11.08.2013 Views

I. FOURIER SERIES, FOURIER TRANSFORM.

I. FOURIER SERIES, FOURIER TRANSFORM.

I. FOURIER SERIES, FOURIER TRANSFORM.

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Computational Methods and advanced Statistics Tools<br />

I. <strong>FOURIER</strong> <strong>SERIES</strong>,<br />

<strong>FOURIER</strong> <strong>TRANSFORM</strong>.<br />

1. UNIFORM CONVERGENCE<br />

The door of faith (Acts 14:27) is always open for us, ushering us into the life<br />

of communion with God and offering entry into his Church. It is possible to<br />

cross that threshold when the word of God is proclaimed and the heart allows<br />

itself to be shaped by transforming grace. To enter through that door is to<br />

set out on a journey that lasts a lifetime.<br />

(Porta fidei - 2012/2013 a Year of Faith)<br />

A series of functions is a series where the general term is a sequence of<br />

functions:<br />

∞<br />

fn(x), x 2 I<br />

n=1<br />

where I is some common domain ½ R.<br />

The series is pointwise convergent to a sum S(x) in I if<br />

n<br />

8x 2 I, 8ǫ > 0, 9¯n(ǫ, x) > 0 : j fk(x) ¡ S(x)j < ǫ, 8n > ¯n(ǫ, x).<br />

The series is uniformly convergent to S(x) in I if<br />

n<br />

8ǫ > 0, 9¯n(ǫ) > 0 : 8x 2 I, j gk(x) ¡ S(x)j < ǫ, 8n > ¯n(ǫ)<br />

k=1<br />

k=1<br />

If the functions fn(x), n 2 N, are continuous for x 2 I, the point<br />

convergence is not sufficient to ensure that the sum S(x) is continuous.<br />

But if the functions fn(x), n 2 N, are continuous for x 2 I, and the<br />

series of functions is uniformly convergent to S(x) for x 2 I, then S(x)<br />

is a continuous function in I.<br />

In turn, a sufficient condition for the uniform convergence of a series of<br />

functions is the ”total convergence” criterion: the series of functions is<br />

uniformly convergent in I if the modulus of the general term is bounded<br />

by the general term of a convergent numerical series.<br />

In other words, if<br />

∞<br />

jfn(x)j · αn, 8x 2 I, with αn < +1<br />

then the series of functions ∞<br />

n=1 fn(x) is uniformly convergent.<br />

1<br />

n=1


2<br />

When such a condition is satisfied one can prove that:<br />

i) the sum of the series of function S(x) is continuous in I;<br />

ii) the definite integral on [a, b] ½ I of the sum of the series is equal to<br />

the sum of the integrals of fn(x);<br />

iii) if, moreover, the general term fn(x) is differentiable and, in the<br />

interval I, jf ′ n (x)j is bounded by the general term of convergent numer-<br />

ical series, the sum S(x) of the series of functions is differentiable and<br />

the sum of the derivatives converges to S ′ (x).<br />

For example, since<br />

<br />

n<br />

1<br />

<br />

j sin(nx)j ·<br />

n3 n<br />

1<br />

<br />

< +1 and<br />

n3 n<br />

1<br />

<br />

jn cos(nx)j ·<br />

n3 n<br />

1<br />

< +1<br />

n2 the sum S(x) of <br />

n sin(nx)/n3 exists, it is continuous, it admits a continuous<br />

derivative S ′ (x), which is the sum of the series of the derivatives<br />

8x 2 R. In such a case we say that the series can be derived term by<br />

term. Besides, for any finite a, b with a < b,<br />

<br />

b b <br />

fn(x)dx = fn(x)dx<br />

n<br />

a<br />

for fn(x) = sin(nx)/n 3 , as well as for fn(x) = cos(nx)/n 2 . In such<br />

cases we say that the series can be integrated term by term.<br />

2. <strong>FOURIER</strong> <strong>SERIES</strong><br />

And that [the unheard-of precision of the processes associated with the ”Big-<br />

