math3306-lecturenotes
math3306-lecturenotes
We continue our study of measure theory and integration started in Math 3301. Topics
covered are the Lebesgue measure on IRN , product of measure spaces, the change of variables
theorem, Fubini theorems, Lp spaces, convolutions and Fourier transforms.
All questions, comments, remarks and suggestions are welcome.
References
3 Lp spaces 15
3.1 Convex functions and inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3
4 CONTENTS
Chapter 1
Our target in this chapter is to give a mathematical meaning to the intuitive notions of area
and volume. We shall assign a measure (a volume) to many subsets of IRN . This measure is
called the Lebesgue measure and the class of subsets having a measure is called the Lebesgue
σ−algebra.
Here is our plan for this chapter.
1. Construct an outer measure λ∗N on IRN and establish its special properties.
2. Use the first part of Caratheodory’s theorem to construct a σ−algebra LN on IRN and
establish its properties.
3. Use the second part of Caratheodory’s theorem to construct a measure λN on (IRN , LN )
called the Lebesgue measure. Establish some important properties λN .
If
P∞ A ⊂ ∪∞ n=1 Pn , then the sequence (Pn ) is called a countable covering of A by N −cells and
vol(P ∗
n=1 n ) is called the total volume of the covering. In words, λN (A) is the the smallest
total volume of countable coverings of A by N -cells. Thus, measuring a subset of IRN involves
approximating this set by N −cells from outside; and this is why we call λ∗N an outer measure.
Some comments are in order. First, λ∗ is well defined. Indeed, let
(∞ ∞
)
X [
XA = volN (Pn )) Pn is an N -cell and A ⊂ Pn .
n=1 n=1
S∞ P∞
Then XA ⊂ [0, ∞]. Since A ⊂ n=1 ] − n, n[N and n=1 (2n)N = +∞, we have +∞ ∈ XA .
Therefore XA is not empty and so it has an infinimum. It is possible that XA = {+∞}. For
example, XIRN = {+∞}.
5
6 CHAPTER 1. THE LEBESGUE MEASURE ON IRN
Definition 1.1 Let E ⊂ IRN and a ∈ IRN . We set E + a = {x + a|x ∈ E}. We say that
E + a is a translate of E. If a ∈ IR, we set aE = {ax|x ∈ E}. aE is the image of E under the
homothety x 7→ ax.
Exercise. Prove the following
1. If E is an N −cell then E + a is an N −cell and volN (E + a) = volN (E).
2. If a ∈ IR and P is an N −cell, then volN (aP ) = |a|N volN (P ).
Lemma 1.1 An open subset of IRN is a countable union of open and bounded N −cells.
Corollary 1.1 The Borel σ−algebra of IRN is generated by the open and bounded N −cells.
Lemma 1.2 The Borel σ−algebra on IRN is generated by the family of open half spaces, i,e.,
by the sets of the form
6. λN is σ−finite.
Here is another important relation between the Lebesgue σ−algebra and the Borel σ−algebra
on IRN .
Proposition 1.5 The Lebesgue σ−algebra is the completion of the Borel σ−algebra. This
means that a subset A ⊂ IRN is Lebesgue measurable if and only if there exists a Borel set B
and a negligible set M , such that A = B ∪ M .
Corollary 1.2 Let f : IRN → IR be Lebesgue measurable. Then there exists a Borel measur-
able function g : IRN → IR that coincides with f almost everywhere.
The proof is difficult and can be found in the book of Rudin (Theorem 8.26).
It follows in particular that the Lebesgue measure is invariant under rotations and symmetries
and more generally under any differentiable homeomorphism with Jacobian ±1.
8 CHAPTER 1. THE LEBESGUE MEASURE ON IRN
Theorem 1.2 (Change of variables) Let U and V be two open sets of IRN and let ϕ :
U → V be a differentiable homeomorphism. Let A ⊂ U be a Lebesgue measurable set and
f : ϕ(A) → IR be Lebesgue measurable. If either f has constant sign or is summable, then
Z Z
f (x) dx = f (ϕ(y))|J(ϕ)|dy.
ϕ(A) A
∂ old variables
In this form, the Jacobian should be thought of as .
∂ new variables
Chapter 2
Proposition 2.3 Let (X, A), (Y, B) and (Z, C) be measurable spaces. Let f : X × Y → Z be
measurable. Then the partial maps f (x, ·) : Y → Z and f (·, y) : X → Z are measurable.
9
10 CHAPTER 2. PRODUCT OF MEASURE SPACES
Proof. Let for every x ∈ X, gx : Y → X × Y be defined by gx (y) = (x, y). Then first,
each component of gx is measurable(the first component is constant and the second component
is the identity). By the previous proposition, gx is measurable . Second f (x, ·) = f ◦ gx .
Therefore f (x, ·) is measurable as a composition of measurable functions. The proof that f (·, y)
is measurable is similar.
Now more generally, let (Xi , Ai ), i = 1, . . . , d, be a sequence of measurable spaces. Then we
denote by A1 ⊗ · · · ⊗ Ad the sigma algebra on X1 × · · · × Xd generated by the sets of the form
A1 × · · · × Ad , where Ai ∈ Ai , that is by the rectangles with measurable sides. This the smallest
sigma algebra on X1 × · · · × Xd that makes the projections pi : (x1 , . . . , xd ) 7→ xi measurable.
Proposition 2.4 [Associativity of ⊗] Let (X1 , A1 ), (X2 , A2 ) and (X2 , A3 ) be measurable spaces.
Then
(A1 ⊗ A2 ) ⊗ A3 = A1 ⊗ (A2 ⊗ A3 ) = A1 ⊗ A2 ⊗ A3 .
Proof. The σ−algebra (A1 ⊗ A2 ) ⊗ A3 is the smallest sigma algebra that makes the maps
((x1 , x2 ), x3 ) 7→ (x1 , x2 ) and ((x1 , x2 ), x3 ) 7→ x3 measurable. These maps are measurable if
and only if the maps (x1 , x2 , x3 ) 7→ x1 , (x1 , x2 , x3 ) 7→ x2 and (x1 , x2 , x3 ) 7→ x3 measurable.
But the smallest sigma algebra that makes these maps measurable is A1 ⊗ A2 ⊗ A3 . Thus,
(A1 ⊗ A2 ) ⊗ A3 = A1 ⊗ A2 ⊗ A3 .
A similar reasoning show that A1 ⊗ (A2 ⊗ A3 ) = A1 ⊗ A2 ⊗ A3 .
Let (X, A) and (Y, B) be two measurable spaces and let E ∈ A ⊗ B. For every x ∈ X, the
section of E by x is defined by
Ex = {y ∈ Y |(x, y) ∈ E}.
Note that Ex is a measurable set, i.e., Ex ∈ B. Indeed, the characteristic function of Ex satisfies
1Ex (y) = 1E (x, y). The result follows from Proposition 2.3 and the fact that a set is measurable
if and only if its characteristic function is measurable.
Similarly, for y ∈ Y , the section of E by y is defined by
E y = {x ∈ X|(x, y) ∈ E}.
Lemma 2.1 Let (X, A, µ) and (Y, B, ν) be measure spaces with ν σ−finite. Then for every
E ∈ A ⊗ B the map α : X → [0, ∞] defined by α(x) = ν(Ex ) is A−measurable.
