Anda di halaman 1dari 137

Fourier Analysis

Translation by Olof Staans of the lecture notes


Fourier analyysi
by
Gustaf Gripenberg
January 5, 2009
0
Contents
0 Integration theory 3
1 Finite Fourier Transform 10
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.2 L
2
-Theory (Energy theory) . . . . . . . . . . . . . . . . . . . . 14
1.3 Convolutions (Faltningar) . . . . . . . . . . . . . . . . . . . . . 21
1.4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
1.4.1 Wirtingers Inequality . . . . . . . . . . . . . . . . . . . . 31
1.4.2 Weierstrass Approximation Theorem . . . . . . . . . . . . 32
1.4.3 Solution of Dierential Equations . . . . . . . . . . . . . . 33
1.4.4 Solution of Partial Dierential Equations . . . . . . . . . . 35
2 Fourier Integrals 36
2.1 L
1
-Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.2 Rapidly Decaying Test Functions . . . . . . . . . . . . . . . . . . 43
2.3 L
2
-Theory for Fourier Integrals . . . . . . . . . . . . . . . . . . . 45
2.4 An Inversion Theorem . . . . . . . . . . . . . . . . . . . . . . . . 48
2.5 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2.5.1 The Poisson Summation Formula . . . . . . . . . . . . . . 52
2.5.2 Is

L
1
(1) = C
0
(1) ? . . . . . . . . . . . . . . . . . . . . . . 53
2.5.3 The Euler-MacLauren Summation Formula . . . . . . . . . 54
2.5.4 Schwartz inequality . . . . . . . . . . . . . . . . . . . . . . 55
2.5.5 Heisenbergs Uncertainty Principle . . . . . . . . . . . . . 55
2.5.6 Weierstrass Non-Dierentiable Function . . . . . . . . . . 56
2.5.7 Dierential Equations . . . . . . . . . . . . . . . . . . . . 59
2.5.8 Heat equation . . . . . . . . . . . . . . . . . . . . . . . . . 63
2.5.9 Wave equation . . . . . . . . . . . . . . . . . . . . . . . . 64
1
CONTENTS 2
3 Fourier Transforms of Distributions 67
3.1 What is a Measure? . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.2 What is a Distribution? . . . . . . . . . . . . . . . . . . . . . . . 69
3.3 How to Interpret a Function as a Distribution? . . . . . . . . . . . 71
3.4 Calculus with Distributions . . . . . . . . . . . . . . . . . . . . . 73
3.5 The Fourier Transform of a Distribution . . . . . . . . . . . . . . 76
3.6 The Fourier Transform of a Derivative . . . . . . . . . . . . . . . 77
3.7 Convolutions (Faltningar) . . . . . . . . . . . . . . . . . . . . . 79
3.8 Convergence in o

. . . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.9 Distribution Solutions of ODE:s . . . . . . . . . . . . . . . . . . . 87
3.10 The Support and Spectrum of a Distribution . . . . . . . . . . . . 89
3.11 Trigonometric Polynomials . . . . . . . . . . . . . . . . . . . . . . 93
3.12 Singular dierential equations . . . . . . . . . . . . . . . . . . . . 95
4 Fourier Series 99
5 The Discrete Fourier Transform 102
5.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.2 FFT=the Fast Fourier Transform . . . . . . . . . . . . . . . . . . 104
5.3 Computation of the Fourier Coecients of a Periodic Function . . 107
5.4 Trigonometric Interpolation . . . . . . . . . . . . . . . . . . . . . 113
5.5 Generating Functions . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.6 One-Sided Sequences . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.7 The Polynomial Interpretation of a Finite Sequence . . . . . . . . 118
5.8 Formal Power Series and Analytic Functions . . . . . . . . . . . . 119
5.9 Inversion of (Formal) Power Series . . . . . . . . . . . . . . . . . . 121
5.10 Multidimensional FFT . . . . . . . . . . . . . . . . . . . . . . . . 122
6 The Laplace Transform 124
6.1 General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
6.2 The Standard Laplace Transform . . . . . . . . . . . . . . . . . . 125
6.3 The Connection with the Fourier Transform . . . . . . . . . . . . 126
6.4 The Laplace Transform of a Distribution . . . . . . . . . . . . . . 129
6.5 Discrete Time: Z-transform . . . . . . . . . . . . . . . . . . . . . 130
6.6 Using Laguerra Functions and FFT to Compute Laplace Transforms131
Chapter 0
Integration theory
This is a short summary of Lebesgue integration theory, which will be used in
the course.
Fact 0.1. Some subsets (=delmangder) E 1 = (, ) are measurable
(=matbara) in the Lebesgue sense, others are not.
General Assumption 0.2. All the subsets E which we shall encounter in this
course are measurable.
Fact 0.3. All measurable subsets E 1 have a measure (=matt) m(E), which
in simple cases correspond to the total length of the set. E.g., the measure of
the interval (a, b) is b a (and so is the measure of [a, b] and [a, b)).
Fact 0.4. Some sets E have measure zero, i.e., m(E) = 0. True for example if
E consists of nitely many (or countably many) points. (mattet noll)
The expression a.e. = almost everywhere (n.o. = nastan overallt) means that
something is true for all x 1, except for those x which belong to some set E
with measure zero. For example, the function
f(x) =
_
1, [x[ 1
0, [x[ > 1
is continuous almost everywhere. The expression f
n
(x) f(x) a.e. means that
the measure of the set x 1 for which f
n
(x) f(x) is zero.
Think: In all but nitely many points (this is a simplication).
3
CHAPTER 0. INTEGRATION THEORY 4
Notation 0.5. 1 = (, ), C = complex plane.
The set of Riemann integrable functions f : I C (I 1 is an interval) such
that
_
I
[f(x)[
p
dx < , 1 p < ,
though much larger than the space C(I) of continuous functions on I, is not
big enough for our purposes. This defect can be remedied by the use of the
Lebesgue integral instead of the Riemann integral. The Lebesgue integral is
more complicated to dene and develop than the Riemann integral, but as a tool
it is easier to use as it has better properties. The main dierence between the
Riemann and the Lebesgue integral is that the former uses intervals and their
lengths while the latter uses more general point sets and their measures.
Denition 0.6. A function f : I C (I 1 is an interval) is measurable if
there exists a sequence of continuous functions f
n
so that
f
n
(x) f(x) for almost all x I
(i.e., the set of points x I for which f
n
(x) f(x) has measure zero).
General Assumption 0.7. All the functions that we shall encounter in this
course are measurable.
Thus, the word measurable is understood throughout (when needed).
Denition 0.8. Let 1 p < , and I 1 an interval. We write f L
p
(I) if
(f is measurable and)
_
I
[f(x)[
p
dx < .
We dene the norm of f in L
p
(I) to be
|f|
L
p
(I)
=
__
I
[f(x)[
p
dx
_
1/p
.
Physical interpretation:
p = 1 |f|
L
1
(I)
=
_
I
[f(x)[dx
CHAPTER 0. INTEGRATION THEORY 5
= the total mass. Probability density if f(x) 0, or a size of the total
population.
p = 2 |f|
L
2
(I)
=
__
I
[f(x)[
2
dx
_
1/2
= total energy (e.g. in an electrical signal, such as alternating current).
These two cases are the two important ones (we ignore the rest). The third
important case is p = .
Denition 0.9. f L

(I) if (f is measurable and) there exists a number


M < such that
[f(x)[ < M a.e.
The norm of f is
|f|
L

(I)
= infM : [f(x)[ M a.e.,
and it is denoted by
|f|
L

(I)
= ess sup
xI
[f(x)[
(essential supremum, vasentligt supremum).
Think: |f|
L

(I)
= the largest value of f in I if we ignore a set of measure
zero. For example:
f(x) =
_

_
0, x < 0
2, x = 0
1, x > 0
|f|
L

(I)
= 1.
Denition 0.10. C

C
(1) = D = the set of (real or complex-valued) functions
on 1 which can be dierentiated as many times as you wish, and which vanish
outside of a bounded interval (such functions do exist!). C

C
(I) = the same
thing, but the function vanish outside of I.
I
= 0 = 0
Infinitely many
derivatives
CHAPTER 0. INTEGRATION THEORY 6
Theorem 0.11. Let I 1 be an interval. Then C

C
(I) is dense in L
p
(I) for
all p, 1 p < (but not in L

(I)). That is, for every f L


p
(I) it is possible
to nd a sequence f
n
C

C
(I) so that
lim
n
|f
n
f|
L
p
(I)
= 0.
Proof. Straightforward (but takes a lot of work).
Theorem 0.12 (Fatous lemma). Let f
n
(x) 0 and let f
n
(x) f(x) a.e. as
n . Then
_
I
f(x)dx lim
n
_
I
f
n
(x)dx
(if the latter limit exists). Thus,
_
I
_
lim
n
f
n
(x)
_
dx lim
n
_
I
f
n
(x)dx
if f
n
0 (f can have no more total mass than f
n
, but it may have less). Often
we have equality, but not always.
Ex.
f
n
(x) =
_
n, 0 x 1/n
0, otherwise.
Homework: Compute the limits above in this case.
Theorem 0.13 (Monotone Convergence Theorem). If
0 f
1
(x) f
2
(x) . . .
and f
n
(x) f(x) a.e., then
_
I
f(x)dx = lim
n
_
I
f
n
(x)dx ( ).
Thus, for a positive increasing sequence we have
_
I
_
lim
n
f
n
(x)
_
dx = lim
n
_
I
f
n
(x)dx
(the mass of the limit is the limit of the masses).
Theorem 0.14 (Lebesgues dominated convergence theorem). (Extremely use-
ful)
If f
n
(x) f(x) a.e. and [f
n
(x)[ g(x) a.e. and
_
I
g(x)dx < (i.e., g L
1
(I)),
CHAPTER 0. INTEGRATION THEORY 7
then
_
I
f(x)dx =
_
I
_
lim
n
f
n
(x)
_
dx = lim
n
_
I
f
n
(x)dx.
Theorem 0.15 (Fubinis theorem). (Very useful for multiple integrals).
If f (is measurable and)
_
I
_
J
[f(x, y)[dy dx <
then the double integral
__
IJ
f(x, y)dy dx
is well-dened, and equal to
=
_
xI
__
yJ
f(x, y)dy
_
dx
=
_
yJ
__
xI
f(x, y)dx
_
dy
If f 0, then all three integrals are well-dened, possibly = , and if one of
them is < , then so are the others, and they are equal.
Note: These theorems are very useful, and often easier to use than the corre-
sponding theorems based on the Rieman integral.
Theorem 0.16 (Integration by parts `a la Lebesgue). Let [a, b]be a nite interval,
u L
1
([a, b]), v L
1
([a, b]),
U(t) = U(a) +
_
t
a
u(s)ds, V (t) = V (a) +
_
t
a
v(s)ds, t [a, b].
Then
_
b
a
u(t)V (t)dt = [U(t)V (t)]
b
a

_
b
a
U(t)v(t)dt.
Proof.
_
b
a
u(t)V (t) =
_
b
a
u(t)
_
t
a
v(s)dsdt
Fubini
=
_
U(b) U(a)
_
V (a) +
_
b
a
__
b
s
u(t)dt
_
v(s)ds.
Since
_
b
s
u(t)dt =
_
_
b
a

_
s
a
_
u(t)dt = U(b) U(a)
_
s
a
u(t)dt = U(b) U(s),
CHAPTER 0. INTEGRATION THEORY 8
we get
_
b
a
u(t)V (t)dt =
_
U(b) U(a)
_
V (a) +
_
b
a
(U(b) U(s)) v(s)ds
=
_
U(b) U(a)
_
V (a) + U(b)
_
V (b) V (a)
_

_
b
a
U(s)v(s)ds
= U(b)V (b) U(a)V (a)
_
b
a
U(s)v(s)ds.
Example 0.17. Sometimes we need test functions with special properties. Let
us take a look how one can proceed.
1 t
b(t)
b(t) =
_
e

1
t(1t)
, 0 < t < 1
0 , otherwise.
Then we can show that b C

(1), and b is a test function with compact support.


Let B(t) =
_
t

b(s)ds and norm it F(t) =


B(t)
B(1)
.
1
1
F(t)
t
F(t) =
_

_
0 , t 0
1 , t 1
increase , 0 < t < 1.
Further F(t) + F(t 1) = 1, t 1, clearly true for t 0 and t 1.
For 0 < t < 1 we check the derivative
d
dt
(F(t) F(1 t)) =
1
B(1)
[B

(t) B

(1 t)] =
1
B(1)
[b(t) b(1 t)] = 0.
CHAPTER 0. INTEGRATION THEORY 9
F(t) F(1t)
Let G(t) = F(Nt). Then G increases from 0 to 1 on the interval 0 t
1
N
.
G(t) + G(
1
N
t) = F(Nt) F(1 Nt) = 1, t 1.
1/N 1
G(t) G(1/Nt)
Chapter 1
The Fourier Series of a Periodic
Function
1.1 Introduction
Notation 1.1. We use the letter with a double meaning:
a) = [0, 1)
b) In the notations L
p
(), C(), C
n
() and C

() we use the letter to


imply that the functions are periodic with period 1, i.e., f(t + 1) = f(t)
for all t 1. In particular, in the continuous case we require f(1) = f(0).
Since the functions are periodic we know the whole function as soon as we
know the values for t [0, 1).
Notation 1.2. |f|
L
p
(T)
=
_
_
1
0
[f(t)[
p
dt
_
1/p
, 1 p < . |f|
C(T)
= max
tT
[f(t)[
(f continuous).
Denition 1.3. f L
1
() has the Fourier coecients

f(n) =
_
1
0
e
2int
f(t)dt, n Z,
where Z = 0, 1, 2, . . . . The sequence

f(n)
nZ
is the (nite) Fourier
transform of f.
Note:

f(n) =
_
s+1
s
e
2int
f(t)dt s 1,
10
CHAPTER 1. FINITE FOURIER TRANSFORM 11
since the function inside the integral is periodic with period 1.
Note: The Fourier transform of a periodic function is a discrete sequence.
Theorem 1.4.
i) [

f(n)[ |f|
L
1
(T)
, n Z
ii) lim
n

f(n) = 0.
Note: ii) is called the RiemannLebesgue lemma.
Proof.
i) [

f(n)[ = [
_
1
0
e
2int
f(t)dt[
_
1
0
[e
2int
f(t)[dt =
_
1
0
[f(t)[dt = |f|
L
1
(T)
(by
the triangle inequality for integrals).
ii) First consider the case where f is continuously dierentiable, with f(0) = f(1).
Then integration by parts gives

f(n) =
_
1
0
e
2int
f(t)dt
=
1
2in
_
e
2int
f(t)

1
0
+
1
2in
_
1
0
e
2int
f

(t)dt
= 0 +
1
2in

(n), so by i),
[

f(n)[ =
1
2n
[

(n)[
1
2n
_
1
0
[f

(s)[ds 0 as n .
In the general case, take f L
1
() and > 0. By Theorem 0.11 we can
choose some g which is continuously dierentiable with g(0) = g(1) = 0 so
that
|f g|
L
1
(T)
=
_
1
0
[f(t) g(t)[dt /2.
By i),
[

f(n)[ = [

f(n) g(n) + g(n)[


[

f(n) g(n)[ +[ g(n)[


|f g|
L
1
(T)
+[ g(n)[
/2 +[ g(n)[.
By the rst part of the proof, for n large enough, [ g(n)[ /2, and so
[

f(n)[ .
This shows that [

f(n)[ 0 as n .
CHAPTER 1. FINITE FOURIER TRANSFORM 12
Question 1.5. If we know

f(n)

n=
then can we reconstruct f(t)?
Answer: is more or less Yes.
Denition 1.6. C
n
() = n times continuously dierentiable functions, periodic
with period 1. (In particular, f
(k)
(1) = f
(k)
(0) for 0 k n.)
Theorem 1.7. For all f C
1
() we have
f(t) = lim
N
M
N

n=M

f(n)e
2int
, t 1. (1.1)
We shall see later that the convergence is actually uniform in t.
Proof. Step 1. We shift the argument of f by replacing f(s) by g(s) = f(s+t).
Then
g(n) = e
2int

f(n),
and (1.1) becomes
f(t) = g(0) = lim
N
M
N

n=M

f(n)e
2int
.
Thus, it suces to prove the case where t = 0 .
Step 2: If g(s) is the constant function g(s) g(0) = f(t), then (1.1) holds since
g(0) = g(0) and g(n) = 0 for n ,= 0 in this case. Replace g(s) by
h(s) = g(s) g(0).
Then h satises all the assumptions which g does, and in addition, h(0) = 0.
Thus it suces to prove the case where both t = 0 and f(0) = 0. For simplicity
we write f instead of h, but we suppose below that t = 0 and f(0) = 0
Step 2: Dene
g(s) =
_
f(s)
e
2is
1
, s ,= integer (=heltal)
if

(0)
2
, s = integer.
For s = n = integer we have e
2is
1 = 0, and by lHospitals rule
lim
sn
g(s) = lim
s0
f

(s)
2ie
2is
=
f

(s)
2i
= g(n)
CHAPTER 1. FINITE FOURIER TRANSFORM 13
(since e
i2n
= 1). Thus g is continuous. We clearly have
f(s) =
_
e
2is
1
_
g(s), (1.2)
so

f(n) =
_
T
e
2ins
f(s)ds (use (1.2))
=
_
T
e
2ins
(e
2is
1)g(s)ds
=
_
T
e
2i(n+1)s
g(s)ds
_
T
e
2ins
g(s)ds
= g(n + 1) g(n).
Thus,
N

n=M

f(n) = g(N + 1) g(M) 0


by the RiemanLebesgue lemma (Theorem 1.4)
By working a little bit harder we get the following stronger version of Theorem
1.7:
Theorem 1.8. Let f L
1
(), t
0
1, and suppose that
_
t
0
+1
t
0
1

f(t) f(t
0
)
t t
0

dt < . (1.3)
Then
f(t
0
) = lim
N
M
N

n=M

f(n)e
2int
0
t 1
Proof. We can repeat Steps 1 and 2 of the preceding proof to reduce the The-
orem to the case where t
0
= 0 and f(t
0
) = 0. In Step 3 we dene the function
g in the same way as before for s ,= n, but leave g(s) undened for s = n.
Since lim
s0
s
1
(e
2is
1) = 2i ,= 0, the function g belongs to L
1
() if and
only if condition (1.3) holds. The continuity of g was used only to ensure that
g L
1
(), and since g L
1
() already under the weaker assumption (1.3), the
rest of the proof remains valid without any further changes.
CHAPTER 1. FINITE FOURIER TRANSFORM 14
Summary 1.9. If f L
1
(), then the Fourier transform

f(n)

n=
of f is
well-dened, and

f(n) 0 as n . If f C
1
(), then we can reconstruct f
from its Fourier transform through
f(t) =

n=

f(n)e
2int
_
= lim
N
M
N

n=M

f(n)e
2int
_
.
The same reconstruction formula remains valid under the weaker assumption of
Theorem 1.8.
1.2 L
2
-Theory (Energy theory)
This theory is based on the fact that we can dene an inner product (scalar
product) in L
2
(), namely
f, g) =
_
1
0
f(t)g(t)dt, f, g L
2
().
Scalar product means that for all f, g, h L
2
()
i) f + g, h) = f, h) +g, h)
ii) f, g) = f, g) C
iii) g, f) = f, g) (complex conjugation)
iv) f, f) 0, and = 0 only when f 0.
These are the same rules that we know from the scalar products in C
n
. In
addition we have
|f|
2
L
2
(T)
=
_
T
[f(t)[
2
dt =
_
T
f(t)f(t)dt = f, f).
This result can also be used to dene the Fourier transform of a function f
L
2
(), since L
2
() L
1
().
Lemma 1.10. Every function f L
2
() also belongs to L
1
(), and
|f|
L
1
(T)
|f|
L
2
(T)
.
CHAPTER 1. FINITE FOURIER TRANSFORM 15
Proof. Interpret
_
T
[f(t)[dt as the inner product of [f(t)[ and g(t) 1. By
Schwartz inequality (see course on Analysis II),
[f, g)[ =
_
T
[f(t)[ 1dt |f|
L
2 |g|
L
2 = |f|
L
2
(T)
_
T
1
2
dt = |f|
L
2
(T)
.
Thus, |f|
L
1
(T)
|f|
L
2
(T)
. Therefore:
f L
2
(t) =
_
T
[f(t)[dt <
=

f(n) =
_
T
e
2int
f(t)dt is dened for all n.
It is not true that L
2
(1) L
1
(1). Counter example:
f(t) =
1

1 + t
2
_

_
L
2
(1)
/ L
1
(1)
C

(1)
(too large at ).
Notation 1.11. e
n
(t) = e
2int
, n Z, t 1.
Theorem 1.12 (Plancherels Theorem). Let f L
2
(). Then
i)

n=
[

f(n)[
2
=
_
1
0
[f(t)[
2
dt = |f|
2
L
2
(T)
,
ii) f =

n=

f(n)e
n
in L
2
() (see explanation below).
Note: This is a very central result in, e.g., signal processing.
Note: It follows from i) that the sum

n=
[

f(n)[
2
always converges if f L
2
()
Note: i) says that

n=
[

f(n)[
2
= the square of the total energy of the Fourier coecients
= the square of the total energy of the original signal f
=
_
T
[f(t)[
2
dt
Note: Interpretation of ii): Dene
f
M,N
=
N

n=M

f(n)e
n
=
N

n=M

f(n)e
2int
.
CHAPTER 1. FINITE FOURIER TRANSFORM 16
Then
lim
M
N
|f f
M,N
|
2
= 0
lim
M
N
_
1
0
[f(t) f
M,N
(t)[
2
dt = 0
(f
M,N
(t) need not converge to f(t) at every point, and not even almost every-
where).
The proof of Theorem 1.12 is based on some auxiliary results:
Theorem 1.13. If g
n
L
2
(), f
N
=

N
n=0
g
n
, g
n
g
m
, and

n=0
|g
n
|
2
L
2
(T)
< ,
then the limit
f = lim
N
N

n=0
g
n
exists in L
2
.
Proof. Course on Analysis II and course on Hilbert Spaces.
Interpretation: Every orthogonal sum with nite total energy converges.
Lemma 1.14. Suppose that

n=
[c(n)[ < . Then the series

n=
c(n)e
2int
converges uniformly to a continuous limit function g(t).
Proof.
i) The series

n=
c(n)e
2int
converges absolutely (since [e
2int
[ = 1), so
the limit
g(t) =

n=
c(n)e
2int
exist for all t 1.
ii) The convergens is uniform, because the error
[
m

n=m
c(n)e
2int
g(t)[ = [

|n|>m
c(n)e
2int
[

|n|>m
[c(n)e
2int
[
=

|n|>m
[c(n)[ 0 as m .
CHAPTER 1. FINITE FOURIER TRANSFORM 17
iii) If a sequence of continuous functions converge uniformly, then the limit is
continuous (proof Analysis II).
proof of Theorem 1.12. (Outline)
0 |f f
M,N
|
2
= f f
M,N
, f f
M,N
)
= f, f)
. .
I
f
M,N
, f)
. .
II
f, f
M,N
)
. .
III
+f
M,N
, f
M,N
)
. .
IV
I = f, f) = |f|
2
L
2
(T)
.
II =
N

n=M

f(n)e
n
, f) =
N

n=M

f(n)e
n
, f)
=
N

n=M

f(n)f, e
n
) =
N

n=M

f(n)

f(n)
=
N

n=M
[

f(n)[
2
.
III = (the complex conjugate of II) = II.
IV =
_
N

n=M

f(n)e
n
,
N

m=M

f(m)e
m
_
=
N

n=M

f(n)

f(m) e
n
, e
m
)
. .

m
n
=
N

n=M
[

f(n)[
2
= II = III.
Thus, adding I II III + IV = I II 0, i.e.,
|f|
2
L
2
(T)

N

n=M
[

f(n)[
2
0.
This proves Bessels inequality

n=
[

f(n)[
2
|f|
2
L
2
(T)
. (1.4)
How do we get equality?
By Theorem 1.13, applied to the sums
N

n=0

f(n)e
n
and
1

n=M

f(n)e
n
,
CHAPTER 1. FINITE FOURIER TRANSFORM 18
the limit
g = lim
M
N
f
M,N
= lim
M
N
N

n=M

f(n)e
n
(1.5)
does exist. Why is f = g? (This means that the sequence e
n
is complete!). This
is (in principle) done in the following way
i) Argue as in the proof of Theorem 1.4 to show that if f C
2
(), then
[

f(n)[ 1/(2n)
2
|f

|
L
1 for n ,= 0. In particular, this means that

n=
[

f(n)[ < . By Lemma 1.14, the convergence in (1.5) is actually


uniform, and by Theorem 1.7, the limit is equal to f. Uniform convergence
implies convergence in L
2
(), so even if we interpret (1.5) in the L
2
-sense,
the limit is still equal to f a.e. This proves that f
M,N
f in L
2
() if
f C
2
().
ii) Approximate an arbitrary f L
2
() by a function h C
2
() so that
|f h|
L
2
(T)
.
iii) Use i) and ii) to show that |f g|
L
2
(T)
, where g is the limit in (1.5).
Since is arbitrary, we must have g = f.
Denition 1.15. Let 1 p < .

p
(Z) = set of all sequences a
n

n=
satisfying

n=
[a
n
[
p
< .
The norm of a sequence a
p
(Z) is
|a|

p
(Z)
=
_

n=
[a
n
[
p
_
1/p
Analogous to L
p
(I):
p = 1 |a|

1
(Z)
= total mass (probability),
p = 2 |a|

2
(Z)
= total energy.
In the case of p = 2 we also dene an inner product
a, b) =

n=
a
n
b
n
.
CHAPTER 1. FINITE FOURIER TRANSFORM 19
Denition 1.16.