Bang”] is supposed to have happened by chance?! What an absurd idea!<br />

(Walter Thirring, b.1927, physicist)<br />

A function f is said to be periodic if it is defined for all real x and if<br />

there is a positive number T such that<br />

(1)<br />

f(x + T ) = f(x), 8x<br />

According to this definition, a constant function is periodic with arbitrary<br />

period. Trigonometric functions, such as sin x, sin 2x, ..., sin nx,<br />

or cos x, ..., cos nx, are periodic with period 2π. If T is a period for f<br />

then f(x + nT ) = f(x), 8n 2 N, so that f is periodic with periods<br />

2T, 3T, ..., too. A trigonometric polynomial is a finite sum of the form<br />

(2)<br />

a◦<br />

2 +<br />

k<br />

(an cos nx + bn sin nx)<br />

n=1<br />

Any trigonometric polynomial is periodic with period 2π.<br />

a<br />

n


Let us assume that f(x) is a periodic function with period 2π that can<br />

be represented by a trigonometric series,<br />

f(x) = a◦<br />

2 +<br />

∞<br />

(3)<br />

(an cos nx + bn sin nx)<br />

−π<br />

−π<br />

n=1<br />

that is, we assume that this series converges and has f(x) as its sum.<br />

Given such a function f(x), we want to determine the cifoefficients an<br />

and bn of the corresponding series (3). We first determine a◦. Integrating<br />

on both sides of (3) from ¡π to π, and assuming that term-by<br />

term integration of the series is allowed, we have<br />

π<br />

f(x) dx = a◦<br />

π ∞<br />

π<br />

π <br />

dx+ an cos nx dx + bn sin nx dx .<br />

2<br />

n=1<br />

The first term on the right equals πa◦. All the other integrals on the<br />

right are zero, as can be readily seen by performing the integrations.<br />

Hence our first result is<br />

(4)<br />

a◦<br />

2<br />

−π<br />

π<br />

1<br />

= f(x) dx.<br />

2π −π<br />

We now determine a1, a2, ... by a similar procedure. We multiply (3) by<br />

cos mx, where m is any fixed positive integer, and integrate from ¡π<br />

to π term by term, we see that π<br />

f(x) cos mx dx becomes:<br />

−π<br />

π<br />

a◦<br />

cos mx dx+<br />

2<br />

(5)<br />

+<br />

∞<br />

·<br />

n=1<br />

an<br />

−π<br />

π<br />

π<br />

cos nx cos mx dx + bn<br />

−π<br />

−π<br />

−π<br />

<br />

sin nx cos mx dx .<br />

The first integral is zero. Now by the trigonometric identities<br />

cos α cos β = 1<br />

[cos(α + β) + cos(α ¡ β)]<br />

2<br />

sin α cos β = 1<br />

[sin(α + β) + sin(α ¡ β)]<br />

2<br />

we obtain<br />

π<br />

cos nx cos mx dx =<br />

−π<br />

1<br />

π<br />

cos(n + m)x dx +<br />

2 −π<br />

1<br />

π<br />

cos(n ¡ m)x dx,<br />

2 −π<br />

π<br />

sin nx cos mx dx =<br />

−π<br />

1<br />

π<br />

sin(n + m)x dx +<br />

2 −π<br />

1<br />

π<br />

sin(n ¡ m)x dx.<br />

2 −π<br />

Integration shows that the four terms on the right are zero, except for<br />

the last term in the first line, which equals π when n = m. Since in (5)<br />

this term is multiplied by am, the whole of (5) is equal to amπ, and our<br />

second result is<br />

am = 1<br />

π<br />

(6)<br />

f(x) cos mx dx, m = 1, 2, ...<br />

π<br />

−π<br />

3


4<br />

We finally determine b1, b2, ... in (3). If we multiply (3) by sin mx,<br />

where m 2 N, and then integrate from ¡π to π, we have<br />

π<br />

<br />

π<br />

a◦<br />

f(x) sin mx dx =<br />

2 +<br />

∞<br />

<br />

(7)<br />

(an cos nx + bn sin nx) sin mx dx<br />

−π<br />

−π<br />

1<br />

Integrating term by term, we see that the right hand side becomes<br />

π<br />

∞<br />

· π<br />

a◦<br />

sin mx dx+ an<br />

2<br />

π<br />

<br />

cos nx sin mx dx + bn sin nx sin mx dx<br />

−π<br />

1<br />

−π<br />

The first integral is zero. The next integral is of the type considered<br />

before, and we know that it is zero for all n 2 N. For the last integral<br />

we need the identity<br />

we obtain<br />

π<br />

sin nx sin mx dx = 1<br />

π<br />

2<br />

−π<br />

−π<br />

sin α sin β = ¡ 1<br />

[cos(α + β) ¡ cos(α ¡ β)]<br />