Theorem 2.1 Let (X, A, µ) and (Y, B, ν) be σ−finite measure spaces. Then there exists a
unique measure denoted by µ ⊗ ν on the product space (X × Y, A ⊗ B) satisfying
(µ ⊗ ν)(A × B) = µ(A)ν(B)
for all A ∈ A and B ∈ B. This measure is σ−finite and is called the product measure of µ and
ν. Moreover, for every E ∈ A ⊗ B, we have
Z Z
(µ ⊗ ν)(E) = ν(Ex ) dµ = µ(E y ) dν.
X Y
Consider now the set IRN for N ≥ 2. Let r and s be two positive integers such that
r + s = N . Then IRN can be identified with IRr × IRs (they have the same algebraic and
topological structures). Four natural σ−algebras can therefore be defined on IRN : the Borel
σ−algebra BN , the product σ−algebras Br ⊗ Bs and Lr ⊗ Ls and the Lebesgue σ−algebra
LN . What are then the relation between these σ−algebras? The answer is given in the next
proposition.
2.2. TONELLI AND FUBINI’S THEOREMS 11
Proposition 2.5 Let r and s be two positive integers such that r + s = N . Then the following
hold.
(i) BN = Br ⊗ Bs Lr ⊗ Ls LN .
(ii) The measure space (IRN , LN , λN ) is the completion of the space (IRN , Lr ⊗ Ls , λr ⊗ λs ).
Fubini’s theorems give sufficient conditions under which the interchange is possible. These
conditions cannot be easily removed as we shall see in the exercises.
Theorem 2.2 (Tonelli) Let (X, A, µ) and (Y, B, ν) be σ−finite measure R spaces and let f :
X
R × Y → [0, ∞] be A ⊗ Bmeasurable. Then, the functions x →
7 Y f (x, y) dν(y) and y 7→
X f (x, y) dµ(x) are measurable and
Z Z Z Z Z
f (x, y)d(µ ⊗ ν) = f (x, y) dν(y) dµ(x) = f (x, y) dµ(x) dν(y). (E)
X×Y X Y Y X
Tonelli theorem states that under certain assumptions, a double integral can be computed as a
nested integral and the order of integration is immaterial. The assumptions are
We shall see in the exercises that the conclusion of the theorem may fail if one of these assump-
tions is not satisfied.
Proof. We prove the theorem in three steps. The first step is not necessary because it is a
particular case of step 2, but we do it because it makes the proof more clear.
Step 1.R Suppose first that
R f is a characteristic
R function. Then f = 1E for some E ∈ A ⊗ B.
Then, Y f (x, y) dν(y) = 1
Y E (x, y) dν = 1
Y Ex (y) dν = ν(Ex ) by definition of the Lebesgue
y
R
integral. Similarly, X f (x, y) dµ(x) = µ(E ). Lemma 2.1 ensures these inner integrals are
measurable. Since they are nonnegative, we can integrate them with respect to the other
measure. Now
Z Z
• f d(µ ⊗ ν) = 1E d(µ ⊗ ν) = (µ ⊗ ν)(E) by definition of the Lebesgue integral,
X×Y X×Y
Z Z Z
• f (x, y) dν(y) dµ(x) = ν(Ex ) dµ(x);
X Y X
Z Z Z
• f (x, y) dµ(x) dµ(y) = µ(E y ) dν(y).
Y X Y
By Theorem 2.1, Z Z
(µ ⊗ ν)(E) = ν(Ex ) dµ = µ(E y ) dν.
X Y
Hence equation (E) when f is a characteristic function.
Pm
Step 2. Suppose now that f is a simple function. Then f = i=1 αi 1Ei . Therefore
12 CHAPTER 2. PRODUCT OF MEASURE SPACES
Z Z m
X m
X
• f d(µ ⊗ ν) = αi 1Ei d(µ ⊗ ν) = αi (µ ⊗ ν)(Ei ) by definition of the Lebesgue
X×Y X×Y i=1 i=1
integral;
Z Z m
X Z
• f (x, y) dν(y) dµ(x) = αi ν((Ei )x ) dµ(x) by linearity of the integral;
X Y i=1 X
Z Z m Z
µ(Eiy ) dν(y) by linearity of the integral.
X
• f (x, y) dµ(x) dµ(y) = αi
Y X i=1 Y
Z Z Z Z Z
fn (x, y)d(µ ⊗ ν) = fn (x, y) dν(y) dµ(x) = fn (x, y) dµ(x) dν(y).
X×Y X Y Y X
Letting n → ∞ and using the monotone convergence theorem (twice for the nested integrals) we
get the result (the inner integrals are measurable as pointwise limits of measurable functions) .
Z Z Z Z
Remark 2.1 Instead of f (x, y) dν(y) dµ(x) we can write dµ(x) f (x, y) dν(y)
Z Z X Y X Y
or dµ f dν.
X Y
Corollary 2.1 Let (X, A, µ) and (Y, B, ν) be σ−finite measure spaces and let f : X × Y → IR
be measurable. Then the following conditions are equivalent
(i) f is µ ⊗ ν summable.
R R
(ii) X Y |f (x, y)| dν dµ < ∞.
R R
(iii) Y X |f (x, y)| dµ dν < ∞.
2
Example. Let f (x, y) = e−x(1+y ) . Then
Z ∞ Z ∞
π
f (x, y) dx dy = ,
0 0 2
whereas 2
Z ∞ Z ∞ Z ∞
−t2
f (x, y) dy dx = 2 e dt .
0 0 0
Z ∞ √
2 π
By Tonelli, e−t dt = (write the details).
0 2
Theorem 2.3 (Fubini) Let (X, A, µ) and (Y, B, ν) be σ−finite measure spaces and let f :
X × Y → IR be µ ⊗ ν summable. Then
1. The function y 7→ f (x, y) is ν−summable for µ−almost every x ∈ X and the function
x 7→ f (x, y) is µ−summable for ν−almost every y ∈ Y .
2.2. TONELLI AND FUBINI’S THEOREMS 13
R R
2. The function x 7→ Y f (x, y) dν is µ−summable and the function y 7→ X f (x, y) dµ is
ν−summable.
3. We have Z Z Z Z Z
f d(µ ⊗ ν) = f dν dµ = f dµ dν.
X×Y X Y Y X
The conclusion of Fubini’s theorem is the same as that of Tonelli. The assumptions however
are
The inner integrals are therefore finite almost everywhere. It follows that
Z Z Z
|f | dν = f + dν + f − dν < ∞
Y Y Y
Z Z Z
for almost every x ∈ X. Similarly, |f | dµ = +
f dµ + f − dµ < ∞ for almost every
X X X
y ∈ Y . This proves 1.
Z Z Z
2. Note that f dν = +
f dν − f − dν. By Tonelli’s theorem, the partial integrals
Z Y Z Y Y
f + dν and f − dν are measurable. Therefore x 7→ Y f dν is measurable on its
R
x 7→
Y Y
domain (that is µ − a.e). Next
Z Z Z Z Z Z
f dν dµ ≤ |f | dν dµ = (f + + f − ) dν dµ
X Y
ZX ZY ZX ZY
= +
f dν dµ + f − dν dµ < ∞.