(Z) = set of all bounded sequences a


n

n=
. The norm
in

(Z) is
|a|

(Z)
= sup
nZ
[a
n
[.
For details: See course in Analysis II.
Denition 1.17. c
0
(Z) = the set of all sequences a
n

n=
satisfying lim
n
a
n
= 0.
We use the norm
|a|
c
0
(Z)
= max
nZ
[a
n
[
in c
0
(Z).
Note that c
0
(Z)

(Z), and that


|a|
c
0
(Z)
= |a|

(Z)
if a

n=
c
0
(Z).
Theorem 1.18. The Fourier transform maps L
2
() one to one onto
2
(Z), and
the Fourier inversion formula (see Theorem 1.12 ii) maps
2
(Z) one to one onto
L
2
(). These two transforms preserves all distances and scalar products.
Proof. (Outline)
i) If f L
2
() then

f
2
(Z). This follows from Theorem 1.12.
ii) If a
n

n=

2
(Z), then the series
N

n=M
a
n
e
2int
converges to some limit function f L
2
(). This follows from Theorem
1.13.
iii) If we compute the Fourier coecients of f, then we nd that a
n
=

f(n).
Thus, a
n

n=
is the Fourier transform of f. This shows that the Fourier
transform maps L
2
() onto
2
(Z).
iv) Distances are preserved. If f L
2
(), g L
2
(), then by Theorem 1.12
i),
|f g|
L
2
(T)
= |

f(n) g(n)|

2
(Z)
,
i.e.,
_
T
[f(t) g(t)[
2
dt =

n=
[

f(n) g(n)[
2
.
CHAPTER 1. FINITE FOURIER TRANSFORM 20
v) Inner products are preserved:
_
T
[f(t) g(t)[
2
dt = f g, f g)
= f, f) f, g) g, f) +g, g)
= f, f) f, g) f, g) +g, g)
= f, f) +g, g) 2Ref, g).
In the same way,

n=
[

f(n) g(n)[
2
=

f g,

f g)
=

f,

f) + g, g) 2

f, g).
By iv), subtracting these two equations from each other we get
f, g) =

f, g).
If we replace f by if, then
Imf, g) = Re if, g) = if, g)
= i

f, g) = Re i

f, g)
= Im

f, g).
Thus, f, g)
L
2
(R)
=

f, g)

2
(Z)
, or more explicitely,
_
T
f(t)g(t)dt =

n=

f(n) g(n). (1.6)


This is called Parsevals identity.
Theorem 1.19. The Fourier transform maps L
1
() into c
0
(Z) (but not onto),
and it is a contraction, i.e., the norm of the image is the norm of the original
function.
Proof. This is a rewritten version of Theorem 1.4. Parts i) and ii) say that

f(n)

n=
c
0
(Z), and part i) says that |

f(n)|
c
0
(Z)
|f|
L
1
(T)
.
The proof that there exist sequences in c
0
(Z) which are not the Fourier transform
of some function f L
1
() is much more complicated.
CHAPTER 1. FINITE FOURIER TRANSFORM 21
1.3 Convolutions (Faltningar)
Denition 1.20. The convolution (faltningen) of two functions f, g L
1
()
is
(f g)(t) =
_
T
f(t s)g(s)ds,
where
_
T
=
_
+1

for all 1, since the function s f(t s)g(s) is periodic.


Note: In this integral we need values of f and g outside of the interval [0, 1), and
therefore the periodicity of f and g is important.
Theorem 1.21. If f, g L
1
(), then (f g)(t) is dened almost everywhere,
and f g L
1
(). Furthermore,
_
_
f g
_
_
L
1
(T)

_
_
f
_
_
L
1
(T)
_
_
g
_
_
L
1
(T)
(1.7)
Proof. (We ignore measurability)
We begin with (1.7)
_
_
f g
_
_
L
1
(T)
=
_
T
[(f g)(t)[ dt
=
_
T

_
T
f(t s)g(s) ds

dt
-ineq.

_
tT
_
sT
[f(t s)g(s)[ ds dt
Fubini
=
_
sT
__
tT
[f(t s)[ dt
_
[g(s)[ ds
Put v=ts,dv=dt
=
_
sT
__
vT
[f(v)[ dv
_
. .
=f
L
1
(T)
[g(s)[ ds
= |f|
L
1
(T)
_
sT
[g(s)[ ds = |f|
L
1
(T)
|g|
L
1
(T)
This integral is nite. By Fubinis Theorem 0.15
_
T
f(t s)g(s)ds
is dened for almost all t.
Theorem 1.22. For all f, g L
1
() we have
(

f g)(n) =

f(n) g(n), n Z
CHAPTER 1. FINITE FOURIER TRANSFORM 22
Proof. Homework.
Thus, the Fourier transform maps convolution onto pointwise multiplication.
Theorem 1.23. If k C
n
() (n times continuously dierentiable) and f L
1
(),
then k f C
n
(), and (k f)
(m)
(t) = (k
(m)
f)(t) for all m = 0, 1, 2, . . . , n.
Proof. (Outline) We have for all h > 0
1
h
[(k f) (t + h) (k f)(t)] =
1
h
_
1
0
[k(t + h s) k(t s)] f(s)ds.
By the mean value theorem,
k(t + h s) = k(t s) + hk

(),
for some [t s, t s+h], and
1
h
[k(t +hs) k(t s)] = f() k

(t s) as
h 0, and

1
h
[k(t +hs) k(t s)]

= [f

()[ M, where M = sup


T
[k

(s)[. By
the Lebesgue dominated convergence theorem (which is true also if we replace
n by h 0)(take g(x) = M[f(x)[)
lim
h0
_
1
0
1
h
[k(t + h s) k(t s)]f(s) ds =
_
1
0
k

(t s)f(s) ds,
so k f is dierentiable, and (k f)

= k

f. By repeating this n times we nd


that k f is n times dierentiable, and that (k f)
(n)
= k
(n)
f. We must still
show that k
(n)
f is continuous. This follows from the next lemma.
Lemma 1.24. If k C() and f L
1
(), then k f C().
Proof. By Lebesgue dominated convergence theorem (take g(t) = 2|k|
C(T)
f(t)),
(k f)(t + h) (k f)(t) =
_
1
0
[k(t + h s) k(t s)]f(s)ds 0 as h 0.
Corollary 1.25. If k C
1
() and f L
1
(), then for all t 1
(k f)(t) =

n=
e
2int

k(n)

f(n).
Proof. Combine Theorems 1.7, 1.22 and 1.23.
Interpretation: This is a generalised inversion formula. If we choose

k(n) so
that
CHAPTER 1. FINITE FOURIER TRANSFORM 23
i)

k(n) 1 for small [n[
ii)

k(n) 0 for large [n[,
then we set a ltered approximation of f, where the high frequences (= high
values of [n[) have been damped but the low frequences (= low values of [n[)
remain. If we can take

k(n) = 1 for all n then this gives us back f itself, but this
is impossible because of the Riemann-Lebesgue lemma.
Problem: Find a good function k C
1
() of this type.
Solution: The Fejer kernel is one possibility. Choose

k(n) to be a triangular
function:
n = 0 n = 4
F (n)
4
Fix m = 0, 1, 2 . . . , and dene

F
m
(n) =
_
m+1|n|
m+1
, [n[ m
0 , [n[ > m
(,= 0 in 2m + 1 points.)
We get the corresponding time domain function F
m
(t) by using the invertion
formula:
F
m
(t) =
m

n=m

F
m
(n)e
2int
.
Theorem 1.26. The function F
m
(t) is explicitly given by
F
m
(t) =
1
m + 1
sin
2
((m + 1)t)
sin
2
(t)
.
Proof. We are going to show that
m

j=0
j

n=j
e
2int
=
_
sin((m + 1)t)
sin t
_
2
when t ,= 0.
CHAPTER 1. FINITE FOURIER TRANSFORM 24
Let z = e
2it
, z = e
2it
, for t ,= n, n = 0, 1, 2, . . . . Also z ,= 1, and
j

n=j
e
2int
=
j

n=0
e
2int
+
j

n=1
e
2int
=
j

n=0
z
n
+
j

n=1
z
n
=
1 z
j+1
1 z
+
z(1 z
j
)
1 z
=
1 z
j+1
1 z
+
=1
..
z z(1 z
j
)
z z z
..
=1
=
z
j
z
j+1
1 z
.
Hence
m

j=0
j

n=j
e
2int
=
m

j=0
z
j
z
j+1
1 z
=
1
1 z
_
m

j=0
z
j

j=0
z
j+1
_
=
1
1 z
_
1 z
m+1
1 z
z
_
1 z
m+1
1 z
__
=
1
1 z
_
1 z
m+1
1 z

=1
..
z z(1 z
m+1
z(1 z)
_
=
1
1 z
_
1 z
m+1
1 z

1 z
m+1
z 1
_
=
z
m+1
+ 2 z
m+1
[1 z[
2
.
sin t =
1
2i
(e
it
e
it
), cos t =
1
2
(e
it
+ e
it
). Now
[1 z[ = [1 e
2it
[ = [e
it
(e
it
e
it
)[ = [e
it
e
it
[ = 2[sin(t)[
and
z
m+1
2 + z
m+1
= e
2i(m+1)
2 + e
2i(m+1)
=
_
e
i(m+1)
e
i(m+1)
_
2
=
_
2i sin((m + 1))
_
2
.
Hence
m

j=0
j

n=j
e
2int
=
4(sin((m + 1)))
2
4(sin(t))
2
=
_
sin((m + 1))
sin(t)
_
2
Note also that
m

j=0
j

n=j
e
2int
=
m

n=m
m

j=|n|
e
2int
=
m

n=m
(m + 1 [n[)e
2int
.
CHAPTER 1. FINITE FOURIER TRANSFORM 25
-2 -1 1 2
1
2
3
4
5
6
Comment 1.27.
i) F
m
(t) C

() (innitely many derivatives).


ii) F
m
(t) 0.
iii)
_
T
[F
m
(t)[dt =
_
T
F
m
(t)dt =

F
m
(0) = 1,
so the total mass of F
m
is 1.
iv) For all , 0 < <
1
2
,
lim
m
_
1

F
m
(t)dt = 0,
i.e. the mass of F
m
gets concentrated to the integers t = 0, 1, 2 . . .
as m .
Denition 1.28. A sequence of functions F
m
with the properties i)-iv) above is
called a (periodic) approximate identity. (Often i) is replaced by F
m
L
1
().)
Theorem 1.29. If f L
1
(), then, as m ,
i) F
m
f f in L
1
(), and
ii) (F
m
f)(t) f(t) for almost all t.
Here i) means that
_
T
[(F
m
f)(t) f(t)[dt 0 as m
Proof. See page 27.
By combining Theorem 1.23 and Comment 1.27 we nd that F
m
f C

().
This combined with Theorem 1.29 gives us the following periodic version of
Theorem 0.11:
CHAPTER 1. FINITE FOURIER TRANSFORM 26
Corollary 1.30. For every f L
1
() and > 0 there is a function g C

()
such that |g f|
L
1
(T)
.
Proof. Choose g = F
m
f where m is large enough.
To prove Theorem 1.29 we need a number of simpler results:
Lemma 1.31. For all f, g L
1
() we have f g = g f
Proof.
(f g)(t) =
_
T
f(t s)g(s)ds
ts=v,ds=dv
=
_
T
f(v)g(t v)dv = (g f)(t)
We also need:
Theorem 1.32. If g C(), then F
m
g g uniformly as m , i.e.
max
tR
[(F
m
g)(t) g(t)[ 0 as m .
Proof.
(F
m
g)(t) g(t)
Lemma 1.31
= (g F
m
)(t) g(t)
Comment 1.27
= (g F
m
)(t) g(t)
_
T
F
m
(s)ds
=
_
T
[g(t s) g(t)]F
m
(s)ds.
Since g is continuous and periodic, it is uniformly continuous, and given > 0
there is a > 0 so that [g(t s) g(t)[ if [s[ . Split the intgral above
into (choose the interval of integration to be [
1
2
,
1
2
])
_ 1
2

1
2
[g(t s) g(t)]F
m
(s)ds =
_
_

1
2
..
I
+
_

..
II
+
_ 1
2

..
III
_
[g(t s) g(t)]F
m
(s)ds
Let M = sup
tR
[g(t)[. Then [g(t s) g(t)[ 2M, and
[I + III[
_
_

1
2
+
_ 1
2

_
2MF
m
(s)ds
= 2M
_
1

F
m
(s)ds
CHAPTER 1. FINITE FOURIER TRANSFORM 27
and by Comment 1.27iv) this goes to zero as m . Therefore, we can choose
m so large that
[I + III[ (m m
0
, and m
0
large.)
[II[
_

[g(t s) g(t)[F
m
(s)ds

_

F
m
(s)ds

_ 1
2

1
2
F
m
(s)ds =
Thus, for m m
0
we have
[(F
m
g)(t) g(t)[ 2 (for all t).
Thus, lim
m
sup
tR
[(F
m
g)(t) g(t)[ = 0, i.e., (F
m
g)(t) g(t) uniformly
as m .
The proof of Theorem 1.29 also uses the following weaker version of Lemma 0.11:
Lemma 1.33. For every f L
1
() and > 0 there is a function g C() such
that |f g|
L
1
(T)
.
Proof. Course in Lebesgue integration theory.
(We already used a stronger version of this lemma in the proof of Theorem 1.12.)
Proof of Theorem 1.29, part i): (The proof of part ii) is bypassed, typically
proved in a course on integration theory.)
Let > 0, and choose some g C() with |f g|
L
1
(T)
. Then
|F
m
f f|
L
1
(T)
|F
m
g g + F
m
(f g) (f g)|
L
1
(t)
|F
m
g g|
L
1
(T)
+|F
m
(f g)|
L
1
(T)
+|(f g)|
L
1
(T)
Thm 1.21
|F
m
g g|
L
1
(T)
+ (|F
m
|
L
1
(T)
+ 1
. .
=2
) |f g|
L
1
(T)
. .

= |F
m
g g|
L
1
(T)
+ 2.
CHAPTER 1. FINITE FOURIER TRANSFORM 28
Now |F
m
g g|
L
1
(T)
=
_
1
0
[(F
m
g(t) g(t)[ dt

_
1
0
max
s[0,1]
[(F
m
g(s) g(s)[ dt
= max
s[0,1]
[(F
m
g(s) g(s)[
_
1
0
dt
. .
=1
.
By Theorem 1.32, this tends to zero as m . Thus for large enough m,
|F
m
f f|
L
1
(T)
3,
so F
m
f f in L
1
() as m .
(Thus, we have almost proved Theorem 1.29 i): we have reduced it to a proof
of Lemma 1.33 and other standard properties of integrals.)
In the proof of Theorem 1.29 we used the trivial triangle inequality in L
1
():
|f + g|
L
1
(T)
=
_
[f(t) + g(t)[ dt
_
[f(t)[ +[g(t)[ dt
= |f|
L
1
(T)
+|g|
L
1
(T)
Similar inequalities are true in all L
p
(), 1 p , and a more sophisticated
version of the preceding proof gives:
Theorem 1.34. If 1 p < and f L
p
(), then F
m
f f in L
p
() as
m , and also pointwise a.e.
Proof. See Gripenberg.
Note: This is not true in L

(). The correct L

-version is given in Theorem


1.32.
Corollary 1.35. (Important!) If f L
p
(), 1 p < , or f C
n
(), then
lim
m
m

n=m
m + 1 [n[
m + 1

f(n)e
2int
= f(t),
where the convergence is in the norm of L
p
, and also pointwise a.e. In the case
where f C
n
() we have uniform convergence, and the derivatives of order n
also converge uniformly.
CHAPTER 1. FINITE FOURIER TRANSFORM 29
Proof. By Corollary 1.25 and Comment 1.27,
m

n=m
m + 1 [n[
m + 1

f(n)e
2int
= (F
m
f)(t)
The rest follows from Theorems 1.34, 1.32, and 1.23, and Lemma 1.31.
Interpretation: We improve the convergence of the sum

n=

f(n)e
2int
by multiplying the coecients by the damping factors
m+1|n|
m+1
, [n[ m. This
particular method is called Cesaro summability. (Other summability methods
use other damping factors.)
Theorem 1.36. (Important!) The Fourier coecients

f(n), n Z of a func-
tion f L
1
() determine f uniquely a.e., i.e., if

f(n) = g(n) for all n, then
f(t) = g(t) a.e.
Proof. Suppose that g(n) =

f(n) for all n. Dene h(t) = f(t) g(t). Then

h(n) =

f(n) g(n) = 0, n Z. By Theorem 1.29,
h(t) = lim
m
m

n=m
m + 1 [n[
m + 1

h(n)
..
=0
e
2int
= 0
in the L
1
-sense, i.e.
|h| =
_
1
0
[h(t)[ dt = 0
This implies h(t) = 0 a.e., so f(t) = g(t) a.e.
Theorem 1.37. Suppose that f L
1
() and that

n=
[

f(n)[ < . Then the


series

n=

f(n)e
2int
converges uniformly to a continuous limit function g(t), and f(t) = g(t) a.e.
Proof. The uniform convergence follows from Lemma 1.14. We must have
f(t) = g(t) a.e. because of Theorems 1.29 and 1.36.
The following theorem is much more surprising. It says that not every sequence
a
n

nZ
is the set of Fourier coecients of some f L
1
().
CHAPTER 1. FINITE FOURIER TRANSFORM 30
Theorem 1.38. Let f L
1
(),

f(n) 0 for n 0, and

f(n) =

f(n) (i.e.

f(n) is an odd function). Then


i)

n=1
1
n

f(n) <
ii)

n=
n=0
[
1
n

f(n)[ < .
Proof. Second half easy: Since

f is odd,

n=0
nZ
[
1
n

f(n)[ =

n>0
[
1
n

f(n)[ +

n<0
[
1
n

f(n)[
= 2

n=1
[
1
n

f(n)[ < if i) holds.


i): Note that

f(n) =

f(n) gives

f(0) = 0. Dene g(t) =
_
t
0
f(s)ds. Then
g(1) g(0) =
_
1
0
f(s)ds =

f(0) = 0, so that g is continuous. It is not dicult to
show (=homework) that
g(n) =
1
2in

f(n), n ,= 0.
By Corollary 1.35,
g(0) = g(0) e
2i00
. .
=1
+ lim
m
m

n=m
m + 1 [n[
m + 1
. .
even
g(n)
..
even
e
2in0
. .
=1
= g(0) +
2
2i
lim
m
m

n=0
m + 1 n
m + 1

f(n)
n
. .
0
.
Thus
lim
m
m

n=1
m + 1 n
m + 1

f(n)
n
= K = a nite pos. number.
In particular, for all nite M,
M

n=1

f(n)
n
= lim
m
M

n=1
m + 1 n
m + 1

f(n)
n
K,
and so

n=1

f(n)
n
K < .
Theorem 1.39. If f C
k
() and g = f
(k)
, then g(n) = (2in)
k

f(n), n Z.
Proof. Homework.
Note: True under the weaker assumption that f C
k1
(), g L
1
(), and
f
k1
(t) = f
k1
(0) +
_
t
0
g(s)ds.
CHAPTER 1. FINITE FOURIER TRANSFORM 31
1.4 Applications
1.4.1 Wirtingers Inequality
Theorem 1.40 (Wirtingers Inequality). Suppose that f L
2
(a, b), and that f
has a derivative in L
2
(a, b), i.e., suppose that
f(t) = f(a) +
_
t
a
g(s)ds
where g L
2
(a, b). In addition, suppose that f(a) = f(b) = 0. Then
_
b
a
[f(t)[
2
dt
_
b a

_
2
_
b
a
[g(t)[
2
dt (1.8)
_
=
_
b a

_
2
_
b
a
[f

(t)[
2
dt
_
.
Comment 1.41. A function f which can be written in the form
f(t) = f(a) +
_
t
a
g(s)ds,
where g L
1
(a, b) is called absolutely continuous on (a, b). This is the Lebesgue
version of dierentiability. See, for example, Rudins Real and Complex Anal-
ysis.
Proof. i) First we reduce the interval (a, b) to (0, 1/2): Dene
F(s) = f(a + 2(b a)s)
G(s) = F

(s) = 2(b a)g(a + 2(b a)s).


