
or find a counterexample to show that it doesn’t.
56 L.W. Marcoux Introduction to Lebesgue measure
Hint for (a): For each n ≥ 1, partition the interval [0, n) into n2n equal subintervals
−1
Ek,n = [ 2kn , k+1 n
2n ), 0 ≤ k < (n2 ) − 1. Set En2 ,n = [n, ∞]. Use the sets f (Ek,n ),
n
n
0 ≤ k ≤ n2 to build φn .
Hint for (b): This should be very short, otherwise you are doing something wrong.
Exercise 4.16.
Let E ⊆ R be a set of measure zero, and let f ∶ E → R be any function whatsoever.
Prove that f is measurable.
5. LEBESGUE INTEGRATION 57
5. Lebesgue integration
I know that there are people who do not love their fellow man, and I
hate people like that.
Tom Lehrer
5.3. Example.
(a) Let φ = 0χ[4,∞) + 17χQ∩[0,4) + 29χ[2,4)∖Q . Then
(b) Let C ⊆ [0, 1] be the Cantor set from Example 3.15, and consider
φ = 1 χC + 2 χ[5,9] . Then
5.5. Lemma. Let E ∈ M(R) and suppose that φ, ψ ∶ E → R are simple, real-
valued, measurable functions. Then there exist
(i) N ∈ N,
(ii) α1 , α2 , . . . , αN , β1 , β2 , . . . , βN ∈ R, and
(iii) H1 , H2 , . . . , HN ∈ M(E)
such that Hi ∩ Hj = ∅ if 1 ≤ i ≠ j ≤ N ,
N N
φ = ∑ αn χHn and ψ = ∑ βn χHn .
n=1 n=1
Remark. The key things to notice here are that the Hn ’s appearing in the decom-
positions of φ and ψ are the same, and the representations are disjoint.
Proof. Let φ = ∑M 1 M2
j=1 aj χEj and ψ = ∑k=1 bk χFk , where Ej , Fk are measurable subsets
of E for all 1 ≤ j ≤ M1 , 1 ≤ k ≤ M2 , the Ej ’s are pairwise disjoint, and the Fk ’s are
pairwise disjoint. (That such a decomposition exists is clear, as we may simply write
φ and ψ in standard form.)
5. LEBESGUE INTEGRATION 59
and similarly
M1 M2
ψ = ∑ ∑ bk χEj ∩Fk .
j=1 k=1
Relabel {Ej ∩ Fk ∶ 1 ≤ j ≤ M1 , 1 ≤ k ≤ M2 } as {Hn ∶ 1 ≤ n ≤ N } to complete the proof.
(The αn ’s and βn ’s are clearly just relabelings of the aj ’s and the bk ’s respectively.)
◻
5.6. Lemma. Let E ∈ M(R) and suppose that φ ∈ Simp(E, [0, ∞]). If φ =
∑N
n=1 αn χHnis any disjoint representation of φ, then
N
∫ φ = ∑ αn mHn .
E n=1
(a) Then κφ + ψ = ∑N
n=1 (καn + βn )χHn is a disjoint representation of κφ + ψ,
and so by Lemma 5.6,
N
∫ (κφ + β) = ∑ (καn + βn )mHn
E n=1
N N
= κ( ∑ αn mHn ) + ∑ βn mHn
n=1 n=1
= κ ∫ φ + ∫ ψ.
E E
5.9. Remarks.
(a) We leave it as an exercise for the reader to show that the above definition
is equivalent to defining
new
∫ f = sup{∫ φ ∶ φ ∈ Simp(E, [0, ∞]), 0 ≤ φ ≤ f }.
E E
(The difference being that we now allow the simple functions to be extended
real-valued and non-negative, instead of just real-valued and non-negative.)
(b) The reason for putting the superscript “new” in the above integral is the
following. Observe that if φ ∈ Simp(E, [0, ∞]), we now have two definitions
for the integral of φ. That is, writing φ = ∑N n=1 αn χHn in standard form,
we have our original definition (Definition 5.2)
N
∫ φ = ∑ αn mHn ,
E n=1
while from Definition 5.8, our new definition of the integral of φ becomes
new
∫ φ = sup{∫ ψ ∶ ψ ∈ Simp(E, [0, ∞)), 0 ≤ ψ ≤ φ}.
E E
5. LEBESGUE INTEGRATION 61
new
(c) It is entirely possible that ∫E f = ∞. For example, if φ = ∞ ⋅ χ[0,1] , then
new
∫[0,1] φ = ∞.
new
Alternatively, if f (x) = x, x ∈ [0, ∞), then ∫[0,∞) f = ∞. The proof of
this is left as an exercise for the reader.
5.10. Let us reconcile these two definitions. Once this is done, we will no longer
need to distinguish between the original and the new integral for non-negative,
simple, measurable functions, and so we shall drop the superscript “new” for the
integrals of non-negative, measurable extended-real valued functions altogether.
On the one hand, note that φ ∈ {ψ ∈ Simp(E, [0, ∞)), 0 ≤ ψ ≤ φ}, and so by
new
definition of ∫E φ, we have that
new
∫ φ≤∫ φ.
E E
On the other hand, if ψ ∈ Simp(E, [0, ∞)) and 0 ≤ ψ ≤ φ, then by Proposition 5.7
(b),
∫ ψ ≤ ∫ φ,
E E
and thus
new
∫ φ = sup{∫ ψ ∶ ψ ∈ Simp(E, [0, ∞)), 0 ≤ ψ ≤ φ} ≤ ∫ φ.
E E E
∫ f
E
for the Lebesgue integral of an element f ∈ L(E, [0, ∞]).
5.11. Remark. We shall see in the Assignments that even when f is a rela-
tively innocuous-looking function (for example f (x) = x on [0, 1]), calculating the
Lebesgue integral of f directly from the definition is an arduous task. Fortunately,
Theorem 5.24 below will provide us with an alternate means of calculating the inte-
grals of a large family of (Riemann integrable) functions, by showing that in many
cases, the Lebesgue integral coincides with the Riemann integral. Of course, when
the function is sufficiently nice, we may apply the Fundamental Theorem of Calculus
to calculate the latter.
Sets of measure zero will play a central role in the theory that follows. The
reason for this lies partly in the fact that the Lebesgue integral “ignores” these sets,
in a sense which we shall now make precise.
62 L.W. Marcoux Introduction to Lebesgue measure
5.13. Example. Let E ∈ M(R). Given f, g ∈ L(E, R), we say that f = g almost
everywhere if
B ∶= {x ∈ E ∶ f (x) ≠ g(x)}
has measure zero.
More specifically, therefore, χQ = 0 = χC a.e. on R, where C is the Cantor set
from Example 3.15.
∫ f = ∫ f1 + ∫ f2 .
E X Y
(b) If g ≤ h, then ∫E g ≤ ∫E h.
(c) If H ⊆ E is a measurable set, then
∫ g = ∫ g ⋅ χH ≤ ∫ g.
H E E
Proof. The proof of this lemma is left as a worthwhile exercise for the reader.
◻
Thus
∫ f =∫ f +∫ f
E E∖B B
=∫ g+0
E∖B
=∫ g+∫ g
E∖B B
= ∫ g.
E
◻
We now come to one of the major results in this course. In dealing with prop-
erties that hold almost everywhere on a measurable set E ∈ M(R) (as will be the
case in the Monotone Convergence Theorem below), we often have recourse to the
following line of argument: we isolate the “bad” set K of measure zero where the
property under consideration fails to hold, and the deal with the “good” set E ∖ K,
where the property holds everywhere. Using Proposition 5.15 and Lemma 5.14, we
can often “glue” these results together. This is a lot of quotation marks, which are
the written equivalent of randomly flailing arms. Let us see the strategy in action,
where it might make more sense.
Step Two.
Next, for each n ≥ 1, set
En ∶= {x ∈ E ∶ fn (x) > fn+1 (x)},
so that mEn = 0, and thus En is measurable. Let B = E0 ∪ (∪∞
n=1 En ). Then
∞
m∗ B ≤ ∑ mEn = 0,
n=0
Step Three. The first two steps were only to reduce the problem to the case where
the interesting properties hold everywhere. Now the real argument begins.
Since gn ≤ gn+1 ≤ g for all n ≥ 1, by Lemma 5.14, we have that
∫ gn ≤ ∫ gn+1 ≤ ∫ g
H H H
for all n ≥ 1, and thus
sup ∫ gn = lim ∫ gn ≤ ∫ g.
n≥1 H n→∞ H H
Conversely, suppose that φ ∈ Simp(H, [0, ∞]) and that φ ≤ g. Let 0 < ρ < 1. We
shall prove that
∫ ρ φ = ρ ∫ φ ≤ sup ∫ gn .
H H n≥1 H
∫ g ≤ sup ∫ gn .
H n≥1 H
Combining this with the reverse inequality from Step Three shows that
∫ g = sup ∫ gn = n→∞
lim ∫ gn .
H n≥1 H H
Step Six. There remains only to “glue” the above results together to get the desired
statement.
By Lemma 5.14,
∫ f = ∫ f + ∫ f = 0 + ∫ g = n→∞
lim ∫ gn = lim ∫ fn + ∫ fn = lim ∫ fn .
n→∞ n→∞
E B H H H B H E
◻
Steps Three to Five of the above proof provide a proof of the Monotone Con-
vergence Theorem in the case where the sequence (fn )∞
n=1 is everywhere increasing
and where the sequence tends to f everywhere.
66 L.W. Marcoux Introduction to Lebesgue measure
∫ χE = lim ∫ fn = lim 0 = 0.
[0,1] n→∞ [0,1] n→∞
0≤∫ f =∫ χE = mE ≤ mQ = 0.
[0,1] [0,1]
The point is that the limit function χE is Lebesgue integrable, even though it is not
Riemann integrable.
The first half the of the following Corollary extends results from Proposition 5.7.
There, we dealt with nonnegative, simple, measurable functions. We remove the
requirement that the functions be simple.
∫ κf + g = κ ∫ f + ∫ g.
E E E
(b) If (hn )∞ N
n=1 is a sequence in L(E, [0, ∞]) and if h(x) ∶= limN →∞ ∑n=1 hn (x)
for all x ∈ E, then h is measurable and
∞
∫ h = ∑ ∫ hn .
E n=1 E
Proof.
5. LEBESGUE INTEGRATION 67
(a) From our work in the Assignments (see Exercise 4.15), we may choose
increasing sequences (φn )∞ ∞
n=1 and (ψn )n=1 in Simp(E, [0, ∞]) such that
f (x) = limn→∞ φn (x) and g(x) = limn→∞ ψn (x) for all x ∈ E.
By the Monotone Convergence Theorem,
∫ f = n→∞
lim ∫ φn
E E
and
∫ g = n→∞
lim ∫ ψn .
E E
Given 0 ≤ κ ∈ R, it follows that (κφn + ψn )∞
n=1 is again an increasing se-
quence of non-negative, simple, measurable, functions converging pointwise
to the function κf + g.
Applying the Monotone Convergence Theorem 5.16 once more, we see
that
∫ (κf + g) = n→∞
lim ∫ (κφn + ψn ) = lim κ ∫ φn + ∫ ψn = κ ∫ f + ∫ g.
n→∞
E E E E E E
(b) For N ≥ 1, set gN ∶= ∑N
n=1 hn .
Then each gN is measurable (exercise), and
N
g =
∫E N ∑n=1 ∫E n h by part (a). Furthermore,
0 ≤ g1 ≤ g2 ≤ ⋯.
Now h(x) = limN →∞ gN (x) for all x ∈ E, and so h is measurable by Corol-
lary 4.19.
By the Monotone Convergence Theorem,
N ∞
∫ h = Nlim
→∞
∫ gN = Nlim
→∞
∑ ∫ hn = ∑ ∫ hn .
E E n=1 E n=1 E
By part (b),
∫ f = ∫ f ⋅ χH
H E
∞
= ∫ ∑ fn
E n=1
∞
= ∑ ∫ fn
n=1 E
∞
= ∑ ∫ f ⋅ χHn
n=1 E
∞
= ∑∫ f.
n=1 Hn
◻
68 L.W. Marcoux Introduction to Lebesgue measure
In Chapter 9 and later, we shall need to specify the variable with respect to
which we are integrating. In analogy to the usual notation for Riemann integration,
we shall write
∫ f = ∫ f (s)dm(s)
E E
to denote the Lebesgue integral of f with respect to the variable s.
∫ f = ∫ g.
E E
This will prove to be more useful than it might first appear to be. One
huge problem with L1 (E, R) is that it is not a vector space!!! One problem
5. LEBESGUE INTEGRATION 69
lies in the fact that if f, g ∈ L1 (E, R), x ∈ E and f (x) = ∞, g(x) = −∞,
then what should (f + g)(x) be?
This is not an issue insofar as L1 (E, R) is concerned. As we shall see
in Chapter 6, we may establish an equivalence relation on L1 (E, R) by set-
ting f ∼ g if f = g a.e. on E. We can then turn the equivalence classes of
elements of L1 (E, R) into a vector space in a natural way. In most texts,
these equivalence classes are denoted by the same notation used to denote
functions, and indeed, they are often referred to as “functions”, although
technically speaking they are not. Being absolute sticklers for detail, and
inspired by our French heritage, we shall exercise as much caution as possi-
ble in the use of language, and will try to be as precise as humanly possible
in keeping the notation and terminology straight.
(e) Suppose that g ∶ E → C is a measurable function. Let us write
g = (g1 − g2 ) + i(g3 − g4 ),
where g1 = (Re g)+ , g2 = (Re g)− , g3 = (Im g)+ , and g4 = (Im g)− .
We shall say that g is Lebesgue integrable if each of g1 , g2 , g3 and g4
are, in which case we define
∫ g = (∫ g1 − ∫ g2 ) + i(∫ g3 − ∫ g4 ).
E E E E E
Of course, this is equivalent to requiring that Re g and Im g be Lebesgue
integrable, in which case we define
The perspicacious reader will observe that we have carefully avoided all
notions of “extended” complex-valued functions.
= ∫ κf − ∫ κf −
+
E E
= κ ∫ f − κ ∫ f− +
E E
= κ ∫ f.
E
Case 3. κ < 0.
Then (κf )+ = −κf − and (κf )− = −κf + , so that
+ −
∫ κf ∶= ∫ (κf ) − ∫ (κf )
E E E
= ∫ (−κ)f − ∫ (−κ)f +−
E E
= (−κ) ∫ f − (−κ) ∫ f + −
E E
= (−κ)(− ∫ f )
E
= κ ∫ f.
E
(b) Let h = f + g, so that h ∈ L(E, R), as the latter is a vector space. Write
h = h+ − h− .
Then
h+ , h− ≤ h+ + h− = ∣h∣ ≤ ∣f ∣ + ∣g∣ = f + + f − + g + + g − ,
so that
+ + − + −
∫ h ≤ ∫ f + ∫ f + ∫ g + ∫ g < ∞,
E E E E E
and
− + − + −
∫ h ≤ ∫ f + ∫ f + ∫ g + ∫ g < ∞.
E E E E E
It follows that h ∈ L1 (E, R).
Furthermore, h = f + g implies that h+ + f − + g − = h− + f + + g + , whence
+ − − − + +
∫ h +∫ f +∫ g =∫ h +∫ f +∫ g .
E E E E E E
From this it easily follows that
+ −
∫ h = ∫ h − ∫ h = ∫ f + ∫ g.
E E E E E
+ −
(c) Note that ∣f ∣ = f + f is measurable, and
+ −
∫ ∣f ∣ = ∫ f + ∫ f < ∞,
E E E
proving that ∣f ∣ ∈ L1 (E, R).
5. LEBESGUE INTEGRATION 71
Finally,
∣∫ f ∣ = ∣∫ f + − ∫ f − ∣
E E E
≤ ∣∫ f + ∣ + ∣∫ f − ∣
E E
= ∫ f + ∫ f−
+
E E
= ∫ ∣f ∣.
E
b
∫ φ=∫ φ.
[a,b] a
N
φ = ∑ αn χ[pn−1 ,pn ) .
n=1
72 L.W. Marcoux Introduction to Lebesgue measure
Then
N
∫ φ = ∑ αn m[pn−1 , pn )
[a,b] n=1
N
= ∑ αn (pn − pn−1 )
n=1
N pn
= ∑∫ αn
n=1 pn−1
b N
=∫ ∑ αn χ[pn−1 ,pn )
a n=1
b
=∫ φ.
a
◻
That is, the Lebesgue and Riemann integrals of f over [a, b] coincide.
Proof. Suppose that ∣f ∣ is bounded above on [a, b] by 0 < M ∈ R. Let g = M χ[a,b] .
Clearly 12 (f + g) is Riemann integrable and
1 b 1 b M (b − a)
∫ (f + g) = ∫ f + .
a 2 2 a 2
If we prove that 12 (f +g) ∈ L1 ([a, b], R), then it is readily seen that f ∈ L1 ([a, b], R)
and – in light of Lemma 5.23 –
1 1 M (b − a)
∫ (f + g) = ∫ f+ .
[a,b] 2 2 [a,b] 2
1
We have shown that by replacing f by 2 (f + g) if necessary, we may assume
from the outset that 0 ≤ f ≤ M on [a, b].
Cast your mind back to the halcyon days when you studied Chapter 1, and more
specifically to the Cauchy Criterion, Theorem 1.13. Recall that it asserts that for
each n ≥ 1 there exists a partition Rn ∈ P([a, b]) such that for all refinements X and
Y of Rn , and for all choices of test values X ∗ and Y ∗ for X and Y respectively,
1
∣S(f, X, X ∗ ) − S(f, Y, Y ∗ )∣ < .
n
N
Let QN ∶= ∪n=1 Rn , 1 ≤ N ∈ N, so that QN is a common refinement of R1 , R2 , . . . , RN .
Write
QN = {a = q0,N < q1,N < q2,N < ⋯ < qmN ,N = b}.
Set Hk,N = [qk−1,N , qk,N ), 1 ≤ k ≤ mN − 1 and HmN ,N = [qmN −1,N , qmN ,N ].
5. LEBESGUE INTEGRATION 73
Define
αk,n = inf{f (t) ∶ t ∈ Hk,n }, 1 ≤ k ≤ mN ,
βk,n = sup{f (t) ∶ t ∈ Hk,n }, 1 ≤ k ≤ mN ,
mN mN
and set φN = ∑N αk,n χHk,n and ψN =
n=1 ∑k=1 βk,n χHk,n . ∑N
n=1 ∑k=1
Since each QN is a refinement of QN −1 , it is not hard to see that
φ1 ≤ φ2 ≤ φ3 ≤ ⋯ ≤ f ≤ ⋯ ≤ ψ3 ≤ ψ2 ≤ ψ1 .
Moreover, using Lemma 5.23 above, we obtain
b b
∫ φN = ∫ φN = inf{S(f, QN , Q∗N ) ∶ Q∗N test values for QN } ≤ ∫ f,
[a,b] a a
and similarly,
b b
∫ ψN = ∫ ψN = sup{S(f, QN , Q∗∗ ∗∗
N ) ∶ QN test values for QN } ≥ ∫ f.
[a,b] a a
But QN is a refinement of RN , and thus
1
∣S(f, QN , Q∗N ) − S(f, QN , Q∗∗
N )∣ <
N
for all choices of test values Q∗N and Q∗∗ N for QN .
b
It follows that ∫[a,b] φN ≤ ∫a f ≤ ∫[a,b] ψN and
1
∣∫ φN − ∫ ψN ∣ ≤ , N ≥ 1.
[a,b] [a,b] N
Let φ(x) = supn≥1 φn (x) = limn→∞ φn (x) ≤ f (x), and ψ(x) = inf n≥1 ψn (x) =
limn→∞ ψn (x) ≥ f (x), x ∈ [a, b]. Since each φn , ψn is measurable, so are φ and ψ,
and by the Monotone Convergence Theorem and Lemma 5.23,
∫ φ = lim ∫ φN
[a,b] n→∞ [a,b]
b
= lim ∫ φN
n→∞ a
b
=∫ f
a
b
= lim ∫ ψN
n→∞ a
= lim ∫ ψN
n→∞ [a,b]
=∫ ψ.
[a,b]
Proof. Observe that f is bounded and Riemann-integrable if and only if its real and
imaginary parts are bounded and Riemann-integrable. The result now immediately
follows by applying Theorem 5.24 to each of these.
◻
Theorem 5.24 required some effort. Let us see that it was worth it.
5.26. Example. Let f (x) = x, x ∈ [0, 1]. In the Assignments, you computed
the Lebesgue integral of f over [0, 1] to be
1
∫ x= .
[0,1] 2
Equipped with Theorem 5.24, it is child’s play. The function f is clearly bounded
and continuous (hence Riemann-integrable) over [0, 1], and so by that Theorem,
x=1
1 x2 1
∫ x=∫ x dx = ] = .