X Y X Y
Proof. Observe first that our assumptions imply that ∆ is Borel measurable (write the
details). Now extend f to IR2 by setting
(
f (x, y) if (x, y) ∈ ∆
f¯(x, y) =
0 if (x, y) ∈
/ ∆.
We know from Math 3301 that f¯ is still Borel measurable. Note that
Z Z
¯
f (x, y) dλ2 = f (x, y) dλ2 .
IR2 ∆
By Tonelli’s theorem (if f has constant sign) or Fubini’s theorem (if f is summable), we have
Z Z Z
f¯(x, y) dλ2 = f¯(x, y) dy dx.
IR2 IR IR
But because f¯ vanishes outside ∆ and coincide with f on ∆, the iterated integral is
:0
Z Z
Z Z Z Z
f¯(x, y) dy f¯(x, y) dy dx + f¯(x,
dx = y) dy dx
IR IR [a,b] IR
IR\[a,b]
IR
:0
Z Z Z
f¯(x, y) dy + ¯
=
f (x, y) dy
dx
[a,b] [ϕ1 (x),ϕ2 (x)]
IR\[ϕ
1 (x),ϕ2 (x)]
Z Z !
= f (x, y) dy dx.
[a,b] [ϕ1 (x),ϕ2 (x)]
Remark 2.2 Indeed, we have a similar result for domains of the form
where D is a Borel measurable subset of IRN and ϕ1 , ϕ2 : D → IR are given Borel measurable
functions. Let f : ∆ → IR be Borel measurable. If either f has constant sign or is summable,
then
Z Z Z ϕ2 (x1 ,...,xN ) !
f (x1 , . . . , xN , xN +1 ) dλN +1 = f (x1 , . . . , xN , xN +1 ) dxN +1 dλN .
∆ D ϕ1 (x1 ,...,xN )
Chapter 3
Lp spaces
Lp spaces are normed spaces of functions that are very important in Analysis and especially in
the theory of partial differential equations. Before we introduce them, we prove some important
inequalities related to the very important notion of convexity.
for every t ∈]0, 1[ and every x, y ∈ I. Geometrically, a function is convex if the part of its graph
between any two points is below the straight line segment joining these points.
A function ϕ : I → IR is said to be concave if −ϕ is convex, or equivalently
for every t ∈]0, 1[ and every x, y ∈ I. Geometrically, a function is concave if the part of its
graph between any two points is above the straight line segment joining these points.
Remark 3.1 The epigraph of a function ϕ : I → IR is the set of points in the plane that lie
above the graph of ϕ. It is denoted by epi(ϕ). So
Proposition 3.1 Let I be an interval of IR and let ϕ : I → IR be a function. Then the following
conditions are equivalent
(i) ϕ is convex.
15
16 CHAPTER 3. LP SPACES
Inequality (ii) means that the slope of the line joining (x, ϕ(x)) and (y, ϕ(y)) is smaller than the
slope of the line joining (y, ϕ(y)) to (z, ϕ(z)). Inequality (iii) means that the slope of the line
joining (y, ϕ(y)) and (z, ϕ(z)) is smaller than the slope of the line joining (x, ϕ(x)) to (z, ϕ(z)).
Inequality (iv) means that the slope of the line joining (y, ϕ(y)) and (x, ϕ(x)) is smaller than
the slope of the line joining (x, ϕ(x)) to (z, ϕ(z)).
x−y
Proof. (i)⇒(ii). Observe that x = (1 − t)y + tz where t = z−y . Since 0 < t < 1, convexity of
ϕ implies that ϕ(x) ≤ (1 − t)ϕ(y) + tϕ(z). Therefore
x−y
ϕ(x) − ϕ(y) ≤ t(ϕ(z) − ϕ(y)) = (ϕ(z) − ϕ(y)).
z−y
Dividing by x − y > 0, we get the result. Going back through the steps, we can prove that
(ii)⇒(i).
(i)⇒(iii). As above, we have ϕ(x) ≤ (1 − t)ϕ(y) + tϕ(z). Therefore
Going back through the steps, you can prove that (iii)⇒(i).
(i)⇒(iv). (iv) follows from (ii) and (iii).
(iv)⇒(i). Let y, z ∈ I such that y < z and let t ∈]0, 1[. Let x = (1 − t)y + tz. Then y < x < z
and so
ϕ(x) − ϕ(y) ϕ(z) − ϕ(x)
≤ .
x−y z−x
It follows that
x−y t
ϕ(x) − ϕ(y) ≤ (ϕ(z) − ϕ(x)) = (ϕ(z) − ϕ(x)) .
z−x 1−t
Therefore
ϕ(x) ≤ (1 − t)ϕ(y) + tϕ(z)
that is
ϕ((1 − t)y + tz) ≤ (1 − t)ϕ(y) + tϕ(z).
Observe now the following. Inequality (ii) implies that for each y ∈ I the function s 7→
ϕ(s)−ϕ(y)
s−y is increasing for s > y andinequality (iii) implies that for each z ∈ I, the function
ϕ(s)−ϕ(z) ϕ(s)−ϕ(a)
s 7→ s−z It follows that for any a ∈ I the function s 7→
is increasing for s < z. s−a is
increasing on I\{a}. Actually this condition is equivalent to the convexity of ϕ.
A convex function need not be differentiable. For example x 7→ |x| is not differentiable at 0.
However, it has a left and right derivative at 0. This is not a coincidence. A convex function
has a left and right derivative at every point inside its domain.
3.1. CONVEX FUNCTIONS AND INEQUALITIES 17
Proposition 3.2 Let I be an open interval of IR and let ϕ : I → IR be convex. Then the
following hold.
(i) At each point x ∈ I, ϕ has a left derivative ϕ0` (x) and right derivative ϕ0r (x) with ϕ0` (x) ≤
ϕ0r (x).
(ii) ϕ is continuous.
(iv) For every x ∈ I, for every β ∈ [ϕ0` (x), ϕ0r (x)], and for every y ∈ I, we have
Now fix x and z and let y vary. Then the function y 7→ ϕ(y)−ϕ(x)
y−x is increasing and bounded
from above. Therefore it has a limit as y % x. This means that ϕ has a left derivative at x.
If now we fix x and y and let z vary, we see that g(z) := ϕ(z)−ϕ(x)
z−x has a limit as z & x.1
This means that ϕ has a right derivative at x.
Now, if we fix x and let y % x and z & x in (3.1), we get ϕ0` (x) ≤ ϕ0r (x).
(ii). Let x ∈ I. Tthe existence of a left derivative at x implies the ϕ is continuous from the
left at x, that is, lim ϕ(y) = ϕ(x). Also the existence of a right derivative at x implies ϕ is
y%x
continuous from the right at x, that is, lim ϕ(y) = ϕ(x). It follows that ϕ is continuous at x.
y&x
(iii). Let x, z ∈ I satisfy x < z. We have to show that ϕ0` (x) ≤ ϕ0` (z). So let y and w satisfy
y < x < w < z. In this reasining, x and z are fixed and y and w vary. By the previous
proposition
ϕ(y) − ϕ(x) ϕ(x) − ϕ(w) ϕ(w) − ϕ(z)
≤ ≤ .
y−x x−w w−z
Letting y % x and w % z, we get ϕ0` (x) ≤ ϕ0` (z). Acompletely similar reasoning shows that
ϕ0r (x) ≤ ϕ0r (z).