Then F(0) = F(1/2) = 0 and F(t) =
_
t
0
G(s)ds. Change variable in the integral:
t = a + 2(b a)s, dt = 2(b a)ds,
and (1.8) becomes
_
1/2
0
[F(s)[
2
ds
1
4
2
_
1/2
0
[G(s)[
2
ds. (1.9)
We extend F and G to periodic functions, period one, so that F is odd and G
is even: F(t) = F(t) and G(t) = G(t) (rst to the interval (1/2, 1/2) and
CHAPTER 1. FINITE FOURIER TRANSFORM 32
then by periodicity to all of 1). The extended function F is continuous since
F(0) = F(1/2) = 0. Then (1.9) becomes
_
T
[F(s)[
2
ds
1
4
2
_
T
[G(s)[
2
ds
|F|
L
2
(T)

1
2
|G|
L
2
(T)
By Parsevals identity, equation (1.6) on page 20, and Theorem 1.39 this is
equivalent to

n=
[

F(n)[
2

1
4
2

n=
[2n

F(n)[
2
. (1.10)
Here

F(0) =
_
1/2
1/2
F(s)ds = 0.
since F is odd, and for n ,= 0 we have (2n)
2
4
2
. Thus (1.10) is true.
Note: The constant
_
ba

_
2
is the best possible: we get equality if we take

F(1) ,= 0,

F(1) =

F(1), and all other



F(n) = 0. (Which function is this?)
1.4.2 Weierstrass Approximation Theorem
Theorem 1.42 (Weierstrass Approximation Theorem). Every continuous func-
tion on a closed interval [a, b] can be uniformly approximated by a polynomial:
For every > 0 there is a polynomial P so that
max
t[a,b]
[P(t) f(t)[ (1.11)
Proof. First change the variable so that the interval becomes [0, 1/2] (see
previous page). Then extend f to an even function on [1/2, 1/2] (see previous
page). Then extend f to a continuous 1-periodic function. By Corollary 1.35,
the sequence
f
m
(t) =
m

n=m

F
m
(n)

f(n)e
2int
(F
m
= Fejer kernel) converges to f uniformly. Choose m so large that
[f
m
(t) f(t)[ /2
for all t. The function f
m
(t) is analytic, so by the course in analytic functions,
the series

k=0
f
(k)
m
(0)
k!
t
k
CHAPTER 1. FINITE FOURIER TRANSFORM 33
converges to f
m
(t), uniformly for t [1/2, 1/2]. By taking N large enough we
therefore have
[P
N
(t) f
m
(t)[ /2 for t [1/2, 1/2],
where P
N
(t) =

N
k=0
f
(k)
m
(0)
k!
t
k
. This is a polynomial, and [P
N
(t) f(t)[
for t [1/2, 1/2]. Changing the variable t back to the original one we get a
polynomial satisfying (1.11).
1.4.3 Solution of Dierential Equations
There are many ways to use Fourier series to solve dierential equations. We
give only two examples.
Example 1.43. Solve the dierential equation
y

(x) + y(x) = f(x), 0 x 1, (1.12)


with boundary conditions y(0) = y(1), y

(0) = y

(1). (These are periodic bound-


ary conditions.) The function f is given, and C is a constant.
Solution. Extend y and f to all of 1 so that they become periodic, period
1. The equation + boundary conditions then give y C
1
(). If we in addition
assume that f L
2
(), then (1.12) says that y

= f y L
2
() (i.e. f

is
absolutely continuous).
Assuming that f C
1
() and that f

is absolutely continuous we have by one


of the homeworks

(y

)(n) = (2in)
2
y(n),
so by transforming (1.12) we get
4
2
n
2
y(n) + y(n) =

f(n), n Z, or (1.13)
( 4
2
n
2
) y(n) =

f(n), n Z.
Case A: ,= 4
2
n
2
for all n Z. Then (1.13) gives
y(n) =

f(n)
4
2
n
2
.
The sequence on the right is in
1
(Z), so y(n)
1
(Z). (i.e.,

[ y(n)[ < ). By
Theorem 1.37,
y(t) =

n=

f(n)
4
2
n
2
. .
= y(n)
e
2int
, t 1.
CHAPTER 1. FINITE FOURIER TRANSFORM 34
Thus, this is the only possible solution of (1.12).
How do we know that it is, indeed, a solution? Actually, we dont, but by working
harder, and using the results from Chapter 0, it can be shown that y C
1
(),
and
y

(t) =

n=
2in y(n)e
2int
,
where the sequence
2in y(n) =
2in y(n)
4
2
n
2
belongs to
1
(Z) (both
2in
4
2
n
2
and y(n) belongs to
2
(Z), and the product of two

2
-sequences is an
1
-sequence; see Analysis II). The sequence
(2in)
2
y(n) =
4
2
n
2
4
2
n
2

f(n)
is an
2
-sequence, and

n=
4
2
n
2
4
2
n
2

f(n) f

(t)
in the L
2
-sense. Thus, f C
1
(), f

is absolutely continuous, and equation


(1.12) holds in the L
2
-sense (but not necessary everywhere). (It is called a mild
solution of (1.12)).
Case B: = 4
2
k
2
for some k Z. Write
4
2
n
2
= 4
2
(k
2
n
2
) = 4
2
(k n)(k + n).
We get two additional necessary conditions:

f(k) = 0. (If this condition is not
true then the equation has no solutions.)
If

f(k) =

f(k) = 0, then we get innitely many solutions: Choose y(k) and
y(k) arbitrarily, and
y(n) =

f(n)
4
2
(k
2
n
2
)
, n ,= k.
Continue as in Case A.
Example 1.44. Same equation, but new boundary conditions: Interval is [0, 1/2],
and
y(0) = 0 = y(1/2).
CHAPTER 1. FINITE FOURIER TRANSFORM 35
Extend y and f to [1/2, 1/2] as odd functions
y(t) = y(t), 1/2 t 0
f(t) = f(t), 1/2 t 0
and then make them periodic, period 1. Continue as before. This leads to a
Fourier series with odd coecients, which can be rewritten as a sinus-series.
Example 1.45. Same equation, interval [0, 1/2], boundary conditions
y

(0) = 0 = y

(1/2).
Extend y and f to even functions, and continue as above. This leads to a solution
with even coecients y(n), and it can be rewritten as a cosinus-series.
1.4.4 Solution of Partial Dierential Equations
See course on special functions.
Chapter 2
Fourier Integrals
2.1 L
1
-Theory
Repetition: 1 = (, ),
f L
1
(1)
_

[f(t)[dt < (and f measurable)


f L
2
(1)
_

[f(t)[
2
dt < (and f measurable)
Denition 2.1. The Fourier transform of f L
1
(1) is given by

f() =
_

e
2it
f(t)dt, 1
Comparison to chapter 1:
f L
1
()

f(n) dened for all n Z
f L
1
(1)

f() dened for all 1
Notation 2.2. C
0
(1) = continuous functions f(t) satisfying f(t) 0 as
t . The norm in C
0
is
|f|
C
0
(R)
= max
tR
[f(t)[ (= sup
tR
[f(t)[).
Compare this to c
0
(Z).
Theorem 2.3. The Fourier transform T maps L
1
(1) C
0
(1), and it is a
contraction, i.e., if f L
1
(1), then

f C
0
(1) and |

f|
C
0
(R)
|f|
L
1
(R)
, i.e.,
36
CHAPTER 2. FOURIER INTEGRALS 37
i)

f is continuous
ii)

f() 0 as
iii) [

f()[
_

[f(t)[dt, 1.
Note: Part ii) is again the Riemann-Lesbesgue lemma.
Proof. iii) The same as the proof of Theorem 1.4 i).
ii) The same as the proof of Theorem 1.4 ii), (replace n by , and prove this
rst in the special case where f is continuously dierentiable and vanishes outside
of some nite interval).
i) (The only new thing):
[

f( + h)

f()[ =

_
R
_
e
2i(+h)t
e
2it
_
f(t)dt

_
R
_
e
2iht
1
_
e
2it
f(t)dt

-ineq.

_
R
[e
2iht
1[[f(t)[ dt 0 as h 0
(use Lesbesgues dominated convergens Theorem, e
2iht
1 as h 0, and
[e
2iht
1[ 2).
Question 2.4. Is it possible to nd a function f L
1
(1) whose Fourier trans-
form is the same as the original function?
Answer: Yes, there are many. See course on special functions. All functions
which are eigenfunctions with eigenvalue 1 are mapped onto themselves.
Special case:
Example 2.5. If h
0
(t) = e
t
2
, t 1, then

h
0
() = e

2
, 1
Proof. See course on special functions.
Note: After rescaling, this becomes the normal (Gaussian) distribution function.
This is no coincidence!
Another useful Fourier transform is:
Example 2.6. The Fejer kernel in L
1
(1) is
F(t) =
_
sin(t)
t
_
2
.
CHAPTER 2. FOURIER INTEGRALS 38
The transform of this function is

F() =
_
1 [[ , [[ 1,
0 , otherwise.
Proof. Direct computation. (Compare this to the periodic Fejer kernel on page
23.)
Theorem 2.7 (Basic rules). Let f L
1
(1), , 1
a) g(t) = f(t ) g() = e
2i

f()
b) g(t) = e
2it
f(t) g() =

f( )
c) g(t) = f(t) g() =

f()
d) g(t) = f(t) g() =

f()
e) g(t) = f(t) g() =

f(

) ( > 0)
f ) g L
1
and h = f g

h() =

f() g()
g)
g(t) = 2itf(t)
and g L
1
_

_

f C
1
(1), and

() = g()
h)
f is absolutely continuous
and f

= g L
1
(1)
_
g() = 2i

f().
Proof. (a)-(e): Straightforward computation.
(g)-(h): Homework(?) (or later).
The formal inversion for Fourier integrals is

f() =
_

e
2it
f(t)dt
f(t)
?
=
_

e
2it

f()d
This is true in some cases in some sense. To prove this we need some
additional machinery.
Denition 2.8. Let f L
1
(1) and g L
p
(1), where 1 p . Then we
dene
(f g)(t) =
_
R
f(t s)g(s)ds
for all those t 1 for which this integral converges absolutely, i.e.,
_
R
[f(t s)g(s)[ds < .
CHAPTER 2. FOURIER INTEGRALS 39
Lemma 2.9. With f and p as above, f g is dened a.e., f g L
p
(1), and
|f g|
L
p
(R)
|f|
L
1
(R)
|g|
L
p
(R)
.
If p = , then f g is dened everywhere and uniformly continuous.
Conclusion 2.10. If |f|
L
1
(R)
1, then the mapping g f g is a contraction
from L
p
(1) to itself (same as in periodic case).
Proof. p = 1: same proof as we gave on page 21.
p = : Boundedness of f g easy. To prove continuity we approximate f by a
function with compact support and show that |f(t) f(t +h)|
L
1 0 as h 0.
p ,= 1, : Signicantly harder, case p = 2 found in Gasquet.
Notation 2.11. B|c(1) = all bounded and continuous functions on 1. We
use the norm
|f|
BUC(R)
= sup
tR
[f(t)[.
Theorem 2.12 (Approximate identity). Let k L
1
(1),

k(0) =
_

k(t)dt =
1, and dene
k

(t) = k(t), t 1, > 0.


If f belongs to one of the function spaces
a) f L
p
(1), 1 p < (note: p ,= ),
b) f C
0
(1),
c) f B|c(1),
then k

f belongs to the same function space, and


k

f f as
in the norm of the same function space, i.e.,
|k

f f|
L
p
(R)
0 as if f L
p
(1)
sup
tR
[(k

f)(t) f(t)[ 0 as
_
if f B|c(1),
or f C
0
(1).
It also conveges a.e. if we assume that
_

0
(sup
s|t|
[k(s)[)dt < .
CHAPTER 2. FOURIER INTEGRALS 40
Proof. The same as the proofs of Theorems 1.29, 1.32 and 1.33. That is,
the computations stay the same, but the bounds of integration change ( 1),
and the motivations change a little (but not much).
Example 2.13 (Standard choices of k).
i) The Gaussian kernel
k(t) = e
t
2
,

k() = e

2
.
This function is C

and nonnegative, so
|k|
L
1 =
_
R
[k(t)[dt =
_
R
k(t)dt =

k(0) = 1.
ii) The Fejer kernel
F(t) =
sin(t)
2
(t)
2
.
It has the same advantages, and in addition

F() = 0 for [[ > 1.


The transform is a triangle:

F() =
_
1 [[, [[ 1
0, [[ > 1
1 1
F( )
iii) k(t) = e
2|t|
(or a rescaled version of this function. Here

k() =
1
1 + ()
2
, 1.
Same advantages (except C

)).
CHAPTER 2. FOURIER INTEGRALS 41
Comment 2.14. According to Theorem 2.7 (e),

k

()

k(0) = 1 as
, for all 1. All the kernels above are low pass lters (non causal).
It is possible to use one-sided (causal) lters instead (i.e., k(t) = 0 for
t < 0). Substituting these into Theorem 2.12 we get approximate identities,
which converge to a -distribution. Details later.
Theorem 2.15. If both f L
1
(1) and

f L
1
(1), then the inversion formula
f(t) =
_

e
2it

f()d (2.1)
is valid for almost all t 1. By redening f on a set of measure zero we can
make it hold for all t 1 (the right hand side of (2.1) is continuous).
Proof. We approximate
_
R
e
2it

f()d by
_
R
e
2it
e

f()d (where > 0 is small)


=
_
R
e
2it
2

2
_
R
e
2is
f(s)dsd (Fubini)
=
_
sR
f(s)
_
R
e
2i(st)
e

2
. .
k(
2
)
dds
. .
()
(Ex. 2.13 last page)
() The Fourier transform of k(
2
) at the point s t. By Theorem 2.7 (e) this
is equal to
=
1

k(
s t

) =
1

k(
t s

)
(since

k() = e

2
is even).
The whole thing is
_
sR
f(s)
1

k
_
t s

_
ds = (f k1

)(t) f L
1
(1)
as 0
+
according to Theorem 2.12. Thus, for almost all t 1,
f(t) = lim
0
_
R
e
2it
e

f()d.
On the other hand, by the Lebesgue dominated convergence theorem, since
[e
2it
e

f()[ [

f()[ L
1
(1),
lim
0
_
R
e
2it
e

f()d =
_
R
e
2it

f()d.
CHAPTER 2. FOURIER INTEGRALS 42
Thus, (2.1) holds a.e. The proof of the fact that
_
R
e
2it

f()d C
0
(1)
is the same as the proof of Theorem 2.3 (replace t by t).
The same proof also gives us the following approximate inversion formula:
Theorem 2.16. Suppose that k L
1
(1),

k L
1
(1), and that

k(0) =
_
R
k(t)dt = 1.
If f belongs to one of the function spaces
a) f L
p
(1), 1 p <
b) f C
0
(1)
c) f B|c(1)
then
_
R
e
2it

k()

f()d f(t)
in the norm of the given space (i.e., in L
p
-norm, or in the sup-norm), and also
a.e. if
_

0
(sup
s|t|
[k(s)[)dt < .
Proof. Almost the same as the proof given above. If k is not even, then we
end up with a convolution with the function k

(t) =
1

k(t/) instead, but we


can still apply Theorem 2.12 with k(t) replaced by k(t).
Corollary 2.17. The inversion in Theorem 2.15 can be interpreted as follows:
If f L
1
(1) and

f L
1
(1), then,

f(t) = f(t) a.e.


Here

f(t) = the Fourier transform of



f evaluated at the point t.
Proof. By Theorem 2.15,
f(t) =
_
R
e
2i(t)

f()d
. .
Fourier transform of

f at the point (t)
a.e.
CHAPTER 2. FOURIER INTEGRALS 43
Corollary 2.18.

f(t) = f(t) (If we repeat the Fourier transform 4 times, then


we get back the original function). (True at least if f L
1
(1) and

f L
1
(1). )
As a prelude (=preludium) to the L
2
-theory we still prove some additional results:
Lemma 2.19. Let f L
1
(1) and g L
1
(1). Then
_
R
f(t) g(t)dt =
_
R

f(s)g(s)ds
Proof.
_
R
f(t) g(t)dt =
_
tR
f(t)
_
sR
e
2its
g(s)dsdt (Fubini)
=
_
sR
__
tR
f(t)e
2ist
dt
_
g(s)ds
=
_
sR

f(s)g(s)ds.
Theorem 2.20. Let f L
1
(1), h L
1
(1) and

h L
1
(1). Then
_
R
f(t)h(t)dt =
_
R

f()

h()d. (2.2)
Specically, if f = h, then (f L
2
(1) and)
|f|
L
2
(R)
= |

f|
L
2
(R)
. (2.3)
Proof. Since h(t) =
_
R
e
2it

h()d we have
_
R
f(t)h(t)dt =
_
tR
f(t)
_
R
e
2it

h()ddt (Fubini)
=
_
sR
__
tR
f(t)e
2ist
dt
_

h()d
=
_
R

f()

h()d.
2.2 Rapidly Decaying Test Functions
(Snabbt avtagande testfunktioner).
Denition 2.21. o = the set of functions f with the following properties
i) f C

(1) (innitely many times dierentiable)


CHAPTER 2. FOURIER INTEGRALS 44
ii) t
k
f
(n)
(t) 0 as t and this is true for all
k, n Z
+
= 0, 1, 2, 3, . . . .
Thus: Every derivative of f 0 at innity faster than any negative power of t.
Note: There is no natural norm in this space (it is not a Banach space).
However, it is possible to nd a complete, shift-invariant metric on this space (it
is a Frechet space).
Example 2.22. f(t) = P(t)e
t
2
o for every polynomial P(t). For example,
the Hermite functions are of this type (see course in special functions).
Comment 2.23. Gripenberg denotes o by C

(1). The functions in o are called


rapidly decaying test functions.
The main result of this section is
Theorem 2.24. f o

f o
That is, both the Fourier transform and the inverse Fourier transform maps this
class of functions onto itself. Before proving this we prove the following
Lemma 2.25. We can replace condition (ii) in the denition of the class o by
one of the conditions
iii)
_
R
[t
k
f
(n)
(t)[dt < , k, n Z
+
or
iv) [
_
d
dt
_
n
t
k
f(t)[ 0 as t , k, n Z
+
without changing the class of functions o.
Proof. If ii) holds, then for all k, n Z
+
,
sup
tR
[(1 + t
2
)t
k
f
(n)
(t)[ <
(replace k by k + 2 in ii). Thus, for some constant M,
[t
k
f
(n)
(t)[
M
1 + t
2
=
_
R
[t
k
f
(n)
(t)[dt < .
Conversely, if iii) holds, then we can dene g(t) = t
k+1
f
(n)
(t) and get
g

(t) = (k + 1)t
k
f
(n)
(t)
. .
L
1
+t
k+1
f
(n+1)
(t)
. .
L
1
,
CHAPTER 2. FOURIER INTEGRALS 45
so g

L
1
(1), i.e.,
_

[g

(t)[dt < .
This implies
[g(t)[ [g(0) +
_
t
0
g

(s)ds[
[g(0)[ +
_
t
0
[g

(s)[ds
[g(0)[ +
_

[g

(s)[ds = [g(0)[ +|g

|
L
1,
so g is bounded. Thus,
t
k
f
(n)
(t) =
1
t
g(t) 0 as t .
The proof that ii) iv) is left as a homework.
Proof of Theorem 2.24. By Theorem 2.7, the Fourier transform of
(2it)
k
f
(n)
(t) is
_
d
d
_
k
(2i)
n

f().
Therefore, if f o, then condition iii) on the last page holds, and by Theorem
2.3,

f satises the condition iv) on the last page. Thus

f o. The same ar-
gument with e
2it
replaced by e
+2it
shows that if

f o, then the Fourier
inverse transform of

f (which is f) belongs to o.
Note: Theorem 2.24 is the basis for the theory of Fourier transforms of distribu-
tions. More on this later.
2.3 L
2
-Theory for Fourier Integrals
As we saw earlier in Lemma 1.10, L
2
() L
1
(). However, it is not true that
L
2
(1) L
1
(1). Counter example:
f(t) =
1

1 + t
2
_

_
L
2
(1)
, L
1
(1)
C

(1)
(too large at ).
So how on earth should we dene

f() for f L
2
(1), if the integral
_
R
e
2int
f(t)dt
CHAPTER 2. FOURIER INTEGRALS 46
does not converge?
Recall: Lebesgue integral converges converges absolutely
_
[e
2int
f(t)[dt < f L
1
(1).
We are saved by Theorem 2.20. Notice, in particular, condition (2.3) in that
theorem!
Denition 2.26 (L
2
-Fourier transform).
i) Approximate f L
2
(1) by a sequence f
n
o which converges to f in
L
2
(1). We do this e.g. by smoothing and cutting (utjamning och
klippning): Let k(t) = e
t
2
, dene
k
n
(t) = nk(nt), and
f
n
(t) = k
_
t
n
_
. .

(k
n
f)(t)
. .