2<br />

−π<br />

cos(n ¡ m)x dx ¡ 1<br />

π<br />

cos(n + m)x dx,<br />

2 −π<br />

The last term is zero. The first term on the right is zero when n 6= m<br />

and is π when n = m. Since in (7) this term is multiplied by bm, the<br />

right hand side in (7) is equal to bmπ, and our last result is<br />

(8)<br />

bm = 1<br />

π<br />

π<br />

f(x) sin mx dx, m 2 N.<br />

−π<br />

The above calculation can be performed for a periodic function with<br />

any period T > 0, (f(x) = f(x + T ), 8x 2 R). Besides, since the<br />

integrals (4,6, 8) exist under wide conditions (e.g. if f is piecewise<br />

continuous), we give the following definition:<br />

DEF. 2.A (Fourier series) If f(x) is a piecewse continuous, periodic<br />

function with period 2π, the Fourier series associated with f(x) is the<br />

(formal) series<br />

(9)<br />

f(x) » a◦<br />

2 +<br />

∞<br />

(an cos nx + bn sin nx)<br />

n=1<br />

where the Fourier coefficients an, bn are given by:<br />

an = 1<br />

π<br />

(10)<br />

f(x) cos nx dx, n = 0, 1, 2, ...<br />

π<br />

(11)<br />

−π<br />

bn = 1<br />

π<br />

f(x) sin nx dx, n = 1, 2, ...<br />

π −π


More generally, for a piecewise continuous, periodic function with period<br />

T , the associated Fourier series is<br />

f(x) » a◦<br />

2 +<br />

∞<br />

·<br />

an cos( 2π<br />

T nx) + bn sin( 2π<br />

T nx)<br />

<br />

(12)<br />

n=1<br />

with Fourier coefficients:<br />

an = 2<br />

(13)<br />

T<br />

T/2<br />

(14)<br />

4<br />

bn = 2<br />

T<br />

−T/2<br />

T/2<br />

−T/2<br />

f(x) cos( 2π<br />

nx) dx, n 2 N◦<br />

T<br />

f(x) sin( 2π<br />

nx) dx, n 2 N<br />

T<br />

REMARK 2.B - a) An equivalent choice is to express the coefficients<br />

by integrals from 0 to T , by virtue of periodicity of the function f:<br />

an = 2<br />

T<br />

f(x) cos(<br />

T 0<br />

2π<br />

(15)<br />

nx) dx, n 2 N◦<br />

T<br />

bn = 2<br />

T<br />

f(x) sin(<br />

T 0<br />

2π<br />

nx) dx, n 2 N<br />

T (16)<br />

Actually the same numbers are obtained by integrating on any interval<br />

[c, c + T ], with fixed c 2 R.<br />

b) The symbol » reminds that the trigonometric expansion is still<br />

formal: by that symbol nothing is said about convergence, and, if the<br />

series is convergent, the sum is not necessarily f(x).<br />

c) The constant term in (12), i.e. a◦/2 = 1<br />

T<br />

f(x) dx, is the mean<br />

T 0<br />

value of f in one period.<br />

REMARK 2.C - Coefficients of the Fourier series in exponential form.<br />

- If the Fourier series of a T ¡ periodic function f(x) is in exponential<br />

form, i.e.<br />

then the coefficients are<br />

f(x) »<br />

cn = 1<br />

T<br />

T<br />

0<br />

+∞<br />

n=−∞<br />

2π<br />

i<br />

cne T nx<br />

2π<br />

−i<br />

f(x)e T nx dx<br />

The exponential Fourier series is equivalent to the trigonometric Fourier<br />

series.<br />

Proof. Take, for simplicity, T = 2π. As above, integrating term by<br />

term 2π<br />

f(x)dx = 2πc◦ + <br />

[ einx<br />

0 = 2πc◦<br />

0<br />

n=0<br />

in ]2π<br />

5


6<br />

so that c◦ is the mean of the periodic function.<br />

Multiplying both sides by e−imx , m 6= 0, and integrating we find:<br />

2π<br />

f(x)e −imx dx = e<br />

[cn<br />

i(n−m)x<br />

i(n ¡ m) ]2π 0 + 2πcm<br />

0<br />

n=m<br />

so that cm = (2π) −1 2π<br />

0 f(x)e−imx dx as stated.<br />

Of course the equivalence depends on the Euler identity: e inx =<br />

cos(nx) + i sin(nx). 4<br />

3. THEOREMS ON <strong>FOURIER</strong> <strong>SERIES</strong><br />

[God] desires all men to be saved and come to the knowledge of the truth.<br />

( 1 Tim 2:4 )<br />

THEOREM 3.A (A priori bound on Fourier coefficients)<br />

If the periodic function f admits continuous r¡th derivative in [0, T ]<br />

with r ¸ 1, then the coefficients of the Fourier exponential series satisfy<br />

jcnj · cjnj −r , 8n 2 Z,<br />

where the constant c is independent of n. Similarly, janj · ˜cn −r , jbnj ·<br />