(iv). Let x, y, z ∈ I satisfy y < x < z. The proof of (i) shows that
Remark 3.2 A convex function ϕ : I → IR need not be continuous at the end point of its
domain. For example the function ϕ : [0, 1] → IR defined by ϕ(0) = 1 and ϕ(x) = 0 for
0 < x ≤ 1 is convex but discontinuous at x = 0. This is why we assumed that I is open in the
previous proposition.
(i) ϕ is convex.
1
As z decreases to x, g(z) decreases and is bounded from below (by g(y)). Therefore it has a limit.
18 CHAPTER 3. LP SPACES
(ii) ϕ(y) ≥ ϕ(x) + ϕ0 (x)(y − x) for every y ∈ I (this means that the graph of ϕ is above any
of its tangent lines).
(iii) ϕ0 is increasing.
If moreover ϕ is twice differentiable, then (i), (ii) and (iii) are equivalent to the condition
ϕ00 ≥ 0.
Proof. (i)⇒(ii) follows from point (iv) of the the previous proposition.
ϕ(y)−ϕ(x)
(ii)⇒(iii). Let x, y ∈ I satisfy y < x. Then ϕ(y) ≥ ϕ(x) + ϕ0 (x)(y − x) and so y−x ≤ ϕ0 (x).
ϕ(x)−ϕ(y)
Similarly we have ϕ(x) ≥ ϕ(y) + ϕ0 (y)(x − y) and so x−y ≥ ϕ0 (y). Thus,
(iii)⇒(i). Suppose that ϕ is not convex. By Proposition 3.1 (iv), there exists x, y, z ∈ I such
that y < x < z and
ϕ(x) − ϕ(y) ϕ(z) − ϕ(x)
> .
x−y z−x
ϕ(x)−ϕ(y)
Now by the value theorem, there exists ξ1 between y and x such that x−y = ϕ0 (ξ1 ) and
there exists ξ2 between x and z such that ϕ(z)−ϕ(x)
z−x = ϕ0 (ξ2 ). But then we have ϕ0 (ξ1 ) > ϕ0 (ξ2 )
although ξ1 < ξ2 and this contradicts the monotonicity of ϕ0 .
Remark 3.3 If I is not necessarily open but ϕ is continuously differentiable, then the previous
corollary still hold.
1 1
+ = 1,
p q
q
or equivalently if pq = p + q, or p = q−1 , or pq = q − 1. We extend this definition to include the
case p = 1 and q = ∞ so that 1 and ∞ are called conjugate.
Proposition 3.4 (Young’s inequality) Let p and q be two conjugate numbers > 1 and let
a, b be nonnegative. Then
ap bq
ab ≤ + .
p q
3.1. CONVEX FUNCTIONS AND INEQUALITIES 19
Proposition 3.5 (Hölder’s inequality) Let (X, A, µ) be a measure space and let f, g : X →
IR be two measurable functions. Let p and q be two conjugate numbers > 1. Then
Z Z 1/p Z 1/q
p q
|f g| dµ ≤ |f | dµ |g| dµ .
X X X
1/p
Proof. Let ||u||p = X |u|p dµ
R
. We shall see shortly that this is a semi-norm. Then the
inequality can be written
||f g||1 ≤ ||f ||p ||g||q .
1 1
|f (x)g(x)| ≤ |f (x)|p + |g(x)|q .
p q
that is
1 1 1 1
||f g||1 ≤ ||f ||pp + ||g||qq = + = 1,
p q p q
since ||f ||p = ||g||q = 1.
Step 3. Assume that ||f ||p 6= 0 and ||g||q 6= 0. Set
f g
f1 = and g1 =
||f ||p ||g||q
so that ||f1 ||p = ||g1 ||q = 1. It follows from Step 2 that ||f1 g1 ||1 ≤ 1, that is
||f g||1
≤ 1.
||f ||p ||g||q
Proposition 3.6 (Minkowski’s inequality) Let (X, A, µ) be a measure space and let f, g :
X → IR be measurable functions. Let p ∈ [1, ∞[. Then
3.2 Definitions
Definition 3.3 Let 1 ≤ p < ∞ and let (X, A, µ) be a measure space. We denote by Lp (X) the
set of measurable functions f : X → IR, such that |f |p is summable on X:
Z
p
L (X) = {f : X → IR | f is measurable and |f |p dµ < ∞}.
X
Proposition 3.7 Let 1 ≤ p < ∞ and let (X, A, µ) be a measure space. Then
Proof. (i) Lp is a subspace of the space of all functions from X → IR equipped with the usual
addition and multiplication by a scalar. Indeed, let f, g ∈ Lp (X) and let α be a constant. Then
first, f + g and αf are measurable. Next, |f + g|p ≤ (|f | + |g|)p ≤ 2p−1 (|f |p + |g|p ). Integrating
over X, we get Z Z Z
|f + g|p ≤ 2p−1 |f |p + |g|p < ∞.
X X X
Z Z
Thus f + g ∈ Lp (X). Next |αf |p = |α|p |f |p < ∞ and so αf ∈ Lp (X).
X X
(ii) We clearly have
1. ||f ||p ≥ 0.
3.2. DEFINITIONS 21
2. ||0||p = 0.
3. ||αf ||p = |α|||f ||p (Homogeneity)
4. ||f + g||p ≤ ||f ||p + ||g||p (Triangle inequality)
The triangle inequality is just Minkowski’s inequality. However, ||f ||p = 0 ⇒ f = 0 a.e. So
|| · ||p is not strictly speaking a norm.
Definition 3.4 Let (X, A, µ) be a measure space and let f : X → IR be measurable. We say
that f is essentially bounded if there exists a constant C ≥ 0 such that |f (x)| ≤ C a.e. We
say that C is an essential bound of |f |. The set of measurable essentially bounded functions is
denoted by L∞ (X), that is,
Remark 3.4 Let (X, A, µ) = (IRN , LN , λN ) and let f : IRN → IR be continuous and bounded
then sup |f (x)| = ||f ||∞ . In fact we have |f (x)| ≤ ||f ||∞ everywhere because the set where
f (x) > ||f ||∞ is open (by continuity of f ) and negligible and so it is empty. Therefore sup |f | ≤
||f ||∞ . Conversely, sup |f | is essential bound for |f | so ||f ||∞ ≤ sup |f |.
On Lp (X), we define the relation f ∼ g if f = g a.e. It should be clear that this is an equivalence
relation which is moreover compatible with the addition an multiplication by scalars. The
quotient set of this relation is denoted by Lp (X). So an element of Lp (X) is a class of functions
that coincide a.e. between each other. But there is a natural structure of vector space on
Lp (X). Indeed, let us denote by [f ] the equivalence class containing the function f . Define and
addition and multiplication by a scalar by
This makes sense because if f1 ∈ [f ] and g1 ∈ [g], then f1 + g1 = f + g a.e. so [f1 + g1 ] = [f + g].
Also αf1 = αf a.e and so [αf1 ] = [αf ]. Now one can easily check that Lp is a vector space on
IR (the neutral element of the addition being [0], the class of all functions that are equal a.e.
to zero). Finally we set ||[f ]||p = ||f ||p . This also makes sense since if f ∼ g then ||f ||p = ||g||p .