. .
the product belongs to S
() this tends to zero faster than any polynomial as t .
() smoothing by an approximate identity, belongs to C

and is bounded.
By Theorem 2.12 k
n
f f in L
2
as n . The functions k
_
t
n
_
tend
to k(0) = 1 at every point t as n , and they are uniformly bounded
by 1. By using the appropriate version of the Lesbesgue convergence we
let f
n
f in L
2
(1) as n .
ii) Since f
n
converges in L
2
, also

f
n
must converge to something in L
2
. More
about this in Analysis II. This follows from Theorem 2.20. (f
n
f
f
n
Cauchy sequence

f
n
Cauchy seqence

f
n
converges.)
iii) Call the limit to which f
n
converges The Fourier transform of f, and
denote it

f.
Denition 2.27 (Inverse Fourier transform). We do exactly as above, but re-
place e
2it
by e
+2it
.
Final conclusion:
CHAPTER 2. FOURIER INTEGRALS 47
Theorem 2.28. The extended Fourier transform which we have dened above
has the following properties: It maps L
2
(1) one-to-one onto L
2
(1), and if

f is the
Fourier transform of f, then f is the inverse Fourier transform of

f. Moreover,
all norms, distances and inner products are preserved.
Explanation:
i) Normes preserved means
_
R
[f(t)[
2
dt =
_
R
[

f()[
2
d,
or equivalently, |f|
L
2
(R)
= |

f|
L
2
(R)
.
ii) Distances preserved means
|f g|
L
2
(R)
= |

f g|
L
2
(R)
(apply i) with f replaced by f g)
iii) Inner product preserved means
_
R
f(t)g(t)dt =
_
R

f() g()d,
which is often written as
f, g)
L
2
(R)
=

f, g)
L
2
(R)
.
This was theory. How to do in practice?
One answer: We saw earlier that if [a, b] is a nite interval, and if f L
2
[a, b]
f L
1
[a, b], so for each T > 0, the integral

f
T
() =
_
T
T
e
2it
f(t)dt
is dened for all 1. We can try to let T , and see what happens. (This
resembles the theory for the inversion formula for the periodical L
2
-theory.)
Theorem 2.29. Suppose that f L
2
(1). Then
lim
T
_
T
T
e
2it
f(t)dt =

f()
in the L
2
-sense as T , and likewise
lim
T
_
T
T
e
2it

f()d = f(t)
in the L
2
-sense.
CHAPTER 2. FOURIER INTEGRALS 48
Proof. Much too hard to be presented here. Another possibility: Use the Fejer
kernel or the Gaussian kernel, or any other kernel, and dene

f() = lim
n
_
R
e
2it
k
_
t
n
_
f(t)dt,
f(t) = lim
n
_
R
e
+2it

k
_

n
_

f()d.
We typically have the same type of convergence as we had in the Fourier inversion
formula in the periodic case. (This is a well-developed part of mathematics, with
lots of results available.) See Gripenbergs compendium for some additional
results.
2.4 An Inversion Theorem
From time to time we need a better (= more useful) inversion theorem for the
Fourier transform, so let us prove one here:
Theorem 2.30. Suppose that f L
1
(1) + L
2
(1) (i.e., f = f
1
+ f
2
, where
f
1
L
1
(1) and f
2
L
2
(1)). Let t
0
1, and suppose that
_
t
0
+1
t
0
1

f(t) f(t
0
)
t t
0

dt < . (2.4)
Then
f(t
0
) = lim
S
T
_
T
S
e
2it
0
f()d, (2.5)
where

f() =

f
1
() +

f
2
().
Comment: Condition (2.4) is true if, for example, f is dierentiable at the point
t
0
.
Proof. Step 1. First replace f(t) by g(t) = f(t + t
0
). Then
g() = e
2it
0
f(),
and (2.5) becomes
g(0) = lim
S
T
_
T
S
g()d,
and (2.4) becomes
_
1
1

g(t t
0
) g(0)
t t
0

dt < .
Thus, it suces to prove the case where t
0
= 0 .
CHAPTER 2. FOURIER INTEGRALS 49
Step 2: We know that the theorem is true if g(t) = e
t
2
(See Example 2.5 and
Theorem 2.15). Replace g(t) by
h(t) = g(t) g(0)e
t
2
.
Then h satises all the assumptions which g does, and in addition, h(0) = 0.
Thus it suces to prove the case where both () t
0
= 0 and f(0) = 0 .
For simplicity we write f instead of h but assume (). Then (2.4) and (2.5)
simplify:
_
1
1

f(t)
t

dt < , (2.6)
lim
S
T
_
T
S

f()d = 0. (2.7)
Step 3: If f L
1
(1), then we argue as follows. Dene
g(t) =
f(t)
2it
.
Then g L
1
(1). By Fubinis theorem,
_
T
S

f()d =
_
T
S
_

e
2it
f(t)dtd
=
_

_
T
S
e
2it
df(t)dt
=
_

_
1
2it
e
2it
_
T
S
f(t)dt
=
_

_
e
2iTt
e
2i(S)t

f(t)
2it
dt
= g(T) g(S),
and this tends to zero as T and S (see Theorem 2.3). This proves
(2.7).
Step 4: If instead f L
2
(1), then we use Parsevals identity
_

f(t)h(t)dt =
_

f()

h()d
in a clever way: Choose

h() =
_
1, S t T,
0, otherwise.
CHAPTER 2. FOURIER INTEGRALS 50
Then the inverse Fourier transform h(t) of

h is
h(t) =
_
T
S
e
2it
d
=
_
1
2it
e
2it
_
T
S
=
1
2it
_
e
2iTt
e
2i(S)t

so Parsevals identity gives


_
T
S

f()d =
_

f(t)
1
2it
_
e
2iTt
e
2i(S)t

dt
= (with g(t) as in Step 3)
=
_

_
e
2iTt
e
2i(S)t

g(t)dt
= g(T) g(S) 0 as
_
T ,
S .
Step 5: If f = f
1
+ f
2
, where f
1
L
1
(1) and f
2
L
2
(1), then we apply Step 3
to f
1
and Step 4 to f
2
, and get in both cases (2.7) with f replaced by f
1
and f
2
.

Note: This means that in most cases where f is continuous at t


0
we have
f(t
0
) = lim
S
T
_
T
S
e
2it
0
f()d.
(continuous functions which do not satisfy (2.4) do exist, but they are dicult
to nd.) In some cases we can even use the inversion formula at a point where
f is discontinuous.
Theorem 2.31. Suppose that f L
1
(1) +L
2
(1). Let t
0
1, and suppose that
the two limits
f(t
0
+) = lim
tt
0
f(t)
f(t
0
) = lim
tt
0
f(t)
exist, and that
_
t
0
+1
t
0

f(t) f(t
0
+)
t t
0

dt < ,
_
t
0
t
0
1

f(t) f(t
0
)
t t
0

dt < .
CHAPTER 2. FOURIER INTEGRALS 51
Then
lim
T
_
T
T
e
2it
0
f()d =
1
2
[f(t
0
+) + f(t
0
)].
Note: Here we integrate
_
T
T
, not
_
T
S
, and the result is the average of the right
and left hand limits.
Proof. As in the proof of Theorem 2.30 we may assume that
Step 1: t
0
= 0 , (see Step 1 of that proof)
Step 2: f(t
0
+) + f(t
0
) = 0 , (see Step 2 of that proof).
Step 3: The claim is true in the special case where
g(t) =
_
e
t
, t > 0,
e
t
, t < 0,
because g(0+) = 1, g(0) = 1, g(0+) + g(0) = 0, and
_
T
T
g()d = 0 for all T,
since f is odd = g is odd.
Step 4: Dene h(t) = f(t) f(0+) g(t), where g is the function in Step 3. Then
h(0+) = f(0+) f(0+) = 0 and
h(0) = f(0) f(0+)(1) = 0, so
h is continuous. Now apply Theorem 2.30 to h. It gives
0 = h(0) = lim
T
_
T
T

h()d.
Since also
0 = f(0+)[g(0+) + g(0)] = lim
T
_
T
T
g()d,
we therefore get
0 = f(0+) + f(0) = lim
T
_
T
T
[

h() + g()]d = lim


T
_
T
T

f()d.
Comment 2.32. Theorems 2.30 and 2.31 also remain true if we replace
lim
T
_
T
T
e
2it

f()d
by
lim
0
_

e
2it
e
()
2

f()d (2.8)
(and other similar summability formulas). Compare this to Theorem 2.16. In
the case of Theorem 2.31 it is important that the cuto kernel (= e
()
2
in
(2.8)) is even.
CHAPTER 2. FOURIER INTEGRALS 52
2.5 Applications
2.5.1 The Poisson Summation Formula
Suppose that f L
1
(1) C(1), that

n=
[

f(n)[ < (i.e.,



f
1
(Z)), and
that

n=
f(t +n) converges uniformly for all t in some interval (, ). Then

n=
f(n) =

n=

f(n) (2.9)
Note: The uniform convergence of

f(t + n) can be dicult to check. One


possible way out is: If we dene

n
= sup
<t<
[f(t + n)[,
and if

n=

n
< , then

n=
f(t + n) converges (even absolutely), and
the convergence is uniform in (, ). The proof is roughly the same as what we
did on page 29.
Proof of (2.9). We rst construct a periodic function g L
1
() with the
Fourier coecients

f(n):

f(n) =
_

e
2int
f(t)dt
=

k=
_
k+1
k
e
2int
f(t)dt
t=k+s
=

k=
_
1
0
e
2ins
f(s + k)ds
Thm 0.14
=
_
1
0
e
2ins
_

k=
f(s + k)
_
ds
= g(n), where g(t) =

n=
f(t + n).
(For this part of the proof it is enough to have f L
1
(1). The other conditions
are needed later.)
So we have g(n) =

f(n). By the inversion formula for the periodic Fourier
transform:
g(0) =

n=
e
2in0
g(n) =

n=
g(n) =

n=

f(n),
CHAPTER 2. FOURIER INTEGRALS 53
provided (=forutsatt) that we are allowed to use the Fourier inversion formula.
This is allowed if g C[, ] and g
1
(Z) (Theorem 1.37). This was part of
our assumption.
In addition we need to know that the formula
g(t) =

n=
f(t + n)
holds at the point t = 0 (almost everywhere is no good, we need it in exactly
this point). This is OK if

n=
f(t + n) converges uniformly in [, ] (this
also implies that the limit function g is continuous).
Note: By working harder in the proof, Gripenberg is able to weaken some of the
assumptions. There are also some counter-examples on how things can go wrong
if you try to weaken the assumptions in the wrong way.
2.5.2 Is

L
1
(1) = C
0
(1) ?
That is, is every function g C
0
(1) the Fourier transform of a function f
L
1
(1)?
The answer is no, as the following counter-example shows. Take
g() =
_

ln 2
, [[ 1,
1
ln(1+)
, > 1,

1
ln(1)
, < 1.
Suppose that this would be the Fourier transform of a function f L
1
(1). As
in the proof on the previous page, we dene
h(t) =

n=
f(t + n).
Then (as we saw there), h L
1
(), and

h(n) =

f(n) for n = 0, 1, 2, . . ..
However, since

n=1
1
n

h(n) = , this is not the Fourier sequence of any h


L
1
() (by Theorem 1.38). Thus:
Not every h C
0
(1) is the Fourier transform of some f L
1
(1).
But:
f L
1
(1)

f C
0
(1) ( Page 36)
f L
2
(1)

f L
2
(1) ( Page 47)
f o

f o ( Page 44)
CHAPTER 2. FOURIER INTEGRALS 54
2.5.3 The Euler-MacLauren Summation Formula
Let f C

(1
+
) (where 1
+
= [0, )), and suppose that
f
(n)
L
1
(1
+
)
for all n Z
+
= 0, 1, 2, 3 . . .. We dene f(t) for t < 0 so that f(t) is even.
Warning: f is continuous at the origin, but f

may be discontinuous! For exam-


ple, f(t) = e
|2t|
f(t)=e
2|t|
We want to use Poisson summation formula. Is this allowed?
By Theorem 2.7,

f
(n)
= (2i)
n

f(), and

f
(n)
is bounded, so
sup
R
[(2i)
n
[[

f()[ < for all n

n=
[

f(n)[ < .
By the note on page 52, also

n=
f(t +n) converges uniformly in (1, 1). By
the Poisson summation formula:

n=0
f(n) =
1
2
f(0) +
1
2

n=
f(n)
=
1
2
f(0) +
1
2

n=

f(n)
=
1
2
f(0) +
1
2

f(0) +
1
2

n=1
_

f(n) +

f(n)
_
=
1
2
f(0) +
1
2

f(0) +

n=1
_

1
2
_
e
2int
+ e
2int
_
. .
cos(2nt)
f(t)dt
=
1
2
f(0) +
_

0
f(t)dt +

n=1
_

0
cos(2nt)f(t)dt
Here we integrate by parts several times, always integrating the cosine-function
and dierentiating f. All the substitution terms containing odd derivatives of
CHAPTER 2. FOURIER INTEGRALS 55
f vanish since sin(2nt) = 0 for t = 0. See Gripenberg for details. The result
looks something like

n=0
f(n) =
_

0
f(t)dt +
1
2
f(0)
1
12
f

(0) +
1
720
f

(0)
1
30240
f
(5)
(0) + . . .
2.5.4 Schwartz inequality
The Schwartz inequality will be used below. It says that
[f, g)[ |f|
L
2|g|
L
2
(true for all possible L
2
-spaces, both L
2
(1) and L
2
() etc.)
2.5.5 Heisenbergs Uncertainty Principle
For all f L
2
(1), we have
__

t
2
[f(t)[
2
dt
___

2
[

f()[
2
d
_

1
16
2
_
_
f
_
_
4
L
2
(R)
Interpretation: The more concentrated f is in the neighborhood of zero, the
more spread out must

f be, and conversely. (Here we must think that |f|
L
2
(R)
is xed, e.g. |f|
L
2
(R)
= 1.)
In quantum mechanics: The product of time uncertainty and space uncer-
tainty cannot be less than a given xed number.
CHAPTER 2. FOURIER INTEGRALS 56
Proof. We begin with the case where f o. Then
16
_
R
[tf(t)[dt
_
R
[

f()[d = 4
_
R
[tf(t)[dt
_
R
[f

(t)[dt
(

() = 2i

f() and Parsevals iden. holds). Now use Scwartz ineq.
4
__
R
[tf(t)[[f

(t)[dt
_
= 4
__
R
[tf(t)[[f

(t)[dt
_
4
__
R
Re[tf(t)f

(t)]dt
_
= 4
__
R
t
_
1
2
_
f(t)f

(t) + f(t)f

(t)
_
_
dt
_
2
=
_
R
t
d
dt
(f(t)f(t))
. .
=|f(t)|
dt (integrate by parts)
=
_
[t[f(t)[]

. .
=0

[f(t)[dt
_
=
__

[f(t)[dt
_
This proves the case where f o. If f L(1), but f o, then we choose a
sequence of functions f
n
o so that
_

[f
n
(t)[dt
_

[f(t)[dt and
_

[tf
n
(t)[dt
_

[tf(t)[dt and
_

[

f
n
()[d
_

[

f()[d
(This can be done, not quite obvious). Since the inequality holds for each f
n
, it
must also hold for f.
2.5.6 Weierstrass Non-Dierentiable Function
Dene (t) =

k=0
a
k
cos(2b
k
t), t 1 where 0 < a < 1 and ab 1.
Lemma 2.33. This sum denes a continuous function which is not dier-
entiable at any point.
CHAPTER 2. FOURIER INTEGRALS 57
Proof. Convergence easy: At each t,

k=0
[a
k
cos(2b
k
t)[

k=0
a
k
=
1
1 a
< ,
and absolute convergence convergence. The convergence is even uniform: The
error is

k=K
a
k
cos(2b
k
t)

k=K
[a
k
cos(2b
k
t)[

k=K
a
k
=
a
K
1 a
0 as K
so by choosing K large enough we can make the error smaller than , and the
same K works for all t.
By Analysis II: If a sequence of continuous functions converges uniformly, then
the limit function is continuous. Thus, is continuous.
Why is it not dierentiable? At least does the formal derivative not converge:
Formally we should have

(t) =

k=0
a
k
2b
k
(1) sin(2b
k
t),
and the terms in this serie do not seem to go to zero (since (ab)
k
1). (If a sum
converges, then the terms must tend to zero.)
To prove that is not dierentiable we cut the sum appropriatly: Choose some
function L
1
(1) with the following properties:
i) (1) = 1
ii) () = 0 for
1
b
and > b
iii)
_

[t(t)[dt < .
0 1/b 1 b
()
We can get such a function from the Fejer kernel: Take the square of the Fejer
kernel ( its Fourier transform is the convolution of

f with itself), squeeze it
(Theorem 2.7(e)), and shift it (Theorem 2.7(b)) so that it vanishes outside of
CHAPTER 2. FOURIER INTEGRALS 58
(
1
b
, b), and (1) = 1. (Sort of like approximate identity, but (1) = 1 instead of
(0) = 1.)
Dene
j
(t) = b
j
(b
j
t), t 1. Then
j
() = (b
j
), so (b
j
) = 1 and
() = 0 outside of the interval (b
j1
, b
j+1
).

3
b
4
b
3
b
2
b 1 1/b
Put f
j
=
j
. Then
f
j
(t) =
_

(t s)
j
(s)ds
=
_

k=0
a
k
1
2
_
e
2ib
k
(ts)
+ e
2ib
k
(ts)
_

j
(s)ds
(by the uniform convergence)
=

k=0
a
k
2
_

_
e
2ib
k
t
. .
=
k
j

j
(b
k
) + e
2ib
k
t

j
(b
k
)
. .
=0
_

_
=
1
2
a
j
e
2ib
k
t
.
(Thus, this particular convolution picks out just one of the terms in the series.)
Suppose (to get a contradiction) that can be dierentiated at some point t 1.
Then the function
(s) =
_
(t+s)(t)
s

(t) , s ,= 0
0 , s = 0
is (uniformly) continuous and bounded, and (0) = 0. Write this as
(t s) = s(s) + (t) s

(t)
CHAPTER 2. FOURIER INTEGRALS 59
i.e.,
f
j
(t) =
_
R
(t s)
j
(s)ds
=
_
R
s(s)
j
(s)ds + (t)
_
R

j
(s)ds
. .
=
j
(0)=0

(t)
_
R
s
j
(s)ds
. .

j
(0)
2i
=0
=
_
R
s(s)b
j
(b
j
s)ds
b
j
s=t
= b
j
_
R
(
s
b
j
)
. .
0 pointwise
s(s)ds
. .
L
1
0 by the Lesbesgue dominated convergence theorem as j .
Thus,
b
j
f
j
(t) 0 as j
1
2
_
a
b
_
j
e
2ib
j
t
0 as j .
As [e
2ib
j
t
[ = 1, this is
_
a
b
_
j
0 as j . Impossible, since
a
b
1. Our
assumption that is dierentiable at the point t must be wrong (t) is not
dierentiable in any point!
2.5.7 Dierential Equations
Solve the dierential equation
u

(t) + u(t) = f(t), t 1 (2.10)


where we require that f L
2
(1), u L
2
(1), u C
1
(1), u

L
2
(1) and that u

is of the form
u

(t) = u

(0) +
_
t
0
v(s)ds,
where v L
2
(1) (that is, u

is absolutely continuous and its generalized


derivative belongs to L
2
).
The solution of this problem is based on the following lemmas:
Lemma 2.34. Let k = 1, 2, 3, . . .. Then the following conditions are equivalent:
i) u L
2
(1)C
k1
(1), u
(k1)
is absolutely continuous and the generalized
derivative of u
(k1)
belongs to L
2
(1).
CHAPTER 2. FOURIER INTEGRALS 60
ii) u L
2
(1) and
_
R
[
k
u(k)[
2
d < .
Proof. Similar to the proof of one of the homeworks, which says that the same
result is true for L
2
-Fourier series. (There ii) is replaced by

[n

f(n)[
2
< .)
Lemma 2.35. If u is as in Lemma 2.34, then

u
(k)
() = (2i)
k
u()
(compare this to Theorem 2.7(g)).
Proof. Similar to the same homework.
Solution: By the two preceding lemmas, we can take Fourier transforms in (2.10),
and get the equivalent equation
(2i)
2
u()+ u() =

f(), 1 (4
2

2
) u() =

f(), 1 (2.11)
Two cases:
Case 1: 4
2

2
,= 0, for all 1, i.e., must not be zero and not a positive
number (negative is OK, complex is OK). Then
u() =

f()
4
2

2
, 1
so u = k f, where k = the inverse Fourier transform of

k() =
1
4
2

2
.
This can be computed explicitly. It is called Greens function for this problem.
Even without computing k(t), we know that
k C
0
(1) (since

k L
1
(1).)
k has a generalized derivative in L
2
(1) (since
_
R
[

k()[
2
d < .)
k does not have a second generalized derivative in L
2
(since
_
R
[
2

k()[
2
d = .)
How to compute k? Start with a partial fraction expansion. Write
=
2
for some C
CHAPTER 2. FOURIER INTEGRALS 61
( = pure imaginary if < 0). Then
1
4
2

2
=
1

2
4
2

2
=
1
2

1
+ 2
=
A
2
+
B
+ 2
=
A + 2A + B 2B
( 2)( + 2)

(A + B) = 1
(AB)2 = 0
_
A = B =
1
2
Now we must still invert
1
+2
and
1
2
. This we do as follows:
Auxiliary result 1: Compute the transform of
f(t) =
_
e
zt
, t 0,
0 , t < 0,
where Re(z) > 0 ( f L
2
(1) L
1
(1), but f / C(1) because of the jump at
the origin). Simply compute:

f() =
_

0
e
2it
e
zt
dt
=
_
e
(z+2i)t
(z + 2i)
_

0
=
1
2i + z
.
Auxiliary result 2: Compute the transform of
f(t) =
_
e
zt
, t 0,
0 , t > 0,
where Re(z) > 0 ( f L
2
(1) L
1
(1), but f / C(1))


f() =
_
0

e
2it
e
zt
dt
=
_
e
(z2i)t
(z 2i)t
_
0

=
1
z 2i
.
Back to the function k:

k() =
1
2
_
1
2
+
1
+ 2
_
=
1
2
_
i
i 2i
+
i
i + 2i
_
.
CHAPTER 2. FOURIER INTEGRALS 62
We dened by requiring
2
= . Choose so that Im() < 0 (possible because
is not a positive real number).
Re(i) > 0, and

k() =
1
2
_
i
i 2i
+
i
i + 2i
_
The auxiliary results 1 and 2 gives:
k(t) =
_
i
2
e
it
, t 0
i
2
e
it
, t < 0
and
u(t) = (k f)(t) =
_

k(t s)f(s)ds
Special case: = negative number = a
2
, where a > 0. Take = ia
i = i(i)a = a, and
k(t) =
_

1
2a
e
at
, t 0

1
2a
e
at
, t < 0 i.e.
k(t) =
1
2a
e
|at|
, t 1
Thus, the solution of the equation
u

(t) a
2
u(t) = f(t), t 1,
where a > 0, is given by
u = k f where
k(t) =
1
2a
e
a|t|
, t 1
This function k has many names, depending on the eld of mathematics you are
working in:
i) Greens function (PDE-people)
ii) Fundamental solution (PDE-people, Functional Analysis)
iii) Resolvent (Integral equations people)
CHAPTER 2. FOURIER INTEGRALS 63
Case 2: = a
2
= a nonnegative number. Then

f() = (a
2
4
2

2
) u() = (a 2)(a + 2) u().
As u() L
2
(1) we get a necessary condition for the existence of a solution: If
a solution exists then
_
R


f()
(a 2)(a + 2)

2
d < . (2.12)
(Since the denominator vanishes for =
a
2
, this forces

f to vanish at
a
2
,
and to be small near these points.)
If the condition (2.12) holds, then we can continue the solution as before.
Sideremark: These results mean that this particular problem has no eigenval-
ues and no eigenfunctions. Instead it has a contionuous spectrum consisting
of the positive real line. (Ignore this comment!)
2.5.8 Heat equation
This equation:
_

t
u(t, x) =

2
x
2
u(t, x) + g(t, x),
_
t > 0
x 1
u(0, x) = f(x) (initial value)
is solved in the same way. Rather than proving everything we proceed in a formal
mannor (everything can be proved, but it takes a lot of time and energy.)
Transform the equation in the x-direction,
u(t, ) =
_
R
e
2ix
u(t, x)dx.
Assuming that
_
R
e
2ix
t
u(t, x) =

t
_
R
e
2ix
u(t, x)dx we get
_

t
u(t, ) = (2i)
2
u(t, ) + g(t, )
u(0, ) =

f()