[0,1] 0 2 x=0 2
1
5.27. Example. Let f (x) = , x ∈ E ∶= [1, ∞). We wish to determine ∫[1,∞) f .
x2
For each n ≥ 1, set fn ∶= f ⋅ χ[1,n] . Then f is measurable (because it is continuous
except at one point of E), and
0 ≤ f1 ≤ f2 ≤ ⋯,
∫ f = lim ∫ fn
[1,∞) n→∞ [1,∞)
1
= lim ∫
n→∞ [1,n] x2
n 1
= lim ∫
n→∞ 1 x2
1 x=n
= lim − ]
n→∞ x x=1
1
= lim (− − (−1))
n→∞ n
= 1.
In this example, the Lebesgue integral of f returns the value of the improper
Riemann integral of f over [1, ∞). Two things are worth noting:
5.28. Example. The Monotone Convergence Theorem 5.16 states that if (fn )∞
n=1
is an (almost everywhere) increasing sequence of measurable functions on a set
E ∈ M(R), then f = limn→∞ fn is measurable and
∫ lim fn = lim ∫ fn .
E n→∞ n→∞ E
In the absence of the adjective “increasing”, we can not expect this result to hold.
For example, consider the sequence (fn )∞n=1 given by
fn ∶ [1, ∞) → R
⎧
⎪1
⎪ if 1 ≤ x ≤ en
x ↦ ⎨ nx .
⎪
⎪0 if x > en
⎩
en 1
=∫
1 nx
n
log x x=e
= ]
n x=1
=1−0
= 1.
Hence limn→∞ ∫[1,∞) fn = 1 ≠ 0 = ∫[1,∞) f .
I have done a bit (but not a great deal) of research to try to determine why
the next result is referred to as Fatou’s Lemma instead of Fatou’s Theorem. One
possible explanation is that it can be used to prove a number of other useful results
very quickly, and as such, is a “facilitator”, to employ the jolly discourse of psycho-
babble. A second possibility (which I have not read anywhere) is that it is petty
jealousy on the part of his peers. In any case, there is a different result from complex
analysis known as Fatou’s Theorem. In order to state it, we first require the notion
of Lp -spaces, and so we defer its statement to the Appendix of Chapter 6.
∫ limninf fn ≤ limninf ∫ fn .
E E
Proof. For each N ≥ 1, set gN = inf{fn ∶ n ≥ N }. By Proposition 4.18, gN is
measurable for all N and
g1 ≤ g2 ≤ g3 ≤ ⋯.
By the Monotone Convergence Theorem 5.16,
∫ gN ≤ ∫ fn for all n ≥ N,
E E
whence
∫ gN ≤ limninf ∫ fn .
E E
But this holds for any N ≥ 1, and so by taking limits and using the above equality,
we find that
∫ limninf fn = Nlim ∫ gN ≤ limninf ∫ fn .
→∞ E
E E
◻
5. LEBESGUE INTEGRATION 77
∫ lim inf fn = ∫ 0 = 0.
[0,1] n [0,1]
5.31. Example. Let E = [0, ∞) ∈ M(R), and for each n ≥ 1, let fn = − n1 χ[n,2n] .
Then each fn is measurable and (fn )n converges uniformly to f ≡ 0 on E. A fortiori,
(fn )n converges pointwise to f .
Neverthess,
and similarly,
−
∫ f =∫ f− ≤ ∫ g = ∫ g < ∞.
E E∖B E∖B E
+ −
Hence f = f − f ∈ L1 (E, R).
The following result is also one of the major results of measure theory.
78 L.W. Marcoux Introduction to Lebesgue measure
∫ f = n→∞
lim ∫ fn
H H
if and only if
lim ∫ fn .
∫ f = n→∞
E E
In other words, by replacing E with H if necessary, we may assume without loss
of generality that ∣fn (x)∣ ≤ g(x) < ∞ and that f (x) = limn→∞ fn (x) for all x ∈ E.
We shall assume that we have done this.
Step Two.
Note that g − fn ≥ 0 on E, and thus by Fatou’s Lemma 5.29,
≤ lim inf ∫ g − fn
n E
= ∫ g − lim sup ∫ fn ,
E n E
∫ g + ∫ limninf fn = ∫ limninf (g + fn )
E E E
≤ lim inf ∫ (g + fn )
n E
= ∫ g + lim inf ∫ fn ,
E n E
or equivalently
∫ f ≤ limninf ∫ fn .
E E
Putting these two inequalities together shows that
∫ f = n→∞
lim ∫ fn .
E E
◻
80 L.W. Marcoux Introduction to Lebesgue measure
Appendix to Section 5.
5.34. Let E ∈ M(R) and g ∈ L1 (E, R). We have observed that the set B ∶=
{x ∈ E ∶ g(x) ∈ {−∞, ∞}} has measure zero. When necessary, as it was in the proof
of the Lebesgue Dominated Convergence Theorem, we were able to simply “excise”
this set from the domain and concentrate our attention to the set E ∖ B. So why
introduce the extended real-numbers at all?
Convenience. Given an increasing sequence (fn )∞ n=1 in L1 (E, [0, ∞)), the point-
wise limit f (x) ∶= limn→∞ fn (x) need not be real-valued. By introducing the ex-
tended real numbers, we are able to treat the limit function f as simply another
measurable function.
When we define the Lp -spaces in Chapter 6, we shall define each Lp (E) as equiv-
alence classes of (extended real-valued) functions. However, each such equivalence
class will always admit a representative which is real-valued function. Truly, fortune
smiles upon us.
5. LEBESGUE INTEGRATION 81
Exercise 5.3.
Prove Lemma 5.14; that is, let E ∈ M(R) and let f, g and h ∶ E → [0, ∞] be
functions. Suppose that g and h are measurable.
(a) Suppose furthermore that E = X ⊍ Y , where X and Y are measurable. Set
f1 ∶= f ∣X and f2 ∶= f ∣Y . Then f is measurable if and only if both f1 and f2
are measurable. When such is the case,
∫ f = ∫ f1 + ∫ f2 .
E X Y
(b) If g ≤ h, then ∫E g ≤ ∫E h.
(c) If H ⊆ E is a measurable set, then
∫ g = ∫ g ⋅ χH ≤ ∫ g.
H E E
Exercise 5.4.
Prove that if f ∈ L1 (E, R), then
m(f −1 ({−∞})) = 0 = m(f −1 ({∞})).
Exercise 5.5.
sin x
Let f (x) = , x ≥ 1.
x
∞
(a) Prove that the improper Riemann integral ∫1 f (x)dx exists.
(b) Prove that the Lebesgue integral ∫[1,∞) f = ∞ does not exist.
Exercise 5.7.
Let E ∈ M(R). Suppose that f, g ∈ L1 (E, C), and that κ ∈ C. Prove that
(a) the function κf + g ∈ L1 (E, C) and
∫ κf + g = κ ∫ f + ∫ g.
E E E
82 L.W. Marcoux Introduction to Lebesgue measure
∣∫ f ∣ ≤ ∫ ∣f ∣.
E E
Exercise 5.8.
Let E ∈ M(R). Show that a measurable function f ∶ E → C lies in L1 (E, C) if
and only if ∣f ∣ ∈ L1 (E, C).
Show that this fails if we do not assume that f is measurable.
Exercise 5.9.
The following special case of the Dominated Convergence Theorem is easily
derived from the Monotone Convergence Theorem.
Let E ∈ M(R) and suppose that (fn )∞
n=1 is a decreasing sequence in L(E, [0, ∞])
with f1 ∈ L1 (E, [0, ∞]). Thus
f1 ≥ f2 ≥ f3 ≥ ⋯ ≥ 0.
Define f ∶ E → [0, ∞] by f (x) = limn→∞ fn (x), x ∈ E.
Prove that f is measurable and that
∫ f = n→∞
lim ∫ fn .
E E
Exercise 5.10.
Let E ∈ M(R). If φ, ψ ∈ Simp(E, [0, ∞]) and κ = ∞, prove or disprove that
∫ κφ + ψ = κ ∫ φ + ∫ ψ.
E E E
6. Lp SPACES 83
6. Lp Spaces
I’ve been on food stamps and welfare. Anybody help me out? No!
Craig T. Nelson
6.1. Functional analysis is the study of normed linear spaces and the continuous
linear maps between them. Amongst the most important examples of Banach spaces
are the so-called Lp -spaces, and it is to these that we now turn our attention. The
reader may wish to refresh his/her memory as to the definition of a seminorm on a
vector space X over K (Definition 1.2).
6.2. Example. Let E ⊆ R be a Lebesgue measurable set, and suppose that
mE > 0. Recall that
L1 (E, K) = {f ∶ E → K ∶ f is measurable and ∫ ∣f ∣ < ∞}.
E
6.3. Proposition. Let W be a vector space over the field K, and suppose that ν
is a seminorm on W. Let N ∶= {w ∈ W ∶ ν(w) = 0}. Then N is a linear manifold in
W and so W/N is a vector space over K, whose elements we denote by [x] ∶= x + N .
Furthermore, the map
∥ ⋅ ∥ ∶ W/N → R
[x] ↦ ν(x)
defines a norm on W/N .
Proof. Clearly 0 ∈ N and thus N ≠ ∅. Suppose that v, w ∈ N , and k ∈ K. Then
ν(kv + w) ≤ ν(kv) + ν(w) = ∣k∣ν(v) + ν(w) = 0,
84 L.W. Marcoux Introduction to Lebesgue measure
6.4. Let us return to Example 6.2, where we determined that ν1 (⋅) defines a
seminorm on L1 (E, K).
Suppose that g ∈ N1 (E, K) ∶= {f ∈ L1 (E, K) ∶ ν1 (f ) = 0}. Then ∫E ∣g∣ = 0, and
therefore g = 0 a.e. on E. Conversely, if g = 0 a.e. on E, then ∫E ∣g∣ = 0 and therefore
g ∈ N1 (E, K). In other words,
N1 (E, K) = {g ∈ L1 (E, K) ∶ g = 0 a. e. on E}.
Thus [g] = [h] in L1 (E, K)/N1 (E, K) if and only if g − h ∈ N1 (E, K), which is to
say that g = h a.e. on E. By Proposition 6.3, the map ∥[f ]∥ ∶= ν1 (f ) defines a norm
on L1 (E, K) ∶= L1 (E, K)/N1 (E, K).
6.5. Definition. The space L1 (E, K) = L1 (E, K)/N1 (E, K) defined above is
referred to as “L1 of E”, and it is a normed linear space.
It is crucial to remember that the elements of L1 (E, K) are cosets of L1 (E, K);
that is to say, they are equivalence classes of functions which are equal a.e. on E.
Given an element [f ] of L1 (E, K), one can not speak of the value of the function
at a point in E, since we are not dealing with functions!
Our next goal is to perform a similar construction on a family of spaces indexed
by positive real numbers 1 < p < ∞.
6.6. Definition. Let E ∈ M(R), so that E is Lebesgue measurable. Let 1 < p <
∞ be a real number, and set
Lp (E, K) ∶= {f ∈ L(E, K) ∶ ∫ ∣f ∣p < ∞} = {f ∈ L(E, K) ∶ ∣f ∣p ∈ L1 (E, K)}.
E
6. Lp SPACES 85
6.7. We would like to verify that Lp (E, K) is a vector space for all 1 < p < ∞,
1/p
and that νp (f ) ∶= (∫E ∣f ∣p ) defines a seminorm on Lp (E, K). Then we can once
again appeal to Proposition 6.3 to obtain a normed linear space as a quotient of
Lp (E, K).
The proof of this is, however, somewhat technical, and will require a couple of
auxiliary results.
When 1 < p < ∞, we see that the above equation is equivalent to each of the
equations
● p(q − 1) = q and
● (p − 1)q = p.
While these are trivial algebraic manipulations, it will prove useful to keep them in
mind in the proofs below.
6.9. Lemma. Young’s Inequality. Let 1 < p < ∞ and denote by q the
Lebesgue conjugate of p. Let 0 < a, b ∈ R.
(a) Then
ap bq
ab ≤ + .
p q
(b) Equality holds in the above expression if and only if ap = bq .
1 1
Proof. Consider the function g ∶ (0, ∞) → R given by g(x) = xp + − x. Clearly g
p q
is differentiable on (0, ∞) with g ′ (x) = xp−1 − 1. Thus g is clearly strictly decreasing
on (0, 1) and it is strictly increasing on (1, ∞). Furthermore,
1 1
g(1) = + − 1 = 0.
p q
In particular, g(x) ≥ 0 for all x ∈ (0, ∞), and g(x) = 0 if and only if x = 1.
a
(a) Set x0 ∶= q−1 > 0. Then
b
ap 1 a
0 ≤ g(x0 ) = (q−1)p + − q−1 ,
pb q b
86 L.W. Marcoux Introduction to Lebesgue measure
so that
a ap 1
≤ + .
bq−1 pbq q
That is,
ap bq
ab ≤ + .
p q
(b) From above, equality holds if and only if g(x0 ) = 0, which happens if and
only if x0 = 1. But this is clearly equivalent to the condition that a = bq−1 ;
i.e. ap = bp(q−1) = bq .
◻
Recall from Proposition 4.10 that if E ∈ M(R) and f ∶ E → K is a measurable
function, then there exists a measurable function u ∶ E → T such that f = u ⋅ ∣f ∣.
Clearly, if K = R, then the range of u is contained in {−1, 1}. Let us denote by u
the function u(x) = u(x) for all x ∈ E.
ν1 (f f ∗ ) = ∫ f f ∗ = νp (f ).
E
Remark. In an unfortunate coincidence of terminology, we shall also refer to f ∗ as
the Lebesgue conjugate function of f .
Proof.
We first observe that by Proposition 4.8, f g is measurable.
(a) Note that if f = 0 a.e or g = 0 a.e. on E, then f g = 0 a.e. on E and there is
nothing to prove. It is easy to verify that given 0 < α, β ∈ K, αf ∈ Lp (E, K)
and βg ∈ Lq (E, K). Suppose that we can find α0 , β0 ≠ 0 such that
∫ ∣f g∣ ≤ νp (f )νq (g).
E
This shows that, by choosing α0 = νp (f )−1 and β0 = νq (g)−1 , we may assume
without loss of generality that νp (f ) = 1 = νq (g).
6. Lp SPACES 87
νq (f ∗ )q = ∫ ∣f ∗ ∣q
E
q
= ∫ (νp (f )1−p ∣f ∣p−1 )
E
= νp (f )(1−p)q ∫ ∣f ∣(p−1)q
E
−p p
= νp (f ) ∫ ∣f ∣
E
= νp (f )−p νp (f )p
= 1.
Finally,
ν1 (f f ∗ ) = ∫ ∣f f ∗ ∣ = ∫ f f ∗ = νp (f )1−p νp (f )p = νp (f ).
E E
◻
Thus
∣f + g∣p ≤ (∣f ∣ + ∣g∣)p ≤ 2p (∣f ∣p + ∣g∣p ).
It follows that
νp (f + g)p = ∫ ∣f + g∣p ≤ 2p (νp (f )p + νp (g)p ) < ∞,
E
and therefore f + g ∈ Lp (E, K). Set h = f + g, and let h∗ denote the conjugate
function of h. Then h∗ ∈ Lq (E, K), νq (h) = 1 and ν1 (h ⋅ h∗ ) = νp (h).
From this and Hölder’s Inequality we deduce that
νp (f + g) = νp (h)
= ν1 (h ⋅ h∗ )
= ν1 ((f + g)h∗ )
≤ ν1 (f ⋅ h∗ ) + ν1 (g ⋅ h∗ )
≤ νp (f )νq (h∗ ) + νp (g)νq (h∗ )
= νp (f ) + νp (g).
◻
6.12. Corollary. Let E ∈ M(R), and 1 < p < ∞. Then Lp (E, K) is a vector
space, and νp (⋅) defines a seminorm on Lp (E, K).
Proof. Clearly Lp (E, K) ⊆ L(E, K) by definition, and since the latter is a vector
space, it suffices that we prove that Lp (E, K) ≠ ∅, and that f, g ∈ Lp (E, K) and
k ∈ K implies that kf + g ∈ Lp (E, K).
Let ζ ∶ E → K be the zero function ζ(x) = 0 for all x ∈ E. Clearly ζ ∈ Lp (E, K)
and hence Lp (E, K) ≠ ∅.
If f ∈ Lp (E, K) and k ∈ K, then kf is measurable by Proposition 4.8 and
p p p
∫ ∣kf ∣ = ∣k∣ ∫ ∣f ∣ < ∞,
E E
so that kf ∈ Lp (E, K). If g ∈ Lp (E, K) as well, then by Minkowski’s inequality,
kf + g ∈ Lp (E, K). Thus Lp (E, K) is a vector space.
With f, g ∈ Lp (E, K) and k ∈ K as above,
● νp (f ) = ∫E ∣f ∣p ≥ 0, and νp (ζ) = ∫E ζ p = ∫E 0 = 0.
●
1/p 1/p
νp (kf ) = (∫ ∣kf ∣p ) = (∣k∣p ∫ ∣f ∣p ) = ∣k∣νp (f ).
E E
● νp (f + g) ≤ νp (f ) + νp (g) by Minkowski’s inequality.
Thus νp (⋅) defines a seminorm on Lp (E, K). As was the case with L1 (E, K), if
∅ ≠ F ⊆ E is a subset of Lebsegue measure zero, then 0 ≠ χF is measurable and
1/p
νp (χF ) = (∫ ∣χF ∣p ) = ∫ χF = m(F ∩ E) = 0.
E E
Thus νp (⋅) is not a norm.
◻
6. Lp SPACES 89
Once again, we shall appeal to Proposition 6.3 to obtain a normed linear space
from a semi-normed linear space.
6.13. Definition. Let E ∈ M(R) and 1 < p < ∞. We define the Lp -space
Lp (E, K) ∶= Lp (E, K)/Np (E, K),
where Np (E, K) = {f ∈ Lp (E, K) ∶ νp (f ) = 0}. The Lp -norm on Lp (E, K) is the
norm defined by
∥ ⋅ ∥p ∶ Lp (E, K) → R
.
[f ] ↦ νp (f )
Suppose that ([fn ])∞ n=1 is a sequence in Lp (E, K) and suppose furthermore that
∞
γ ∶= ∑n=1 ∥[fn ]∥p < ∞. Our strategy will be to use the representatives fn , n ≥ 1 of
the elements [fn ] ∈ Lp (E, K) to produce a measurable function h ∈ Lp (E, K) such
that h(x) = ∑∞ n=1 fn (x) almost everywhere on E. Then we shall show that in fact,
[h] = ∑∞n=1 n in the sense of norm convergence in Lp (E, K).
[f ]
Step One. First we must show that ∑∞ n=1 fn (x) converges almost everywhere on E.
To that end, for each N ≥ 1, we set gN ∶= ∑Nn=1 ∣fn ∣, and observe that gN ∈ Lp (E, R)
by Corollary 6.12. Observe also that
0 ≤ g1 ≤ g2 ≤ g3 ≤ ⋯.
p
Set g∞ = supN ≥1 gN , so that 0 ≤ g∞ ∈ L(E, [0, ∞]). Then g∞ = supN ≥1 gNp , and so by
the Monotone Convergence Theorem 5.16,
p
∫ g∞ = Nlim ∫ gNp
E →∞ E
and so g ∈ Lp (E, K) – i.e., [g] ∈ Lp (E, K) – and ∥[g]∥p ≤ γ. Next, note that for
x ∈ H,
∞ ∞
∣ ∑ fn (x)∣ ≤ ∑ ∣fn (x)∣ = g∞ (x) = g(x) < ∞,
n=1 n=1
and so ∑∞
n=1 fn (x) ∈ K exists, by the completeness of K.
6. Lp SPACES 91
N
∣hN (x)∣ ≤ ∑ ∣fn (x)∣ ≤ g(x),
n=1
while for x ∈ B, ∣hN (x)∣ = 0 = g(x). Thus ∣hN ∣ ≤ g on E, and as a trivial consequence,
∣hN ∣p ≤ g p on E. From this we conclude that for each N ≥ 1,
p p p
∫ ∣hN ∣ ≤ ∫ g ≤ γ .
E E
p p p
∫ ∣h∣ ≤ ∫ g ≤ γ < ∞.
E E
Step Three. Recall that we are trying to prove that ∑∞ n=1 [fn ] converges in
Lp (E, K). We chose a specific set of representatives, namely the fn ’s themselves,
and we showed that there exists an element h ∈ Lp (E, K) such that almost every-
where on E, namely on H ⊆ E, h = ∑∞ n=1 fn as a pointwise limit of functions. If we
can show that
N
lim ∥[h] − [hN ]∥p = lim ∥[h] − ∑ [fn ]∥p = 0,
N →∞ N →∞ n=1
6.19. Examples.
(a) Let E = R and f = χQ be the characteristic function of the rationals. For
any γ > 0, m{x ∈ R ∶ ∣χQ (x)∣ > γ} ≤ mQ = 0, and so ν∞ (χQ ) = 0.
Clearly there was nothing special about Q in this example, other than
the fact that this set has Lebesgue measure zero.
(b) Suppose that a < b ∈ R and that f ∈ C([a, b], K).