˜cn −r , if an, bn are the coefficients of the trigonometric series.<br />

Proof. - Let f 2 Cr ([0, T ]). The derivative of a periodic function is<br />

still periodic, so integrating by parts r times the expression (?? )<br />

cn = 1<br />

·<br />

f(x)<br />

T<br />

T<br />

2π<br />

e−i T<br />

¡i2πn nx<br />

T T<br />

¡ f ′ (x) T<br />

2π<br />

e−i T<br />

¡i2πn nx <br />

dx<br />

= + 1<br />

T<br />

Therefore<br />

where<br />

T<br />

T<br />

f<br />

i2πn 0<br />

′ 2π<br />

−i<br />

(x)e T nx dx = + 1<br />

T<br />

jcnj · 1<br />

T<br />

0<br />

0<br />

· T<br />

i2πn<br />

r T<br />

·<br />

T<br />

r T<br />

2πjnj 0<br />

· 1<br />

T jnjr · r T<br />

T<br />

jf<br />

2π 0<br />

(r) (x)j dx · c<br />

jnjr c =<br />

· r T<br />

2π<br />

jf (r) 2π<br />

−i<br />

(x)jje T nx j dx<br />

max<br />

x∈[0,T ] jf (r) (x)j.<br />

0<br />

f (r) 2π<br />

−i<br />

(x)e T nx dx.<br />

Since an ¡ ibn = 2cn, choosing ˜c = 2c the proof is complete. 4


Def. A function f(x) is said piecewise continuous in (a, b) ½ R if f<br />

is defined and continuous except in a finite number of points xj 2 (a, b),<br />

j = 1, ..., N, where there exist the finite limits:<br />

f(xj + 0) = limx→xj+(x), f(xj ¡ 0) = limx→xj−f(x).<br />

Such type of regularity is part of the Dirichlet condition, assumed in<br />

the most usual convergence result, that we state without proof:<br />

THEOREM 3.B (Dirichlet’s Criterion) Let the function f(x) be periodic<br />

(with period T ). Let both f(x) and f ′ (x) be piecewise continuous.<br />

Then the trigonometric series associated with f is convergent to f(x)<br />

in each point where f is continuous; while in each point x◦ where f is<br />

discontinuous, the series is convergent to the half-sum<br />

f(x◦ + 0) + f(x◦ ¡ 0)<br />

.<br />

2<br />

EXAMPLE 3.C (Fourier expansion of a periodic step function) - Let<br />

<br />

1 x 2 [2nπ, (2n + 1)π]<br />

f(x) =<br />

0 elsewhere.<br />

which satisfies the Dirichlet conditions, with period T = 2π. The<br />

Fourier coefficients a◦, an, bn are directly calculated:<br />

π<br />

a◦ 1<br />

= f(x) dx =<br />

2 2π<br />

1<br />

π<br />

1 dx =<br />

2π<br />

1<br />

2 .<br />

−π<br />

For n 2 N,<br />

an = 1<br />

π<br />

f(x) cos(nx) dx =<br />

π −π<br />

1<br />

π<br />

cos(nx) dx =<br />