_

t
u(t, ) = 4
2

2
u(t, ) + g(t, )
u(0, ) =

f()
CHAPTER 2. FOURIER INTEGRALS 64
We solve this by using the standard variation of constants lemma:
u(t, ) =

f()e
4
2

2
t
. .
+
_
t
0
e
4
2

2
(ts)
g(s, )ds
. .
= u
1
(t, ) + u
2
(t, )
We can invert e
4
2

2
t
= e
(2

t)
2
= e
(/)
2
where = (2

t)
1
:
According to Theorem 2.7 and Example 2.5, this is the transform of
k(t, x) =
1
2

t
e
(
x
2

t
)
2
=
1
2

t
e

x
2
4t
We know that

f()

k() =

k f(), so
u
1
(t, x) =
_

1
2

t
e
(xy)
2
/4t
f(y)dy,
(By the same argument:
s and t s are xed when we transform.)
u
2
(t, x) =
_
t
0
(k g)(s)ds
=
_
t
0
_

1
2

(ts)
e
(xy)
2
/4(ts)
g(s, y)dyds,
u(t, x) = u
1
(t, x) + u
2
(t, x)
The function
k(t, x) =
1
2

t
e

x
2
4t
is the Greens function or the fundamental solution of the heat equation on the
real line 1 = (, ), or the heat kernel.
Note: To prove that this solution is indeed a solution we need to assume that
- all functions are in L
2
(1) with respect to x, i.e.,
_

[u(t, x)[
2
dx < ,
_

[g(t, x)[
2
dx < ,
_

[f(x)[
2
dx < ,
- some (weak) continuity assumptions with respect to t.
2.5.9 Wave equation
_

2
t
2
u(t, x) =

2
x
2
u(t, x) + k(t, x),
_
t > 0,
x 1.
u(0, x) = f(x), x 1

t
u(0, x) = g(x), x 1
CHAPTER 2. FOURIER INTEGRALS 65
Again we proceed formally. As above we get
_

2
t
2
u(t, ) = 4
2

2
u(t, ) +

k(t, ),
u(0, ) =

f(),

t
u(0, ) = g().
This can be solved by the variation of constants formula, but to simplify the
computations we assume that k(t, x) 0, i.e.,

h(t, ) 0. Then the solution is
(check this!)
u(t, ) = cos(2t)

f() +
sin(2t)
2
g(). (2.13)
To invert the rst term we use Theorem 2.7, and get
1
2
[f(x + t) + f(x t)].
The second term contains the Dirichlet kernel , which is inverted as follows:
Ex. If
k(x) =
_
1/2, [t[ 1
0, otherwise,
then

k() =
1
2
sin(2).
Proof.

k() =
1
2
_
1
1
e
2it
dt = . . . =
1
2
sin(t).
Thus, the inverse Fourier transform of
sin(2)
2
is k(x) =
_
1/2, [x[ 1,
0, [x[ > 1,
(inverse transform = ordinary transform since the function is even), and the
inverse Fourier transform (with respect to ) of
sin(2t)
2
= t
sin(2t)
2t
is
k(
x
t
) =
_
1/2, [x[ t,
0, [x[ > t.
This and Theorem 2.7(f), gives the inverse of the second term in (2.13): It is
1
2
_
x+t
xt
g(y)dy.
CHAPTER 2. FOURIER INTEGRALS 66
Conclusion: The solution of the wave equation with h(t, x) 0 seems to be
u(t, x) =
1
2
[f(x + t) + f(x t)] +
1
2
_
x+t
xt
g(y)dy,
a formula known as dAlemberts formula.
Interpretation: This is the sum of two waves: u(t, x) = u
+
(t, x) +u

(t, x), where


u
+
(t, x) =
1
2
f(x + t) +
1
2
G(x + t)
moves to the left with speed one, and
u

(t, x) =
1
2
f(x t)
1
2
G(x t)
moves to the right with speed one. Here
G(x) =
_
x
0
g(y)dy, x 1.
Chapter 3
Fourier Transforms of
Distributions
Questions
1) How do we transform a function f / L
1
(1), f / L
2
(1), for example
Weierstrass function
(t) =

k=0
a
k
cos(2b
k
t),
where b ,= integer (if b is an integer, then is periodic and we can use
Chapter I)?
2) Can we interpret both the periodic T-transform (on L
1
()) and the Fourier
integral (on L
1
(1)) as special cases of a more general Fourier transform?
3) How do you dierentiate a discontinuous function?
The answer: Use distribution theory, developed in France by Schwartz in
1950s.
3.1 What is a Measure?
We start with a simpler question: what is a -function? Typical denition:
_

_
(x) = 0, x ,= 0
(0) =
_

(x)dx = 1, (for > 0).


67
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 68
We observe: This is pure nonsense. We observe that (x) = 0 a.e., so
_

(x)dx = 0.
Thus: The -function is not a function! What is it?
Normally a -function is used in the following way: Suppose that f is continuous
at the origin. Then
_

f(x)(x)dx =
_

[f(x) f(0)]
. .
=0
when x=0
(x)
..
=0
when x=0
dx + f(0)
_

(x)dx
= f(0)
_

(x)dx = f(0).
This gives us a new interpretation of :
The -function is the operator which evaluates a continuous function at the
point zero.
Principle: You feed a function f(x) to , and gives you back the number
f(0) (forget about the integral formula).
Since the formal integral
_

f(x)(x)dx resembles an inner product, we often


use the notation , f). Thus
, f) = f(0)
Denition 3.1. The -operator is the (bounded linear) operator which maps
f C
0
(1) into the number f(0). Also called Diracs delta .
This is a special case of measure:
Denition 3.2. A measure is a bounded linear operator which maps func-
tions f C
0
(1) into the set of complex numbers C (or real). We denote this
number by , f).
Example 3.3. The operator which maps f C
0
(1) into the number
f(0) + f(1) +
_
1
0
f(s)ds
is a measure.
Proof. Denote G, f) = f(0) + f(1) +
_
1
0
f(s)ds. Then
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 69
i) G maps C
0
(1) C.
ii) G is linear:
G, f + g) = f(0) + g(0) + f(1) + g(1)
+
_
1
0
(f(s) + g(s))ds
= f(0) + f(1) +
_
1
0
f(s)ds
+g(0) + g(1) +
_
1
0
g(s)ds
= G, f) + G, g).
iii) G is continuous: If f
n
f in C
0
(1), then max
tR
[f
n
(t) f(t)[ 0 as
n , so
f
n
(0) f(0), f
n
(1) f(1) and
_
1
0
f
n
(s)ds
_
1
0
f(s)ds,
so
G, f
n
) G, f) as n .
Thus, G is a measure.
Warning 3.4. G, f) is linear in f, not conjugate linear:
G, f) = G, f), and not = G, f).
Alternative notation 3.5. Instead of G, f) many people write G(f) or Gf
(for example , Gripenberg). See Gasquet for more details.
3.2 What is a Distribution?
Physisists often also use the derivative of a -function, which is dened as

, f) = f

(0),
here f

(0) = derivative of f at zero. This is not a measure: It is not dened


for all f C
0
(1) (only for those that are dierentiable at zero). It is linear,
but it is not continuous (easy to prove). This is an example of a more general
distribution.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 70
Denition 3.6. A tempered distribution (=tempererad distribution) is a
continuous linear operator from o to C. We denote the set of such distributions
by o

. (The set o was dened in Section 2.2).


Theorem 3.7. Every measure is a distribution.
Proof.
i) Maps o into C, since o C
0
(1).
ii) Linearity is OK.
iii) Continuity is OK: If f
n
f in o, then f
n
f in C
0
(1), so , f
n
) , f)
(more details below!)
Example 3.8. Dene

, ) =

(0), o. Then

is a tempered distribu-
tion
Proof.
i) Maps o C? Yes!
ii) Linear? Yes!
iii) Continuous? Yes!
(See below for details!)
What does
n
in o mean?
Denition 3.9.
n
in o means the following: For all positive integers k, m,
t
k

(m)
n
(t) t
k

(m)
(t)
uniformly in t, i.e.,
lim
n
max
tR
[t
k
(
(m)
n
(t)
(m)
(t))[ = 0.
Lemma 3.10. If
n
in o, then

(m)
n

(m)
in C
0
(1)
for all m = 0, 1, 2, . . .
Proof. Obvious.
Proof that

is continuous: If
n
in o, then max
tR
[

n
(t)

(t)[ 0
as n , so

,
n
) =

n
(0)

(0) =

, ).
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 71
3.3 How to Interpret a Function as a Distribu-
tion?
Lemma 3.11. If f L
1
(1) then the operator which maps o into
F, ) =
_

f(s)(s)ds
is a continuous linear map from o to C. (Thus, F is a tempered distribution).
Note: No complex conjugate on !
Note: F is even a measure.
Proof.
i) For every o, the integral converges (absolutely), and denes a number
in C. Thus, F maps o C.
ii) Linearity: for all , o and , C,
F, + ) =
_
R
f(s)[(s) + (s)]ds
=
_
R
f(s)(s)ds +
_
R
f(s)(s)ds
= F, ) + F, ).
iii) Continuity: If
n
in o, then
n
in C
0
(1), and by Lebesgues
dominated convergence theorem,
F,
n
) =
_
R
f(s)
n
(s)ds
_
R
f(s)(s)ds = F, ).
The same proof plus a little additional work proves:
Theorem 3.12. If
_

[f(t)[
1 +[t[
n
dt <
for some n = 0, 1, 2, . . ., then the formula
F, ) =
_

f(s)(s)ds, o,
denes a tempered distribution F.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 72
Denition 3.13. We call the distribution F in Lemma 3.11 and Theorem 3.12
the distribution induced by f, and often write f, ) instead of F, ). Thus,
f, ) =
_

f(s)(s)ds, o.
This is sort of like an inner product, but we cannot change places of f and : f
is the distribution and is the test function in f, ).
Does the distribution f determine the function f uniquely? Yes!
Theorem 3.14. Suppose that the two functions f
1
and f
2
satisfy
_
R
[f
i
(t)[
1 +[t[
n
dt < (i = 1 or i = 2),
and that they induce the same distribution, i.e., that
_
R
f
1
(t)(t)dt =
_
R
f
2
(t)(t)dt, o.
Then f
1
(t) = f
2
(t) almost everywhere.
Proof. Let g = f
1
f
2
. Then
_
R
g(t)(t)dt = 0 for all o
_
R
g(t)
(1 + t
2
)
n/2
(1 + t
2
)
n/2
(t)dt = 0 o.
Easy to show that (1 + t
2
)
n/2
(t)
. .
(t)
o o. If we dene h(t) =
g(t)
(1+t
2
)
n/2
,
then h L
1
(1), and
_

h(s)(s)ds = 0 o.
If o then also the function s (t s) belongs to o, so
_
R
h(s)(t s)ds = 0
_
o,
t 1.
(3.1)
Take
n
(s) = ne
(ns)
2
. Then
n
o, and by 3.1,

n
h 0.
On the other hand, by Theorem 2.12,
n
h h in L
1
(1) as n , so this
gives h(t) = 0 a.e.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 73
Corollary 3.15. If we know the distribution f, then from this knowledge we
can reconstruct f(t) for almost all t.
Proof. Use the same method as above. We know that h(t) L
1
(1), and that
(
n
h)(t) h(t) =
f(t)
(1 + t
2
)
n/2
.
As soon as we know the distribution f, we also know the values of
(
n
h)(t) =
_

f(s)
(1 + s
2
)
n/2
(1 + s
2
)
n/2

n
(t s)ds
for all t.
3.4 Calculus with Distributions
(=Rakneregler)
3.16 (Addition). If f and g are two distributions, then f +g is the distribution
f + g, ) = f, ) +g, ), o.
(f and g distributions f o

and g o

).
3.17 (Multiplication by a constant). If is a constant and f o

, then f is
the distribution
f, ) = f, ), o.
3.18 (Multiplication by a test function). If f o

and o, then f is the


distribution
f, ) = f, ) o.
Motivation: If f would be induced by a function, then this would be the natural
denition, because
_
R
[(s)f(s)](s)ds =
_
R
f(s)[(s)(s)]ds = f, ).
Warning 3.19. In general, you cannot multiply two distributions. For example,

2
= is nonsense
( = -function)
=Diracs delta
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 74
However, it is possible to multiply distributions by a larger class of test func-
tions:
Denition 3.20. By the class C

pol
(1) of tempered test functions we mean the
following:
C

pol
(1) f C

(1),
and for every k = 0, 1, 2, . . . there are two numbers M and n so that
[
(k)
(t)[ M(1 +[t[
n
), t 1.
Thus, f C

pol
(1) f C

(1), and every derivative of f grows at most as


a polynomial as t .
Repetition:
_

_
o = rapidly decaying test functions
o

= tempered distributions
C

pol
(1) = tempered test functions.
Example 3.21. Every polynomial belongs to C

pol
. So do the functions
1
1 + x
2
, (1 + x
2
)
m
(m need not be an integer)
Lemma 3.22. If C

pol
(1) and o, then
o.
Proof. Easy (special case used on page 72).
Denition 3.23. If C

pol
(1) and f o

, then f is the distribution


f, ) = f, ), o
(O.K. since o).
Now to the big surprise : Every distribution has a derivative, which is another
distribution!
Denition 3.24. Let f o

. Then the distribution derivative of f is the


distribution dened by
f

, ) = f,

), o
(This is O.K., because o =

o, so f,

) is dened).
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 75
Motivation: If f would be a function in C
1
(1) (not too big at ), then
f,

) =
_

f(s)

(s)ds (integrate by parts)


= [f(s)(s)]

. .
=0

(s)(s)ds
= f

, ).
Example 3.25. Let
f(t) =
_
e
t
, t 0,
e
t
, t < 0.
Interpret this as a distribution, and compute its distribution derivative.
Solution:
f

, ) = f,

) =
_

f(s)

(s)ds
=
_
0

e
s

(s)ds
_

0
e
s

(s)ds
= [e
s
(s)]
0

_
0

e
s
(s)ds
_
e
s
(s)

0

_

0
e
s
(s)ds
= 2(0)
_

e
|s|
(s)ds.
Thus, f

= 2 + h, where h is the function h(s) = e


|s|
, s 1, and = the
Dirac delta (note that h L
1
(1) C(1)).
Example 3.26. Compute the second derivative of the function in Example 3.25!
Solution: By denition, f

, ) = f

). Put

= , and apply the rule


f

, ) = f,

). This gives
f

, ) = f,

).
By the preceding computation
f,

) = 2

(0)
_

e
|s|

(s)ds
= (after an integration by parts)
= 2

(0) +
_

f(s)(s)ds
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 76
(f = original function). Thus,
f

, ) = 2

(0) +
_

f(s)(s)ds.
Conclusion: In the distribution sense,
f

= 2

+ f,
where

, ) =

(0). This is the distribution derivative of Diracs delta. In


particular: f is a distribution solution of the dierential equation
f

f = 2

.
This has something to do with the dierential equation on page 59. More about
this later.
3.5 The Fourier Transform of a Distribution
Repetition: By Lemma 2.19, we have
_

f(t) g(t)dt =
_

f(t)g(t)dt
if f, g L
1
(1). Take g = o. Then o (See Theorem 2.24), so we can
interpret both f and

f in the distribution sense and get
Denition 3.27. The Fourier transform of a distribution f o is the distribu-
tion dened by

f, ) = f, ), o.
Possible, since o o.
Problem: Is this really a distribution? It is well-dened and linear, but is it
continuous? To prove this we need to know that

n
in o
n
in o.
This is a true statement (see Gripenberg or Gasquet for a proof), and we get
Theorem 3.28. The Fourier transform maps the class of tempered distributions
onto itself:
f o



f o

.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 77
There is an obvious way of computing the inverse Fourier transform:
Theorem 3.29. The inverse Fourier transform f of a distribution

f o

is
given by
f, ) =

f, ), o,
where = the inverse Fourier transform of , i.e., (t) =
_

e
2it
()d.
Proof. If = the inverse Fourier transform of , then =

and the formula
simply says that f,

) =

f, ).
3.6 The Fourier Transform of a Derivative
Problem 3.30. Let f o

. Then f

. Find the Fourier transform of f

.
Solution: Dene (t) = 2it, t 1. Then C

pol
, so we can multiply a
tempered distribution by . By various denitions (start with 3.27)

(f

), ) = f

, ) (use Denition 3.24)


= f, ( )

) (use Theorem 2.7(g))


= f,

) (where (s) = 2is(s))
=

f, ) (by Denition 3.27)


=

f, ) (see Denition above of )


=

f, ) (by Denition 3.23).
Thus,

(f

) =

f where () = 2i, 1.
This proves one half of:
Theorem 3.31.

(f

) = (i2)

f and

(2itf) = (

f)

More precisely, if we dene (t) = 2it, then C

pol
, and

(f

) =

f,

(f) =

.
By repeating this result several times we get
Theorem 3.32.

(f
(k)
) = (2i)
k

f k Z
+
(

(2it)
k
f) =

f
(k)
.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 78
Example 3.33. Compute the Fourier transform of
f(t) =
_
e
t
, t > 0,
e
t
, t < 0.
Smart solution: By the Examples 3.25 and 3.26.
f

= 2

+ f (in the distribution sense).


Transform this:
[(2i)
2
1]

f = 2

) = 2(2i)

(since

is the derivative of ). Thus, we need



:

, ) = , ) = (0) =
_
R
(s)ds
=
_
R
1 (s)ds =
_
R
f(s)(s)ds,
where f(s) 1. Thus

is the distribution which is induced by the function
f(s) 1, i.e., we may write

1 .
Thus, (4
2

2
+ 1)

f = 4i, so

f is induced by the function
4i
(1+4
2

2
)
. Thus,

f() =
4i
(1 + 4
2

2
)
.
In particular:
Lemma 3.34.

() 1 and

1 = .
(The Fourier transform of is the function 1, and the Fourier transform of
the function 1 is the Dirac delta.)
Combining this with Theorem 3.32 we get
Lemma 3.35.

(k)
= (2i)
k
, k Z
+
= 0, 1, 2, . . .
_

(2it)
k
_
=
(k)
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 79
3.7 Convolutions (Faltningar)
It is sometimes (but not always) possible to dene the convolution of two distri-
butions. One possibility is the following: If , o, then we know that

( ) =

,
so we can dene to be the inverse Fourier transform of

. The same idea


applies to distributions in some cases:
Denition 3.36. Let f o

and suppose that g o

happens to be such that


g C

pol
(1) (i.e., g is induced by a function in C

pol
(1), i.e., g is the inverse
T-transform of a function in C

pol
). Then we dene
f g = the inverse Fourier transform of

f g,
i.e. (cf. page 77):
f g, ) =

f g, )
where is the inverse Fourier transform of :
(t) =
_

e
2it
()d.
This is possible since g C

pol
, so that

f g o

; see page 74
To get a direct interpretation (which does not involve Fourier transforms) we
need two more denitions:
Denition 3.37. Let t 1, f o

, o. Then the translations


t
fand
t

are given by
(
t
)(s) = (s t), s 1

t
f, ) = f,
t
)
Motivation:
t
translates to the right by the amount t (if t > 0, to the left if
t < 0).
For ordinary functions f we have
_

(
t
f)(s)(s)ds =
_

f(s t)(s)ds (s t = v)
=
_

f(v)(v + t)dv
=
_

f(v)
t
(v)dv,
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 80
t

t

so the distribution denition coincides with the usual denition for functions
interpreted as distributions.
Denition 3.38. The reection operator R is dened by
(R)(s) = (s), o,
Rf, ) = f, R), f o

, o
Motivation: Extra homework. If f L
1
(1) and o, then we can write f
0
0
f
Rf
in the form
(f )(t) =
_
R
f(s)(t s)ds
=
_
R
f(s)(R)(s t)ds
=
_
R
f(s)(
t
R)(s)ds,
and we get an alternative formula for f in this case.
Theorem 3.39. If f o

and o, then f as dened in Denition 3.36,


is induced by the function
t f,
t
R),
and this function belongs to C

pol
(1).
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 81
We shall give a partial proof of this theorem (skipping the most complicated
part). It is based on some auxiliary results which will be used later, too.
Lemma 3.40. Let o, and let

(t) =
(t + ) (t)

, t 1.
Then

in o as 0.
Proof. (Outline) Must show that
lim
0
sup
tR
[t[
k
[
(m)

(t)
(m+1)
(t)[ = 0
for all t, m Z
+
. By the mean value theorem,

(m)
(t + ) =
(m)
(t) +
(m+1)
()
where t < < t + (if > 0). Thus
[
(m)

(t)
(m+1)
(t)[ = [
(m+1)
()
(m+1)
(t)[
= [
_
t

(m+2)
(s)ds[
_
where t < < t + if > 0
or t + < < t if < 0
_

_
t+||
t||
[
(m+2)
(s)[ds,
and this multiplied by [t[
k
tends uniformly to zero as 0. (Here I am skipping
a couple of lines).
Lemma 3.41. For every f o

there exist two numbers M > 0 and N Z


+
so
that
[f, )[ M max
0j,kN
tR
[t
j

(k)
(t)[. (3.2)
Interpretation: Every f o

has a nite order (we need only derivatives


(k)
where k N) and a nite polynomial growth rate (we need only a nite power
t
j
with j N).
Proof. Assume to get a contradiction that (3.2) is false. Then for all n Z
+
,
there is a function
n
o so that
[f,
n
)[ n max
0j,kn
tR
[t
j

(k)
n
(t)[.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 82
Multiply
n
by a constant to make f,
n
) = 1. Then
max
0j,kn
tR
[t
j

(k)
n
(t)[
1
n
0 as n ,
so
n
0 in o as n . As f is continuous, this implies that f,
n
) 0
as n . This contradicts the assumption f,
n
) = 1. Thus, (3.2) cannot be
false.
Theorem 3.42. Dene (t) = f,
t
R). Then C

pol
, and for all n Z
+
,

(n)
(t) = f
(n)
,
t
R) = f,
t
R
(n)
).
Note: As soon as we have proved Theorem 3.39, we may write this as
(f )
(n)
= f
(n)
= f
(n)
.
Thus, to dierentiate f it suces to dierentiate either f or (but not both).
The derivatives may also be distributed between f and :
(f )
(n)
= f
(k)

(nk)
, 0 k n.
Motivation: A formal dierentiation of
(f )(t) =
_
R
f(t s)(s)ds gives
(f )

=
_
R
f

(t s)(s)ds = f

,
and a formal dierentiation of
(f )(t) =
_
R
f(s)(t s)ds gives
(f )

=
_
R
f(s)

(t s)ds = f

.
Proof of Theorem 3.42.
i)
1

[(t + ) (t)] = f,
1

(
t+
R
t
R)). Here
1

(
t+
R
t
R)(s) =
1

[(R)(s t ) R(s t)]


=
1

[(t + s) (t s)] (by Lemma 3.40)


(t s) = (R

)(s t) =
t
R

.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 83
Thus, the following limit exists:
lim
0
1

[(t + ) (t)] = f,
t
R

).
Repeating the same argument n times we nd that is n times dieren-
tiable, and that

(n)
= f,
t
R
(n)
)
(or written dierently, (f )
(n)
= f
(n)
.)
ii) A direct computation shows: If we put
(s) = (t s) = (R)(s t) = (
t
R)(s),
then

(s) =

(t s) =
t
R

. Thus f,
t
R

) = f,

) = f

, ) =
f

,
t
R) (by the denition of distributed derivative). Thus,

= f,
t
R

) =
f

,
t
R) (or written dierently, f

= f

). Repeating this n times


we get
f
(n)
= f
(n)
.
iii) The estimate which shows that C

pol
: By Lemma 3.41,
[
(n)
(t)[ = [f
(n)
,
t
R)[
M max
0j,kN
sR
[s
j
(
t
R)
(k)
(s)[ ( as above)
= M max
0j,kN
sR
[s
j

(k)
(t s)[ (t s = v)
= M max
0j,kN
vR
[(t v)
j

(k)
(s)[
a polynomial in [t[.
To prove Theorem 3.39 it suces to prove the following lemma (if two distribu-
tions have the same Fourier transform, then they are equal):
Lemma 3.43. Dene (t) = f,
t
R). Then =

f .
Proof. (Outline) By the distribution denition of :
, ) = ,

) for all o.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 84
We compute this:
,

) =
_

(s)
..
function
in C

pol

(s)ds
=
_

f,
s
R)

(s)ds
= (this step is too dicult: To show that we may move
the integral to the other side of f requires more theory
then we have time to present)
= f,
_

s
R (s)ds) = ()
Here
s
R is the function
(
s
R)(t) = (R)(t s) = (s t) = (
t
)(s),
so the integral is
_

(s t)

(s)ds =
_

(
t
)(s)

(s)ds (see page 43)
=
_

(
t
)(s)(s)ds (see page 38)
=
_

e
2its
(s)(s)ds
. .
F-transform of
() = f,

) =

f, )
=

f , ). Thus, =

f .
Using this result it is easy to prove:
Theorem 3.44. Let f o

, , o. Then
(f )
. .
in C

pol

..
in S
. .
in C

pol
= f
..
in S

( )
. .
in S
. .
in C

pol
Proof. Take the Fourier transforms:
(f )
. .

f

..