We claim that f ∈ L∞ ([a, b], K) and that ν∞ (f ) = ∥f ∥sup . Indeed,
every continuous function is measurable, so f ∈ L([a, b], K). Also, [a, b]
being a compact set, there exists x0 ∈ [a, b] such that ∣f (x0 )∣ = ∥f ∥sup . By
6. Lp SPACES 93
continuity of f , given ε > 0, there exists δ > 0 such that x ∈ [a, b] and
∣x − x0 ∣ < δ implies ∣f (x)∣ > ∣f (x0 )∣ − ε. Whether or not x0 is one of the
endpoints of the interval, it follows that there exist c < d ∈ (a, b) such that
x ∈ [c, d] implies that ∣f (x)∣ > ∣f (x0 )∣ − ε. But then
m{x ∈ [a, b] ∶ ∣f (x)∣ > ∣f (x0 )∣ − ε} ≥ d − c > 0,
and so ν∞ (f ) ≥ ∥f ∥sup − ε. Since ε > 0 was arbitrary, ν∞ (f ) ≥ ∥f ∥sup .
If γ > ∥f ∥sup , then m{x ∈ E ∶ ∣f (x)∣ > γ} = m∅ = 0, and thus ν∞ (f ) ≤ γ.
Hence ν∞ (f ) ≤ ∥f ∥sup . Combining these inequalities,
ν∞ (f ) = ∥f ∥sup .
In particular, f ∈ L∞ ([a, b], K).
For each N ≥ 1, set hN = ∑N n=1 fN ∈ L∞ (E, K). Since f is the pointwise limit of the
measurable functions hN , we see that f ∈ L(E, K). Moreover, the above estimate
shows that ν∞ (f ) ≤ γ < ∞, and thus f ∈ L∞ (E, K).
Let ε > 0 and choose N0 > 0 such that ∑∞ n=N0 +1 ∥[fn ]∥∞ < ε. For all N > N0 , we
have that
∥[f ] − [hN ]∥∞ = ν∞ (f − hN ).
But for x ∈ E,
∞
∣f (x) − hN (x)∣ = ∣ ∑ fn (x)∣
n=N +1
∞
≤ ∑ ∣fn (x)∣
n=N +1
∞
≤ ∑ ∥[fn ]∥∞
n=N +1
< ε.
Thus ∥[f ] − [hN ]∥∞ < ε, N ≥ N0 .
This shows that [f ] = limN →∞ [hN ] = limN →∞ ∑N
n=1 [fn ], as required.
◻
6.25. Recall that if E ∈ M(R), 1 < p < ∞, f ∈ Lp (E, K) and g ∈ Lq (E, K), where
q is the Lebesgue conjugate of p, then Hölder’s Inequality (Theorem 6.10) states
that f g ∈ L1 (E, K) and that
ν1 (f g) ≤ νp (f ) νq (g).
Let us obtain a version of this inequality for p = 1, which will prove especially
useful when we examine Fourier series in later chapters.
Let E ∈ M(R) and suppose that mE < ∞. Let [f ] ∈ L∞ (E, K). Then f ∈
L(E, K) and ∣f (x)∣ ≤ ∥[f ]∥∞ almost everywhere on E. For 1 ≤ p < ∞,
p p p
∫ ∣f ∣ ≤ ∫ ∥[f ]∥∞ = ∥[f ]∥∞ mE < ∞,
E E
proving that [f ] ∈ Lp (E, K), with
∥[f ]∥p ≤ ∥[f ]∥∞ (mE)1/p .
Thus L∞ (E, K) ⊆ Lp (E, K), 1 ≤ p < ∞ when mE < ∞.
Next, suppose that 1 ≤ p < r < ∞, and that [g] ∈ Lr (E, K). Again, g ∈ L(E, K)
and
p r p/r r r r
∫ ∣g∣ = ∫ (∣g∣ ) ≤ ∫ max(1, ∣g∣ ) ≤ ∫ (1 + ∣g∣ ) = mE + ∥[g]∥r < ∞.
E E E E
6. Lp SPACES 97
Our next major goal is to prove that the space [C([a, b], K)] is dense in each
of the spaces Lp ([a, b), K), 1 ≤ p < ∞. We shall accomplish this through a series of
approximations.
6.31. Before proving our next result, we introduce a bit of notation. Given
E ∈ M(R) and 1 ≤ p ≤ ∞, we set
Simpp (E, K) = Simp(E, K) ∩ Lp (E, K).
Since Simp(E, K) and Lp (E, K) are vector spaces over K, so is Simpp (E, K).
We leave it to the exercises for the reader to prove that if mE < ∞ or if p = ∞,
then Simpp (E, K) = Simp(E, K).
6.34. Lemma. Let (X, ∥⋅∥) be a Banach space, and suppose that B ⊆ X satisfies
span B = X. Suppose also that L ⊆ X is a linear manifold and that B ⊆ L. Then
X = L.
Proof. The key is to observe that since L is a linear manifold, L is a closed subspace
of X. Indeed, since X is a vector space, we may apply the Subspace Test from first-
year linear algebra. That is, it suffices to prove that L ≠ ∅, and that if x, y ∈ L and
κ ∈ K, then kx + y ∈ L.
6.35. Proposition. Let a < b ∈ R. If 1 ≤ p < ∞, then [Step([a, b], K)] is dense
in (Lp ([a, b], K), ∥ ⋅ ∥p ).
Proof. Let B ∶= {[χH ] ∶ H ⊆ E measurable}. Then span B = [Simp([a, b], K)], and
by Proposition 6.32, this is dense in Lp ([a, b], K). By Lemma 6.34, it suffices to
show that every [χH ] can be approximated arbitrarily well in the ∥ ⋅ ∥p -norm by
elements of [Step([a, b], K)].
Let H ⊆ [a, b] be a measurable set and ε > 0. Recall from Theorem 3.14 that
we can find an open set G ⊆ R such that H ⊆ G, and m(G ∖ H) < 2ε . Write G
as a disjoint union of open intervals G = ∪∞ n=1 (an , bn ). (Note that each interval
is finite, since mH ≤ m[a, b] < ∞, and m(G ∖ H) < ∞), implying that m(G) =
m(H) + m(G ∖ H) < ∞.)
Thus m(G) = ∑∞ n
n=1 (bn − an ) < ∞. Set Gn = ∪k=1 (ak , bk ), n ≥ 1, and choose N ≥ 1
such that
∞
ε
m(G ∖ GN ) = ∑ (bn − an ) < .
n=N +1 2
Set ψ = χGN ∩[a,b] and observe that ψ ∈ Step([a, b], R).
Moreover,
⎧
⎪1 = ∣1 − 0∣ if x ∈ H ∖ GN
⎪
⎪
⎪
⎪
⎪1 = ∣0 − 1∣
⎪ if x ∈ (GN ∩ [a, b]) ∖ H
∣χH (x) − ψ(x)∣ = ⎨
⎪
⎪
⎪0 = ∣0 − 0∣ if x ∈/ (GN ∪ H)
⎪
⎪
⎪
⎩0 = ∣1 − 1∣
⎪ if x ∈ (GN ∩ H).
It follows that
νp (χH − ψ) = ∫ ∣χH − ψ∣p
E
= ∫ ∣χH − ψ∣
E
= m(H ∖ GN ) + m((GN ∩ [a, b]) ∖ H)
≤ m(H ∖ GN ) + m(GN ∖ H)
≤ m(G ∖ GN ) + m(G ∖ H)
ε ε
< +
2 2
= ε.
It follows that ∥[χH ] − [ψ]∥p < ε, thus showing that [Step([a, b], R)] is dense in
(Lp ([a, b], R), ∥ ⋅ ∥p ).
◻
Observe that Lemma 6.34 greatly simplified the proof; instead of approximating
an arbitrary element in Lp ([a, b], K), or even an arbitrary element of [Simp(E, K)]
by (equivalence classes of) step functions, we reduced the problem to that of ap-
proximating characteristic functions of measurable sets. In the same way, in proving
that [C([a, b], K)] is dense in (Lp ([a, b], K), ∥⋅∥p ), 1 ≤ p < ∞, Lemma 6.34 will reduce
6. Lp SPACES 101
6.36. Theorem. Let a < b ∈ R. If 1 ≤ p < ∞, then [C([a, b], K)] is dense in
(Lp ([a, b], K), ∥ ⋅ ∥p ).
Proof. Let B ∶= {[χ[r,s] ∶ a ≤ r < s ≤ b}. Then span B = [Step([a, b], K)], which is
dense in Lp ([a, b], K) by Proposition 6.35 above.
We may therefore appeal once again to Lemma 6.34, which implies that we
need only show that every such [χ[r,s] ] can be approximated arbitrarily well in the
∥ ⋅ ∥p -norm by elements of [C([a, b], K)].
To begin, choose a ≤ r < s ≤ b, and to dispense with a technicality, choose M > 0
2
such that r + M < s. For n ≥ M , define
⎧
⎪0 if a ≤ x ≤ r
⎪
⎪
⎪
⎪
⎪
⎪n(x − r) if r < x ≤ r + n1
⎪
⎪
⎪
fn (x) = ⎨1 if r + n1 < x ≤ s − n1
⎪
⎪
⎪
⎪
⎪
⎪n(s − x) if s − n1 < x ≤ s
⎪
⎪
⎪
⎩0
⎪ if s < x ≤ b.
(If it not entirely clear from the outset why we picked such a sequence of func-
tions, the reader would be well-advised to graph them. How could you approximate
a step function by a continuous function in a simpler way? The choice of only
defining fn for n ≥ M is to ensure that fn ≤ χ[r,s] for all such n.)
Observe that fn is continuous for each n ≥ M , being piecewise linear. Also,
x ∈/ [r, r + n1 ] ∪ [s − n1 , s] implies that
∣fn (x) − χ[r,s] (x)∣ = 0,
and for all x ∈ [a, b], ∣fn (x) − χ[r,s] (x)∣ ≤ 1. It follows that for all n ≥ M ,
1
p
νp (fn − χ[r,s] ) = (∫ ∣fn (x) − χ[r,s] (x)∣p )
[a,b]
1
p
≤ (∫ 1)
[r,r+ n
1
]∪[s− n
1
,s]
1
2 p
=( ) .
n
Thus 1
2 p
0 ≤ lim ∥[fn ] − [χ[r,s] ]∥p ≤ lim ( ) = 0,
n→∞ n→∞ n
and hence
lim [fn ] = [χ[r,s] ]
n→∞
in (Lp ([a, b], R), ∥ ⋅ ∥p ), completing the proof.
◻
102 L.W. Marcoux Introduction to Lebesgue measure
6.37. This leads to the following interesting result. Recall first that a topological
space is said to be separable if it admits a countable dense subset.
Secondly, recall that if (X, d) is a separable metric space, δ > 0 is a positive real
number and {xλ ∶ λ ∈ Λ} ⊆ X satisfies d(xα , xβ ) ≥ δ for all α ≠ β ∈ Λ, then Λ is
countable. The proof is left to the exercises.
6.38. Corollary. Let a < b ∈ R.
(a) If 1 ≤ p < ∞, then (Lp ([a, b], K), ∥ ⋅ ∥p ) is separable.
(b) The space (L∞ ([a, b], K), ∥ ⋅ ∥∞ ) is not separable.
Proof.
(a) First fix 1 ≤ p < ∞.
Recall from Section 6.29 that L∞ ([a, b], K) ⊆ Lp ([a, b], K). Moreover,
the proof of this assertion showed that if [f ], [g] ∈ L∞ ([a, b], K), then
∥[f ] − [g]∥p = ∥[f − g]∥p ≤ ∥[f ] − [g]∥∞ (b − a)1/p .
Let ε > 0, and let [h] ∈ Lp ([a, b], K). We know from Theorem 6.36 that
[C([a, b], K)] is dense in Lp ([a, b], K) with respect to the p-norm.
Thus we can find g ∈ C([a, b], K) such that
ε
∥[h] − [g]∥p < .
3
Of course, by the Weierstraß Approximation Theorem and Example 6.19
above, we know that we can find a polynomial p(x) = p0 + p1 x + ⋯ + pm xm
such that
ε
∥[g] − [p]∥∞ = ∥g − p∥sup < .
3(b − a)1/p
By Exercise 6.9 below, we can find a polynomial q(x) = q0 + q1 x + ⋯ + qm xm
such that qk ∈ Q + iQ for all 0 ≤ k ≤ m and
ε
∥[p] − [q]∥∞ = ∥p − q∥sup < .
3(b − a)1/p
Finally,
∥[h] − [q]∥p ≤ ∥[h] − [g]∥p + ∥[g] − [p]∥p + ∥[p] − [q]∥p
≤ ∥[h] − [g]∥p + ∥[g] − [p]∥∞ (b − a)1/p + ∥[p] − [q]∥∞ (b − a)1/p
ε ε ε
≤ +[ 1/p
] (b − a)1/p + [ ] (b − a)1/p
3 3(b − a) 3(b − a)1/p
ε ε ε
= + +
3 3 3
= ε.
Thus the family (Q + iQ)[x] of all polynomials with (complex) rational
coefficients, when viewed as continuous functions on [a, b], has the property
that [(Q + iQ)[x]] is dense in (Lp ([a, b], K), ∥ ⋅ ∥p ). Since there are at most
countably many elements in (Q + iQ)[x] and therefore in [(Q + iQ)[x]], we
see that Lp ([a, b], K) is separable, 1 ≤ p < ∞.
6. Lp SPACES 103
Appendix to Section 6.
6.39. If the reader consults almost any other text on measure theory (and the
reader will absolutely fall in this author’s estimation if they do not), they will
almost assuredly observe that we have been incredibly pedantic in our approach
to these notes. Most texts will use the same notation, namely “f ”, to refer to
both a measurable function, as well as to its equivalence class in “Lp (E, K)”, where
E ∈ M(R) is a measurable set. It is left to the reader to keep track of when
they are dealing with a function, and when they are dealing with its image in
Lp (E, K). Statements such as “f (x) = x a.e. on [0, 1]” are used to hint that we
are talking about an equivalence class rather than a function. Truth be told (and
what’s the point of not telling the truth in a set of course notes?), we have always
felt ambivalent about this approach. Without having taken any courses in math
pedagogy, experience has taught us that people inevitably make the mistake of
treating an element of Lp (E, K) as a function – for example, referring to f (x), for
some x ∈ E, when this concept is no longer valid (for the equivalence class of f ).
For this reason, we have attempted to consistently denote functions by simple
letters, e.g. f and g, while denoting their equivalence classes using the bracket
notation [f ] and [g]. While this is more cumbersome, it has the advantage of
being more precise, and our hope is that for the person learning about measure
theory for the first time, it will help to keep the two concepts separate. Once
sufficient mathematical maturity is acquired (oh, maybe a month from now), the
reader should be able to consult other texts with sufficient sophistication to handle
any notation thrown at them.
6.40. Which brings us to another point. In order to deal with pointwise lim-
its of increasing functions for the Monotone Convergence Theorem 5.16, we in-
troduced the notion of extended real numbers, and then of measurable functions
f ∶ E → R (where again, E ∈ M(R) is a measurable set). In many textbooks deal-
ing with Lebesgue measure, the equivalence classes of functions in “Lp (E)” consist
of extended real-valued measurable functions. (We have restricted our attention to
equivalence classes of real-valued measurable functions.)
In fact, both approaches lead to essentially the same theory. For any 1 ≤ p ≤ ∞,
if f ∶ E → R is measurable and νp (f ) < ∞, then the set B ∶= {x ∈ E ∶ ∣f (x)∣ = ∞} has
measure zero, and so we can always find g ∶ E → R such that g = f a.e. on E; for
example, we can choose g = f ⋅ χE∖B . Thus [f ] = [g].
So why not start with the vector space Lp (E, R) and use Proposition 6.3 to
define the quotient space Lp (E, R)? Well, perhaps the best answer we can give is
that Lp (E, R) does not define a vector space over R! Suppose
⎧ ⎧
⎪
⎪∞ if x ∈ Q ⎪
⎪−∞ if x = 0
f (x) = ⎨ and g(x) = ⎨ .
⎪
⎪0 if x ∈/ Q ⎪
⎪0 if x ≠ 0
⎩ ⎩
What should f + g(0) be? We haven’t defined ∞ + (−∞), and we won’t. Rats.
6. Lp SPACES 105
So then where do these other authors get off defining “Lp (E)” in terms of equiv-
alence classes of extended real-valued functions? Well, for one thing, their construc-
tion of “Lp (E)” as often as not does not pass through Proposition 6.3. Instead,
they define two extended real-valued functions f and g to be equivalent if f = g a.e.
on E. While the set of extended real-valued functions does not form a vector space,
the equivalence classes defined do! In fact, up to isometric isomorphism, they form
exactly the same Lp -spaces we have defined. We like the seminorm-to-norm-via-
quotient-spaces approach we have used, in part because we do not have to wave our
hands at any point of our construction. There you go.
6.41. One obvious reason for studying Hölder’s Inequality is that it was required
to obtain Minkowski’s Inequality, which we needed to prove that each Lp (E, K) was
a linear space. The construction of the conjugate function f ∗ ∈ Lq (E, K) from
f ∈ Lp (E, K) might at first glance seem rather arcane, and the eager novice might
wonder why we would be interested in such a thing.
The answer lies in part in Functional Analysis - which is the study of normed
linear spaces and the continuous linear maps between them. Given a Banach space
(X, ∥ ⋅ ∥), one defines the dual space of X to be
X∗ ∶= B(X, K) = {x∗ ∶ X → K ∶ x∗ is linear and continuous}.
Elements of the dual space are called continuous linear functionals on X.
As we have seen in the Assignments, linear maps between normed linear spaces
are continuous if and only if they are bounded. The dual space carries a great deal of
information about the space X itself, and perhaps the most famous and important
theorem from Functional Analysis is the Hahn-Banach Theorem. This is not
actually a single result, but rather two classes of results, all referred to by that same
name. It is usually left up to the reader to recognize which version of the Hahn-
Banach Theorem is being invoked in any given application. We invite the reader to
consult the (free!) reference [3] for more details.
One of the important consequences of the Hahn-Banach Theorem is that the dual
space X∗ of X has sufficiently many functionals to separate points of X, meaning
that if x1 ≠ x2 ∈ X, then there exists x∗ ∈ X∗ such that x∗ (x1 ) ≠ x∗ (x2 ). In fact,
one can do better – given y ∈ X, one can find a linear functional y ∗ ∈ X∗ such that
∥y ∗ ∥X∗ = 1 and y ∗ (y) = ∥y∥. (Applying this to y = x1 − x2 above shows that the
corresponding y ∗ will separate x1 and x2 .)
What Hölder’s Inequality (Theorem 6.15) tells us is that with 1 < p < ∞ and
1 1
p + q = 1, given [g] ∈ Lq (E, K), we may define a linear functional
Φ[g] ∶ Lp (E, K) → K
[f ] ↦ ∫E f g
and that this linear map will be bounded with norm at most ∥[g]∥q . In the Assign-
ments, we verify that every linear functional φ ∈ Lp (E, K) is of the form φ = Φ[g]
for some [g] ∈ Lq (E, K), and that in fact ∥φ∥ = ∥[g]∥q . In other words, we are iden-
tifying Lq (E, K) with the dual space (Lp (E, K))∗ in an isometrically isomorphic
manner.
106 L.W. Marcoux Introduction to Lebesgue measure
The second half of Theorem 6.15 is just an example verifying the Corollary
to the Hahn-Banach Theorem that we mentioned above, namely: the functional
Φ[f ∗ ] ∈ (Lp (E, K))∗ is the norm-one functional which sends [f ] to ∥[f ]∥p .
For more general measure spaces (Z, µ) – which we have not discussed at all in
these notes – we still have that the dual of Lp (Z, µ) may be isometrically isomorphi-
cally identified with Lq (Z, µ) when 1 < p < ∞, but problems start to arise with the
dual spaces of L1 (Z, µ) and L∞ (Z, µ). Without getting into the details at all (the
interested reader may find them online), in the case where the space (Z, µ) is de-
composable, which includes the case where (Z, µ) is a σ-finite measure space, we
do have that (L1 (Z, µ))∗ = L∞ (Z, µ). In particular, this applies to the case of ℓ1 (I),
where I is any set equipped with counting measure. Thus we can always identify
(ℓ1 (I))∗ with ℓ∞ (I), which is a happy, happy situation. The dual of L∞ (Z, µ) tends
to be a real can of worms. For example, the dual of ℓ∞ (N) may be identified with
the so-called regular Borel measures on the Stone-Čech compactification βN
of the natural numbers. It’s big – very, very big.
6.42. The following diagram illustrates the relationship between the Lp -spaces
when the underlying measure space is a bounded interval. We use 1 to denote
constant function 1(x) = x for all x ∈ [a, b].
Inclusions of Lp -spaces
L1 ([a, b], K)
L∞ ([a, b], K)
[K1]
6. Lp SPACES 107
Exercise 6.2.