And now ||f ||p is a norm on Lp since ||[f ]|| = 0 ⇒ f = 0 a.e., and so [f ] = [0].
Remark 3.5 In practise we write ||f ||p or ||f ||Lp instead of ||[f ]||p and of we think of elements
of Lp as functions that are defined a.e. and (not as classes). With this interpretation it makes
no sense to speak about the value of a function on a set of measure zero.
3.3 Properties
Theorem 3.1 (Riesz-Fisher) (Lp (X), || · ||p ) is a Banach space for every 1 ≤ p ≤ +∞.
Proof.
1) Suppose first that p = +∞. Let (fn ) be a Cauchy sequence in L∞ . For any k ∈ IN, there
exists Nk such that
1
||fm − fn ||∞ ≤ for all m, n ≥ Nk .
k
So there exists Ak of measure zero such that
1
|fm (x) − fn (x)| ≤ for all x ∈ X\Ak , m, n ≥ Nk (3.2)
k
S
Let A = k Ak . Then A is of measure zero and for all x ∈ X\A, the sequence (fn (x)) is a
Cauchy sequence in IR. It converges therefore to a function f (x). Letting m → ∞ in (3.2), we
get
1
|f (x) − fn (x)| ≤ for all x ∈ X\A, n ≥ Nk .
k
Setting f (x) = 0 for x ∈ A, we see that f ∈ L∞ (because |f (x)| ≤ |fN1 (x)| + 1 for all x ∈ X)
and ||f − fn ||∞ ≤ k1 for all n ≥ Nk . Consequently, ||f − fn ||∞ → 0.
2) Suppose now that 1 ≤ p < ∞. Let (fn ) be a Cauchy sequence in Lp . It is enough to show
that (fn ) contains a convergent subsequence. Since (fn ) is a Cauchy sequence, there exists n1
such that ||fm − fn ||p ≤ 21 for m, n ≥ n1 . Next, there exists n2 > n1 such that ||fm − fn ||p ≤ 212
for m, n ≥ n2 . Next, there exists n3 > n2 such that ||fm − fn ||p ≤ 213 for m, n ≥ n3 and so on.
This defines a subsequence (fnk ) such that
1
||fnk+1 − fnk ||p ≤ ∀k ≥ 1. (3.3)
2k
3.3. PROPERTIES 23
We show that (fnk ) converges in Lp . To simplify the notation we write fk instead of fnk , so
that
1
||fk+1 − fk ||p ≤ ∀k ≥ 1. (3.4)
2k
Let
n−1
X
gn (x) = |fk+1 (x) − fk (x)|.
k=1
Then by Minkowski’s inequality and (3.4)
n−1 n−1
X X 1
||gn ||p ≤ ||fk+1 − fk ||p ≤ < 1. (3.5)
2k
k=1 k=1
|fm (x) − fn (x)| ≤ |fm (x) − fm−1 (x)| + · · · + |fn+1 (x) − fn (x)| ≤ gm (x) − gn−1 (x) (3.6)
≤ gm (x). (3.7)
It follows that for almost every x ∈ X, (fn (x)) is a Cauchy sequence in IR. Since IR is complete
(fn (x)) converges almost everywhere to a limit f (x). It follows by letting m → ∞ in inequality
(3.7), that
|f (x) − fn (x)| ≤ g(x) a.e. (3.8)
Then first, f ∈ Lp since |f (x)| ≤ |fn (x)| + g(x). Second, |fn (x) − f (x)|p → 0 a.e. with
|fn (x) − f (x)|p ≤ g(x)p . Thus, by the dominated convergence theorem, ||fn − f ||p → 0.
Proposition 3.9 Let (fn ) be a sequence of Lp that converges in Lp to some function f . Then
there exists a subsequence (fnk ) such that
(a) fnk → f a.e.
Remark 3.6 We can formulate the following Lp version of the dominated convergence theorem.
Let p ∈ [1, ∞[ and let fn be a sequence of measurable functions such that 1. fn converges to
f a.e. and 2. |fn | ≤ g a.e. where g ∈ Lp . Then fn converges to f in Lp . Indeed, |fn − f |p
converges to zero a.e. and |fn − f |p ≤ 2p g p (because |fn − f | ≤ |fn | + |f | ≤ 2g). The classical
dominated convergence theorem yields the result. However the result is not true in L∞ (consider
for instance the sequence 1[n,∞[ ).
Proposition 3.10 Let (X, A, µ) be a measure space with µ(X) < ∞. If 1 ≤ p ≤ q ≤ ∞ then
Lq (X) ⊂ Lp (X) and the natural injection is continuous.
Suppose first that q = ∞ and p < ∞. Then X |f |p dµ ≤ ||f ||p∞ µ(X) and so
R
Proof.
||f ||p ≤ ||f ||∞ µ(X)1/p . This estimate proves the result. Next we assume that p < q < ∞. Let
q
f ∈ Lq . Applying Hölder’s inequality to |f |p and 1 with exponents α = q/p and β = q−p , we
obtain q−p
||f ||p ≤ ||f ||q µ(X) pq .
Remark 3.7 If µ(X) = ∞ then the conclusion need not be true. For instance the function f
defined by f (x) = √x12 +1 belongs to L2 (IR) but not to L1 so L2 " L1 .
Theorem 3.2 (Density theorem) Let Ω ⊂ IRN be an open set and let 1 ≤ p < ∞. Then
Cc (Ω) is dense in Lp (Ω).
To prove this theorem we need a lemma which hold in a more general context.
Lemma 3.1 Let (X, A, µ) be a measurable space. Then any function in Lp (X) is the limit in
Lp of a sequence of simple functions in Lp .
Proof. Let (gn ) be an increasing sequence of simple nonnegative functions that converges
(pointwise) to f + and let (hn ) be an increasing sequence of of simple nonnegative functions
that converges to f − . Let ϕn = gn − hn . Then ϕn → f pointwise and |f − ϕn | ≤ |f | + |ϕn | ≤
|f | + hn + gn ≤ 2|f |. Then |f − ϕn |p ≤ 2p |f |p . It follows first that ϕn ∈ Lp and second that
||f − ϕn ||p → 0 by the Lebesgue dominated convergence theorem.
Proof of Theorem 3.2. Thanks to Lemma 3.1, it is enough to prove that any simple
function in Lp (Ω) can be approximated in the Lp norm by a function in Cc (Ω). We prove this
in two steps.
Step 1. For simplicity, we will assume that Ω = IRN (we will prove a more general result in
the next chapter). Let f = 1A ∈ Lp where A is a Lebesgue measurable subset of IRN . Then
3.3. PROPERTIES 25
necessarily λN (A) = ||1A ||pp < +∞. Let > 0 be given. We know from an exercise of chapter 1
that
λN (A) = inf{λN (O)|O open and A ⊂ O}.
From a property of the inf, there exists an open set O such that A ⊂ O and λN (O) < λN (A)+p
Then
||1O − 1A ||p = ||1O\A ||p = λ(O\A)1/p < .