. .
(

f )

= f
..

f
( )
. .

. .

f(

)
.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 85
The transforms are the same, hence so are the original distributions (note that
both (f ) and f ( ) are in C

pol
so we are allowed to take distribution
Fourier transforms).
3.8 Convergence in o

We dene convergence in o

by means of test functions in o. (This is a special


case of weak or weak*-convergence).
Denition 3.45. f
n
f in o

means that
f
n
, ) f, ) for all o.
Lemma 3.46. Let o with (0) = 1, and dene

(t) = (t), t 1, > 0.


Then, for all o,

in o as .
Note: We had this type of -sequences also in the L
1
-theory on page 36.
Proof. (Outline.) The Fourier transform is continuous o o (which we have
not proved, but it is true). Therefore

in o

in o

in o
(/) () () in o as .
Thus, we must show that
sup
R

k
_
d
d
_
j
[ (/) 1] ()

0 as .
This is a straightforward mechanical computation (which does take some time).
Theorem 3.47. Dene

as in Lemma 3.46. Then

in o

as .
Comment: This is the reason for the name -sequence.
Proof. The claim (=pastaende) is that for all o,
_
R

(t)(t)dt , ) = (0) as .
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 86
(Or equivalently,
_
R
(t)(t)dt (0) as ). Rewrite this as
_
R

(t)(R)(t)dt = (

R)(0),
and by Lemma 3.46, this tends to (R)(0) = (0) as . Thus,

, ) , ) for all o as ,
so

in o

.
Theorem 3.48. Dene

as in Lemma 3.46. Then, for all f o

, we have

f f in o

as .
Proof. The claim is that

f, ) f, ) for all o.
Replace with the reected
= R =

f, R) f, R) for all o
(by Thm 3.39) ((

f) )(0) (f )(0) (use Thm 3.44)


f (

)(0) (f )(0) (use Thm 3.39)


f, R(

)) f, R).
This is true because f is continuous and

in o, according to Lemma
3.46.
There is a General Rule about distributions:
Metatheorem: All reasonable claims about distribution convergence are true.
Problem: What is reasonable?
Among others, the following results are reasonable:
Theorem 3.49. All the operations on distributions and test functions which we
have dened are continuous. Thus, if
f
n
f in o

, g
n
g in o

n
in C

pol
(which we have not dened!),

n
in o,

n
in C, then, among others,
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 87
i) f
n
+ g
n
f + g in o

ii)
n
f
n
f in o

iii)
n
f
n
f in o

iv)

n
f
n


f in o

(

=inverse T-transform of )
v)
n
f
n
f in C

pol
vi) f

n
f

in o

vii)

f
n


f in o

etc.
Proof. Easy but long.
3.9 Distribution Solutions of ODE:s
Example 3.50. Find the function u L
2
(1
+
) C
1
(1
+
) with an absolutely
continuous derivative u

which satises the equation


_
u

(x) u(x) = f(x), x > 0,


u(0) = 1.
Here f L
2
(1
+
) is given.
Solution. Let v be the solution of homework 22. Then
_
v

(x) v(x) = f(x), x > 0,


v(0) = 0.
(3.3)
Dene w = u v. Then w is a solution of
_
w

(x) w(x) = 0, x 0,
w(0) = 1.
(3.4)
In addition we require w L
2
(1
+
).
Elementary solution. The characteristic equation is

2
1 = 0, roots = 1,
general solution
w(x) = c
1
e
x
+ c
2
e
x
.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 88
The condition w(x) L
2
(1
+
) forces c
1
= 0. The condition w(0) = 1 gives
w(0) = c
2
e
0
= c
2
= 1. Thus: w(x) = e
x
, x 0.
Original solution: u(x) = e
x
+v(x), where v is a solution of homework 22, i.e.,
u(x) = e
x
+
1
2
e
x
_

0
e
y
f(y)dy
1
2
_

0
e
|xy|
f(y)dy.
Distribution solution. Make w an even function, and dierentiate: we
denote the distribution derivatives by w
(1)
and w
(2)
. Then
w
(1)
= w

(since w is continuous at zero)


w
(2)
= w

+ 2w

(0)
. .
due to jump
discontinuity
at zero in w

0
(Dirac delta at the point zero)
The problem says: w

= w, so
w
(2)
w = 2w

(0)
0
. Transform:
((2i)
2
1) w() = 2w

(0) (since

0
1)
= w() =
2w

(0)
1+4
2

2
,
whose inverse transform is w

(0)e
|x|
(see page 62). We are only interested in
values x 0 so
w(x) = w

(0)e
x
, x > 0.
The condition w(0) = 1 gives w

(0) = 1, so
w(x) = e
x
, x 0.
Example 3.51. Solve the equation
_
u

(x) u(x) = f(x), x > 0,


u

(0) = a,
where a =given constant, f(x) given function.
Many dierent ways exist to attack this problem:
Method 1. Split u in two parts: u = v + w, where
_
v

(x) v(x) = f(x), x > 0


v

(0) = 0,
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 89
and
_
w

(x) w(x) = 0, x > 0


w

(0) = a,
We can solve the rst equation by making an even extension of v. The second
equation can be solved as above.
Method 2. Make an even extension of u and transform. Let u
(1)
and u
(2)
be
the distribution derivatives of u. Then as above,
u
(1)
= u

(u is continuous)
u
(2)
= u

+ 2 u

(0)
..
=a

0
(u

discontinuous)
By the equation: u

= u + f, so
u
(2)
u = 2a
0
+ f
Transform this:
[(2i)
2
1] u = 2a +

f, so
u =
2a
1+4
2

2


f
1+4
2

2
Invert:
u(x) = ae
|x|

1
2
_

e
|xy|
f(y)dy.
Since f is even, this becomes for x > 0:
u(x) = ae
x

1
2
e
x
_

0
e
y
f(y)dy
1
2
_

0
e
|xy|
f(y)dy.
Method 3. The method to make u and f even or odd works, but it is a dirty
trick which has to be memorized. A simpler method is to dene u(t) 0 and
f(t) 0 for t < 0, and to continue as above. We shall return to this method in
connection with the Laplace transform.
Partial Dierential Equations are solved in a similar manner. The computations
become slightly more complicated, and the motivations become much more com-
plicated. For example, we can replace all the functions in the examples on page
63 and 64 by distributions, and the results stay the same.
3.10 The Support and Spectrum of a Distribu-
tion
Support = the piece of the real line on which the distribution stands
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 90
Denition 3.52. The support of a continuous function is the closure (=slutna
holjet) of the set x 1 : (x) ,= 0.
Note: The set x 1 : (x) ,= 0 is open, but the support contains, in addition,
the boundary points of this set.
Denition 3.53. Let f o

and let | 1 be an open set. Then f vanishes


on | (=forsvinner pa |) if f, ) = 0 for all test functions o whose
support is contained in |.
U

Interpretation: f has no mass in |, no action on |.


Example 3.54. vanishes on (0, ) and on (, 0). Likewise vanishes
(k)
(k Z
+
= 0, 1, 2, . . .) on (, 0) (0, ).
Proof. Obvious.
Example 3.55. The function
f(t) =
_
1 [t[, [t[ 1,
0, [t[ > 1,
vanishes on (, 1) and on (1, ). The support of this function is [1, 1]
(note that the end points are included).
Denition 3.56. Let f o

. Then the support of f is the complement of the


largest set on which f vanishes. Thus,
supp(f) = M
_

_
M is closed, f vanishes on 1 M, and
f does not vanish on any open set
which is strictly bigger than 1 M.
Example 3.57. The support of the distribution
(k)
a
is the single point a. Here
k Z
+
, and
a
is point evaluation at a:

a
, ) = (a).
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 91
Denition 3.58. The spectrum of a distribution f o

is the support of

f.
Lemma 3.59. If M 1 is closed, then supp(f) M if and only if f vanishes
on 1 M.
Proof. Easy.
Example 3.60. Interpret f(t) = t
n
as a distribution. Then

f =
1
(2i)
n

(n)
, as
we saw on page 78. Thus the support of

f is 0, so the spectrum of f is 0.
By adding such functions we get:
Theorem 3.61. The spectrum of the function f(t) 0 is empty. The spectrum
of every other polynomial is the single point 0.
Proof. f(t) 0 spectrum is empty follows from denition. The other
half is proved above.
The converse is true, but much harder to prove:
Theorem 3.62. If f o

and if the spectrum of f is 0, then f is a polynomial


(, 0).
This follows from the following theorem by taking Fourier transforms:
Theorem 3.63. If the support of f is one single point a then f can be written
as a nite sum
f =
n

k=0
a
n

(k)
a
.
Proof. Too dicult to include. See e.g., Rudins Functional Analysis.
Possible homework: Show that
Theorem 3.64. The spectrum of f is a f(t) = e
2iat
P(t), where P is a
polynomial, P , 0.
Theorem 3.65. Suppose that f o

has a bounded support, i.e., f vanishes on


(, T) and on (T, ) for some T > 0 ( supp(f) [T, T]). Then

f
can be interpreted as a function, namely as

f() = f, (t)e
2it
),
where o is an arbitrary function satifying (t) 1 for t [T 1, T +1] (or,
more generally, for t | where | is an open set containing supp(f)). Moreover,

f C

pol
(1).
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 92
Proof. (Not quite complete)
Step 1. Dene
() = f, (t)e
2it
),
where is as above. If we choose two dierent
1
and
2
, then
1
(t)
2
(t) = 0
is an open set | containing supp(f). Since f vanishes on 1|, we have
f,
1
(t)e
2it
) = f,
2
(t)e
2it
),
so () does not depend on how we choose .
Step 2. For simplicity, choose (t) so that (t) 0 for [t[ > T + 1 (where T as
in the theorem). A simple but boring computation shows that
1

[e
2i(+)t
e
2it
](t)

e
2it
(t) = 2ite
2it
(t)
in o as 0 (all derivatives converge uniformly on [T1, T+1], and everything
is 0 outside this interval). Since we have convergence in o, also the following
limit exists:
lim
0
1

(( + ) ()) =

()
= lim
0
f,
1

(e
2i(+)t
e
2it
)(t))
= f, 2ite
2it
(t)).
Repeating the same computation with replaced by (2it)(t), etc., we nd
that is innitely many times dierentiable, and that

(k)
() = f, (2it)
k
e
2it
(t)), k Z
+
. (3.5)
Step 3. Show that the derivatives grow at most polynomially. By Lemma 3.41,
we have
[f, )[ M max
0t,lN
tR
[t
j

(l)
(t)[.
Apply this to (3.5) =
[
(k)
()[ M max
0j,lN
tR

t
j
_
d
dt
_
l
(2it)
k
e
2it
(t)

.
The derivative l = 0 gives a constant independent of .
The derivative l = 1 gives a constant times [[.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 93
The derivative l = 2 gives a constant times [[
2
, etc.
Thus, [
(k)
()[ const. x[1 +[w[
N
], so C

pol
.
Step 4. Show that =

f. That is, show that
_
R
()()d =

f, )(= f, )).
Here we need the same advanced step as on page 83:
_
R
()()d =
_
R
f, e
2it
(t)())d
= (why??) = f,
_
R
e
2it
(t)()d)
= f, (t) (t))
_
since (t) 1 in a
neighborhood of supp(f)
_
= f, ).
A very short explanation of why why?? is permitted: Replace the integral by
a Riemann sum, which converges in o, i.e., approximate
_
R
e
2it
()d = lim
n

k=
e
2i
k
t
(
k
)
1
n
,
where
k
= k/n.
3.11 Trigonometric Polynomials
Denition 3.66. A trigonometric polynomial is a sum of the form
(t) =
m

j=1
c
j
e
2i
j
t
.
The numbers
j
are called the frequencies of .
Theorem 3.67. If we interpret as a polynomial then the spectrum of is

1
,
2
, . . . ,
m
, i.e., the spectrum consists of the frequencies of the polynomial.
Proof. Follows from homework 27, since supp(

j
) =
j
.
Example 3.68. Find the spectrum of the Weierstrass function
(t) =

k=0
a
k
cos(2b
k
t),
where 0 < a < 1, ab 1.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 94
To solve this we need the following lemma
Lemma 3.69. Let 0 < a < 1, b > 0. Then
N

k=0
a
k
cos(2b
k
t)

k=0
a
k
cos(2b
k
t)
in o

as N .
Proof. Easy. Must show that for all o,
_
R
_
N

k=0

k=0
_
a
k
cos(2b
k
t)(t)dt 0 as N .
This is true because
_
R

k=N+1
[a
k
cos(2b
k
t)(t)[dt
_
R

k=N+1
a
k
[(t)[dt

k=N+1
a
k
_

[(t)[dt
=
a
N+1
1 a
_

[(t)[dt 0 as N .
Solution of 3.68: Since

N
k=0

k=0
in o

, also the Fourier transforms converge


in o

, so to nd it is enough to nd the transform of

N
k=0
a
k
cos(2b
k
t) and
to let N . This transform is

0
+
1
2
[a(
b
+
b
) + a
2
(
b
2 +
b
2 ) + . . . + a
N
(
b
N +
b
N)].
Thus,
=
0
+
1
2

n=1
a
n
(
b
n +
b
n),
where the sum converges in o

, and the support of this is 0, b, b


2
, b
3
, . . .,
which is also the spectrum of .
Example 3.70. Let f be periodic with period 1, and suppose that f L
1
(),
i.e.,
_
1
0
[f(t)[dt < . Find the Fourier transform and the spectrum of f.
Solution: (Outline) The inversion formula for the periodic transform says that
f =

n=

f(n)e
2int
.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 95
Working as on page 86 (but a little bit harder) we nd that the sum converges
in o

, so we are allowed to take transforms:

f =

n=

f(n)
n
(converges still in o

).
Thus, the spectrum of f is n N :

f(n) ,= 0. Compare this to the theory of
Fourier series.
General Conclusion 3.71. The distribution Fourier transform contains all the
other Fourier transforms in this course. A universal transform.
3.12 Singular dierential equations
Denition 3.72. A linear dierential equation of the type
n

k=0
a
k
u
(k)
= f (3.6)
is regular if it has exactly one solution u o

for every f o

. Otherwise it is
singular.
Thus: Singular means: For some f o

it has either no solution or more than


one solution.
Example 3.73. The equation u

= f. Taking f = 0 we get many dierent so-


lutions, namely u =constant (dierent constants give dierent solutions). Thus,
this equation is singular.
Example 3.74. We saw earlier on page 59-63 that if we work with L
2
-functions
instead of distributions, then the equation
u

+ u = f
is singular i > 0. The same result is true for distributions:
Theorem 3.75. The equation (3.6) is regular

k=0
a
k
(2i)
k
,= 0 for all 1.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 96
Before proving this, let us dene
Denition 3.76. The function D() =

n
k=0
a
k
(2i)
k
is called the symbol of
(3.6)
Thus: Singular the symbol vanishes for some 1.
Proof of theorem 3.75. Part 1: Suppose that D() ,= 0 for all 1.
Transform (3.6):
n

k=0
a
k
(2i)
k
u =

f D() u =

f.
If D() ,= 0 for all , then
1
D()
C

pol
, so we can multiply by
1
D()
:
u =
1
D()

f u = K f
where K is the inverse distribution Fourier transform of
1
D()
. Therefore, (3.6)
has exactly one solution u o

for every f o

.
Part 2: Suppose that D(a) = 0 for some a 1. Then
D
a
, ) =
a
, D) = D(a)(a) = 0.
This is true for all o, so D
a
is the zero distribution: D
a
= 0.

k=0
a
k
(2i)
k

a
= 0.
Let v be the inverse transform of
a
, i.e.,
v(t) = e
2iat

k=0
a
k
v
(k)
= 0.
Thus, v is one solution of (3.6) with f 0. Another solution is v 0. Thus,
(3.6) has at least two dierent solutions the equation is singular.
Denition 3.77. If (3.6) is regular, then we call K =inverse transform of
1
D()
the Greens function of (3.6). (Not dened for singular problems.)
How many solutions does a singular equation have? Find them all! (Solution
later!)
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 97
Example 3.78. If f C(1) (and [f(t)[ M(1 +[t[
k
) for some M and k), then
the equation
u

= f
has at least the solutions
u(x) =
_
x
0
f(x)dx + constant
Does it have more solutions?
Answer: No! Why?
Suppose that u

= f and v

= f u

= 0. Transform this (2i)( u v) =


0.
Let be a test function which vanishes in some interval [, ]( the support
of is in (, ] [, )). Then
() =
()
2i
is also a test function (it is 0 in [, ]), since (2i)( u v) = 0 we get
0 = (2i)( u v), )
= u v, 2i()) = u v, ).
Thus, u v, ) = 0 when supp() (, ] [, ), so by denition,
supp( u v) 0. By theorem 3.63, u v is a polynomial. The only polynomial
whose derivative is zero is the constant function, so u v is a constant.
A more sophisticated version of the same argument proves the following theorem:
Theorem 3.79. Suppose that the equation
n

k=0
a
k
u
(k)
= f (3.7)
is singular, and suppose that the symbol D() has exactly r simple zeros
1
,
2
, . . . ,
r
.
If the equation (3.7) has a solution v, then every other solution u o

of (3.7)
is of the form
u = v +
r

j=1
b
j
e
2i
j
t
,
where the coecients b
j
can be chosen arbitrarily.
CHAPTER 3. FOURIER TRANSFORMS OF DISTRIBUTIONS 98
Compare this to example 3.78: The symbol of the equation u

= f is 2i, which
has a simple zero at zero.
Comment 3.80. The equation (3.7) always has a distribution solution, for all
f o

. This is proved for the equation u

= f in [GW99, p. 277], and this can


be extended to the general case.
Comment 3.81. A zero
j
of order r 2 of D gives rise to terms of the type
P(t)e
2i
j
t
, where P(t) is a polynomial of degree r 1.
Chapter 4
The Fourier Transform of a
Sequence (Discrete Time)
From our earlier results we very quickly get a Fourier transform theory for se-
quences a
n

n=
. We interpret this sequence as the distribution

n=
a
n

n
(
n
= Diracs delta at the point n)
For example, this converges in o

if
[a
n
[ M(1 +[n[
N
) for some M, N
and the Fourier transform is:

n=
a
n
e
2in
=

k=
a
k
e
2ik
which also converges in o

. This transform is identical to the inverse transform


discussed in Chapter 1 (periodic function!), except for the fact that we replace i
by i (or equivalently, replace n by n). Therefore:
Theorem 4.1. All the results listed in Chapter 1 can be applied to the theory
of Fourier transforms of sequences, provided that we intercharge the Fourier
transform and the inverse Fourier transform.
Notation 4.2. To simplify the notations we write the original sequence as f(n),
n Z, and denote the Fourier transform as

f. Then

f is periodic (function or
99
CHAPTER 4. FOURIER SERIES 100
distribution, depending on the size of [f(n)[ as n ), and

f() =

n=
f(n)e
2in
.
From Chapter 1 we can give e.g., the following results:
Theorem 4.3.
i) f
2
(Z)

f L
2
(),
ii) f
1
(Z)

f C() (converse false),
iii) (

fg) =

f g if e.g.
_

f L
1
()
g L
1
()
or
_
f
2
(Z)
g
2
(Z)
iv) Etc.
We can also dene discrete convolutions:
Denition 4.4. (f g)(n) =

k=
f(n k)g(k).
This is dened whenever the sum converges absolutely. For example, if f(k) ,= 0
only for nitely many k or if
f
1
(Z), g

(Z), or if
f
2
(Z), g
2
(Z), etc.
Lemma 4.5.
i) f
1
(Z), g L
p
(Z), 1 p , f g
p
(Z)
ii) f
1
(Z), g c
0
(Z) f g c
0
(Z).
Proof. Same as in Chapter 1 (replace all integrals by sums).
Theorem 4.6. If f
1
(Z) and g
1
(Z), then
(

f g)() =

f() g().
Also true if e.g. f
2
(Z) and g
2
(Z).
Proof.
1
-case: Same as proof of Theorem 1.21 (replace integrals by sums).
In the
2
-case we rst approximate by an
1
-sequence, use the
1
-theory, and pass
to the limit.
CHAPTER 4. FOURIER SERIES 101
Notation 4.7. Especially in the engineering literature, but also in mathematical
literature, one often makes a change of variable: we have

f() =

n=
f(n)e
2in
=

n=
f(n)
_
e
2i
_
n
=

n=
f(n)z
n
,
where z = e
2i
.
Denition 4.8. Engineers dene F(z) =

n=
f(n)z
n
as the (bilateral)
(=dubbelsidig) Z-transformation of f.
Denition 4.9. Most mathematicians dene F(z) =

n=
f(n)z
n
instead.
Note: If f(n) = 0 for n < 0 we get the onesided (=unilateral) transform
F(z) =

n=0
f(n)z
n
(or

n=0
f(n)z
n
).
Note: The Z-transform is reduced to the Fourier transform by a change of vari-
able
z = e
2i
, so [0, 1] [z[ = 1
Thus, z takes values on the unit circle. In the case of one-sided sequences we can
also allow [z[ > 1 (engineers) or [z[ < 1 (mathematicians) and get power series
like those studied in the theory of analytic functions.
All Fourier transform results apply
Chapter 5
The Discrete Fourier Transform
We have studied four types of Fourier transforms:
i) Periodic functions on 1

f dened on Z.
ii) Non-periodic functions on 1

f dened on 1.
iii) Distributions on 1

f dened on 1.
iv) Sequences dened on Z

f periodic on 1.
The nal addition comes now:
v) f a periodic sequence (on Z)

f a periodic sequence.
5.1 Denitions
Denition 5.1.
N
= all periodic sequences F(m) with period N, i.e., F(m+
N) = F(m).
Note: These are in principle dened for all n Z, but the periodicity means
that it is enough to know F(0), F(1), . . . , F(N 1) to know the whole sequence
(or any other set of N consecutive (= pa varandra foljande) values).
Denition 5.2. The Fourier transform of a sequence F
N
is given by

F(m) =
1
N
N1

k=0
e

2imk
N
F(k), m Z.
102
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 103
Warning 5.3. Some people replace the constant
1
N
in front of the sum by
1

N
or
omit it completely. (This aects the inversion formula.)
Lemma 5.4.