Let E ∈ M(R) satisfy mE < ∞ and φ ∈ Simp(E, K). Prove that for all 1 ≤ p ≤ ∞,
φ ∈ Lp (E, K).
Exercise 6.3.
Let E ∈ M(R) and 1 < p < ∞. Suppose that [f ] ∈ Lp (E, K) and [g] ∈ Lq (E, K).
Then
[f ] ⋅ [g] ∶= [f g]
is well-defined, and [f g] ∈ L1 (E, K).
Exercise 6.5.
Paragraph 6.29 and the subsequent paragraphs establish a set of relationships
between Lp -spaces for various values of p ∈ [1, ∞], as well as the spaces of (equiva-
lence classes) of step functions and simple functions, in the case where the underlying
measurable domain E ∈ M(R) has finite measure.
Determine what relations hold between these spaces in the case where mE = ∞.
Exercise 6.6.
Let E ∈ M(R), and 1 ≤ p ≤ ∞. Let D ⊆ Lp (E, K). Prove that the following are
equivalent:
(a) The set [D] ∶= {[f ] ∶ f ∈ D} is dense in Lp (E, K).
(b) For each g ∈ Lp (E, K) and ε > 0, there exists f ∈ D such that
νp (f − g) < ε.
108 L.W. Marcoux Introduction to Lebesgue measure
Exercise 6.7.
Recall that given E ∈ M(R) and 1 ≤ p ≤ ∞, we defined
Simpp (E, K) = Simp(E, K) ∩ Lp (E, K).
Prove that if mE < ∞ or if p = ∞, then Simpp (E, K) = Simp(E, K).
Exercise 6.8.
Let (X, d) be a metric space. Suppose that there exists an uncountable set
{xλ ∶ λ ∈ Λ} in X and a positive real number δ > 0 such that d(xα , xβ ) ≥ δ for all
α ≠ β ∈ Λ. Prove that (X, d) is not separable.
Exercise 6.9.
Let a < b ∈ R and let p(x) = p0 + p1 x + ⋯ + pm xm be a polynomial of degree
m in (C([a, b], R), ∥ ⋅ ∥sup ). Prove that given any ε > 0, there exists a polynomial
q(x) = q0 + q1 x + ⋯ + qm xm in (C([a, b], R), ∥ ⋅ ∥sup ) such that qk ∈ Q, 0 ≤ k ≤ m, and
∥p − q∥sup < ε.
Conclude that given a polynomial r(x) = r0 + r1 x + ⋯ + rm xm of degree m in
(C([a, b], C), ∥ ⋅ ∥sup ) and ε > 0, we can find a polynomial s(x) = s0 + s1 x + ⋯ + sm xm
with sk ∈ Q + iQ, 0 ≤ k ≤ m, such that
∥r − s∥sup < ε.
Exercise 6.10.
Recall that if A and B are sets, we define the symmetric difference of A and
B to be
A∆B ∶= (A ∪ B) ∖ (B ∩ A).
Let a < b be real numbers and suppose that a ≤ r1 < s1 ≤ b and a ≤ r2 < s2 ≤ b.
Suppose furthermore that either r1 ≠ r2 or that s1 ≠ s2 . Prove that the symmetric
difference
[r1 , s1 ] ∆ [r2 , s2 ]
contains a non-degenerate interval [u, v] ⊆ [a, b]. (By non-degenerate, we simply
mean that u < v.)
Exercise 6.11.
Let E ∈ M(R). Prove that the map:
Ω ∶ (C(E, K), ∥ ⋅ ∥sup ) → (L∞ (E, K), ∥ ⋅ ∥∞ )
f ↦ [f ]
is a linear isometry, and deduce that [C(E, K)] ∶= {[f ] ∶ f ∈ C(E, K)} is a closed
subspace Banach space of L∞ (E, K) which is (isometrically) isomorphic to C(E, K).
In other words, as Banach spaces, we can identify C(E, K) with its image
[C(E, K)] in L∞ (E, K).
7. HILBERT SPACES 109
7. Hilbert spaces
7.1. We have seen in the previous section that if E ∈ M(R) and 1 ≤ p ≤ ∞, then
(Lp (E, R), ∥ ⋅ ∥p ) is a Banach space. The case where p = 2 is very special and merits
individual attention.
Let us recall the following definition.
7.3. Theorem. The Cauchy-Schwarz Inequality. Suppose that (H, ⟨⋅, ⋅⟩)
is an inner product space over K. Then
∣⟨x, y⟩∣ ≤ ⟨x, x⟩1/2 ⟨y, y⟩1/2
for all x, y ∈ H.
Proof. Let x, y ∈ H. If ⟨x, y⟩ = 0, then there is nothing to prove.
Suppose therefore that ⟨x, y⟩ ≠ 0. For any κ ∈ K,
0 ≤ ⟨x − κy, x − κy⟩
= ⟨x, x⟩ − κ⟨y, x⟩ − κ⟨x, y⟩ + ∣κ∣2 ⟨y, y⟩.
⟨x, y⟩
Setting κ = yields
⟨y, y⟩
∣⟨x, y⟩∣2 ∣⟨x, y⟩∣2 ∣⟨x, y⟩∣2
0 ≤ ⟨x, x⟩ − − + ,
⟨y, y⟩ ⟨y, y⟩ ⟨y, y⟩
which is equivalent to
∣⟨x, y⟩∣2 ≤ ⟨x, x⟩ ⟨y, y⟩,
as required.
◻
110 L.W. Marcoux Introduction to Lebesgue measure
7.4. Proposition. Let (H, ⟨⋅, ⋅⟩) be an inner product space. Then the map
∥x∥ ∶= ⟨x, x⟩1/2 , x∈H
defines a norm on H, called the norm induced by the inner product.
Proof. We shall prove that it defines a norm, and leave it to the reader to prove
that this is what it is called.
Let x, y ∈ H and κ ∈ K.
(a) Clearly ∥x∥ ≥ 0, as ⟨x, x⟩ ≥ 0.
(b) Note that ∥x∥ = 0 if and only if ∥x∥2 = ⟨x, x⟩ = 0, which happens if and only
if x = 0.
(c) ∥κx∥2 = ⟨κx, κx⟩ = ∣κ∣2 ⟨x, x⟩ = ∣κ∣2 ∥x∥2 , and thus
∥κx∥ = ∣κ∣ ∥x∥.
(d) From the Cauchy-Schwarz Inequality,
∥x + y∥2 = ⟨x + y, x + y⟩
= ⟨x, x⟩ + ⟨x, y⟩ + ⟨y, x⟩ + ⟨y, y⟩
≤ ∥x∥2 + ∣⟨x, y⟩∣ + ∣⟨y, x⟩∣ + ∥y∥2
= ∥x∥2 + 2∥x∥ ∥y∥ + ∥y∥2
= (∥x∥ + ∥y∥)2 .
Thus ∥x + y∥ ≤ ∥x∥ + ∥y∥.
This completes the proof.
◻
It follows that every inner product space (H, ⟨⋅, ⋅⟩) is also a normed linear space
(H, ∥⋅∥), using the norm induced by the inner product. Unless we explicitly mention
a different norm for H (which is highly unlikely), we shall always assume that this
is the norm to which we are referring. Also - let us not forget that every normed
linear space is also a metric space, using the metric induced by the norm.
As such, every inner product space is a metric space, under the metric induced
by the norm induced by the inner product.
Although a Hilbert space is technically an order pair consisting of a vector space
H and an inner product ⟨⋅, ⋅⟩, we shall typically speak informally of the Hilbert space
H.
7.6. Examples.
(a) Let N ≥ 1 be an integer. Consider x = (xn )N N N
n=1 and y = (yn )n=1 ∈ H ∶= C .
The map
N
⟨x, y⟩ ∶= ∑ xn yn
n=1
7. HILBERT SPACES 111
∫ f1 g1 = ∫ f g ∈ K,
E E
and so the map is well-defined, as claimed.
Let [f ], [g], [h] ∈ L2 (E) and κ ∈ K. Then
(a)
⟨[f ], [f ]⟩ = ∫ ∣f ∣2 ≥ 0,
E
and equality occurs if and only if f = 0 a.e. on E, i.e. if and only if [f ] = 0.
1
(This also shows that ⟨[f ], [f ]⟩ 2 = ∥[f ]∥2 .)
(b)
⟨κ[f ] + [g], [h]⟩ = ∫ (κf + g)h
E
= κ ∫ f h + ∫ gh
E E
= κ⟨[f ], [h]⟩ + ⟨[g], [h]⟩.
112 L.W. Marcoux Introduction to Lebesgue measure
(c)
⟨[f ], [g]⟩ = ∫ f g = ∫ f g = ⟨[g], [f ]⟩.
E E
Thus ⟨⋅, ⋅⟩ is an inner product (this is the standard inner product on L2 (E, K)),
and the norm induced by this inner product is the ∥ ⋅ ∥2 -norm. The last statement
is clear.
◻
7.8. Recall that a subset E of an inner product space (H, ⟨⋅, ⋅⟩) is said to be or-
thogonal if x ≠ y in E implies that ⟨x, y⟩ = 0. Also, E is said to be an orthonormal
1
set if E is orthogonal and x ∈ E implies that ∥x∥ = 1 = ⟨x, x⟩ 2 .
7.10. Remarks.
(a) By Zorn’s Lemma, every orthonormal set in H can be extended to an onb
for H.
(b) If H is infinite-dimensional, then an onb for H is not a Hamel basis for H.
7.11. Examples.
(a) Let N ≥ 1 be an integer and consider H = CN , equipped with the standard
inner product ⟨⋅, ⋅⟩. For 1 ≤ n ≤ N , define en = (δn,k )N
k=1 , where δa,b denotes
the Kronecker delta function. Then {en }N n=1 is an onb for H.
(b) Let N ≥ 1 be an integer and ρk = k, 1 ≤ k ≤ N . Set en = ( √1 δn,k )N k=1 , where
k
δa,b denotes the Kronecker delta function. Then {en }N N
n=1 is an onb for C ,
equipped with the inner product from Example 7.6 (b), namely
N
⟨x, y⟩ρ ∶= ∑ ρn xn yn .
n=1
⟨[f ], [g]⟩ ∶= ∫ f g.
[0,2π]
Then [ξn ] ∈ L2 ([0, 2π], C) for all n ∈ Z. In the Assignments, we shall see
that {[ξn ]}n∈Z is an onb for L2 ([0, 2π], C).
7.16. Remarks.
(a) Given any non-empty subset S ⊆ H, let
S ⊥ ∶= {y ∈ H ∶ ⟨x, y⟩ = 0 for all x ∈ S}.
It is routine to show that S ⊥ is a norm-closed subspace of H. In particular,
⊥
(S ⊥ ) ⊇ span S,
the norm closure of the linear span of S.
(b) Recall from Linear Algebra that if V is a vector space and W is a (vector)
subspace of V, then there exists a (vector) subspace X ⊆ V such that
(i) W ∩ X = {0}, and
(ii) V = W + X ∶= {w + x ∶ w ∈ W, y ∈ X }.
We say that W is algebraically complemented by X . The existence of
such a X for each W says that every vector subspace of a vector space is
algebraically complemented. We shall write V = W +̇X to denote the fact
that X is an algebraic complement for W in V.
If X is a Banach space and Y is a closed subspace of X, we say that Y
is topologically complemented if there exists a closed subspace Z of X
such that Z is an algebraic complement to Y. The issue here is that both Y
and Z must be closed subspaces. It can be shown that the closed subspace
7. HILBERT SPACES 115
◻
Before stating our next result, we recall the following from your previous Analysis
course.
7.20. Lemma. Let (X, d) be a separable metric space. Let δ > 0 be a real
number and Y ⊆ X be a set with the property that y, z ∈ Y with y ≠ z implies that
d(y, z) ≥ δ. Then Y is countable.
Proof. The proof is left to the exercises.
◻
converges in H.
Proof. First note that if x, y ∈ E with x ≠ y, then
1/2 √
∥x − y∥ = ⟨x − y, x − y⟩1/2 = (∥x∥2 + ∥y∥2 ) = 2.
By Lemma 7.20 above, E must be countable. Thus we may write E = {en }∞
n=1 .
(The astute reader will note that with MN ∶= span{e1 , e2 , . . . , eN } and PN defined
as the orthogonal projection of H onto MN , we have that yN ∶= PN x, N ≥ 1.
It results from Bessel’s Inequality that we can find N0 > 0 so that
∞
2 2
∑ ∣⟨x, ek ⟩∣ < ε .
k=N0 +1
This shows that (yN )∞N =1 is a Cauchy sequence. Since H is complete, this Cauchy
sequence converges, i.e.
∞
∑ ⟨x, en ⟩en = lim yN ∈ H.
n=1 N →∞
◻
We also refer to the linear maps implementing the above isomorphism as unitary
operators. Note that
∥U x∥2 = ⟨U x, U x⟩ = ⟨x, x⟩ = ∥x∥2
for all x ∈ H1 , so that unitary operators are isometries. Moreover, the inverse map
U −1 ∶ H2 → H1 defined by U −1 (U x) ∶= x is also linear, and
⟨U −1 (U x)U −1 (U y)⟩ = ⟨x, y⟩ = ⟨U x, U y⟩,
so that U −1 is also a unitary operator.
Note that if L ⊆ H1 is a closed subspace, then L is complete, whence U L is also
complete and hence closed in H2 .
Appendix to Section 7.
7.25. As mentioned in the Appendix to Chapter 6, given E ∈ M(R) and 1 <
p < ∞, one may identify the dual space of Lp (E, R) with Lq (E, R) via a linear,
isometric isomorphism. Since the Lebesgue conjugate of p = 2 is q = 2, this gives
us an identification of the dual space of L2 (E, R) with itself, via a linear, isometric
isomorphism.
Once again, the case where p = 2 is special. In this case, we obtain a second
identification of the dual of a Hilbert H space with itself. Given y ∈ H, we define
the linear function Φy ∈ H∗ by setting
Φy (x) = ⟨x, y⟩, x ∈ H.
As we shall see in the Assignments, the map
ϱ ∶ H → H∗
y ↦ Φy
is a conjugate-linear isometric isomorphism. Here, conjugate-linear means that
ϱ(y + kz) = ϱ(y) + kϱ(z) for all y, z ∈ H and k ∈ K. Of course, when K = R, ϱ is linear.
There are reasons for preferring this identification of H∗ with H to the identi-
fication that we obtained in the Appendix to Chapter 6. Given Banach spaces X
and Y and an operator T ∈ B(X, Y), we may define an operator T ∗ ∈ B(Y∗ , X∗ ) via
T ∗ y ∗ (x) = y ∗ (T x) for all x ∈ X, y ∗ ∈ Y∗ . The identification of H with H∗ mentioned
in the paragraph above provides us with an involution on B(H), that is, a map
∗ ∶ B(H) → B(H) which satisfies
● (T ∗ )∗ = T for all T ∈ B(H);
● (T1 + kT2 )∗ = T1∗ + kT2∗ for all T1 , T2 ∈ B(H), k ∈ K;
● (T1 T2 )∗ = T2∗ T1∗ for all T1 , T2 ∈ B(H).
Moreover, one may check that ∥T ∗ ∥ = ∥T ∥ for all T ∈ B(H), and finally that
∥T ∗ T ∥ = ∥T ∥2 for all T ∈ B(H).
Thus B(H) becomes what is called an involutive Banach algebra using this
operation ∗, and the last equality is referred to as the C ∗ -equation. Involutive
Banach algebras whose involution satisfies the C ∗ -equation are called C ∗ -algebras.
The C ∗ -equation, as anodyne as it might at first appear, has untold consequences
for the structure and representation theory of the algebra. People have spent their
lives studying C ∗ -algebras. If you haven’t been moved to pity yet, your heart is
made of stone.
7.26. Theorem. The Riesz Representation Theorem. Let {0} =/ H be a
Hilbert space over K, and let φ ∈ H∗ . Then there exists a unique vector y ∈ H so
that
φ(x) = ⟨x, y⟩ for all x ∈ H.
Moreover, ∥φ∥ = ∥y∥.
Proof. Given a fixed y ∈ H, let us denote by βy the map βy (x) = ⟨x, y⟩. Our goal
is to show that H∗ = {βy ∶ y ∈ H}. First note that if y ∈ H, then βy (kx1 + x2 ) =
7. HILBERT SPACES 121
7.27. Remark. The fact that the map Θ defined in the proof the Riesz Rep-
resentation Theorem above induces an isometric, conjugate-linear automorphism of
H is worth remembering.
122 L.W. Marcoux Introduction to Lebesgue measure
Exercise 7.1. Let H = ℓ2 , equipped with standard inner product. For n ≥ 1, let
en = (δn,k )∞ ∞
k=1 , where δa,b is the Kronecker delta function. Prove that {en }n=1 is an
onb for H.
8. FOURIER ANALYSIS - AN INTRODUCTION 123
8.1. One of the main achievements of linear algebra is that – in particular when
dealing with finite-dimensional vector spaces – one is able to reduce a great many
questions about abstract linear maps to very concrete and computation-friendly
questions about matrices. As we know, to each linear map T from an n-dimensional
vector space V over a field F to a m-dimensional vector space W over F we may
associate an m × n matrix [T ] ∈ Mm,n (F) as follows: we let BV ∶= {v1 , v2 , . . . , vn } and
BW = {w1 , w2 , . . . , wm } be bases for V and W respectively. Given x ∈ V, we may
express x as a linear combination of the elements of BV in a unique way, say
x = α1 v1 + α2 v2 + ⋯ + αn vn ,
where αi ∈ F, 1 ≤ i ≤ n. We write [x]BV to denote the corresponding n-tuple (αk )nk=1 .
Recall that the map
ΓV ∶ V → Fn
x ↦ [x]BV ,
n
is an isomorphism of V and F .
Similarly, every y ∈ W corresponds to a unique m-tuple [y]BW = (βj )m m
j=1 ∈ F ,
m m
so that y = ∑j=1 βj wj , and the map ΓW ∶ W → F defined by ΓW (y) = [y]BW is an
isomorphism of vector spaces.
The matrix that we then associate to T is [T ] ∶= [tij ], where, for 1 ≤ j ≤ n,
(ti,j )m
i=1 = [T vj ]BW . Obviously the matrix for T depends upon our choice of bases
BV for V and BW for W.
While T acts upon the elements of V, the matrix [T ] acts upon the coordinates
of vectors x ∈ V, and returns the coordinates of the vector T x; that is,
[T ][x]BV = [T x]BW ,
where [x]BV (resp. [T x]BW ) represents the coordinates of x (resp. T x) with respect
to the basis BV for V (resp. BW for W.)
In practice, finding the coordinates of a vector in a vector space V with re-
spect to a basis for a given vector space often reduces to solving a system of linear
equations. The situation is greatly simplified, however, if the vector space H is a
finite-dimensional Hilbert space over the field K, and the basis E ∶= {e1 , e2 , . . . , en }
in question is an orthonormal basis.
In this case, the coordinates of x ∈ H are deduced from the formula
x = ⟨x, e1 ⟩e1 + ⟨x, e2 ⟩e2 + ⋯ + ⟨x, en ⟩en .
Thus, for T ∶ H → H linear, the matrix of T corresponding to the basis E (for both
the domain and the codomain) is simply [T ] = [tij ], where tij = ⟨T ej , ei ⟩, 1 ≤ i, j ≤ n.
124 L.W. Marcoux Introduction to Lebesgue measure
8.2. When the vector spaces V and W above are infinite-dimensional, it is still
possible to associate to each linear map T ∶ V → W a (generalized) matrix as above;
the problem now lies in the fact that the bases involved are obviously infinite, and
there is no need for them to be countable. For example, it can be shown that if
(X, ∥ ⋅ ∥) is an infinite-dimensional Banach space, then any (Hamel) basis for X must
be uncountable. This greatly mitigates the usefulness of the (generalized) matrix
representation of linear maps from X to X.
Analysis, however, distinguishes itself from Algebra in its ubiquitous recourse to
approximation. We may ask whether or not we can find a suitable subset B of X,
preferably countable, such that every x ∈ X can be approximated in norm by finite
linear combinations of elements of B.
In the case of separable Hilbert spaces, we have already seen that this is the
case. When H is an infinite-dimensional separable Hilbert space over K, we have
seen that H admits a countable, maximal orthonormal set E = {en }∞ n=1 (which we
have dubbed an “orthonormal basis” for H – or a “Hilbert space basis” for H), such
that x ∈ H implies that
∞ N
x = ∑ ⟨x, en ⟩en ∶= lim ∑ ⟨x, en ⟩en .
n=1 N →∞ n=1
While E is certainly linearly independent, it is never a Hamel basis (see Exercise 1).