π 0<br />

1<br />

· π sin(nx)<br />

= 0,<br />

π n 0<br />

bn = 1<br />

π<br />

f(x) sin(nx) dx =<br />

π −π<br />

1<br />

π<br />

sin(nx) dx =<br />

π 0<br />

1<br />

·<br />

¡<br />

π<br />

cos(nx)<br />

π =<br />

n 0<br />

= 1<br />

nπ [1 ¡ (¡1)n <br />

0, n even<br />

] = 2 , n odd .<br />

nπ<br />

Thus bn vanishes for even n, it equals 2<br />

nπ<br />

is:<br />

1<br />

2 +<br />

∞<br />

[an cos(nx) + bn sin(nx)] = 1<br />

2 +<br />

n=1<br />

0<br />

for odd n. The Fourier series<br />

∞<br />

n=0<br />

2<br />

sin [(2n + 1)x] .<br />

(2n + 1)π<br />

REMARK 3.D - (On the computation of Fourier coefficients)<br />

7


8<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

−5 0 5<br />

Figure 1. Graph of the Fourier series calculated by the<br />

first 25 harmonics, i.e. for n from 0 to 12.<br />

a) In order to compute the coefficients of the Fourier series, several<br />

remarks are useful.<br />

A function f(x) is said to be even if f(¡x) = f(x), 8x; it is odd if<br />

f(¡x) = ¡f(x), 8x. If the function f(x) is even, then all the Fourier<br />

coefficients bn vanish: indeed, under a substitution y = ¡x,<br />

bn = 2<br />

T<br />

T 0<br />

T/2<br />

2<br />

f(¡y) sin(¡<br />

T<br />

2π<br />

2<br />

ny) dy = ¡<br />

T T<br />

−T/2<br />

f(x) sin( 2π 2<br />

nx) dx =<br />

T T<br />

x<br />

T/2<br />

−T/2<br />

T/2<br />

−T/2<br />

f(x) sin( 2π<br />

nx) dx =<br />

T<br />

f(y) sin( 2π<br />

ny) dy = ¡bn<br />

T<br />

Thus 2bn = 0, so that bn = 0. Similarly, it turns out that a◦ = an = 0<br />

if the function is odd.<br />

b) Therefore an even periodic function admits a series expansion with<br />

only cosine terms; while an odd periodic function admits a series<br />

expansion with only sine terms. The next remark regards an arbitrary<br />

function f(x), defined in some interval, for example in (0, π). We can<br />

extend it both to an even function and to an odd function. Indeed,<br />

we can define it in the interval (¡π, 0) so that it becomes even:<br />

⎧<br />

⎤<br />

⎨ f(x), x 2 (0, π)<br />

f◦(x) = limx→0+ f(x), x = 0 ⎦<br />

⎩<br />

f(¡x), x 2 (¡π, 0)<br />

Another choice is to define an odd extension of f(x):<br />

⎧<br />

⎤<br />

⎨ f(x), x 2 (0, π)<br />

f1(x) = 0, x = 0 ⎦<br />

⎩<br />

¡f(¡x), x 2 (¡π, 0)