Set now, for each n ∈ IN∗ , On = O ∩ B(0, n). Then
1. On is open.
2. On is compact (closed and bounded).
S∞
3. The sequence (On ) is increasing and O = n=1 On .
This implies that 1O = lim 1On . Using the Lebesgue dominated convergence theorem we
see that ||1O − 1On ||p → 0. Therefore we can choose some n such that ||1O − 1On ||p < . Let
V = On . Then V is an open set whose closure is compact and
Now for each k ∈ IN∗ , set ϕk (x) = min(1, kd(x, V c )). Then first, ϕk is Lipshitz continuous with
constant k (check this). Second, supp ϕk ⊂ V . Therefore ϕk has a compact support; and so
ϕk ∈ Cc (IRN ). Third, ϕk → 1V pointwise. Again using the Lebesgue dominated convergence
theorem we see that ||ϕk − 1V ||p → 0. Thus we can choose k such that
Remark 3.8 The above theorem does not hold for p = +∞. We can prove for example that
the closure of Cc (IR) in L∞ (IR) is the space of continuous functions that vanish at infinity,
which is a proper subspace of L∞ (IR). Prove this if you want.
We end this chapter by a useful result that we shall use later.
Proposition 3.11 (Continuity in the mean) Let 1 ≤ p < +∞ and let f ∈ Lp (IRN ). Then
Z
lim |f (x + h) − f (x)|p dx = 0.
h→0 IRN
Otherwise sated, if τh f is the function defined by τh f (x) = f (x + h), then ||τh f − f ||p → 0 as
h → 0.
Proof. See the exercises.
26 CHAPTER 3. LP SPACES
Chapter 4
4.1 Convolutions
The convolution of two functions f, g : IRN → IR is formally defined by
Z
(f ∗ g)(x) = f (x − y)g(y) dy
IRN
There are conditions under which this integral is well defined. This integral exists if (1) f and g
are both Lebesgue measurable and nonnegative or (2) f ∈ L1 (IRN ) and g ∈ Lp (IRN ). We shall
deal with these cases, but first an example.
2 2 2
Example. e−x ∗ e−x = π2 e−x /2 .
p
27
28 CHAPTER 4. CONVOLUTIONS AND FOURIER TRANSFORMS
R
Definition 4.1 The function x 7→ IRN f (x − y)g(y) dy is denoted by f ∗ g, so that
Z
(f ∗ g)(x) = f (x − y)g(y) dy
IRN
1. f ∗ g = g ∗ f (commutativity).
2. f ∗ (g ∗ h) = (f ∗ g) ∗ h (associativity).
5. {f ∗ g 6= 0} ⊂ {f 6= 0} + {g 6= 0}.
The above two integrals are equal by the change of variable u = y + z; v = z whose Jacobian
is 1 (write the details).
4) Follows from Tonelli theorem and a change of variables (write the details).
5) Let x ∈ IRN be such that x ∈ / {f 6= 0} + {g 6= 0}. We claim that for any y ∈ IRN , either
g(y) = 0 or f (x−y) = 0. Indeed, otherwise, there exists y0 such that g(y0 ) 6= 0 and f (x−y0 ) 6= 0.
This means that y0 ∈ {g 6= 0} and x−y0 ∈ {f 6= 0}. Then x = (x−y0 )+y0 ∈ {f 6= 0}+{g 6= 0},
contrary to the assumption. It follows from the claim that g(y)f (x − y) = 0 for every y ∈ IRN .
This implies that (f ∗ g)(x) = 0.
4.1. CONVOLUTIONS 29
3. The class of f ∗ g in Lp (IRN ) depends only on the classes of f and g, that is, if f1 = f2
a.e. and g1 = g2 a.e, then f1 ∗ g1 = f2 ∗ g2 a.e.
By Fubini’s theorem, the partial function y 7→ f (x − y)g(y) is integrable for almost every x.
Therefore the convolution is well defined almost
R everywhere.
R
The inequality in point 2. follows from | h| ≤ |h| and Tonelli theorem (interchange of
the order of integration). Indeed,
Z Z Z Z Z
||f ∗ g||1 = |f ∗ g| dx = f (x − y)g(y) dy dx ≤ |f (x − y)||g(y)| dy dx
Z Z Z Z
= |f (x − y)||g(y)| dx dy = |g| |f | = ||f ||1 ||g||1 .
By case 1, the term on the right is finite for almost every x. This proves the integrability of
y 7→ f (x − y)g(y) for almost every x.
Raising to the power p both sides of the above inequality, we get
Z p
p p/p0
|(f ∗ g)(x)| ≤ |f (x − y)||g(y)| dy ≤ (|f | ∗ |g|p )(x)||f ||1 .
Integrating with respect to x and applying the estimate ||u ∗ v||1 ≤ ||u||1 ||v||1 established in
case 1, we see that f ∗ g ∈ Lp and
p/p0
||f ∗ g||pp ≤ ||f ||1 ||g p ||1 ||f ||1 ,
30 CHAPTER 4. CONVOLUTIONS AND FOURIER TRANSFORMS
that is,
p/p0
||f ∗ g||pp ≤ ||f ||1 ||g||p ||f ||1 .
Raising to 1/p, we finally get ||f ∗ g||p ≤ ||f ||1 ||g||p .
Case 3. p = +∞. This case is straightforward (write the details). But note that in this case
f ∗ g is defined everywhere. Indeed, we have for every x and almost every y, |f (x − y)g(y)| ≤
|f (x − y)|||g||∞ . Therefore
Z Z
|f (x − y)g(y)| dy ≤ ||g||∞ |f (x − y)| dy = ||g||∞ ||f ||1 .
Therefore the function x 7→ f (x − y)g(y) is summable for every x and not only for almost every
x.
3. Let f1 = f2 a.e. and g1 = g2 a.e, and let
Then
Proposition 4.4 Let p ∈ [1, ∞]. The convolution satisfies the following properties.
(iv) supp (f ∗ g) ⊂ supp (f ) + supp (g). In particular, if f and g have compact support, then
f ∗ g has also compact support.
Proof. The proof of the (i), (ii) and (iii) is similar to the proof of Proposition 4.2. As in
that proposition, we have {f ∗ g 6= 0} ⊂ {f 6= 0} + {g 6= 0} ⊂ supp (f ) + supp (g). Taking the
closure, we get the result. Finally, the sum of two compact sets of IRN is a compact set.
Corollary 4.1 For every p ∈ [1, ∞], the convolution defines a bilinear continuous map from
L1 × Lp to Lp and from Lp × L1 to Lp .
∂ |α|
Dα f = .
∂xα1 1 · · · ∂xαNN
4.1. CONVOLUTIONS 31
∂7f
For example, D(1,4,2) f = . The space of infinitely many times differentiable functions
∂x∂y 4 ∂z 2
with compact support is denoted by Cc∞ (IRN ), that is, Cc∞ (IRN ) = C ∞ (IRN ) ∩ Cc (IRN ). This
space is not trivial for it contains the function ρ defined by
( 1
1
e ||x||2 −1 if ||x|| < 1
2
ρ(x) = e ||x|| −1 1B(0,1) =
0 if ||x|| ≥ 1
where ||x|| denotes the Euclidean norm of x ∈ IRN . We will prove in fact that for any open
subset Ω ⊂ IRN , the space Cc∞ (Ω) is dense in Lp (Ω).
Proposition 4.5 Let f ∈ L1 (IRN ) and g ∈ Cck (IRN ). Then f ∗ g ∈ C k (IRN ) and for every
multi-index α ∈ INN such that |α| ≤ k, we have
Dα (f ∗ g) = f ∗ Dα g.