We leave it as an exercise for the reader to prove that L2 (T, C) is a vector space
and that the function
ν2 ∶ L2 (T, C) → R
1/2
1 2
f ↦ ( 2π ∫[−π,π) ∣f ∣ )
defines a seminorm on L2 (T, C). Set N (T, C) ∶= {f ∈ L2 (T, C) ∶ ν2 (f ) = 0}. Arguing
as in Section 6.4, we see that [f ] = [g] in L2 (T, C) ∶= L2 (T, C)/N (T, C) if and only if
f = g a.e. on R, or equivalently f = g a.e. on [−π, π) (given that f, g are 2π-periodic
on R). By Proposition 6.3, we obtain a norm on L2 (T, C) by setting ∥[f ]∥2 ∶= ν2 (f ),
[f ] ∈ L2 (T, C).
Furthermore, the function
⟨⋅, ⋅⟩T ∶ L2 (T, C) × L2 (T, C) → C
1
([f ], [g]) ↦ 2π ∫[−π,π) f g
8. FOURIER ANALYSIS - AN INTRODUCTION 125
defines an inner product on L2 (T, C), and ∥ ⋅ ∥2 is precisely the norm induced by
this inner product. A fortiori, L2 (T, C) is complete with respect to this norm, and
is therefore a Hilbert space. Finally, we leave it to the reader to verify that with
ξn (θ) = einθ , θ ∈ R, n ∈ Z, the set {[ξn ]}n∈Z forms an onb for L2 (T, C).
[f ]
Given n ∈ Z, we shall refer to the complex number αn ∶= ⟨[f ], [ξn ]⟩T as the
nth -Fourier coefficient of [f ] relative to the onb ([ξn ])n∈Z .
As we shall see in the Assignments, the map
U ∶ L2 (T, C) ↦ ℓ2 (Z, C)
[f ]
[f ] ↦ (αn )n∈Z
defines a unitary operator from the Hilbert space L2 (T, C) to ℓ2 (Z, C). In particular,
[f ] [g]
therefore, it is injective. Thus, if [f ], [g] ∈ L2 (T, C) and αn = αn for all n ∈ Z, then
f = g a.e. on R. In other words, an element [f ] ∈ L2 (T, C) is entirely determined
by its Fourier coefficients. Moreover, given any sequence (βn )n∈Z ∈ ℓ2 (Z, C), there
[f ]
exists [f ] ∈ L2 (T, C) such that αn = βn , n ∈ Z.
Let [f ] ∈ L2 (T, C). For each 1 ≤ N ∈ N, set
N
∆N ([f ]) = ∑ αn[f ] [ξn ].
n=−N
th
We shall say that ∆N ([f ]) is the N partial sum of the Fourier series of [f ]. It
follows from Theorem 7.22 that
[f ] = lim ∆N ([f ]),
N →∞
where the convergence is relative to the ∥ ⋅ ∥2 -norm defined above.
This is an entirely satisfactory state of affairs, and demonstrates clearly just how
well-behaved Hilbert spaces are. Our next goal is to try to extend this theory to
the L1 -setting. Here, we will quickly discover that things are far more complicated,
and for that reason perhaps also that much more interesting!
8.4. In trying to extend this theory beyond the Hilbert space setting, we shall
once again require some notations and definitions. We define:
● Trig(T, C) ∶= span{ξn ∶ n ∈ Z} = {∑N n=−N αn ξn ∶ αn ∈ C, 1 ≤ N ∈ N};
● C(T, C) ∶= {f ∶ R → C ∶ f is continuous and 2π-periodic};
● Simp(T, C) ∶= {f ∶ R → C ∶ f ∣[−π,π) is a simple function and f is 2π-periodic};
● Step(T, C) ∶= {f ∶ R → C ∶ f ∣[−π,π) is a step function and f is 2π-periodic};
● for 1 ≤ p < ∞,
Lp (T, C) ∶= {f ∶ R → C ∶ f is measurable, 2π-periodic, and ∫ ∣f ∣p < ∞};
[−π,π)
● and for p = ∞,
L∞ (T, C) = {f ∶ R → C ∶ f is measurable, 2π-periodic, and essentially bounded}.
Observe that
Trig(T, C) ⊆ C(T, C) ⊆ Lp (T, C), 1 ≤ p ≤ ∞.
126 L.W. Marcoux Introduction to Lebesgue measure
As was the case with p = 2 above, for each 1 ≤ p < ∞, the set Lp (T, C) forms a
vector space over C, and the map
νp ∶ Lp (T, C) → R
1/p
1 p
f ↦ ( 2π ∫[−π,π) ∣f ∣ )
defines a seminorm on Lp (T, C).
Moreover, if p = ∞, we repeat the arguments of Chapter 6. That is, for f ∈
L∞ (T, C), we set
ν∞ (f ) ∶= ess sup(f ) ∶= inf{δ > 0 ∶ m{θ ∈ [−π, π) ∶ ∣f (θ)∣ > δ} = 0},
and verify that this defines a seminorm on L∞ (T, C).
Appealing to Proposition 6.3, for each 1 ≤ p ≤ ∞, we obtain a norm ∥ ⋅ ∥p
on Lp (T, C) ∶= Lp (T, C)/Np (T, C), where Np (T, C) ∶= {f ∈ Lp (T, C) ∶ νp (f ) = 0}.
Perhaps unsurprisingly, we find that [f ] = [g] ∈ Lp (T, C) if and only if f = g a.e. on
R (equivalently f = g a.e. on [−π, π), because of 2π-periodicity). The details are
left to the reader (see the Exercises below).
It is also a simple but important exercise (which we again leave to the reader)
to verify that for f ∈ C(T, C),
∥[f ]∥∞ = ∥f ∥sup ∶= sup{∣f (θ)∣ ∶ −π ≤ θ < π}.
Note that the supremum on the right hand side of this equation exists as a finite
number since f ∈ C(T, C) implies that f is continuous on R, and hence f is bounded
on [−π, π] ⊇ [−π, π).
a function f which lies in C(T, C) must satisfy f (−π) = limθ→π f (θ). Despite this,
[Step(T, C)] and [C(T, C)] are dense in Lp (T, C) for all such p, and this is also left
to the exercises. These facts will prove useful below.
Of course, this raises the question of why we even bother with Lp (T, C), given
that it is isomorphic to Lp ([−π, π), C). This would be a good time for the reader to
consult the Appendix, even if this is not something the reader typically does.
∑ f̂(n) ξn
n∈Z
8.7. Remark. Observe that if f, g ∈ L1 (T, C) and f = g a.e. on [−π, π), then
f̂(n) = ̂
g (n) for all n ∈ Z.
αn[f ] ∶= f̂(n), n ∈ Z,
then this is well-defined. We then define
[f ]
∑ αn [ξn ]
n∈Z
[f ]
8.8. In the case p = 2, we have more than once seen that (αn )n∈Z ∈ ℓ2 (Z, C).
While nothing so nice holds for [f ] ∈ L1 (T, C), the situation is not altogether hope-
less.
First note that ∣ξr (θ)∣ = 1 for all θ ∈ R and all r ∈ R. As such, for f ∈ L1 (T, C),
we have
1
∣f̂(r)∣ = ∣ ∫ f ξr ∣
2π [−π,π)
1
≤ ∫ ∣f ξr ∣
2π [−π,π)
1
= ∫ ∣f ∣
2π [−π,π)
= ν1 (f )
= ∥[f ]∥1 .
In particular,
Proof. The key to the proof of this result is to notice that it is really quite simple
to prove when f ∣[−π,π) is the characteristic function of an interval. But Lebesgue
integration is linear, and the span of these (2π-periodic extensions of) characteristic
functions of intervals is Step(T, C). So the result will hold for this class as well.
Given this, one simply appeals to the density of [Step(T, C)] in L1 (T, C) to obtain
the full result. As a rule, life is not great, but some parts of it really are.
● Suppose first that f0 is the characteristic function of an interval, say
f0 = χ[s,t] ,
and that the Lebesgue and Riemann integrals coincide (see Theorem 5.24),
1
f̂(r) = ∫ χ[s,t] ξr
2π [−π,π)
1
= ∫ e−irθ
2π [s,t]
1 t
−irθ
= ∫ e dθ
2π s
θ=t
1 e−irθ
= ]
2π −ir θ=s
1 e−irt − e−irs
= [ ].
2π −ir
From this it easily follows that
2 1
∣f̂(r)∣ ≤ = ,
2π∣r∣ π∣r∣
whence
lim f̂(r) = 0 = lim f̂(r).
r→∞ r→−∞
● Next, suppose that f ∈ Step(T, C). Let f0 ∶= f ∣[−π,π) , and write f0 =
∑Mk=1 βk χHk as a disjoint representation, where each Hk = [sk , tk ] is a subin-
terval of [−π, π).
The result is now a simple consequence of the argument above, com-
bined with the linearity of the Lebesgue integral, and is left to the reader.
● Finally, let [f ] ∈ L1 (T, C), ε > 0 be arbitrary, and choose g ∈ Step(T, C)
such that ∥[f ] − [g]∥1 < ε/2.
Then
1
f̂(r) = ∫ f ξr
2π [−π,π)
1 1
= ∫ (f − g)ξr + ∫ gξr
2π [−π,π) 2π [−π,π)
= f̂ − g(r) + ̂
g (r).
But as we have seen, ∣f̂ − g(r)∣ ≤ ν1 (f − g) = ∥[f − g]∥1 = ∥[f ] − [g]∥1 < ε/2
for all r ∈ R. Since g ∈ Step(T, C), from the previous case, we see that
we may choose N > 0 such that ∣r∣ ≥ N implies that ∣̂ g (r)∣ < ε/2, and thus
∣r∣ ≥ N implies that
ε ε
∣f̂(r)∣ ≤ ∣f̂
− g(r)∣ + ∣̂
g (r)∣ < + = ε.
2 2
This proves that limr→∞ f (r) = 0 = limr→−∞ f̂(r), completing the proof
̂
of the first statement of the Theorem. The second statement is an easy
consequence of the first, and is left to the reader.
◻
130 L.W. Marcoux Introduction to Lebesgue measure
[f ]
8.10. We began by recalling that [f ] ∈ L2 (T, C) if and only if (αn )n∈Z ∈
ℓ2 (Z, C). We then defined the Fourier coefficients of [f ] ∈ L1 (T, C) in exactly the
same way as for elements of L2 (T, C), and we have succeeded in showing that
(αn[f ] )n∈Z ∈ c0 (Z, C).
So far, however, we have not shown that this is an “if and only if” statement, and
one reason for this is that it is not. We shall see by Chapter 12 that the map
Λ ∶ (L1 (T, C), ∥ ⋅ ∥1 ) → (c0 (Z, C), ∥ ⋅ ∥∞ )
[f ]
[f ] ↦ (αn )n∈Z
is a continuous, injective linear map, but that it is not surjective. Indeed, linearity
is a simple result which is left to the exercises, while continuity of Λ is an easy
consequence of the estimate of Section 8.8. In analogy to the situation for L2 (T, C),
it is tempting to ask whether or not the range of Λ is ℓ1 (Z, C). In Chapter 12, we
shall discover that this is overly optimistic.
We are left with a number of questions. Let [f ] ∈ L1 (T, C).
[f ]
(a) Does the Fourier series ∑n∈Z αn [ξn ] converge, and if so, in which sense?
Pointwise (a.e.)? Uniformly? In the L1 -norm?
(b) If the Fourier series does converge in some sense, does it converge back to
f?
(c) Is [f ] completely determined by its Fourier series? That is, if [f ], [g] ∈
[f ] [g]
L1 (T, C) and αn = αn for all n ∈ Z, is [f ] = [g]? (We know that this was
true for [f ], [g] ∈ L2 (T, C).)
These are some of the questions we will consider in the following sections.
8. FOURIER ANALYSIS - AN INTRODUCTION 131
Appendix to Section 8.
8.11. So where does the notation L1 (T, C) come from, given that we are dealing
with 2π-periodic functions on R? The issue lies in the fact that we are really inter-
ested in studying functions on T ∶= {z ∈ C ∶ ∣z∣ = 1}, but that we have not yet defined
what we mean by a measure on that set. We are therefore identifying [−π, π) with
T via the bijective function ψ(θ) = eiθ . Thus, an alternative approach to this would
be to say that a subset E ⊆ T is measurable if and only if ψ −1 (E) ⊆ [−π, π) is
Lebesgue measurable. In order to “normalise” the measure of T (i.e. to make its
measure equal to 1), we simply divide Lebesgue measure on [−π, π) by 2π.
This still doesn’t quite explain why we are interested in 2π-periodic functions
on R, rather than just functions on [−π, π), though. Here is the “kicker ”. The
unit circle T ⊆ C has a very special property, namely, that it is a group. Given
θ0 ∈ T, we can “rotate” a function f ∶ T → C in the sense that we set g(θ) = f (θ ⋅ θ0 ).
Observe that rotation along T corresponds to translation (modulo 2π) of the interval
[−pi, π). The key is the irritating “modulo 2π” problem. If we don’t use modular
arithmetic, and if a function g is only defined on [−π, π), we can not “translate” it,
since the new function need no longer have as its domain: [−π, π). We get around
this by extending the domain of g to R and making g 2π-periodic. Then we may
translate g by any real number τs○ (g)(θ) ∶= g(θ − s), which has the effect that if we
set f (eiθ ) = g(θ), then g(θ −s) = f (eiθ ⋅e−is ). That is, translation of g under addition
corresponds to rotation of f under multiplication.
The last thing that we need to know is that such translations of functions will
play a crucial role in our study of Fourier series of elements of L1 (T, C). Aside from
being a Banach space, L1 (T, C) can be made into an algebra under convolution.
While our analysis will not take us as far as that particular result, we will still need
to delve into the theory of convolutions of continuous functions with functions in
L1 (T, C). This will provide us with a way of understanding how and why various
series associated to the Fourier series of an element [f ] ∈ L1 (T, C) converge or
diverge. Since convolutions are defined as averages under translation by the group
action, and since T is a group under multiplication and R is a group under addition,
our identification of (T, ⋅) with ([−π, π), +) (using modular arithmetic) is not an
unreasonable way of doing things.
We have been speaking in vague generalities. The next four sections will hope-
fully add meaning to the above statements.
132 L.W. Marcoux Introduction to Lebesgue measure
Exercise 8.1.
Let H be an infinite-dimensional Hilbert space over K, and let E be an onb for
H.
(a) Prove that E is linearly independent.
(b) Prove that E is not a Hamel basis for H.
Exercise 8.2.
Let 1 ≤ p ≤ ∞. Prove that [f ] = [g] in Lp (T, C) if and only if f = g a.e. on R.
Exercise 8.3.
(a) Prove that the map
Φ ∶ L2 ([−π, π], C) → L2 ([−π, π), C)
[f ] ↦ [f ∣[−π,π) ]
is an isomorphism of Hilbert spaces.
(b) Recall from Example[5.11] that if we set
1
ξn (θ) = √ einθ , θ ∈ [−π, π], n ∈ Z,
2π
then ([ξn ])n∈Z is an onb for L2 ([−π, π], C). Let ψn ∶= ξn ∣[−π,π) , n ∈ Z, and
prove that ([ψn ])n∈Z is an onb for L2 ([−π, π), C).
Exercise 8.4.
Let 1 ≤ p ≤ ∞. Prove that the map
Ξp ∶ Lp ([−π, π), C) → Lp (T, C)
[f ] ↦ [fˇ]
is a well-defined, isometric isomorphism of Banach spaces.
Exercise 8.5.
Let 1 ≤ p < ∞.
(a) Prove that [Simp(T, C)] is dense in (Lp (T, C), ∥ ⋅ ∥p ).
(b) Prove that [Step(T, C)] is dense in (Lp (T, C), ∥ ⋅ ∥p ).
(c) Prove that [C(T, C)] is dense in (Lp (T, C), ∥ ⋅ ∥p ).
(d) Prove that [Simp(T, C)] is dense in (L∞ (T, C), ∥ ⋅ ∥∞ ).
Exercise 8.6.
Let f ∈ C(T, C). Prove that
∥[f ]∥∞ = ∥f ∥sup ∶= sup{∣f (θ)∣ ∶ θ ∈ [−π, π)}.
8. FOURIER ANALYSIS - AN INTRODUCTION 133
Exercise 8.8.
Let f ∈ L1 (T, C). Prove (or disprove) that the function
f̂ ∶ R → C
1
r → 2π ∫[−π,π) f ξr
is continuous on R.
134 L.W. Marcoux Introduction to Lebesgue measure
9. Convolution
9.1. Recall that an algebra B is a vector space over a field F which also happens
to be a ring. A Banach algebra A is a Banach space over K which is simultane-
ously an algebra, and for which multiplication is jointly continuous by virtue of its
satisfying the inequality
∥ab∥ ≤ ∥a∥ ∥b∥
for all a, b ∈ A.
For example, (C(X, K), ∥ ⋅ ∥sup ) is a Banach algebra for each locally compact,
Hausdorff topological space X, as is Mn (K) ≃ B(Kn ) (for each n ≥ 1), when equipped
with the operator norm.
So far, we have seen that L1 (T, C) is a Banach space, but we have not investi-
gated any multiplicative structure on it. One can equip L1 (T, C) with an operation
∗ under which it becomes a Banach algebra. Indeed, given f, g ∈ L1 (T, C), we set
1
g ◇ f (θ) ∶= ∫ g(s)f (θ − s)dm(s),
2π [−π,π)
and we refer to this as the convolution of g and f . One shows that [f1 ] = [f2 ] and
[g1 ] = [g2 ] in L1 (T, C) yields that g1 ◇ f1 = g2 ◇ f2 almost everywhere, allowing one
to define
[g] ∗ [f ] ∶= [g ◇ f ]
for all [f ], [g] ∈ L1 (T, C).
It is not entirely clear, a priori, that g ◇ f (θ) ∈ C for any θ ∈ R, and it is
even less clear that g ◇ f ∈ L1 (T, C). Nevertheless, it is true. The proof of this,
however, requires a bit more measure theory than we have developed so far. The
key ingredient we are missing is Fubini’s Theorem, which is stated in the Appendix
to this Chapter.
What is easier to prove, however, and what we shall prove is that we can turn
L1 (T, C) (and consequently L1 (T, C)) into a left module over C(T, C) using con-
volution. That is, given g ∈ C(T, C) and f ∈ L1 (T, C), we shall set
1
g ◇ f (θ) ∶= ∫ g(s)f (θ − s)dm(s),
2π [−π,π)
and we shall prove that g ◇ f ∈ C(T, C) ⊆ L1 (T, C). Assuming this for the moment,
if f1 ∈ L1 (T, C) and f1 = f a.e. on R, then clearly g ◇ f (θ) = g ◇ f1 (θ) for all θ ∈ R,
whence g ◇ f = g ◇ f1 , and this allows us to define
g ∗ [f ] = [g ◇ f ], [f ] ∈ L1 (T, C).
9. CONVOLUTION 135
∫ h=∫ f.
[−π,π) [−π,π)
(c) Define φf,θ ∶ R → C by φf,θ (s) = f (θ − s). Then φf,θ ∈ L1 (T, C) and
ν1 (φf,θ ) = ν1 (f ).
That is,
1 1
∫ ∣f (θ − s)∣dm(s) = ∫ ∣f (t)∣dm(t).
2π [−π,π) 2π [−π,π)
Proof. This is an Assignment question.
◻
9.3. Definition. Let f ∈ L1 (T, C) and g ∈ C(T, C). We define the convolution
of f by g to be the function
g◇f ∶ R → C
1
θ ↦ 2π ∫[−π,π) g(s)f (θ − s)dm(s).
The alert reader (hopefully you) will have observed that there is a problem with
this definition. For one thing - how do we know that g ◇ f (θ) exists as a complex
number for each θ ∈ R? Let us resolve this issue immediately.
Fix θ ∈ R. Then
1
∣g ◇ f (θ)∣ = ∣∫ g(s)f (θ − s)dm(s)∣
2π [−π,π)
1
≤ ∫ ∣g(s)∣∣φf,θ (s)∣dm(s)
2π [−π,π)
≤ ∥g∥sup ν1 (φf,θ )
= ∥g∥sup ν1 (f ) < ∞.
Thus g ◇ f is indeed a complex-valued function.
136 L.W. Marcoux Introduction to Lebesgue measure
9.5. Remark. In light of Lemma 9.4, for f ∈ L1 (T, C) and g ∈ C(T, C), we shall
define the convolution of g by f to be
1
f ◇ g(θ) = ∫ g(θ − t)f (t)dm(t).
2π [−π,π)
In so doing, we have guaranteed that f ◇ g(θ) = g ◇ f (θ) for all θ ∈ R, and so
henceforth we shall simply refer to this function as the convolution of f and g.
9.6. Proposition. Let g ∈ C(T, C) and f ∈ L1 (T, C). Then g ◇ f ∈ C(T, C).
Proof. Note that since g is continuous on R and 2π-periodic, it is in fact uniformly
continuous on R. (See the exercises below.)
Let ε > 0 and choose δ > 0 such that ∣x − y∣ < δ implies that ∣g(x) − g(y)∣ < ε.