Finally, one can extend the functions f◦(x) and f1(x) to the whole of<br />

R, so that the resulting function is periodic with period 2π. Assuming<br />

the Dirichlet conditions to hold, it follows that<br />

(17)<br />

where<br />

and<br />

(18)<br />

f◦(x) = 1<br />

2 a◦ +<br />

∞<br />

an cos(nx)<br />

n=1<br />

an = 1<br />

π<br />

f◦(x) cos(nx) dx =<br />

π −π<br />

2<br />

π<br />

f(x) cos(nx) dx<br />

π 0<br />

f1(x) =<br />

∞<br />

bn sin(nx)<br />

n=1<br />

where<br />

bn = 1<br />

π<br />

f1(x) sin(nx) dx =<br />

π −π<br />

2<br />

π<br />

f(x) sin(nx) dx<br />

π 0<br />

(indeed f1(x)sin(nx) is even, as a product of two odd functions). Now<br />

under a restriction of (17), (18) to the interval (0, π), it follows that a<br />

function with domain (0, π) can be expanded in a cosine-series as well<br />

as in a sine-series.<br />

4. APPLICATION OF THE <strong>FOURIER</strong> <strong>SERIES</strong> TO<br />

DIFFERENTIAL EQUATIONS<br />

To believe in a God means to realize that the facts of the world are not the<br />

whole story. To believe in a God means to realize that life has a meaning.<br />

(Ludwig Wittgenstein, 1889-1951)<br />

EX. 4.A - The damped harmonic oscillator with a periodic force<br />

Let us compute a particular solution of the damped harmonic oscillator<br />

with a forcing term<br />

(19)<br />

¨x + µ ˙x + ω 2 x = f(t)<br />

when f(t) is a real valued periodic function with period T . Although<br />

not real-valued, it is convenient first to consider<br />

where Ω = 2π<br />

T<br />

f(t) = ρ ¢ e iΩt<br />

is the pulsation of the forcing term. In this case a<br />

particular solution in the form const. ¢ eiΩt is easily found:<br />

x ∗ (t) =<br />

ρe iΩt<br />

ω 2 ¡ Ω 2 + iΩµ<br />

9


10<br />

As a second step let us consider f(t) given by a finite Fourier expansion<br />

in exponential form:<br />

N<br />

f(t) = cne iΩnt<br />

n=−N<br />

where cn = ¯c−n since f(t) is real valued. A particular solution with<br />

period T is given by<br />

x ∗ (t) =<br />

N<br />

x ∗ n(t), x ∗ n(t) =<br />

n=−N<br />

cne iΩnt<br />

ω 2 ¡ n 2 Ω 2 + inΩµ<br />

Notice that x∗ is real since the addends of the sum with opposite indices<br />

n, ¡n are complex conjugate. Moreover x∗ n (t) is a particular solution<br />

of the differential equation<br />

¨x + µ ˙x + ω 2 x = cne iΩnt .<br />

Thus we can immediately verify:<br />

¨x ∗ + µ ˙x ∗ + ω 2 x ∗ N<br />

= (¨x ∗ n + µ ˙x ∗ n + ω 2 x ∗ n) =<br />

n=−N<br />

N<br />

n=−N<br />

cne iΩnt = f(t).<br />

Finally, let f(t) admit a Fourier expansion with infinitely many nonzero<br />

coefficients:<br />

f(t) =<br />

∞<br />

n=−∞<br />

cne iΩnt<br />

A particular solution is looked for in the form:<br />

x ∗ (t) =<br />

∞ cn<br />

ω2 ¡ n2Ω2 + inΩµ eiΩnt<br />

(20)<br />

n=−∞<br />

To this end we have to check the convergence of (20); then, provided<br />

its convergence, to check whether is it a solution of the differential<br />

equation. Now a sufficient condition to derive k times term by term<br />

is the convergence of the series of the k¡th order derivatives:<br />

(21)<br />

∞<br />

n=−∞<br />

cn(iΩn) k<br />

ω 2 ¡ n 2 Ω 2 + inΩµ eiΩnt<br />

uniformly with respect to t. On the other hand we know that f 2 C r<br />

implies the bound on the Fourier coefficients jcnj · cn −r . Therefore<br />

the general term of (23) is bounded by<br />

cΩ k<br />

n k<br />

nr (ω2 ¡ n2Ω2 ) 2 + n2 µ 2 · Cnk−r−2<br />

Ω2 for some constant C > 0 independent of n. So the series (23) is uniformly<br />

convergent with respect to t if<br />

r + 2 ¡ k > 1;