Proof. Note first that our assumptions imply that f ∗g is defined everywhere because g ∈ L∞ .
We prove first the result for k = 0. We need to show the continuity of f ∗ g. Since g is
continuous, the function x 7→ f (y)g(x − y) is continuous and |f (y)g(x − y)| ≤ M |f (y)| where
M =Rsup |g|. By a theorem on the continuity of integrals depending on a parameter, the function
x 7→ IRN f (y)g(x − y) is continuous, that is f ∗ g is continuous.
∂F ∂g
We prove next the result for k = 1. Let F (x, y) = f (y)g(x − y). Then ∂xi
= f (y) ∂x i
(x − y)
∂F ∂g
and so | ∂x i
| ≤ M |f (y)| where M = sup | ∂x i
| and f ∈ L1 . The result follows from the theorem
on differentiation under the integral sign. The relation
∂(f ∗ g)
Z
∂g
(x) = f (y) (x − y) dy
∂xi IRN ∂xi
∗g)
shows that ∂(f
∂xi is continuous (by the theorem on the continuity of integrals depending on a
parameter). The general case follows by induction on k.
Remark 4.1 The result of the previous proposition remains true if f ∈ Lp for an arbitrary
p ∈ [1, +∞]. Indeed, let x0 ∈ IRN be given. We will prove that f ∗ g is differentiable on a
neighborhood of x0 . Since g has compact support, the set B(x0 , 1) − supp g is bounded and so
it is contained in a compact set K. Now we can write
Z
(f ∗ g)(x) = f (y)g(x − y) dy ∀x ∈ B(x0 , 1)
K
because if y ∈
/ K, then x − y ∈ / supp g and so g(x − y) = 0. Keeping the same notation as
in the above proof, we have | ∂xi | ≤ M |f (y)| and now f ∈ Lp (IRN ) ⊂ Lp (K) ⊂ L1 (K). The
∂F
Corollary 4.2 Let f ∈ L1 (IRN ) and g ∈ Cc∞ (IRN ). Then f ∗ g ∈ C ∞ (IRN ) ∩ Lp (IRN ) for any
p ∈ [1, ∞].
Definition 4.2 A sequence of mollifiers1 is a sequence of functions (ρn ) ⊂ Cc∞ (IRN ) such that
1. ρn ≥ 0.
2. supp ρn ⊂ B 0 (0, n1 ).
R
3. IRN ρn = 1.
1
Suite régularisante.
32 CHAPTER 4. CONVOLUTIONS AND FOURIER TRANSFORMS
1
A sequence of mollifiers always exist. Indeed, let for example ρ be defined by e ||x||2 −1 1B(0,1) and
let Z −1
ρn (x) = ρ nN ρ(nx).
Theorem 4.1 Let (ρn ) be a sequence of mollifiers and let f ∈ Lp (IRN ) for some p ∈ [1, +∞[.
Then
lim ||f ∗ ρn − f ||p = 0,
n→∞
that is, the sequence (f ∗ ρn ) converges to f in Lp .
We prove this theorem in two steps. Let us record the first step in a lemma.
Lemma 4.1 Let (ρn ) be a sequence of mollifiers and let f ∈ Cc (IRN ). Then (f ∗ ρn ) converges
to f uniformly on compact subsets of IRN . In particular for any compact subset K ⊂ IRN ,
||ρn ∗ f − f ||Lp (K) → 0 as n → ∞.
Proof. Let K ⊂ IRN be a fixed compact set and let ε > 0 be given. Uniform continuity of f
on the compact set K + B 0 (0, 1) implies that there exists δ > 0 (depending on K and ε) such
that
|f (x − y) − f (x)| < ε ∀x ∈ K, ∀y ∈ B(0, δ).
Z
Since ρn (y) dy = 1, we can write
IRN
Z Z
(f ∗ ρn )(x) − f (x) = f (x − y)ρn (y) dy − f (x) ρn (y) dy
N IRN
ZIR
= [f (x − y) − f (x)] ρn (y) dy
N
ZIR
= [f (x − y) − f (x)] ρn (y) dy
1
B(0, n )
Let therefore n > 1δ and x ∈ K. Then n1 < δ and so B(0, n1 ) ⊂ B(0, δ). If follows that
|f (x − y) − f (x)| < ε when y ∈ B(0, 1/n) and so
Z
|(f ∗ ρn )(x) − f (x)| ≤ ε ρn (y) dy = ε.
1
B(0, n )
and so
||f ∗ ρn − f ||Lp (K) ≤ λN (K)1/p ε
This means that ||ρn ∗ f − f ||Lp (K) → 0.
Proof of Theorem 4.1. Let ε > 0 be given. By Theorem 3.2, there exists f1 ∈ Cc (IRN )
such that ||f − f1 ||p < ε. Now we know from a previous proposition that
supp (f1 ∗ ρn ) ⊂ supp (f1 ) + B 0 (0, 1/n) ⊂= supp (f1 ) + B 0 (0, 1/n) ⊂ supp (f1 ) + B 0 (0, 1).
4.2. FOURIER TRANSFORMS 33
If we set K := supp (f1 ) + B 0 (0, 1), we see that K is a fixed compact set (independent of n)
that contains both supp (f1 ∗ ρn ) and supp (f1 ).
By the previous lemma
Z
|(f1 ∗ ρn )(x) − f1 (x)|p dx → 0.
K
But Z Z
|(f1 ∗ ρn )(x) − f1 (x)|p = |(f1 ∗ ρn )(x) − f1 (x)|p dx.
IRN K
Therefore ||f1 ∗ ρn − f1 ||p → 0 and so ||f1 ∗ ρn − f1 ||p < ε for all n large enough.
Now by the triangle inequality
By Young’s inequality, ||(f − f1 ) ∗ ρn ||p ≤ ||f − f1 ||p ||ρn ||1 = ||f − f1 ||p < ε. Therefore
Corollary 4.4 Let Ω ⊂ IRN be open and let 1 ≤ p < ∞. Then Cc∞ (Ω) is dense in Lp (Ω).
Proof. Let f ∈ Lp (Ω). We will find a sequence (fn ) in Cc∞ (Ω) such that ||fn − f ||Lp (Ω) → 0.
Let (
f (x) if x ∈ Ω
f¯(x) =
0 otherwise.
It follows from the dominated convergence theorem that ||gn − f¯||p → 0 (check this). By the
previous theorem ||f¯ ∗ ρn − f¯||p → 0. Hence ||fn − f ||Lp (Ω) → 0.
Definition 4.3 Let f ∈ L1 (IRN ). The Fourier transform of f is the complex valued function
fˆ defined by Z
fˆ(ω) = e−2πiω·x f (x) dx.
IRN
∗ g = fˆĝ.
Proposition 4.7 Let f, g ∈ L1 . Then f[
Proof. Recall that f ∗ g ∈ L1 and so f[ ∗ g exists. Recall also that, by Tonelli’s theorem, the
function (x, y) 7→ f (x − y)g(y) is integrable on IRN × IRN . Therefore, by Fubini’s theorem,
Z
f ∗ g(ω) = (f ∗ g)(x)e−2πiω·x dx
[
Z Z
= f (x − y)g(y) dy e−2πiω·x dx
Z Z
= f (x − y)e−2πiω·(x−y) g(y)e−2πiω·y dx dy
Z Z
= g(y)e−2πiω·y f (x − y)e−2πiω·(x−y) dx dy
= fˆ(ω)ĝ(ω).