Let θ0 , θ ∈ R and suppose that ∣θ − θ0 ∣ < δ. Then, by noting that
9.7. Remark. Suppose that g ∈ C(T, C), f1 , f2 ∈ L1 (T, C) and that [f1 ] = [f2 ] ∈
L1 (T, C); i.e. that f1 = f2 a.e. on R. Then, since φg,θ f1 = φg,θ f2 a.e. on R (where
φg,θ is defined as in Lemma 9.2(c) above), we find that for each θ ∈ R,
g ◇ f1 (θ) = f1 ◇ g(θ)
1
= ∫ g(θ − t)f1 (t)dm(t)
2π [−π,π)
1
= ∫ g(θ − t)f2 (t)dm(t)
2π [−π,π)
= f2 ◇ g(θ)
= g ◇ f2 (θ).
9.9. Homogeneous Banach spaces. Let f ∈ L1 (T, C), and let s ∈ R. Con-
sider the function
τs○ (f ) ∶ R → C
θ ↦ f (θ − s).
One should think of τs○ as translating f by s. The superscript ○ above the τs is to
indicate that we are acting on functions. When acting on elements of L1 (T, C), we
shall drop this superscript.
As we shall see in the Assignments, the fact that M(R) is invariant under transla-
tion, that Lebesgue measure is translation-invariant, and that the set of 2π-periodic
functions is again invariant under translation implies that
τs○ (f ) ∈ L1 (T, C)
as well. Furthermore, if [f ] = [g] ∈ L1 (T, C), then [τs○ (f )] = [τs○ (g)], as is easily
verified. Thus we define the operation of translation by s on L1 (T, C) via
τs ([f ]) ∶= [τs○ (f )].
9.11. Example. Recall that [C(T, C)] ⊆ L∞ (T, C) is a subset of L1 (T, C) and
that it is clearly a linear manifold. Furthermore, for f ∈ C(T, C), it follows from
Example 6.19 (b) (see also Exercise 6.11) that
∥[f ]∥∞ = ∥f ∥sup ∶= sup{∣f (θ)∣ ∶ θ ∈ [−π, π)},
and that ([C(T, C)], ∥ ⋅ ∥∞ ) is a Banach space. We claim that it is in fact a homo-
geneous Banach space over T.
9. CONVOLUTION 139
9.12. Example. Let 1 ≤ p < ∞. We claim that (Lp (T, C), ∥ ⋅ ∥p ) is a homoge-
neous Banach space over T.
(a) Let f ∈ Lp (T, C) and let q denote the Lebesgue conjugate of p; i.e. p1 + 1q = 1.
Recall from Proposition 4.10 that there exists a measurable function
u ∶ R → T such that f = u ⋅ ∣f ∣. Note that u ∈ Lq (T, C); indeed, the fact
that f is 2π-periodic implies that u is, and u has already been seen to be
140 L.W. Marcoux Introduction to Lebesgue measure
measurable. Moreover,
1/q 1/q
1 1
∥[u]∥q = ( ∫ ∣u∣q ) = ( ∫ 1) = 1.
2π [−π,π) 2π [−π,π)
1
∥[f ]∥1 = ∫ ∣f ⋅ u∣ ≤ ∥[f ]∥p ∥[u]∥q ≤ ∥[f ]∥p .
2π [−π,π)
(b) Observe that [Trig(T, C)] ⊆ [C(T, C)] ⊆ Lp (T, C) ⊆ L1 (T, C).
(c) and (d) Let [f ] ∈ Lp (T, C) and s ∈ R. As noted in Section 9.9, τs [f ] ∈
L1 (T, C), and in particular τs○ (f ) is measurable.
Moreover, using Lemma 9.2:
1/p
1
∥τs [f ]∥p = ( ∫ ∣f (θ − s)∣p dm(s))
2π [−π,π)
1/p
1
=( ∫ ∣f (θ)∣p dm(θ))
2π [−π,π)
= ∥[f ]∥p < ∞.
∥τs [f ] − τs0 [f ]∥p ≤ ∥τs [f ] − τs [h]∥p + ∥τs [h] − τs0 [h]∥p + ∥τs0 [h] − τs0 [f ]∥p .
Let us estimate each of the terms on the right-hand side of this inequality.
Now (since translation is isometric on Lp (T, C) from above)
ε
∥τs [f ] − τs [h]∥p = ∥τs [f − h]∥p = ∥[f − h]∥p = ∥[f ] − [h]∥p < ,
3
and similarly,
ε
∥τs0 [f ] − τs0 [h]∥p < .
3
9. CONVOLUTION 141
Moreover,
1/p
1
∥τs [h] − τs0 [h]∥p = ( ∫ ∣τs○ (h) − τs○0 (h)∣p )
2π [−π,π)
1/p
1
≤( ∫ ∥τs [h] − τs0 [h]∥p∞ )
2π [−π,π)
1/p
1
≤( ∫ (ε/3)p )
2π [−π,π)
ε
= .
3
Substituting these estimates into the inequality above shows that for
∣s − s0 ∣ < δ,
ε ε ε
∥τs [f ] − τs0 [f ]∥p < + + = ε,
3 3 3
and thus translation is continuous on (Lp (T, C), ∥ ⋅ ∥p ).
Hence (Lp (T, C), ∥ ⋅ ∥p ) is a homogeneous Banach space over T when 1 ≤ p < ∞.
∥τs [f ] − τ0 [f ]∥∞ ≥ 1.
9.14. Let g ∈ C(T, C) and [f ] ∈ L1 (T, C). We have defined (see Definition 9.8)
the convolution of g and [f ] to be g ∗ [f ] ∶= [g ◇ f ], where
1
g ◇ f (θ) = ∫ g(s)f (θ − s) dm(s).
2π [−π,π)
That is, we define g ∗[f ] by first defining g ◇f pointwise, using Lebesgue integration.
Now consider that Example 9.12 shows that (L1 (T, C), ∥ ⋅ ∥1 ) is a homogeneous
Banach space over T. As such, the function
β ∶ R → L1 (T, C)
s ↦ g(s)τs [f ]
is continuous. By Theorem 1.14,
1 π 1 π
∫ β(s) ds = ∫ g(s)τs [f ] ds
2π −π 2π −π
exists in L1 (T, C), and is obtained as an ∥ ⋅ ∥1 -limit of Riemann sums (β, PN , PN∗ ) ∈
L1 (T, C) using partitions PN of [−2π, 2π] with corresponding choices PN∗ of test
values for PN .
If we fix g ∈ C(T, C), then we obtain a map
Γg ∶ L1 (T, C) → L1 (T, C)
1 π
[f ] ↦ 2π ∫−πg(s)τs [f ] ds.
We leave it to the reader to verify that Γg is linear.
The reader may have noticed that there is a striking resemblance between the
operators Cg and Γg . After all, τs [f ] = [τs○ (f )], where τs○ (f )(θ) = f (θ − s), θ ∈ R.
Our next goal is to show that in fact, Γg = Cg , so that for each [f ] ∈ L1 (T, C),
1 π
Cg [f ] = g ∗ [f ] = [g ◇ f ] = ∫ g(s)τs [f ] ds = Γg [f ].
2π −π
This is not an obvious nor a trivial result: the two constructions are entirely different.
Quite frankly, some of us – and you know who you are – don’t deserve this.
9.15. In Example 9.11, we showed that ([C(T, C)], ∥ ⋅ ∥∞ ) is a homogeneous Ba-
nach space over T. It follows that given f ∈ C(T, C), the map s ↦ τs [f ] is continuous,
or equivalently, the map s ↦ τs○ (f ) is continuous from (R, ∣ ⋅ ∣) to (C(T, C), ∥ ⋅ ∥sup ).
a Banach space and the map β ∶ R → C(T, C) defined by β(s) ∶= g(s)τs○ (f ) ∈ C(T, C)
is continuous. By Theorem 1.14, Γ○g (f ) ∈ C(T, C) exists as a ∥ ⋅ ∥sup -limit of Riemann
sums S(β, PN , PN∗ ) ∈ (C(T, C), ∥ ⋅ ∥sup ). In fact, as we saw there, we may suppose
without loss of generality that for each N ≥ 1, PN ∈ P([−π, π]) is a regular partition
2π
of [−π, π] into 2N subintervals of equal length N , and PN∗ = PN ∖ {−π}, so that PN∗
2
is a set of test values for PN .
Meanwhile, g◇f is the convolution of g and f defined pointwise through Lebesgue
integration, as in Definition 9.3.
Let us temporarily fix θ0 ∈ R and define a function γ(= γθ0 ) ∶ R → K via:
In the present case where g and f are both continuous, the map γ is easily seen to
also be continuous, and thus both bounded and Riemann integrable on [−π, π). By
Theorem 5.24 therefore,
1
g ◇ f (θ0 ) ∶= ∫ g(s)f (θ0 − s) dm(s)
2π [−π,π)
1
= ∫ γ(s) dm(s)
2π [−π,π)
1 π
= ∫ γ(s) ds
2π −π
1 π
= ∫ g(s)f (θ0 − s) ds,
2π −π
Finally,
2N
= ∣Γ○g (f )(θ0 ) − ∑ γ(pn )(pn − pn−1 )∣
n=1
= ∣Γg (f )(θ0 ) − S(γ, PN , PN∗ )∣.
○
Step One.
First we shall show that Γg [fm ] = g ∗ [fm ] for all m ≥ 1.
By Lemma 9.16, Γ○g (fm ) = g ◇ fm for all m ≥ 1. Thus
Step Two.
Next we show that g ∗ [f ] = limm→∞ g ∗ [fm ] in (L1 (T, C), ∥ ⋅ ∥1 ).
Now, for all m ≥ 1 and all θ ∈ R,
Step Three.
We now show that Γg [f ] = limm→∞ Γg [fm ] in (L1 (T, C), ∥ ⋅ ∥1 ).
Indeed,
1 π
∥Γg [f ] − Γg [fm ]∥1 = ∥ ∫ g(s)τs ([f − fm ]) ds∥
2π −π 1
1 π
≤ ∫ ∣g(s)∣ ∥τs ([f − fm ])∥1 ds
2π −π
1 π
≤ ∥g∥sup ∫ ∥[f ] − [fm ]∥1 ds
2π −π
= ∥g∥sup ∥[f ] − [fm ]∥1 .
Step Four.
Finally (!) we see that in (L1 (T, C), ∥ ⋅ ∥1 ) we have that
◻
146 L.W. Marcoux Introduction to Lebesgue measure
L (T,C)
Phrased another way, ΓB
g [f ] = Γg
1
[f ]. By Theorem 9.17,
L1 (T,C)
ΓB
g [f ] = Γg [f ] = g ∗ [f ] = Cg [f ].
(b) Recall that – in any homogeneous Banach space over T – we defined the
continuous map Φ[f ] ∶ R → B via Ψ[f ] (s) = τs [f ], and we have that
∥Ψ[f ] (s)∥B = ∥[f ]∥B for all s ∈ R.
Next, consider that
1 π
∥g ∗ [f ]∥B = ∥∫ g(s)τs [f ] ds∥
2π −π B
1 π
= ∥∫ g(s)Ψ[f ] (s) ds∥
2π −π B
1 π
≤ ∫ ∣g(s)∣ ∥Ψ[f ] (s)∥B ds
2π −π
1 π
= ∫ ∣g(s)∣ ∥[f ]∥B ds
2π −π
= ∥[f ]∥B ν1 (g).
◻
Next, using the fact that the Lebesgue and Riemann integrals of bounded, continu-
ous functions are equal, we find that
1
g ◇ f (0) = ∫ g(s)f (0 − s)dm(s)
2π [−π,π)
1 π
= ∫ g(s)f (−s)ds.
2π −π
Case One. Suppose that g is invertible; i.e. g(s) ≠ 0 for all s ∈ R. Choose
g(−s)
f (s) = , s ∈ R,
∣g(−s)∣
so that f is continuous and ∥[f ]∥∞ = ∥f ∥sup = 1.
Furthermore,
∥Cg ∥ ≥ ∥Cg [f ]∥∞
≥ ∣Cg (f )(0)∣
1 π
= ∣∫ g(s)f (−s)ds∣
2π −π
1 π
= ∫ ∣g(s)∣ds
2π −π
= ∥[g]∥1 .
Case Two. We shall make use of the following result, which was once a bonus
question for the Assignments, but has now been relegated to the Appendix (see
Theorem 9.40): let g ∈ C(T, C) and ε > 0. Then there exists h ∈ C(T, C) such that
h(s) ≠ 0 for all s ∈ R and ∥h − g∥sup < ε.
9.21. Theorem. Let g ∈ C(T, C), and let Cg ∶ L1 (T, C) → L1 (T, C) be the
convolution operator corresponding to g, so that Cg [f ] = g ∗ [f ]. Then ∥Cg ∥ =
ν1 (g) = ∥[g]∥1 .
Proof. By Theorem 9.18 (b), we see that for all [f ] ∈ L1 (T, C),
∥Cg [f ]∥1 = ∥g ∗ [f ]∥1 ≤ ν1 (g) ∥[f ]∥1 ,
whence ∥Cg ∥ ≤ ν1 (g).
There remains to show that ∥Cg ∥ ≥ ν1 (g), or equivalently, that ∥Cg ∥ ≥ ∥[g]∥1 .
To that end, we consider the functions fn = nπχ[−1/n,1/n] , n ≥ 1.
Clearly [fn ] ∈ L1 (T, C) and ∥[fn ]∥1 = 1 for all n.
Moreover, for all θ ∈ [−π, π), and hence for all θ ∈ R, we have (using Lemma 9.2)
1
(g ◇ fn )(θ) = ∫ g(s)fn (θ − s)dm(s)
2π [−π,π)
1
= ∫ g(θ − t)fn (t)dm(t)
2π [−π,π)
1
= (nπ) ∫ g(θ − t)dm(t)
2π [−1/n, 1/n]
n
= ∫ g(θ − t)dm(t).
2 [−1/n, 1/n]
Note also that
n
g(θ) = ∫ g(θ) dm(t),
2 [−1/n,1/n]
since g(θ) acts as a constant in this integral.
Thus
n
∣g ◇ fn (θ) − g(θ)∣ = ∣ ∫ g(θ − t) − g(θ) dm(t)∣
2 [−1/n, 1/n]
n
≤ ∫ ∣g(θ − t) − g(θ)∣dm(t)
2 [−1/n, 1/n]
Let ε > 0 and choose 1 ≤ N ∈ N such that ∣x−y∣ < N1 implies that ∣g(x)−g(y)∣ < ε.
This is possible because g is uniformly continuous on R (by virtue of being continuous
on [−π, π) and 2π-periodic).
For n > N , t ∈ [− n1 , n1 ] implies that ∣(θ − t) − θ∣ ≤ n1 < δ, and so ∣g(θ − t) − g(θ)∣ < ε.
But then for all n ≥ N and for all θ ∈ R,
n
∣g ◇ fn (θ) − g(θ)∣ ≤ ∫ ε dm(t) = ε.
2 [−1/n, 1/n]
Hence
1
ν1 (g ◇ fn − g) = ∫ ∣g ◇ fn (θ) − g(θ)∣ dm(θ)
2π [−π,π)
1
≤ ∫ ε dm(θ)
2π [−π,π)
= ε.
150 L.W. Marcoux Introduction to Lebesgue measure
In the next couple of chapters, we shall explore the connection between convo-
lution operators and convergence of Fourier series.
9. CONVOLUTION 151
Appendix to Section 9.
9.29. Example. Let A = M3 (C) and B = M7 (C). Then M ∶= M3×7 (C) becomes
an A − B bimodule, using usual matrix multiplication on the left by elements of A,
and usual matrix multiplication on the right by elements of B.
9.30. Example.
(a) Let A = ℓ∞ (N, C) and M = c0 (N, C). Then M is a A bimodule, where we
define (an )n ⋅ (mn )n ∶= (an mn )n = (mn )n ⋅ (an )n for all a = (an )n ∈ A and
(mn )n ∈ M.
(b) We can also set M = ℓ∞ (N, C) and A = c0 (N, C). Using the same operations
as above, ℓ∞ (N, C) becomes a c0 (N, C) bimodule.
9.31. Our approach to convolution has been to show that L1 (T, C) is a bimodule
over C(T, C), using convolution ∗ as our “multiplication” operation. It is relatively
straightforward to prove that conditions (i) and (ii) of Definition 9.34 hold. What is
left obvious (and what is left as an assignment exercise) is that condition (iii) holds
as well.
For this, it is worth observing that given g, h ∈ C(T, C) and f ∈ L1 (T, C), h ◇ f ∈
C(T, C), and thus g ◇ (h ◇ f ) may be realised as a Riemann integral, instead of
9. CONVOLUTION 153
The norm condition for the product of a and b above ensures that multiplication
is jointly continuous; i.e. the map
µ∶ A×A → A
(a, b) ↦ ab
is continuous.
Some of the above examples of algebras are actually examples of Banach alge-
bras.
9.34. We leave it as an exercise for the reader to prove that (C(K, C), ∥⋅∥sup ) is a
unital Banach algebra, whenever K is a compact, Hausdorff space. Here, functions
are added and multiplied pointwise, and (κf )(x) = κ(f (x)) for all κ ∈ C, f ∈
C(K, C).
9.35. As mentioned above, one can extend the notion of convolution to obtain
a product operation on L1 (T, C), under which the latter becomes a Banach algebra.
Alas, this algebra is non-unital: that is, it does not admit an identity element under
this operation.
In the study of Banach algebras, and more specifically of C ∗ -algebras of oper-
ators on a Hilbert space, one often comes across the situation where the algebra
is non-unital, but does admit the “next best thing” to a unit, namely a bounded
approximate unit.
154 L.W. Marcoux Introduction to Lebesgue measure
for all a ∈ A.
We say that (eλ )λ∈Λ is a bounded approximate unit if there exists M > 0
such that ∥eλ ∥ ≤ M for all λ ∈ Λ.
Typically, when the algebra in question is separable (i.e. A admits a countable
dense set), the approximate unit may be chosen to be a sequence, rather than a net.
We shall come across examples of these in Chapter 11, hidden under the guise
of summability kernels.
Then A is a Banach space using pointwise operations, and we can also give
it a multiplication operation in a similar way: that is, we set (wn )∞ ∞
n=1 ⋅ (zn )n=1 =
∞
(wn zn )n=1 . It is routine to verify that (A, ∥ ⋅ ∥∞ ) is a Banach algebra.
It is clearly non-unital. However, if we set en = (1, 1, . . . , 1, 0, 0, 0, . . .) for each
n ≥ 1 (where there are n terms which are equal to 1 and all remaining terms are
equal to 0), then (en )∞ n=1 is easily seen to be a bounded approximate identity for A.
9.38. There are also discrete versions of convolution; let g ∈ ℓ∞ (Z, C) and f ∈
ℓ1 (Z, C). We may define the discrete convolution of g and f as follows:
∞ ∞
g ∗ f (n) = ∑ g(m)f (n − m) = ∑ g(n − m)f (m).
m=−∞ m=−∞
For those who are interested, this operation turns (ℓ1 (Z, C), ∥⋅∥1 ) into a commu-
tative Banach algebra, using discrete convolution as the multiplication operation.
Is it unital?
9.39. Remark. There is a minor subtlety in the proof of Theorem 9.17 that
goes along the following lines:
Fix 1 ≤ m an integer. We defined Γ○g (fm ) as a Riemann integral in the Banach
space (C(T, C), ∥ ⋅ ∥sup ). Thus, for an appropriate sequence (PN )N of partitions of
[−π, π], we have that
Γ○g (fm ) = lim S(βm , PN , PN∗ ),
N →∞
where βm (s) = g(s)τs○ (fm ), s ∈ [−π, π).
But given any Riemann sum S(βm , Q, Q∗ ) of the form
M M
∗ ∗ ○
∑ βm (qk )(qk − qk−1 ) = ∑ g(qk )τqk∗ (fm )(qk − qk−1 )
k=1 k=1
9. CONVOLUTION 155
Since the map h → [h] from (C(T, C), ∥ ⋅ ∥sup ) to ([C(T, C)], ∥ ⋅ ∥∞ ) is a bijective
linear isometry, it follows that the image [Γ○g (fm )] of Γ○g (fm ) is
[Γ○g (fm )] = lim [S(βm , PN , PN∗ )],
N →∞
and that this convergence is with respect to the ∥ ⋅ ∥∞ norm. On the other hand,
looking how each [S(βm , PN , PN∗ )] is defined, we see that the latter limit is precisely
how we defined Γg ([fm ]) ∈ ([C(T, C), ∥ ⋅ ∥∞ ), and thus
[Γ○g (fm )] = Γg [fm ], m ≥ 1.
Next, each [S(βm , PN , PN∗ )] ∈ [C(T, C)] ⊆ L1 (T, C), and Γg [fm ] ∈ [C(T, C)] ⊆
L1 (T, C) as well. As remarked in the proof of Theorem 9.17, ∥[h]∥1 ≤ ∥[h]∥∞ for all
[h] ∈ [C(T, C)], and thus
0 ≤ lim ∥[Γ○g (fm )] − [S(βm , PN , PN∗ )]∥1 ≤ lim ∥[Γ○g (fm )] − [S(βm , PN , PN∗ )]∥∞ = 0,
N →∞ N →∞
proving that
Γg ([fm ]) = [Γ○g (fm )] = lim [S(βm , PN , PN∗ )],
N →∞
with the convergence taking place in (L1 (T, C), ∥ ⋅ ∥1 ).