F is periodic with the same period N as F.
Proof.

F(m + N) =
1
N

one period
e

2i(m+N)k
N
F(k)
=
1
N

one period
e
2ik
. .
=1
e

2imk
N
F(k)
=

F(m).
Thus, F
N


F
N
.
Theorem 5.5. F can be reconstructed from

F by the inversion formula
F(k) =
N1

m=0
e
2imk
N

F(m).
Note: No
1
N
in front here.
Note: Matlab puts the
1
N
in front of the inversion formula instead!
Proof.

m
e
2imk
N
1
N

l
e

2iml
N
F(l) =
1
N
N1

l=0
F(l)
N1

m=0
e
2im(kl)
N
. .
=
8
>
>
<
>
>
:
N, if l = k
0, if l ,= k
= F(k)
We know that
_
e
2i
N
_
N
= 1, so e
2i
N
is the N:th root of 1:
2i/N
e
e
2*2i/N
2/ 2/
1
We add N numbers, whose absolute value is one, and who point symmetrically
in all the dierent directions indicated above. For symmetry reasons, the sum
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 104
must be zero (except when l = k). (You always jump an angle
2(kl)
N
for each
turn, and go k l times around before you are done.)
Denition 5.6. The convolution F G of two sequences in
N
are dened by
(F G)(m) =

one period
F(mk)G(k)
(Note: Some indeces get out of the interval [0, N1]. You must use the periodicity
of F and G to get the corresponding values of F(mk)G(k).).
Denition 5.7. The (ordinary) product F G is dened by
(F G)(m) = F(m)G(m), m Z.
Theorem 5.8. (

F G) =

F

G and (

F G) = N

F

G (note the extra factor
N).
Proof. Easy. (Homework?)
Denition 5.9. (RF)(n) = F(n) (reection operator).
As before: The inverse transform = the ususal transform plus reection:
Theorem 5.10.

F = N(

RF) (note the extra factor N), where = Fourier


transform and = Inverse Fourier transform.
Proof. Easy. We could have avoided the factor N by a dierent scaling (but
then it shows up in other places instead).
5.2 FFT=the Fast Fourier Transform
Question 5.11. How many ops do we need to compute the Fourier transform
of F
N
?
FLOP=FLoating Point Operation=multiplication or addition or combination
of both.
1 Megaop = 1 million ops/second (10
6
)
1 Gigaop = 1 billion ops/second (10
9
)
(Used as speed measures of computers.)
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 105
Task 5.12. Compute

F(m) =
1
N

N1
k=0
e

2imk
N
F(k) with the minimum amount
of ops (=quickly).
Good Idea 5.13. Compute the coecients
_
e

2i
N
_
k
=
k
only once, and store
them in a table. Since
k+N
=
k
, we have e

2imk
N
=
mk
=
r
where r =
remainder when we divide mk by N. Thus, only N numbers need to be stored.
Thus: We can ignore the number of ops needed to compute the coecients
e

2imk
N
(done in advance).
Trivial Solution 5.14. If we count multiplication and addition separetely, then
we need to compute N coecients (as m = 0, 1, . . . , N 1), and each coecient
requires N muliplications and N 1 additions. This totals
N(2N 1) = 2N
2
N 2N
2
ops .
This is too much.
Brilliant Idea 5.15. Regroup (=omgruppera) the terms, using the symmetry.
Start by doing even coecients and odd coecients separetely:
Suppose for simplicity that N is even. Then, for even m, (put N = 2n)

F(2m) =
1
N
N1

k=0

2mk
F(k)
=
1
N
_
n1

k=0

2mk
F(k) +
2n1

k=n

2mk
F(k)
. .
Replace k by k+n
_
=
1
N
_
n1

k=0

2mk
F(k) +
2m(k+n)
F(k + n)
_
=
2
N
n1

k=0
e

2imk
(N/2)
1
2
[F(k) + F(k + n)] .
This is a new discrete time periodic Fourier transform of the sequence G(k) =
1
2
[F(k) + F(n + k)] with period n =
N
2
.
A similar computation (see Gripenberg) shows that the odd coecients can be
computed from

F(2m + 1) =
1
n
n1

k=0
e

2imk
n
H(k),
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 106
where H(k) =
1
2
e

ik
n
[F(k) F(k + n)]. Thus, instead of one transform of order
N we get two transforms of order n =
N
2
.
Number of ops: Computing the new transforms by brute force (as in 5.14 on
page 105) we need the following ops:
Even: n(2n 1) =
N
2
2

N
2
+ n additions =
N
2
2
ops.
Odd: The numbers e

ik
n
= e

2ik
N
are found in the table already computed.
We essentially again need the same amount, namely
N
2
2
+
N
2
(n extra multiplica-
tions).
Total:
N
2
2
+
N
2
2
+
N
2
= N
2
+
N
2
N
2
. Thus, this approximately halfed the number
of needed ops.
Repeat 5.16. Divide the new smaller transforms into two halfs, and again, and
again. This is possible if N = 2
k
for some integer k, e.g., N = 1024 = 2
10
.
Final conclusion: After some smaller adjustments we get down to
3
2
2
k
k ops.
Here N = 2
k
, so k = log
2
N, and we get
Theorem 5.17. The Fast Fourier Transform with radius 2 outlined above needs
approximately
3
2
N log
2
N ops.
This is much smaller than 2N
2
N for large N. For example N = 2
10
= 1024
gives
3
2
N log
2
N 15000 2000000 = 2N
2
N.
Denition 5.18. Fast Fourier transform with
_

_
radius 2 : split into 2 parts at each step N = 2
k
radius 3 : split into 3 parts at each step N = 3
k
radius m : split into m parts at each step N = m
k
Note: Based on symmetries. The same computations repeat themselves, so by
combining them in a clever way we can do it quicker.
Note: The FFT is so fast that it caused a minor revolution to many branches of
numerical analysis. It made it possible to compute Fourier transforms in practice.
Rest of this chapter: How to use the FFT to compute the other transforms dis-
cussed earlier.
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 107
5.3 Computation of the Fourier Coecients of
a Periodic Function
Problem 5.19. Let f C(). Compute

f(k) =
_
1
0
e
2ikt
f(t)dt
as eciently as possible.
Solution: Turn f into a periodic sequence and use FFT!
Conversion 5.20. Choose some N Z, and put
F(m) = f(
m
N
), m Z
(equidistant sampling). The periodicity of f makes F periodic with period N.
Thus, F
N
.
Theorem 5.21 (Error estimte). If f C() and

f
1
(Z) (i.e.,

f(k)[ < ),
then

F(m)

f(m) =

k=0

f(m + kN).
Proof. By the inversion formula, for all t,
f(t) =

jZ
e
2ijt

f(j).
Put t
k
=
k
N

f(t
k
) = F(k) =

jZ
e
2ikj
N

f(j)
(this series converges uniformly by Lemma 1.14). By the denition of

F:

F(m) =
1
N
N1

k=0
e

2imk
N
F(k)
=
1
N

jZ

f(j)
N1

k=0
e
2i(jm)k
N
. .
=
8
>
>
<
>
>
:
N, if
jm
N
= integer
0, if
jm
N
,= integer
=

lZ

f(m + Nl).
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 108
Take away the term

f(m) (l = 0) to get

F(m) =

f(m) +

l=0

f(m + Nl).
Note: If N is large and if

f(m) 0 quickly as m , then the error

l=0

f(m + Nl) 0.
First Method 5.22. Put
i)

f(m)

F(m) if [m[ <
N
2
ii)

f(m)
1
2

F(m) if [m[ =
N
2
(N even)
iii)

f(m) 0 if [m[ >
N
2
.
Here ii) is not important. We could use

f(
N
2
) = 0 or

f(
N
2
) =

F(m) instead.
Here

F(m) =
1
N
N1

k=0
e

2imk
N
F(k).
Notation 5.23. Let us denote (note the extra star)

|k|N/2
a
k
=

|k|N/2
a
k
= the usual sum of a
k
if N odd (then we have exactly N terms), and

|k|N/2
a
k
=
a sum where the rst and last terms have been
divided by two (these are the same if the sequence
is periodic with period N, there is one term too
many in this case).
First Method 5.24 (Error). The rst method gives the error:
i) [m[ <
N
2
gives the error
[

f(m)

F(m)[

k=0
[

f(m + kN)[
ii) [m[ =
N
2
gives the error
[

f(m)
1
2

F(m)[
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 109
iii) [m[ >
N
2
gives the error [

f(m)[.
we can simplify this into the following crude (=grov) estimate:
sup
mZ
[

f(m)

F(m)[

|m|N/2
[

f(m)[ (5.1)
(because this sum is the actual error).
First Method 5.25 (Drawbacks).
1

Large error.
2

Inaccurate error estimate (5.1).


3

The error estimate based on



f and not on f.
We need a better method.
Second Method 5.26 (General Principle).
1

Evaluate t at the points t


k
=
k
N
(as before), F(k) = f(t
k
)
2

Use the sequence F to construct a new function P C(T) which approx-


imates f.
3

Compute the Fourier coecients of P.


4

Approximate

f(n) by

P(n).
For this to succeed we must choose P in a smart way. The nal result will be
quite simple, but for later use we shall derive P from some basic principles.
Choice of P 5.27. Clearly P depends on F. To simplify the computations we
require P to satisfy (write P = P(F))
A) P is linear: P(F + G) = P(F) + P(G)
B) P is translation invariant: If we translate F, then P(F) is translated by
the same amount: If we denote
(
j
F)(m) = F(mj), then
P(
j
F) =
j/N
P(F)
(j discrete steps a time dierence of j/N).
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 110
This leads to simple computations: We want to compute

P(m) (which we use as
approximations of

f(m)) Dene a -sequence:
D(n) =
_
1, for n = 0, N, 2N, . . .
0, otherwise.
Then
N 2N
period
zero
(
k
D)(n) =
_
1, if n = k + jN, j Z
0, otherwise,
so
[F(k)
k
D](n) =
_
F(k), n = k + jN
0, otherwise.
and so
F =
N1

k=0
F(k)
k
D
Therefore, the principles A) and B) give
P(F) =
N1

k=0
F(k)P(
k
D)
=
N1

k=0
F(k)
k/N
P(D),
Where P(D) is the approximation of D = unit pulse at time zero D.
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 111
We denote this function by p. Let us transform P(F):
(

P(F))(m) =
_
1
0
N+1

k=0
F(k)(
k/N
p)(s)e
2ism
ds
=
N+1

k=0
F(k)
_
1
0
e
2ism
p(s
k
N
)ds (s
k
N
= t)
=
N+1

k=0
F(k)
_
one
period
e
2im(t+
k
N
)
p(t)dt
=
N+1

k=0
F(k)e

2imk
N
_
one
period
e
2imt
p(t)dt
. .
p(m)
= p(m)
N+1

k=0
F(k)e

2imk
N
. .
=N

F(m)
= N p(m)

F(m).
We can get rid of the factor N by replacing p by Np. This is our approximation
of the pulse of size N at zero
_
N, n = 0 + jN
0, otherwise.
Second Method 5.28. Construct F as in the First Method, and compute

F.
Then the approximation of

f(m) is

f(m)

F(m) p(m),
where p is the Fourier transform of the function that we get when we apply our
approximation procedure to the sequence
ND(n) =
_
N, n = 0(+jN)
0, otherwise.
Note: The complicated proof of this simple method will pay o in a second!
Approximation Method 5.29. Use any type of translation-invariant inter-
polation method, for example splines. The simplest possible method is linear
interpolation: If we interpolate the pulse ND in this way we get Thus,
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 112
N points
N
p(t) =
_
N(1 N[t[), [t[
1
N
0,
1
N
[t[ 1
1
N
(periodic extension)
This is a periodic version of the kernel. A direct computation gives
p(m) =
_
sin(m/N)
m/N
_
2
.
We get the following interesting theorem:
Theorem 5.30. If we rst discretize f, i.e. we replace f by the sequence F(k) =
f(k/N), the compute

F(m), and nally multiply

F(m) by
p(m) =
_
sin(m/N)
m/N
_
2
.
then we get the Fourier coecients for the function which we get from f by linear
interpolation at the points t
k
= k/N.
(This corresponds to the computation of the Fourier integral
_
1
0
e
2imt
f(t)dt by
using the trapetsoidal rule. Other integration methods have similar interpreta-
tions.)
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 113
5.4 Trigonometric Interpolation
Problem 5.31. Construct a good method to approximate a periodic function
f C(T) by a trigonometric polynomial
N

m=N
a
m
e
2imt
(a nite sum, resembles inverse Fourier transformation).
Useful for numerical computation etc.
Note: The earlier Second Method gave us a linear interpolation, not trigono-
metric approximation.
Note: This trigonometric polynomial has only nitely many Fourier coecients
,= 0 (namely a
m
, [m[ N).
Actually, the First Method gave us a trigonometric polynomial. There we had
_

f(m)

F(m) for [m[ <
N
2
,

f(m)
1
2

F(m) for [m[ =


N
2
,

f(m) 0 for [m[ >


N
2
.
By inverting this sequence we get a trigonometric approximation of f: f(t)
g(t), where
g(t) =

|m|N/2

F(m)e
2imt
. (5.2)
We have two dierent errors:
i)

f(m) is replaced by

F(m) =
1
N

N1
k=0
f(
k
N
)e
2ikm
N
,
ii) The inverse series was truncated to N terms.
Strange fact: These two errors (partially) cancel each other.
Theorem 5.32. The function g dened in (5.2) satises
g(
k
N
) = f(
k
N
), n Z,
i.e., g interpolates f at the points t
k
(which were used to construct rst F and
then g).
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 114
Proof. We dened F(k) = f(
k
N
), and

F(m) =

|k|N/2
F(k)e

2imk
N
.
By the inversion formula on page 103,
g(
k
N
) =

|m|N/2

F(m)e
2imk
N
(use periodicity)
=
N1

m=0

F(m)e
2imk
N
= F(k) = f(
k
N
)
Error estimate: How large is [f(t)g(t)[ between the mesh points t
k
=
k
N
(where
the error is zero)? We get an estimate from the computation in the last section.
Suppose that

f
1
(Z) and f C(T) so that the inversion formula holds for all
t (see Theorem 1.37). Then
f(t) =

mZ

f(m)e
2imt
, and
g(t) =

|m|N/2

F(m)e
2imt
(Theorem 5.21)
=

|m|N/2
_

f(m) +

k=0

f(m + kN)
_
e
2imt
= f(t)

|m|N/2

f(m)e
2imt
+

|m|N/2

k=0

f(m + kN)e
2imt
.
Thus
[g(t) f(t)[

|m|N/2
[

f(m)[ +

|m|N/2

k=0
[

f(m + kM)[
. .
=
P

|l|N/2
|

f(l)|
= 2

|m|N/2
[

f(m)[
(take l = m + kN, every [l[ >
N
2
appears one time, no [l[ <
N
2
appears, and
[l[ =
N
2
two times).
This leads to the following theorem:
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 115
Theorem 5.33. If

m=
[

f(m)[ < , then


[g(t) f(t)[ 2

|m|N/2
[

f(m)[,
where
g(t) =

|m|N/2

F(m)e
2imt
, and

F(m) =
1
N

|m|N/2
e

2imk
N
f(
k
N
).
This is nice if

f(m) 0 rapidly as m . Better accuracy by increasing N.
5.5 Generating Functions
Denition 5.34. The generating function of the sequence J
n
(x) is the func-
tion
f(x, z) =

n
J
n
(x)z
n
,
where the sum over n Z or over n Z
+
, depending on for which values of n
the functions J
n
(x) are dened.
Note: We did this in the course on special functions. E.g., if J
n
= Bessels
function of order n, then
f(x, z) = e
x
2
(z1/2)
.
Note: For a xed value of x, this is the mathematicians version of the Z-
transform described on page 101.
Make a change of variable:
z = e
2it
f(x, e
2it
) =

nZ
J
n
(x)(e
2it
)
n
=

nZ
J
n
(x)e
2int
,
Comparing this to the inversion formula in Chapter 1 we get
Theorem 5.35. For a xed x, the n:th Fourier coecient of the function t
f(x, e
2it
) is equal to J
n
(x).
Thus, we can compute J
n
(x) by the method descibed in Section 5.3 to compute
the coecients a
n
= J
n
(x) (x = xed, n varies):
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 116
1) Discretize F(k) = f(x, e
2ik
N
)
2)

F(m) =
1
N

|k|N/2
e

2imk
N
F(k)
3)

F(m) J
n
(x) =

k=0
a
m+kN
, (Theorem 5.21)
where a
m+kN
= J
m+kN
(x).
5.6 One-Sided Sequences
So far we have been talking about periodic sequences (in
N
). Instead one often
wants to discuss
A) Finite sequences A(0), A(1), . . . , A(N 1) or
B) One-sided sequences A(n), n Z
+
= 0, 1, 2 . . .
Note: 5.36. A nite sequence is a special csse of a one-sided sequence: put
A(n) = 0 for n N.
Note: 5.37. A one-sided sequence is a special case of a two-sided sequence: put
A(n) = 0 for n < 0.
Problem: These extended sequences are not periodic. We cannot use the Fast
Fourier Transform directly.
Notation 5.38. C
Z
+
= all complex valued sequences A(n), n Z
+

Denition 5.39. The convolution of two sequences A, B C


Z
+
is
(A B)(m) =
m

k=0
A(mk)B(k), m Z
+
Note: The summation boundaries are the natural ones if we think that A(k) =
B(k) = 0 for k < 0.
Notation 5.40.
A
|n
(k) =
_
A(k), 0 k < n
0, k n.
Thus, this restricts the sequence A(k) to the n rst terms.
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 117
Lemma 5.41. (A B)
|n
= (A
|n
B
|n
)
|n
Proof. Easy.
Notation 5.42. A = 0
n
means that A(k) = 0 for 0 k < n 1, i.e., A
|n
= 0.
Lemma 5.43. If A = 0
n
and B = 0
m
, then A B = 0
n+m
.
Proof. Easy.
Computation of A B 5.44 (One-sided convolution).
1) Choose a number N 2n (often a power of 2).
2) Dene
F(k) =
_
A(k), 0 k < n,
0, n k < N,
and extend F to be periodic, period N.
3) Dene
G(k) =
_
B(k), 0 k < n,
0, n k < N,
periodic extension: G(k + N) = G(k).
Then, for all m, 0 m < n,
(F G)(m)
. .
periodic convolution
=
N1

k=0
F(mk)G(k)
=
m

k=0
F(mk)G(k)
=
m

k=0
A(mk)G(k) = (A B)(m)
. .
one-sided convolution
Note: Important that N 2n.
Thus, this way we have computed the n rst coecients of (A B).
Theorem 5.45. The method described below allows us to compute (A B)
|n
(=the rst n coecients of A B) with a number of FLOP:s which is
C nlog
2
n, where C is a constant.
Method: 1)-3) same as above
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 118
4) Use FFT to compute

F

G(= N(

F G)).
5) Use the inverse FFT to compute
F G =
1
N
(

F

G)
Then (A B)
|n
= (F G)
|n
.
Note: A naive computation of A B
|n
requires C
1
n
2
FLOPs, where C
1
is
another constant.
Note: Use naive method if n small. Use FFT-inverse FFT if n is large.
Note: The rest of this chapter applies one-sided convolutions to dierent situa-
tions. In all cases the method described in Theorem 5.45 can be used to compute
these.
5.7 The Polynomial Interpretation of a Finite
Sequence
Problem 5.46. Compute the product of two polynomials:
p(x) =
n

k=0
a
k
x
k
q(x) =
m

l=0
b
l
x
l
.
Solution: Dene a
k
= 0 for k > 0 and b
l
= 0 for l > m. Then
p(x)q(x) =
_

k=0
a
k
x
k
__

l=0
b
l
x
l
_
. .
sums are actually nite
=

k,l
a
k
b
l
x
k+l
(k + l = j, k = j l)
=

j
x
j
j

l=0
a
jl
b
l
=
m+n

j=0
c
j
x
j
,
where c
j
=

j
l=0
a
jl
b
l
. This gives
Theorem 5.47.
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 119
i) Multiplication of two polynomials corresponds to a convolution of their
coecients: If
p(x) =
n

k=0
a
k
x
k
, q(x) =
m

l=0
b
l
x
l
,
then p(x)q(x) =

m+n
j=0
c
j
x
j
, where c = a b.
ii) Addition of two polynomials corresponds to addition of the coecients:
p(x) + q(x) =

c
j
x
j
, where c
j
= a
j
+ b
j
.
iii) Multiplication of a polynomial by a complex constant corresponds to mul-
tiplication of the coecients by the same constant.
Operation Polynomial Coecients
Addition p(x) + q(x) a
k
+ b
k

max{m,n}
k=0
Multiplication by
C
p(x) a
k

n
k=0
Multiplication p(x)q(x) (a b)(k)
Thus there is a one-to-one correspondence between
polynomials nite sequences,
where the operations correspond as described above. This is used in all symbolic
computer computations of polynomials.
Note: Two dierent conventions are in common use:
A) rst coecient is a
0
(= lowest order),
B) rst coecient is a
n
(= highest order).
5.8 Formal Power Series and Analytic Functions
Next we extend polynomials so that they may contain innitely many terms.
Denition 5.48. A Formal Power Series (FPS) is a sum of the type

k=0
A(k)x
k
which need not converge for any x ,= 0. (If it does converge, then it denes an
analytic function in the region of convergence.)
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 120
Example 5.49.