Proposition 4.8 (Exchange theorem) Let f, g ∈ L1 . Then f ĝ and fˆg are integrable and
Z Z
f ĝ dλN = fˆg dλN .
Proof. f ĝ is integrable because f is integrable and ĝ is bounded. Similarly for fˆg. By Tonelli’s
theorem the function (x, ω) 7→ f (ω)g(x)e−2iπω·x is integrable and so by Fubini’s theorem,
Z Z Z
−2πiω·x
f (ω)ĝ(ω)dω = f (ω) g(x)e dx dω
Z Z
−2πiω·x
= g(x) f (ω)e dω dx
Z
= g(x)fˆ(x) dx.
Then F has exactly the same properties as the operator F. Observe also the following: given
a function f , let f˜ be defined by f˜(x) = f (−x). Then F(f ) = F(f˜).
36 CHAPTER 4. CONVOLUTIONS AND FOURIER TRANSFORMS
4.2.2 Differentiability
Lemma 4.3 Let f : IR → C be continuously differentiable, integrable and such that f 0 is
integrable. Then limx→±∞ f (x) = 0.
Rx
Proof. RBy the fundamental theorem of calculus, f (x) = f (0) + 0 f 0 (t) dt. Since f 0 is
x
integrable, 0 f 0 (t) dt has a limit as x → ±∞. Therefore f has a limit as x → ±∞. Since f is
integrable, the limit is necessarily zero.
Remark 4.3 If g : IR → C is integrable then g need not have a limit at ±∞ (give an example).
However, if g is integrable and has a limit at +∞ or −∞, then necessarily this limit is zero.
Indeed, let L = limx→∞ g(x). Then, for every ε > 0 there exists a ∈ IR such that |g(x) − L| < ε
for Rall x ≥ a. If L 6= 0, take ε = |L|/2 > 0. Then, by the triangle inequality, |g(x)| > |L|/2 and
∞ R∞
so a |g(x)| dx ≥ a (|L|/2) dx = ∞. But this contradicts the integrability of g. Therefore
L = 0. A similar reasoning yields limx→−∞ g(x) = 0.
Proposition 4.10 Let k ∈ IN∗ and let f : IRN → C be a function such that M α f ∈ L1 for
every multi-index α such that |α| ≤ k. Then fˆ ∈ C k and
Dα fˆ = (−2πi)|α| M
[ αf .
Remark 4.4 In the literature, there are other definitions of the Fourier transform. For example
some authors define the Fourier transform by
Z
1
fˆ(ω) = f (x)e−iω·x dx.
(2π)N/2 IRN
With this definition we have
4.2. FOURIER TRANSFORMS 37
∗ g = (2π)N/2 fˆĝ.
1. f[
Proof. We prove the theorem for N = 1 because it is easier to write and understand.
1. By the Leibniz rule (f g)(n) = nk=0 Cnk f (k )g (n−k) . Therefore
P
n
X
xm (f g)n = Cnk xm f (k) g (n−k) .
k=0
is therefore enough (and necessary) to prove that S ∈ L1 . So let f ∈ S Then |f (x)| ≤ A and
A+B
|x2 f (x)| ≤ B. Therefore |f (x)| ≤ 1+x 1
2 and the last function is in L .
Then the inversion theorem implies that f = FF(f ) if both f and F(f ) are in L1 . Now if
f ∈ S then by the above F(f ) ∈ S ⊂ L1 and so we have f = FF(f ) . This identity proves
that F is injective. Now let f ∈ S. Then, f and fˆ belong to S ⊂ L1 . Let h(ω) = fˆ(−ω). Then
h ∈ S. By the inversion theorem,
Z Z
f (x) = ˆ
f (ω)e+2πix·ω
dω = fˆ(−ω)e−2πix·ω dω by the change of variables ω → −ω
IRN IRN
Z
= h(ω)e−2πix·ω dω = F(h)(x)
IRN
Thus, f = F(h). This proves that F is surjective. The proof shows that in fact F −1 = F.
Proposition 4.12 For all f, g ∈ S, we have F(f ∗ g) = F(f )F(g) and F(f g) = F(f ) ∗ F(g).
Proof. The first identity holds for functions in L1 and so also for functions in S. Exactly in
the same way F −1 takes convolutions into products. Let now f, g ∈ S, and let ϕ = F(f ) and
ψ = F(g). Then
Remark 4.5 It follows from the computations above that S is stable under convolutions.
In the following, L2 denotes the space L2 (IRN , C) equipped with the complex dot product
Z
hf, gi = f ḡ dλ.
IRN
Proposition 4.13 For all f, g ∈ S, we have hfˆ, ĝi = hf, gi. In particular ||fˆ||2 = ||f ||2 .
4.2. FOURIER TRANSFORMS 39
Proof. Let f, g ∈ S. Using the exchange theorem and the inversion theorem we get
Z Z
¯
ˆ ˆ¯
hf , ĝi = f (t)g((t) dt = fˆ(t)g((−t)
ˆ ˆ dt
Z Z Z
ˆ
ˆ
= f (t)ḡ(−t) dt = f (−t)ḡ(−t) dt = f (t)ḡ(t) dt = hf, gi.
It follows that F is an isometry between (S, || · ||2 ) into (L2 , || · ||2 ). Since S is dense in L2
and L2 is complete, there is a unique continuous extension of F to L2 . Indeed, let f ∈ L2 ,
then there exists a sequence (fn ) in S that converges to f in L2 . Then F(fn ) is a Cauchy
sequence L2 because ||F(fn ) − F(fm )|| = ||fn − fm ||. Therefore F(fn ) is convergent in L2 . We
set G(f ) = lim F(fn ). This makes sense because the result is independent of the sequence fn .
Indeed, if gn is another sequence of S that converges to f , then ||F(fn ) − F(gn )|| = ||fn − gn ||.
To prove uniqueness of this extension, let G 0 be another continuous extension to L2 of F.
Then G and G 0 coincide on S. Let f ∈ S and let (fn ) be a sequence in S that converges to f in
L2 . Then G(fn ) = G 0 (fn ). Letting n → ∞, we get G(f ) = G 0 (f ).
It is not difficult to see that G is also linear and bounded.
In the sequel we still denote this extension by F.
It follows that ||F(f )||2 = ||F(f˜)||2 = ||f˜||2 = ||f ||2 . This means that F is also bounded. Next,
let fn be a sequence of S that converges to f in L2 . We know that FFfn = fn . By continuity
of F and F we get FFf = f . Similarly we get FF(f ) = f .
sin(2πω)
Example of use of Parseval’s identity. Let f = 1[−1,1] . We know that fˆ(ω) = πω .
Parseval’s identity implies that
sin(2πω) 2
Z ∞ Z ∞
(1[−1,1] )2 dx = dω.
−∞ −∞ πω
sin(2πω) 2
Z ∞
Therefore dω = 2. By the change of variables t = 2πω, we get
−∞ πω
sin t 2 sin t 2
Z ∞ Z ∞
π
dt = π and so dt = .
−∞ t 0 t 2