The following result was required for Case Two of Theorem 9.20.
9.40. Theorem. If f ∈ C(T, C) and ε > 0, then there exists g ∈ C(T, C) so that
g(x) =/ 0 for all x ∈ [−π, π] and ∥f − g∥sup < ε.
Remark: Just in case you’ve forgotten - here’s a note to remind you that this is
very much a complex phenomenon. If we replace complex functions by real-valued
functions, the corresponding assertion is false.
A Banach algebra with the property that the invertible elements are dense is
said to have topological stable rank one. As you might imagine, there is a
notion of higher topological stable ranks. When X is a compact topological space,
the topological stable rank of C(X, K) is supposed to be a measure of the dimension
of X.
Proof. (I) Let ε > 0. By the Stone-Weierstrass Theorem, we can find a polynomial
p0 so that
sup ∣p0 (x) − f (x)∣ < ε/100.
x∈[a,b]
51
By adding a constant function κ1 to p0 with ∣κ∣ ≤ 100 ε if necessary, we can assume
that for p = p0 + κ1,
min(∣p(a)∣, ∣p(b)∣) ≥ ε/2.
156 L.W. Marcoux Introduction to Lebesgue measure
Observe that
1
∣p(a) − p(b)∣ = ∣p0 (a) − p0 (b)∣ ≤ ∣p0 (a) − f (a)∣ + ∣f (a) − f (b)∣ + ∣f (b) − p0 (b)∣ ≤ ε,
50
and that
52
sup ∣f (x) − p(x)∣ ≤ ε.
x∈[a,b] 100
(II)
Let p(x) = α0 (x−α1 )(x−α2 )⋯(x−αn ) for the appropriate choices of αi , 0 ≤ i ≤ n.
If we impose the norm ∥b∥∞ ∶= supx∈[a,b] ∣b(x)∣ for b ∈ C[x] ⊆ C([a, b], K), then it is
clear that the map
Φ∶ (Cn , ∥ ⋅ ∥∞ ) → (C[x], ∥ ⋅ ∥∞ )
(β1 , β2 , ..., βn ) ↦ (x − β1 )(x − β2 )⋯(x − βn )
is continuous. Thus, given η > 0 we can find β1 , β2 , ..., βn ∈ C ∖ R so that βk − αk is
sufficiently small, 1 ≤ k ≤ n, to guarantee that with q(x) = α0 (x−β1 )(x−β2 )⋯(x−βn ),
we have ∥p − q∥∞ < η. Since βk ∈/ R for all 1 ≤ k ≤ n, it follows that q has no real
roots - i.e. q(x) =/ 0 for all x ∈ [a, b].
Observe that ∥p − q∥∞ < η implies that
1
∣q(a) − q(b)∣ ≤ ∣q(a) − p(a)∣ + ∣p(a) − p(b)∣ + ∣p(b) − q(b)∣ ≤ η + ε + η.
50
1
In particular, if we choose η = 100 ε, then
1
∣q(a) − q(b)∣ ≤ ε.
25
(III)
Now we are in the situation where
● q ∶ [a, b] → C is a polynomial and q(x) =/ 0 for all x ∈ [a, b],
1
● ∣q(a)∣ ≥ ∣p(a)∣ − η ≥ 12 ε − 100 49
ε = 100 49
ε, and similarly ∣q(b)∣ ≥ 100 ε, and
●
1
∣q(a) − q(b)∣ ≤ ε.
25
Since q is continuous at b, we can find δ > 0 so that b − δ ≤ x ≤ b implies that
1
∣q(x) − q(b)∣ < ε.
25
Let r ∶ [a, b] → C be the continuous function on [a, b] defined by
● r(x) = 0 if x ∈ [a, b − δ],
x − (b − δ)
● r(x) = (q(a) − q(b)) , x ∈ [b − δ, b].
δ
9. CONVOLUTION 157
1
Observe that ∣r(x)∣ ≤ ∣q(a) − q(b)∣ ≤ 25 ε for all x ∈ [b − δ, b], and hence for all
x ∈ [a, b]. Let g(x) = q(x) + r(x), x ∈ [a, b]. Clearly g is continuous since each of q
and r are continuous, and for x ∈ [b − δ, b] we have
1 1 41
∣g(x)∣ = ∣q(x) + r(x)∣ ≥ ∣q(x)∣ − ∣r(x)∣ ≥ (∣q(b)∣ − ε) − ε ≥ ε.
25 25 100
In particular, g(x) =/ 0 for all x ∈ [b − δ, b]. Of course, g(x) = q(x) =/ 0 for all
x ∈ [a, b − δ], and thus g(x) =/ 0 for all x ∈ [a, b].
Also,
g(a) = q(a) = q(b) + r(b) = g(b).
1
Finally, as we have seen above, ∣r(x)∣ ≤ ∣q(a) − q(b)∣ ≤ 25 ε for all x ∈ [a, b], so
sup ∣f (x) − g(x)∣ ≤ sup ∣f (x) − p(x)∣ + sup ∣p(x) − q(x)∣ + sup ∣q(x) − g(x)∣
x∈[a,b] x∈[a,b] x∈[a,b] x∈[a,b]
52 1 1
≤ ε+ ε+ ε
100 100 25
< ε.
This completes the proof.
◻
∫ (∫ f (x, y)dµY (y)) dµX (x) = ∫ (∫ f (x, y)dµX (x)) dµY (y),
X Y Y X
and these coincide with
∫ f (x, y)dµX×Y (x, y).
X×Y
158 L.W. Marcoux Introduction to Lebesgue measure
Exercise 9.3. Consider the function χ[0,π) ∈ L∞ ([−π, π)), and let f = χ
̂ [0,π) ∈
L∞ (T, C) be its 2π-periodic extension, as defined in Paragraph 8.4.
If −π < s < 0, prove that
∥τs [f ] − τ0 [f ]∥∞ = 1.
Exercise 9.5. Let g ∈ C(T, C). Prove that the convolution operator Cg ∶ L1 (T, C) →
L1 (T, C) defined by Cg [f ] = g ∗ [f ] is linear.
Yogi Berra
10.1. We began our discussion of Fourier series by noting that if [f ] ∈ L2 (T, C),
then the sequence (∆N ([f ]))∞ N =1 of partial sums of the Fourier series of [f ] converges
in the ∥ ⋅ ∥2 -norm to [f ] (see Paragraph 8.3). Our goal was to see to what extent we
could extend these results to elements [f ] ∈ L1 (T, C). Somewhere along the way,
we seem to have been distracted by the concept of convolution. Let us show that
all roads lead to the Dirichlet kernel, which we now define.
N
DN = ∑ ξn .
n=−N
We mention in passing that this use of the word “kernel ” has nothing to do with
the null space of any linear map. It is just another example of the overuse of certain
terminology in mathematics.
N N
∆○N (f ) = ∑ αn[f ] ξn = ∑ f̂(n)ξn .
n=−N n=−N
It is clear that ∆○N (f ) ∈ C(T, C), being a finite linear combination of {ξn }N
n=−N .
[f ] [g]
If f = g a.e. on R, then αn = αn for all n ∈ Z, and thus ∆N (f ) = ∆○N (g) for
○
Thus ∆N ([f ]) is the N th -partial sum of the Fourier series of [f ]. Note that when
[f ] ∈ L2 (T, C), this definition coincides with our previous definition.
160 L.W. Marcoux Introduction to Lebesgue measure
Proof.
(a) For all θ ∈ R and n ≥ 1, ξ−n (θ) + ξn (θ) = e−inθ + einθ = 2 cos(nθ). Thus
N N
DN (θ) = ∑ ξn (θ) = 1 + 2 ∑ cos(nθ) ∈ R.
n=−N n=1
2
If 0 ≠ θ ∈ [−π, π), then
(−2i) sin ((N + 12 )θ) sin((N + 12 )θ)
DN (θ) = = .
ρ(θ) sin( 12 θ)
Meanwhile,
N
DN (0) = 1 + 2 ∑ cos(0) = 2N + 1.
n=1
(d) Since DN is an even, continuous function, and since ∣ sin(x)∣ ≤ ∣x∣, 0 ≤ x ≤ π,
1
ν1 (DN ) = ∫ ∣DN ∣
2π [−π,π)
1 π
= ∫ ∣DN (θ)∣ dθ
2π −π
1 π
= ∫ ∣DN (θ)∣ dθ
π 0
1 π sin((N + 1 )θ)
2
= ∫ ∣ 1
∣ dθ
π 0 sin( 2 θ)
1 π ∣ sin((N + 1 )θ)∣
2
≥ ∫
π 0 ∣ 12 θ∣
2 π ∣ sin((N + 1 )θ)∣
2
= ∫ .
π 0 ∣θ∣
162 L.W. Marcoux Introduction to Lebesgue measure
2 N π ∣ sin λ∣
≥ ∫ dλ
π 0 ∣λ∣
2 N nπ ∣ sin λ∣
= ∑∫ dλ
π n=1 (n−1)π ∣λ∣
2 N nπ ∣ sin λ∣
≥ ∑∫ dλ
π n=1 (n−1)π nπ
2 N 1 nπ
≥ 2 ∑ ∫ ∣ sin λ∣ dλ
π n=1 n (n−1)π
But on any interval of the form [(n − 1)π, nπ], the sine function does not
change sign (i.e. it is either always non-positive on such an interval, or
always non-negative), and so an easy calculation shows that
nπ nπ
∫ ∣ sin λ∣ dλ = ∣∫ sin λ dλ∣ = 2.
(n−1)π (n−1)π
From this we see that
4 N 1
ν1 (DN ) ≥ ∑ ,
π 2 n=1 n
as required.
◻
We invite the reader to skip to the Appendix to see the graphs of D2 , D5 and
D10 .
The next result follows immediately from Theorems 9.18 and 9.20, together with
Theorem 10.4 (d) and the fact that the harmonic series ∑∞ 1
n=1 n diverges.
10.5. Corollary. For each N ≥ 1, let DN denote the Dirichlet kernel of order
N.
(a) If CDN ∈ B(([C(T, C)], ∥ ⋅ ∥∞ )) is the convolution operator corresponding to
DN , N ≥ 1, then
lim ∥CDN ∥ = ∞.
N →∞
(b) If CDN ∈ B((L1 (T, C), ∥ ⋅ ∥1 )) is the convolution operator corresponding to
DN , N ≥ 1, then
lim ∥CDN ∥ = ∞.
N →∞
10. THE DIRICHLET KERNEL 163
10.10. Examples.
(a) It follows from the Baire Category Theorem that (R, d) is of the second
category, where d represents the standard metric d(x, y) = ∣x − y∣, x, y ∈ R.
(b) Let δ denote the discrete metric on Q, so that
⎧
⎪1 if p ≠ q
⎪
δ(p, q) = ⎨
⎪
⎪0 if p = q.
⎩
Writing Q = {qn }∞ n=1 , which we may do since Q is denumerable, we find
that for each n ≥ 1, Fn ∶= {qn } is a closed set, and clearly Q = ∪∞
n=1 Fn .
However, Fn is not nowhere dense, since as well as being closed, Fn is
open, and qn ∈ Fn = int(Fn ) ≠ ∅.
This is just as well, since (Q, δ) is a complete metric space, and hence
of the second category by the Baire Category Theorem. In other words,
we knew that Q was not a countable union of closed, nowhere dense sets
in (Q, δ).
164 L.W. Marcoux Introduction to Lebesgue measure
(c) The sets Fn = {qn } from (b) are closed and nowhere dense in (R, d), where
d is the standard metric from (a). Since Q = ∪∞
n=1 Fn , we see that Q is is of
the first category in (R, d).
10.11. Remark. An alternate form of the Baire Category Theorem says that if
(X, d) is a complete metric space, and if (Gn )∞
n=1 is a countable collection of dense,
open sets in X, then
∩∞
n=1 Gn ≠ ∅.
We shall not require this below.
The second result from real analysis which we shall recall is the following.
10.14. Corollary. Let (X, ∥⋅∥X ) and (Y, ∥⋅∥Y ) be Banach spaces and let (Tn )∞
n=1
be an unbounded sequence in B(X, Y), i.e. supn≥1 ∥Tn ∥ = ∞.
Let H = {x ∈ X ∶ supn≥1 ∥Tn x∥ < ∞}. Then H is of the first category in X, and
J ∶= X ∖ H is of the second category.
Proof. Note that 0 ∈ H, so that H ≠ ∅.
If H were of the second category, then – by the Banach-Steinhaus Theorem for
Operators, Theorem 10.13 above – {Tn }∞ n=1 would be bounded, a contradiction.
Thus H is of the first category. Let J = X ∖ H. By definition, for all x ∈ J,
sup ∥Tn x∥Y = ∞.
n≥1
We claim that J is of the second category.
Indeed, suppose otherwise. Then we could choose sequences (Kn )∞ ∞
n=1 and (Ln )n=1
of closed, nowhere dense sets in X such that
H ⊆ ∪∞
n=1 Kn and J ⊆ ∪∞
n=1 Ln .
But then
X = H ∪ J = ∪∞ ∞
n=1 Kn ⋃ ∪n=1 Ln
must be of the first category, contradicting the Baire Category Theorem, as X is a
complete metric space.
◻
166 L.W. Marcoux Introduction to Lebesgue measure
10.15. It is perhaps worth noting that this result is much, much better than it
might appear on the surface. The statement that supn≥1 ∥Tn ∥ = ∞ is the statement
that for each n ≥ 1, there exists xn ∈ X with ∥xn ∥X = 1 such that limn→∞ ∥Tn xn ∥Y =
∞. A priori, it is not clear that there should exist any x ∈ X such that
lim ∥Tn x∥Y = ∞.
n→∞
The above Corollary not only says that such a vector x ∈ X exists; it asserts that
this is true for a very large set of x’s, in the sense that the set H of x’s for which it
fails is a set of the first category in X.
We are finally prepared to answer the question of whether or not the partial
sums of the Fourier series of an element [f ] of L1 (T, C) necessarily converge to [f ]
in the ∥ ⋅ ∥1 -norm. As we shall now see, an easy application of Corollary 10.14 shows
that this almost never happens. (Here we use “almost never” informally, to refer
to the notion of sets of the first category, and not in the sense of sets of Lebesgue
measure zero!). What is more, essentially the same argument shows that the partial
sums of the Fourier series of an element [f ] of [C(T, C)] rarely converge in the ∥ ⋅ ∥∞
norm. These are dark times indeed for the sequence of partial sums of a Fourier
series.
Then K1 is a set of the first category in (L1 (T, C), ∥⋅∥1 ), whose complement
L1 (T, C) ∖ K1 is a set of the second category in L1 (T, C).
Proof. The proofs of both of these parts are almost identical. We shall prove (b),
and leave the proof of (a) as an exercise.
Recall that ∆N ([f ]) = DN ∗ [f ] = CDN ([f ]), N ≥ 1, where CDN ∈ B(L1 (T, C))
is the convolution operator corresponding to DN , as described in Theorem 9.20.
Furthermore, by Corollary 10.5,
lim ∥CDN ∥ = ∞.
N →∞
Of course, if [h] ∈ L1 (T, C) and [h] = limN →∞ ∆N ([h]) = limN →∞ CDN ([h]),
then (CDN ([h]))∞N =1 is bounded in L1 (T, C), and thus it is clear that K1 ⊆ H1 ,
where
H1 ∶= {[f ] ∈ L1 (T, C) ∶ sup ∥CDN ([f ])∥1 < ∞}.
N ≥1
10. THE DIRICHLET KERNEL 167
By Corollary 10.14, H1 is a set of the first category in L1 (T, C), and J1 ∶= L1 (T, C) ∖
H1 is a set of the second category in L1 (T, C).
In particular, for any [f ] ∈ J1 , we have that the partial sums (∆N ([f ]))∞N =1 fail
to converge to [f ], as the sequence is not even bounded.
◻
168 L.W. Marcoux Introduction to Lebesgue measure
Exercise 10.1.
Prove that the Cantor set is nowhere dense in [0, 1], where [0, 1] is equipped
with the standard metric d(x, y) = ∣x − y∣ inherited from R.
Exercise 10.2.
Fill in the details of the proof of Theorem 10.4(c) by proving that with ρ(θ) ∶=
(−2i) sin( 12 θ),
1
ρ(θ)DN (θ) = e−i(N + 2 )θ − ei(N + 2 )θ = (−2i) sin ((N + )θ) .
1 1
Exercise 10.3.
Fill in the details of the proof of Theorem 10.16 (a), namely: let
K∞ ∶= {[f ] ∈ [C(T, C)] ∶ [f ] = lim ∆N ([f ]) in ([C(T, C)], ∥ ⋅ ∥∞ )}.
N →∞
Prove that K∞ is a set of the first category in ([C(T, C)], ∥ ⋅ ∥∞ ), whose complement
[C(T, C)] ∖ K∞ is a set of the second category.
11. THE FÉJER KERNEL 171
11.1. We have seen in the last Chapter that if [f ] ∈ L1 (T, C), and if ∆N ([f ]) =
[f ]
∑N
n=−N αn [ξn ], then (∆N ([f ]))∞N =1 almost never converges to [f ].
Not all is lost. In this Chapter we shall replace partial sums (∆N ([f ]))∞ N =1 of
the Fourier series of [f ] by weighted partial sums which will converge to [f ].
The next Proposition is routine, and its proof is left to the exercises.
11.3. Proposition. Suppose that X is a Banach space and (xn )∞ n=0 is a se-
quence in X. Let (σN )∞
N =1 denote the sequence of Cesàro means of (x n ) ∞
n=0 .
If x = limn→∞ xn exists, then x = limN →∞ σN .
11.4. Definition. Let f ∈ L1 (T, C). The N th -Cesàro sum of the Fourier
series of f is the N th -Cesàro mean of the sequence (∆○n (f ))∞
n=0 . Thus
○ 1
σN (f ) = (D0 ◇ f + D1 ◇ f + ⋯ + DN −1 ◇ f )
N
= FN ◇ f,
1
σN [f ] ∶= (D0 ∗ [f ] + D1 ∗ [f ] + ⋯ + DN −1 ∗ [f ])
N
= FN ∗ [f ]
= [FN ◇ f ]
○
= [σN (f )].
172 L.W. Marcoux Introduction to Lebesgue measure
11.5. Remark. We remark that the fact that Dn ∈ C(T, C) for all n ≥ 0 implies
○
that FN ∈ C(T, C) for all N ≥ 1. By Proposition 9.6, it follows that σN (f ) ∈ C(T, C) ⊆
L1 (T, C) for all f ∈ L1 (T, C).
Furthermore, for all θ ∈ R,
○ 1
σN (f )(θ) = ∫ FN (s)f (θ − s) dm(s)
2π [−π,π)
1
= ∫ FN (θ − s)f (s) dm(s)
2π [−π,π)
As seen in Theorem 9.18, the fact that FN ∈ C(T, C) also implies that for every
homogeneous Banach space B and [f ] ∈ B, we have
σN [f ] = FN ∗ [f ] ∈ B.
As a special case of this phenomenon, σN [f ] = FN ∗ [f ] ∈ Lp (T, C) for all [f ] ∈
Lp (T, C).
1 N −1
FN = ∑ Dn
N n=0
1 N −1 n
= ∑ ( ∑ ξk )
N n=0 k=−n
1
= (ξ−N +1 + 2ξ−N +2 + ⋯ + (N − 1)ξ−1 + N ξ0
N
+(N − 1)ξ1 + ⋯ + 2ξN −2 + ξN −1 ) .
Let ρ = (2 − (ξ−1 + ξ1 ))N . As was the case with the Dirichlet kernel, we
observe that the product ρ ⋅ FN involves a telescoping sum, and that
That is,
1 1 − cos (N θ)
FN (θ) =
N 1 − cos θ
2
i θ N
−i θN
1 (e 2 − e 2 )
=
N (ei θ2 − e−i θ2 )2
N 2
1 ⎛ sin( 2 θ) ⎞
= .
N ⎝ sin( 21 θ) ⎠
In particular, FN ≥ 0.
π
(c) Keeping in mind that 0 ≠ n implies that ∫−π ξn (θ) dθ = 0, we get:
1 π
ν1 (FN ) = ∫ ∣FN (θ)∣ dθ
2π −π
1 π
= ∫ FN (θ) dθ
2π −π
1 1 π N −1 n
= ∫ ( ∑ ( ∑ ξk (θ))) dθ
2π N −π n=0 k=−n
1 N −1 n 1 π
= ∑ ∑ ∫ ξk (θ) dθ
N n=0 k=−n 2π −π
1 N −1 1 π
= ∑ ∫ ξ0 (θ) dθ
N n=0 2π −π
1 N −1
= ∑1
N n=0
= 1.