k=0
x
k
k!
converges for all x (and the sum is e
x
).
Example 5.50.

k=0
x
k
converges for [x[ < 1 (and the sum is
1
1x
).
Example 5.51.

k=0
k!x
k
converges for no x ,= 0.
All of these are formal power series (and the rst two are ordinary power
series).
Calculus with FPS 5.52. We borrow the calculus rules from the polynomials:
i) We add two FPS:s by adding the coecients:
_

k=0
A(k)x
k
_
+
_

k=0
B(k)x
k
_
=

k=0
[A(k) + B(k)]x
k
.
ii) We multiply a FPS by a constant by multiplying each coecients by :

k=0
A(k)x
k
=

k=0
[A(k)]x
k
.
iii) We multiply two FPS:s with each other by taking the convolution of the
coecients:
_

k=0
A(k)x
k
__

k=0
B(k)x
k
_
=

k=0
C(k)x
k
,
where C = A B.
Notation 5.53. We denote

A(x) =

k=0
A(k)x
k
.
Conclusion 5.54. There is a one-to-one correspondence between all Formal
Power Series and all one-sided sequences (bounded or not). We denoted these by
C
Z
+
on page 116.
Comment 5.55. In the sequence (=fortsattningen) we operate with FPS:s.
These power series often converge, and then they dene analytic functions, but
this fact is not used anywhere in the proofs.
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 121
5.9 Inversion of (Formal) Power Series
Problem 5.56. Given a (formal) power series

A(x) =

A(k)x
k
, nd the in-
verse formal power series

B(x) =

B(k)x
k
.
Thus, we want to nd

B(x) so that

A(x)

B(x) = 1, i.e.,
_

k=0
A(k)x
k
__

l=0
B(l)x
l
_
= 1 + 0x + 0x
2
+ . . .
Notation 5.57.
0
= 1, 0, 0, . . . = the sequence whose power series is 1 +
0x + 0x
2
+ . . .. This series converges, and the sum is 1. More generally:

k
= 0, 0, 0, . . . , 0, 1, 0, 0, . . .
= 0 + 0x + 0x
2
+ . . . + 0x
k1
+ 1x
k
+ 0x
k+1
+ 0x
k+2
+ . . .
= x
k
Power Series Sequence

k
x
k
Solution. We know that AB =
0
, or equivalently,

A(x)

B(x) = 1. Explicitly,

A(x)

B(x) = A(0)B(0) (times x
0
) (5.3)
+[A(0)B(1) + A(1)B(0)]x (5.4)
+[A(0)B(2) + A(1)B(1) + A(2)B(0)]x
2
(5.5)
+. . . (5.6)
From this we can solve:
i) A(0)B(0) = 1 = A(0) ,= 0 and B(0) =
1
A(0)
.
ii) A(0)B(1) + A(1)B(0) = 0 = B(1) =
A(1)B(0)
A(0)
(always possible)
iii) A(0)B(2) + A(1)B(1) + A(2)B(0) = 1 = B(2) =
1
A(0)
[A(1)B(1) +
A(2)B(0)], etc.
we get a theorem:
Theorem 5.58. The FPS

A(x) can be inverted if and only if A(0) ,= 0. The
inverse series [A(x)]
1
is obtained recursively by the procedure described above.
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 122
Recursive means:
i) Solve B(0)
ii) Solve B(1) using B(0)
iii) Solve B(2) using B(1) and B(0)
iv) Solve B(3) using B(2), B(1) and B(0), etc.
This is Hard Work. For example
sin(x) = x
x
3
3!
+
x
5
5!
. . .
cos(x) = 1
x
2
2!
+
x
4
4!
. . .
tan(x) =
sin(x)
cos(x)
= sin(x)
1
cos(x)
. .
convolution
=???
Hard Work means: Number of FLOPS is a constant times N
2
. Better method:
Use FFT.
Theorem 5.59. Let A(0) ,= 0. and let

B(x) be the inverse of

A(x). Then, for
every k 1,
B
|2k
= (B
|k
(2
0
A B
|k
))
|2k
(5.7)
Proof. See Gripenberg.
Usage: First compute B
|1
=
1
A(0)
, 0, 0, 0, . . .
Then B
|2
= B(0), B(1), 0, 0, 0, . . . (use (5.7))
Then B
|4
= B(0), B(1), B(2), B(3), 0, , 0, 0, . . .
Then B
|8
= 8 terms ,= 0 etc.
Use the method on page 117 for the convolutions. (Useful only if you need lots
of coecients).
5.10 Multidimensional FFT
Especially in image processing we also need the discrete Fourier transform in
several dimensions. Let d = 1, 2, 3, . . . be the space dimension. Put
d
N
=
sequences x(k
1
, k
2
, . . . , k
d
) which are N-periodic in each variable separately.
CHAPTER 5. THE DISCRETE FOURIER TRANSFORM 123
Denition 5.60. The d-dimensional Fourier transform is obtained by transform-
ing d successive (=efter varandra) ordinary Fourier transformations, one for
each variable.
Lemma 5.61. The d-dimensional Fourier transform is given by
x(m
1
, m
2
, . . . , m
d
) =
1
N
d

k
1

k
2

k
d
e
2i(k
1
m
1
+k
2
m
2
+...+k
d
m
d
)
N
x(k
1
, k
2
, . . . , k
d
).
Proof. Easy.
All 1-dimensional results generalize easy to the d-dimensional case.
Notation 5.62. We call k = (k
1
, k
2
, . . . , k
d
) and m = (m
1
, m
2
, . . . , m
d
) multi-
indeces (=pluralis av multi-index), and put
k m = k
1
m
1
+ k
2
m
2
+ . . . + k
d
m
d
(=the inner product of k and m).
Lemma 5.63.
x(m) =
1
N
d

k
e

2imk
N
x(k),
x(k) =

m
e
2imk
N
x(m).
Denition 5.64.
(F G)(m) = F(m)G(m)
(F G)(m) =

k
F(mk)G(k),
where all the components of m and k run over one period.
Theorem 5.65.
(F G) =

F

G,
(F G) = N
d

F

G.
Proof. Follows from Theorem 5.8.
In practice: Either use one multi-dimensional, or use d one-dimensional trans-
forms (not much dierence, multi-dimensional a little faster).
Chapter 6
The Laplace Transform
6.1 General Remarks
Example 6.1. We send in a signal u into an amplier, and get an output
signal y:
Black box u y
Under quite general assumptions it can be shown that
y(t) = (K u)(t) =
_
t

K(t s)u(s)ds,
i.e., the output is the convolution (=faltningen) of u with the inpulse re-
sponse K.
Terminology 6.2. Impulse response (=pulssvar) since y = K if u = a delta
distribution.
Causality 6.3. The upper bound in the integral is t, i.e., (Ku)(t) depends only
on past values of u, and not on future values. This is called causality.
If, in addition u(t) = 0 for t < 0, then y(t) = 0 for t < 0, and
y(t) =
_
t
0
K(t s)u(s)ds,
which is a one-sided convolution.
Classication 6.4. Approximately: The Laplace-transform is the Fourier trans-
form applied to one-sided signals (dened on 1
+
). In addition there is a change
of variable which rotate the complax plane.
124
CHAPTER 6. THE LAPLACE TRANSFORM 125
6.2 The Standard Laplace Transform
Denition 6.5. Suppose that
_

0
e
t
[f(t)[dt < for some 1. Then we
dene the Laplace transform

f(s) of f by

f(s) =
_

0
e
st
f(t)dt, (s) .
Lemma 6.6. The integral above converges absolutely for all s C with (s)
(i.e.,

f(s) is well-dened for such s).
Proof. Write s = + i. Then
[e
st
f(t)[ = [e
t
e
it
f(t)[
= e
t
[f(t)[
e
t
[f(t)[, so
_

0
[e
st
f(t)[dt
_

0
e
t
[f(t)[dt < .
Theorem 6.7.

f(s) is analytic in the open half-plane Re(s) > , i.e.,

f(s) has
a complex derivative with respect to s.
Proof. (Outline)

f(z)

f(s)
z s
=
_

0
e
zt
e
st
z s
f(t)dt
=
_

0
e
(zs)t
1
z s
e
st
f(t)dt (put z s = h)
=
_

0
1
h
[e
ht
1]
. .
t as h0
e
st
f(t)dt
As Re(s) > we nd that
_

0
[te
st
f(t)[d < and a shortcomputation (about
1
2
page) shows that the Lebesgue dominated convergence theorem can be applied
(show that [
1
h
(e
ht
1)[ const. t e
t
, where =
1
2
[ + (s)] (this is true
for some small enough h), and then show that
_

0
te
t
[e
st
f(t)[dt < ). Thus,
d
ds

f(s) exists, and


d
ds

f(s) =
_

0
e
st
tf(t)dt, (s) >
Corollary 6.8.
d
ds

f(s) is the Laplace transform of g(t) = tf(t), and this


Laplace transform converges (at least) in the half-plane Re(s) > .
CHAPTER 6. THE LAPLACE TRANSFORM 126
Theorem 6.9.

f(s) is bounded in the half-plane (s) .
Proof. (cf. proof of Lemma 6.6)
[

f(s)[ = [
_

0
e
st
f(t)dt[
_

0
[e
st
f(t)[dt
=
_

0
e
(s)t
[f(t)[dt
_

0
e
t
[f(t)[dt < .
Denition 6.10. A bounded analytic function on the half-plane Re(s) >
is called a H

-function (over this half-plane).


Theorem 6.11. If f is absolutely continuous and
_

0
e
t
[g(t)[dt < (i.e.,
f(t) = f(0) +
_
t
0
g(s)ds, where
_

0
e
t
[g(t)[dt < ), then
(

)(s) = s

f(s) f(0), (s) > .


Proof. Integration by parts (a la Lebesgue) gives
lim
T
_
T
0
e
st
f(t)
. .
=

f(s)
dt = lim
T
_
_
e
st
s
f(t)

T
0
+
1
s
_

0
e
st
f

(t)dt
_
=
1
s
f(0) +
1
s

(s), so
(

)(s) = s

f(s) f(0).
6.3 The Connection with the Fourier Transform
Let Re(s) > , and make a change of variable:
_

0
e
st
f(t)dt (t = 2v; dt = 2dv)
=
_

0
e
2sv
f(2v)2dv (s = + i)
=
_

0
e
2iv
e
2v
f(2v)2dv (put f(t) = 0 for t < 0)
=
_

e
2it
g(t)dt,
where
g(t) =
_
2e
2t
f(2t) , t 0
0 , t < 0.
(6.1)
Thus, we got
CHAPTER 6. THE LAPLACE TRANSFORM 127
Theorem 6.12. On the line Re(s) = (which is a line parallell with the imagi-
nary axis < < )

f(s) coincides (=sammanfaller med) with the Fourier
transform of the function g dened in (6.1).
Thus, modulo a change of variable, the Laplace transform is the Fourier transform
of a function vanishing for t < 0. From Theorem 6.12 and the theory about
Fourier transforms of functions in L
1
(1) and L
2
(1) we can derive a number of
results. For example:
Theorem 6.13. (Compare to Theorem 2.3, page 36) If f L
1
(1
+
) (i.e.,
_

0
[f(t)[dt <
), then
lim
|s|
(s)0
[

f(s)[ = 0
(where s in the half plane Re(s) > 0 in an arbitrary manner)
Re
Im
Combining Theorem 6.12 with one of the theorems about the inversion of the
Fourier integral we get formulas of the type
1
2
_

e
2it

f( + i)d =
_
e
2t
f(t), t > 0,
0, t < 0.
This is often written as a complex line integral: We integrate along the line
Re(s) = , and replace 2t t and multiply the formulas by e
2t
to get
(s = + i, ds = id)
f(t) =
1
2i
_
+i
i
e
st

f(s)ds (6.2)
=
1
2i
_

=
e
(+i)t

f( + i)id
CHAPTER 6. THE LAPLACE TRANSFORM 128
Warning 6.14. This integral seldom converges absolutely. If it does converge
absolutely, then (See Theorem 2.3 with the Fourier theorem replaced by the in-
verse Fourier theorem) the function
g(t) =
_
2e
2t
f(t), t 0,
0, t < 0
must be continuous. In other words:
Lemma 6.15. If the integral (6.2) converges absolutely, then f must be contin-
uous and satisfy f(0) = 0.
Therefore, the inversion theorems given in Theorem 2.30 and Theorem 2.31 are
much more useful. They give (under the assumptions given there)
1
2
[f(t+) + f(t)] = lim
T
1
2i
_
+iT
iT
e
st

f(s)ds
(and we interpret f(t) = 0 for t < 0). By Theorem 6.11, if f is absolutely
continuous and f

L
1
(1
+
), then (use also Theorem 6.13)

f(s) =
1
s
[

(f

)(s) + f(0)],
where

(f

)(s) 0 as [s[ , (s) 0. Thus, for large values of ,



f(+i)
f(0)
i
, so the convergence is slow in general. Apart from the space H

(see page
126) (over the half plane) another much used space (especially in Control theory)
is H
2
.
Theorem 6.16. If f L
2
(1
+
), then the Laplace transform

f of f is analytic in
the half-plane (s) > 0, and it satisfy, in addition
sup
>0
_

f( + i)[
2
d < ,
i.e., there is a constant M so that
_

f( + i)[
2
d M (for all > 0).
Proof. By Theorem 6.12 and the L
2
-theory for Fourier integrals (see Section
2.3),
_

f( + i)[
2
d =
_

0
[2e
2t
f(2t)[
2
dt (2t = v)
= 2
_

0
[e
v
f(v)[
2
dv
2
_

0
[f(v)[
2
dv = 2|f|
L
2
(0,)
.
CHAPTER 6. THE LAPLACE TRANSFORM 129
Converesly:
Theorem 6.17. If is analytic in (s) > 0, and satises
sup
>0
_

[( + i)[
2
d < , (6.3)
then is the Laplace transform of a function f L
2
(1
+
).
Proof. Not too dicult (but rather long).
Denition 6.18. An H
2
-function over the half-plane (s) > 0 is a function
which is analytic and satises (6.3).
6.4 The Laplace Transform of a Distribution
Let f o

(tempered distribution), and suppose that the support of f is con-


tained in [0, ) = 1
+
(i.e., f vanishes on (, 0)). Then we can dene the
Laplace transform of f in two ways:
i) Make a change of variables as on page 126 and use the Fourier transform
theory.
ii) Dene

f(s) as f applied to the test function e
st
, t > 0. (Warning: this
is not a test function!)
Both methods lead to the same result, but the second method is actually simpler.
If (s) > 0, then t e
st
behaves like a test function on [0, ) but not on
(, 0). However, f is supported on [0, ), so it does not matter how e
st
behaves for t < 0. More precisely, we take an arbitrary cut o function
C

pol
satisfying
_
(t) 1 for t 1,
(t) 0 for t 2.
Then (t)e
st
= e
st
for t [1, ), and since f is supported on [0, ) we can
replace e
st
by (t)e
st
to get
Denition 6.19. If f o

vanishes on (, 0), then we dene the Laplace


transform

f(s) of f by

f(s) = f, (t)e
st
), (s) > 0.
CHAPTER 6. THE LAPLACE TRANSFORM 130
(Compare this to what we did on page 84).
Note: In the same way we can dene the Laplace transform of a distribution that
is not necessarily tempered, but which becomes tempered after multiplication by
e
t
for some > 0. In this case the Laplace transform will be dened in the
half-plane s > .
Theorem 6.20. If f vanishes on (, 0), then

f is analytic on the half-plane
s > 0.
Proof omitted.
Note:

f need not be bounded. For example, if f =

, then

)(s) =

, (t)e
st
) = , (t)e
st
)
=
d
dt
e
st
|t=0
= s.
(which is unbounded). On the other hand

(s) = , (t)e
st
) = e
st
|t=0
= 1.
Theorem 6.21. If f o

vanishes on (, 0), then


i) [

tf(t)](s) = [

f(s)]

ii)

f

(s) = s

f(s)
_
(s) > 0
Proof. Easy (homework?)
Warning 6.22. You can apply this distribution transform also to functions, but
remember to put f(t) = 0 for t < 0. This automatically leads to a -term in the
distribution derivative of f: after we dene f(t) = 0 for t < 0, the distribution
derivative of f is
f(0)
0
. .
dervatives of jump at zero
+ f

(t)
..
usual derivative
6.5 Discrete Time: Z-transform
This is a short continuation of the theory on page 101.
In discrete time we also run into one-sided convolutions (as we have seen), and
it is possible to compute these by the FFT. From a mathematical point of view
the Z-tranform is often simpler than the Fourier transform.
CHAPTER 6. THE LAPLACE TRANSFORM 131
Denition 6.23. The Z-transform of a sequence f(n)

n=0
is given by

f(z) =

n=0
f(n)z
n
,
for all these z C for which the series converges absolutely.
Lemma 6.24.
i) There is a number [0, ] so that

f(z) converges for [z[ > and

f(z)
diverges for [z[ < .
ii)

f is analytic for [z[ > .
Proof. Course on analytic functions.
As we noticed on page 101, the Z-transform can be converted to the discrete
time Fourier transform by a simple change of variable.
6.6 Using Laguerra Functions and FFT to Com-
pute Laplace Transforms
We start by recalling some results from the course in special functions:
Denition 6.25. The Laguerra polynomials /
m
are given by
/
m
(t) =
1
m!
e
t
_
d
dt
_
m
(t
m
e
t
), m 0,
and the Laguerra functions
m
are given by

,
(t) =
1
m!
e
t
2
_
d
dt
_
m
(t
m
e
t
), m 0.
Note that
m
(t) = e

t
2
/
m
(t).
Lemma 6.26. The Laguerra polynomials can be computed recusively from the
formula
(m + 1)/
m+1
(t) + (t 2m1)/
m
(t) + m/
m1
(t) = 0,
with starting values /
1
0 and /
1
1.
CHAPTER 6. THE LAPLACE TRANSFORM 132
We saw that the sequence
m

m=0
is an ortonormal sequence in L
2
(1
+
), so that
if we dene, for some f L
2
(1
+
),
f
m
=
_

0
f(t)
m
(t)dt,
then
f(t) =

m=0
f
m

m
(t) (in the L
2
-sense). (6.4)
Taking Laplace transforms in this equation we get

f(s) =

m=0
f
m

m
(s).
Lemma 6.27.
i)

m
(s) =
(s1/2)
m
(s+1/2)
m+1
,
ii)

f(s) =

m=0
f
m
(s1/2)
m
(s+1/2)
m+1
, where f
m
=
_

0
f(t)
m
(t)dt.
Proof. Course on special functions.
The same method can be used to compute inverse Laplace transforms, and this
gives a possibility to use FFT to compute the coecients f
m

m=0
if we know

f(s). The argument goes as follows.


Suppose for simplicity that f L
1
(1), so that

f(s) is dened and bounded on
C
+
= s C[Re s > 0. We want to expand

f(s) into a series of the type

f(s) =

m=0
f
m
(s 1/2)
m
(s + 1/2)
m+1
. (6.5)
Once we know the cocients f
m
we can recover f(t) from formula (6.4). To nd
the coecients f
m
we map the right half-plane C
+
into the unit disk | = z
C : [z[ < 1. We dene
z =
s 1/2
s + 1/2
sz +
1
2
z = s
1
2

s =
1
2
1 + z
1 z
and s + s + 1/2 =
1
2
(1 +
1 + z
1 z
) =
1
1 z
, so
1
s + 1/2
= 1 z
Lemma 6.28.
CHAPTER 6. THE LAPLACE TRANSFORM 133
i) (s) > 0 [z[ < 1 item[ii)](s) = 0 [z[ = 1
iii) s = 1/2 z = 0
iv) s = z = 1
v) s = 0 z = 1
vi) s = 1/2 z =
Proof. Easy.
Conclusion: The function

f(
1
2
1+z
1z
) is analytic inside the unit disc |, (and bounded
if

f is bounded on C
+
).
Making the same change of variable as in (6.5) we get
1
1 z

f(
1
2
1 + z
1 z
) =

m=0
f
m
z
m
.
Let us dene
g(z) =
1
1 z

f(
1
2
1 + z
1 z
), [z[ < 1.
Then
g(z) =

m=0
f
m
z
m
,
so g(z) is the mathematical version of the Z-transform of the sequence f
m

m=0
(in the control theory of the Z-transform we replace z
m
by z
m
).
If we know

f(s), then we know g(z), and we can use FFT to compute the
coecients f
m
: Make a change of variable: Put
N
= e
2i/N
. Then
g(
k
N
) =

m=0
f
m

mk
N
=

m=0
f
m
e
2imk/N

m=0
f
m
e
2imk/N
(if N is large enough). This is the inverse discrete Fourier transform of a periodic
extension of the sequence f
m

N1
m=0
. Thus, f
m
the discrete transformation of
the sequence g(
k
N
)
N1
k=0
. We put
G(k) = g(
k
N
) =
1
1
k
N

f(
1
2
1 +
k
N
1
k
N
),
and get f
m


G(m), which can be computed with the FFT.
CHAPTER 6. THE LAPLACE TRANSFORM 134
Error estimate: We know that f
m
= g(m) (see page 115) and that g(m) = 0 for
m < 0. By the error estimate on page 108 we get
[

G(m) f
m
[ =

k=0
[f
m+kN
[
(where we put f
m
= 0 for m < 0).
Bibliography
[Gri04] Gustaf Gripernberg, Fourier-muunnosten perusteet (lecture notes),
http://math.tkk./teaching/fourier/, 2004.
[GW99] C. Gasquet and P. Witomski, Fourier analysis and applications, Texts
in Applied Mathematics, vol. 30, Springer-Verlag, New York, 1999, Fil-
tering, numerical computation, wavelets, Translated from the French
and with a preface by R. Ryan. MR MR1657104 (99h:42003)
[Wal02] David F. Walnut, An introduction to wavelet analysis, Applied and Nu-
merical Harmonic Analysis, Birkhauser Boston Inc., Boston, MA, 2002.
MR MR1854350 (2002f:42039)
135

Anda mungkin juga menyukai