174 L.W. Marcoux Introduction to Lebesgue measure
11.9. Examples.
(a) For each 1 ≤ n ∈ N, consider the piecewise linear function
kn● ∶ [−π, π) → R
⎧
⎪
⎪ 0 if θ ∈ [−π, −1 1
n ] ∪ [ n , π)
⎪
⎪
θ ↦ ⎨n + n2 θ if θ ∈ ( −1 n , 0]
⎪
⎪
⎪ 2 1
⎩n − n θ if θ ∈ (0, n ).
⎪
For 1 ≤ n ∈ N, let kn be the 2π-periodic function on R whose restriction to
the interval [−π, π) coincides with kn● .
Then (kn )∞n=1 is a positive summability kernel. The details are left to
the reader.
(b) For each 1 ≤ n ∈ N, consider the piecewise linear function
rn● ∶ [−π, π) → R
⎧
⎪
⎪ 0 if θ ∈ [−π, 0] ∪ [ n2 , π)
⎪
⎪
θ ↦ ⎨n2 θ if θ ∈ (0, n1 ]
⎪
⎪
⎪ 2 1 1 2
⎩n − n (θ − n ) if θ ∈ ( n , n ).
⎪
For 1 ≤ n ∈ N, let rn be the 2π-periodic function on R whose restriction to
the interval [−π, π) coincides with rn● .
Then (rn )∞ n=1 is a positive summability kernel. The details are left to
the reader.
and so [f ] = limn→∞ kn ∗ [f ] in B.
Proof. The result is trivial if [f ] = 0. Let 0 ≠ [f ] ∈ B. Recall that
(a) the function
Ψ[f ] ∶ R → B
s ↦ τs [f ]
is continuous, and that
(b) τs is isometric for all s ∈ R. In particular, ∥τs [f ]∥B = ∥[f ]∥B for all s ∈ R.
That τ0 [f ] = [f ] is clear from the definition of τ0 . Let M ∶= supn≥1 ν1 (kn ) < ∞, as
(kn )∞
n=1 is a summability kernel.
Let ε > 0 and choose δ > 0 such that ∣s − 0∣ < δ implies that
ε
∥τs [f ] − τ0 [f ]∥B < .
2M
176 L.W. Marcoux Introduction to Lebesgue measure
This is possible by (a) above. Next, choose 1 ≤ N ∈ N such that n ≥ N implies that
1 −δ 1 π ε
∫ ∣kn (s)∣ds + ∫ ∣kn (s)∣ds < .
2π −π 2π δ 4∥[f ]∥B
Then n ≥ N implies that
1 π
∥kn ∗ [f ] − [f ]∥B = ∥ ∫ kn (s)(τs [f ] − τ0 [f ])ds∥B
2π −π
1 −δ
≤ ∫ ∣kn (s)∣ ∥τs [f ] − τ0 [f ]∥B ds
2π −π
1 δ
+ ∫ ∣kn (s)∣ ∥τs [f ] − τ0 [f ]∥B ds
2π −δ
1 π
+ ∫ ∣kn (s)∣ ∥τs [f ] − τ0 [f ]∥B ds.
2π δ
Now ∥τs [f ] − τ0 [f ]∥B ≤ ∥τs [f ]∥B + ∥τ0 [f ]∥B = 2∥[f ]∥B , and thus for n ≥ N , we have
1 −δ
∥kn ∗ [f ] − [f ]∥B ≤ ∫ ∣kn (s)∣(2∥[f ]∥B ) ds
2π −π
1 δ ε
+ ∫ ∣kn (s)∣ ds
2π −δ 2M
1 −δ
+ ∫ ∣kn (s)∣(2∥[f ]∥B ) ds
2π −π
ε ε
≤ 2∥[f ]∥B +M
4∥[f ]∥B 2M
= ε.
In other words,
lim kn ∗ [f ] = [f ].
n→∞
◻
11.11. Corollary.
(a) For each f ∈ (C(T, C), ∥ ⋅ ∥sup ),
○
lim σN (f ) = f.
N →∞
(b) Again, by Example 9.12, for each 1 ≤ p < ∞, (Lp (T, C), ∥ ⋅ ∥p ) is a homo-
geneous Banach space over T. Let 1 ≤ p < ∞ and [f ] ∈ Lp (T, C). Since
(FN )∞N =1 is a (positive) summability kernel, and since σN ([f ]) = FN ∗ [f ]
for all N ≥ 1,
by Theorem 11.10.
◻
We are now in a position to show that the Fourier coefficients of an Lp (T, C)-
function completely determine that function (almost everywhere).
[f ] [g]
11.12. Corollary. Let 1 ≤ p < ∞. If [f ], [g] ∈ Lp (T, C) and αn = αn for all
n ∈ Z, then [f ] = [g].
[f ] [g]
Proof. It is clear that if αn = αn for all n ∈ Z, then σN [f ] = σN ([g]) for all N ≥ 1.
By Corollary 11.11,
11.13. Local structure and Féjer kernels. Corollaries 11.11 and 11.12 tell
us that Féjer kernels are nice enough to recover an element of Lp (T, C) from its
Fourier coefficients. As we have emphasized in these notes, however, these elements
are equivalence classes of functions in Lp (T, C), and not functions themselves. In
other words, the aforementioned Corollaries say that we can recover functions in
Lp (T, C) almost everywhere on R. We now turn our attention to the functions
themselves, and study in what sense (if any) the convolution of Lp (T, C) functions
with Féjer kernels converge pointwise.
(b) Suppose that there exists a closed interval [a, b] ⊆ [−π, π) such that f is
continuous on [a, b] (in particular, f is continuous from the right at a and
from the left at b), then
(FN ◇ f )∞
N =1
f (θ − s) + f (θ + s)
∣ωf (θ) − ∣ < ε.
2
Now
1
∣σN (f )(θ) − ωf (θ)∣ = ∣ ∫ FN (s)f (θ − s) dm(s) − ωf (θ)∣
2π [−π,π)
1
=∣ ∫ FN (s)(f (θ − s) − ωf (θ)) dm(s)∣
2π [−π,π)
1
≤∣ ∫ FN (s)(f (θ − s) − ωf (θ)) dm(s)∣
2π [−δ,δ]
1
+∣ ∫ FN (s)(f (θ − s) − ωf (θ)) dm(s)∣ .
2π [−π,−δ)∪(δ,π]
1
∫ FN (s)(f (θ − s) − ωf (θ)) dm(s) =
2π [−δ,δ]
1 f (θ − s) + f (θ + s)
∫ FN (s) ( − ωf (θ)) dm(s).
2π [−δ,δ] 2
11. THE FÉJER KERNEL 179
Thus
1
∣ ∫ FN (s)(f (θ − s) − ωf (θ)) dm(s)∣ ≤
2π [−δ,δ]
1 f (θ − s) + f (θ + s)
∫ FN (s) ∣ − ωf (θ)∣ dm(s)
2π [−δ,δ] 2
1
≤ ∫ FN (s)ε dm(s)
2π [−δ,δ]
1
≤ ε( ∫ FN (s) dm(s))
2π [−π,π)
= ε.
Meanwhile, for δ ≤ ∣s∣ ≤ π,
π2 π2
0 ≤ FN (s) ≤ ≤ ,
N s2 N δ 2
and so
1
∣ ∫ FN (s)(f (θ − s) − ωf (θ)) dm(s)∣
2π [−π,−δ)∪(δ,π]
π2 1
≤ 2
( ∫ ∣f (θ − s)∣ + ∣ωf (θ)∣ dm(s))
N δ 2π [−π,−δ)∪(δ,π]
π2 1
≤ 2
( ∫ ∣f (θ − s)∣ dm(s) + ∣ωf (θ)∣)
N δ 2π [−π,π)
π2
≤ (∥[f ]∥1 + ∣ωf (θ)∣),
N δ2
which converges to 0 as N tends to ∞.
Thus limN →∞ σN (f )(θ) = ωf (θ).
(b) The proof is essentially identical to that above.
First note that the continuity of f at θ for θ ∈ [a, b] implies that ωf (θ) =
f (θ) for all θ ∈ [a, b]. But f continuous on [a, b] implies that f is uniformly
continuous on [a, b], and so we may find a single δ > 0 such that 0 ≤ ∣s∣ > δ
implies that
f (θ − s) + f (θ + s)
∣f (θ) − ∣ < ε, θ ∈ [a, b].
2
For this δ > 0, and for all θ ∈ [a, b], the estimates from part (a) above show
that
1 f (θ − s) + f (θ + s)
∣σN (f )(θ) − f (θ)∣ ≤ ∫ FN (s) ∣ − f (θ)∣ dm(s)
2π [−δ,δ] 2
1
+ ∫ FN (s)(∣f (θ − s)∣ + ∣f (θ)∣) dm(s)
2π [−π,−δ)∪(δ,π]
π2
<ε+ (∥[f ]∥1 + ∣f (θ)∣) ,
N δ2
180 L.W. Marcoux Introduction to Lebesgue measure
11.16. Corollary.
(a) If f ∈ L1 (T, C) and f is continuous at θ0 ∈ R, then
lim FN ◇ f (θ0 ) = f (θ0 ).
N →∞
In particular,
f (θ) = lim FN ◇ f (θ)
N →∞
almost everywhere on R.
11. THE FÉJER KERNEL 181
11.18. So far we have seen that while it is extremely rare for the partial sums
of the Fourier series of a continuous function f , or of an element [g] of L1 (T, C) to
converge to f uniformly (or to [g] in the ∥ ⋅ ∥1 -norm), nevertheless, this is always
the case regarding the Cesàro sums of the Fourier series.
Furthermore, Féjer’s Theorem shows that if f ∈ L1 (T, C) and if the average
value ωf (θ) exists at θ ∈ R, then
π 2
Now, since 0 ≤ ∣s∣ < δ ≤ , we have that ∣s∣ ≤ ∣ sin s∣, and thus
2 π
1 s
∣s∣ ≤ ∣ sin ∣, 0 ≤ ∣s∣ < δ.
π 2
Thus
1
∫ ∣f (θ0 + s) − f (θ0 )∣∣DN (s)∣ dm(s)
2π [−δ,δ]
1 ∣f (θ0 + s) − f (θ0 )∣ 1
≤ ∫ ∣ sin(N + )s∣ dm(s)
2π [−δ,δ] ∣s∣/π 2
1 M ∣s∣
≤ ∫ dm(s)
2 [−δ,δ] ∣s∣
= M δ,
independent of N !!!
Next we consider δ ≤ ∣s∣ ≤ π. Observe that
f (θ0 + s) − f (θ0 ) 1
(f (θ0 + s) − f (θ0 ))DN (s) = sin((N + )s)
sin(s/2) 2
f (θ0 + s) − f (θ0 ) eis/2 iN s e−is/2 −iN s
= ( e − e ).
sin(s/2) 2i 2i
11. THE FÉJER KERNEL 183
Define
f (θ0 + s) − f (θ0 ) eis/2
g1 (s) = χ[−π,−δ)∪(δ,π)
sin(s/2) 2i
and
f (θ0 + s) − f (θ0 ) e−is/2
g2 (s) = χ[−π,−δ)∪(δ,π) .
sin(s/2) 2i
Then g1 , g2 are measurable (why?).
Also,
1 1 ∣f (θ0 + s) − f (θ0 )∣ 1
∫ ∣g1 ∣ = ∫
2π [−π,π) 2π [−π,−δ)∪(δ,π) ∣ sin s/2∣ 2
1 ∣f (θ0 + s)∣ + ∣f (θ0 )∣
≤
4π ∣ sin δ/2∣
1
= (∥[f ]∥1 + 2π∣f (θ0 )∣)
2π∣ sin δ/2∣
< ∞.
Thus g1 ∈ L1 (T, C), and similarly, g2 ∈ L1 (T, C).
By the Riemann-Lebesgue Lemma 8.9, we have that
1
lim ∫ g1 (s)eiN s dm(s) = 0
N →∞ 2π [−π,π)
and
1
lim ∫ g2 (s)e−iN s dm(s) = 0.
N →∞ 2π [−π,π)
From this it easily follows that
1
lim ∫ (f (θ0 + s) − f (θ0 ))DN (s) dm(s) = 0.
N →∞ 2π [−π,−δ)∪(δ,π)
11.22. Example. Consider the function f (θ) = ∣θ∣, θ ∈ [−π, π), extended 2π-
periodically to all of R. Clearly f is continuous on R.
Since f is continuous and linear on (−π, 0) ∪ (0, π), it is locally Lipschitz there.
It is also easy to see that f is locally Lipschitz at θ = nπ, n ∈ Z with Lipschitz
constant M = 1.
We are preparing to do the unthinkable: we will calculate sN (f ), N ≥ 1. To
[f ]
do this, we must first calculate αn , n ∈ Z. To quote Fourier himself, “buckle up,
cupcake”.
We first point out that f is Riemann integrable and bounded over [−π, π),
and that multiplying f by ξn = einθ (for any value of n ∈ Z) doesn’t change this
fact. Because of this, and by virtue of Theorem 5.24, in calculating the Fourier
coefficients of [f ], we may always replace the Lebesgue integrals which appear by
Riemann integrals.
Case 1: n = 0. Observe that
1 1 π 1 π 1 2 π
[f ]
α0 = ∫ f= ∫ f (θ) dθ = ∫ ∣θ∣ dθ = π = .
2π [−π,π) 2π −π 2π −π 2π 2
Case 1: n ≠ 0.
In this case,
1
αn[f ] = ∫ f ξn
2π [−π,π)
1 π
−inθ
= ∫ ∣θ∣e dθ
2π −π
1 0 1 π
−inθ −inθ
= ∫ (−θ)e dθ + ∫ θe dθ.
2π −π 2π 0
These integrals are easily found using integration by parts, and we leave them
as exercises for the reader. The answers are:
11. THE FÉJER KERNEL 185
⎧
⎪0 0 ≠ n an even integer
⎪
⎪
⎪
αn[f ] = ⎨
⎪
⎪
⎪
⎪ −2
⎪
⎪ 0 ≠ n an odd integer.
⎩ n2 π
Note that e−inθ + einθ = 2 cos(nθ) for all n ≥ 1. Thus – taking into account that
we only wish to sum over odd n’s below –
π −2
sN ([f ]) = + ∑ ( ) 2 cos((2n − 1)θ)
2 1≤2n−1≤N (2n − 1)2 π
π 4 1
= − ∑ cos((2n − 1)θ).
2 π 1≤2n−1≤N (2n − 1)2
But
1 1
∣ ∑ 2
cos((2n − 1)θ)∣ ≤ ∑ 2
1≤2n−1≤N (2n − 1) 1≤2n−1≤N (2n − 1)
N
1
≤∑ 2
.
n=1 n
But this last series converges, which implies that (sN ([f ]))∞ n=1 converges in the
∥ ⋅ ∥∞ -norm, as ([C(T, C)], ∥ ⋅ ∥∞ ) is complete. In other words,
N
sN (f ) = ∑ αn[f ] ξn
n=−N
converges uniformly in (C(T, C), ∥ ⋅ ∥sup ) to a continuous function.
11.23. It is also instructive to look at the graph of the Féjer kernels of various
orders. Two things worth noticing are that
● first, the amplitude of the function is increasing near 0; this is clear since
each FN is continuous, and FN (0) = N , N ≥ 1.
● For each δ > 0, the functions are becoming uniformly close to zero when
δ < ∣θ∣ < π.
11.25. Examples.
(a) For 1 ≤ n ∈ N, let kn be the 2π-periodic function on R whose restriction to
the interval [−π, π) coincides with
nπχ[− 1 , 1 ] .
n n
Then (kn )∞
n=1 is a positive summability kernel. The details are left to
the reader.
11. THE FÉJER KERNEL 189
Then (kn )∞
n=1 is a positive summability kernel. Again, the details are
left to the reader.
190 L.W. Marcoux Introduction to Lebesgue measure
Exercise 11.1.
Prove the claim of Proposition 11.3, namely: suppose that X is a Banach space
and (xn )∞ ∞
n=0 is a sequence in X. Let (σN )N =1 denote the sequence of Cesàro means
∞
of (xn )n=0 .
Prove that if x = limn→∞ xn exists, then x = limN →∞ σN .
Exercise 11.2.
Prove that the sequences (kn )n and (rn )n listed in Examples 11.9 are indeed
positive summability kernels.
12. WHICH SEQUENCES ARE SEQUENCES OF FOURIER COEFFICIENTS? 191
I’ve learned about his illness. Let’s hope it’s nothing trivial.
Irving S. Cobb
Our approach to this problem will be via Operator Theory. Recall from Chap-
ter 8 that we defined the map
Λ ∶ (L1 (T, C), ∥ ⋅ ∥1 ) → (c0 (Z, C), ∥ ⋅ ∥∞ )
[f ]
[f ] ↦ (αn )n∈Z .
Since Lebesgue integration is linear, so is Λ. Also, as was shown in paragraph 8.8
∣αn[f ] ∣ ≤ ∥[f ]∥1 for all n ∈ Z,
so
∥Λ([f ])∥∞ = sup{∣αn[f ] ∣ ∶ n ∈ Z} ≤ ∥[f ]∥1 .
This is precisely the statement that the operator Λ is bounded, with ∥Λ∥ ≤ 1.
By Corollary 11.12, if [f ], [g] ∈ L1 (T, C) and Λ([f ]) = Λ([g]), then [f ] = [g],
and thus Λ is injective.
The question of whether or not every sequence in c0 (Z, C) is the sequence of
Fourier coefficients of some element of L1 (T, C) is therefore the question of whether
or not Λ is surjective.
The result we shall need from Functional Analysis is the Inverse Mapping The-
orem. To get this result, we will first require a lemma, and some notation.
Given a Banach space (Z, ∥ ⋅ ∥Z ) and a real number r > 0, we denote the closed
ball of radius r centred at the origin by
Zr = {z ∈ Z ∶ ∥z∥Z ≤ r}.
For z0 ∈ Z and ε > 0, we shall denote by B Z (z0 , ε) = {z ∈ Z ∶ ∥z − z0 ∥ < ε} the open
ball of radius ε in Z, centred at z0 .
192 L.W. Marcoux Introduction to Lebesgue measure
12.2. Lemma. Let X and Y be Banach spaces and suppose that T ∈ B(X, Y).
If Y1 ⊆ T Xm for some m ≥ 1, then Y1 ⊆ T X2m .
Proof. First observe that Y1 ⊆ T Xm implies that Yr ⊆ T Xrm for all r > 0.
Choose y ∈ Y1 . Then there exists x1 ∈ Xm so that ∥y − T x1 ∥ < 1/2. Since
y − T x1 ∈ Y1/2 ⊆ T Xm/2 , there exists x2 ∈ Xm/2 so that ∥(y − T x1 ) − T x2 ∥ < 1/4. More
generally, for each n ≥ 1, we can find xn ∈ Xm/2n−1 so that
n
1
∥y − ∑ T xj ∥ < .
j=1 2n
Since X is complete and ∑∞
n=1 ∥xn ∥ ≤ ∑∞
n=1
m
2n−1
= 2m, we have x = ∑∞
n=1 xn ∈ X2m .
By the continuity of T ,
∞ N
T x = T ( ∑ xn ) = lim T ( ∑ xn ) = y.
n=1 N →∞ n=1
Thus y ∈ T X2m ; i.e. Y1 ⊆ T X2m .
◻
So - much like the enigma surrounding the Cadbury Caramilk bar, the mystery
persists.
12. WHICH SEQUENCES ARE SEQUENCES OF FOURIER COEFFICIENTS? 195
Exercise 12.1.
∞
ξn (θ) − ξ−n (θ) sin nθ
Keeping in mind that sin nθ = , θ ∈ R, the series ∑ can be
2i n=1 n
expressed as the Fourier series of a function in L2 (T, C). Which function?
Bibliography
[1] I.N. Herstein. Topics in Algebra, 2nd ed. John Wiley and Sons, New York, Chichester, Brisbane,
Toronto, Singapore, 1975.
[2] Y. Katznelson. An introduction to harmonic analysis, Third edition. Cambridge Mathematical
Library. Cambridge University Press, Cambridge, 2004.
[3] L.W. Marcoux. An introduction to Functional Analysis. Online notes – available at
http://www.math.uwaterloo.ca/∼lwmarcou/, 2012.
[4] J.C. Oxtoby. Measure and Category: A survey of the analogies between topological and measure
spaces, volume 2 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1971.
[5] R.M. Solovay. A model of set-theory in which every set of reals is lebesgue measurable. Ann. of
Math., 92:1–56, 1970.
[6] G. Vitali. Sul problema della misura dei gruppi di punti di una retta. Bologna, Tip. Gamberini
e Parmeggiani, 1905.
[7] R. Whitley. Projecting m onto c0 . The Amer. Math. Monthly, 73:285–286, 1966.
[8] Stephen Willard. General Topology. Addison-Wesley Publishing Co., Reading, Mass.-London-
Don Mills, Ont., 1970.
197
Index
199
200 INDEX