In particular, the series (20) is a solution of the differential equation if<br />

r + 2 ¡ 2 > 1 (k = 2), i.e. if the function f(t) is at least in C 2 .<br />

Notice: for µ small enough, the particular solution x ∗ (t) amplifies the<br />

harmonics n = §[ω/Ω], (where [¢] denotes the integral part), of the<br />

function f(t) for which ω 2 ¡n 2 Ω 2 attains its minimum, while the other<br />

harmonics are damped.<br />

EXAMPLE 4.B Let us consider the one-dimensional heat equation<br />

with boundary and initial conditions:<br />

⎧<br />

∂u ⎨ ∂t<br />

⎩<br />

= 2 ∂2u ∂x2 (22) u(0, t) = u(3, t) = 0,<br />

u(x, 0) = 5 sin(4πx) ¡ 3 sin(8πx) + 2 sin(10πx),<br />

⎤<br />

8t ¸ 0 ⎦<br />

x 2 [0, 3].<br />

Besides, the solution is required to remain bounded 8(x, t) 2 [0, 3]£R + .<br />

Let us solve the equation by separation of variables. By such a method,<br />

we look for a solution (if possible) of the form<br />

u(x, t) = f(x)g(t)<br />

. Then the equation reads f(x) ˙g(t) = 2f ′′ (x)g(t). After separating<br />

variables,<br />

2 ¢ f ′′ (x) ˙g(t)<br />

=<br />

f(x) g(t) .<br />

Since the left hand side depends only on x, the right hand side depends<br />

only on t, and the equation holds for all (x, t) where x, t are independent<br />

variables,<br />

9λ 2 R : 2 f ′′ (x) ˙g(t)<br />

= 2λ =<br />

f(x) g(t) .<br />

Therefore we obtain two ordinary differential equations depending on<br />

an arbitrary parameter λ:<br />

with general solutions<br />

(23)<br />

f ′′ = λf, ˙g = 2λg<br />

f(x) = c1e √ λx + c2e −√ λx , g(t) = ce 2λt<br />

if λ 6= 0<br />

f(x) = c1x + c2, g(t) = c if λ = 0.<br />

Now, the boundedness requirement 8t 2 R + implies λ < 0. Therefore,<br />

setting λ = ¡ω2 and c1 = 1<br />

2 (a + ib = ¯c2, the general solutions (23) can<br />

be written:<br />

f(x) = a cos(ωx) + b sin(ωx), g(t) = ce −2ω2 t .<br />

Hence a solution of the heat equation turns out to be<br />

u(x, t) = f(x)g(t) = ce −2ω2 t [a cos(ωx) + b sin(ωx)] =<br />

= e −2ω2 t [A cos(ωx) + B sin(ωx)]<br />

11


12<br />

u(x,t)<br />

−5 0 5<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0<br />

Figure 2. Graph of the function u(x, t) for some values of<br />

t: t = 0 (continuous line), t = 0.001 and t = 0.01<br />

with constants A, B, ω to be determined by using the boundary conditions<br />

in (??):<br />

u(0, t) = 0 =) A = 0, i.e. u(x, t) = e −2ω2 t B sin(ωx)<br />

u(3, t) = 0 =) B sin(3ω) = 0<br />

This last equality, in turn, implies<br />

either B = 0, arbitrary ω, or arbitrary B, ω = nπ/3, n 2 Z.<br />

Now B = 0 is immediately rejected since it gives an identically vanishing<br />

solution. Therefore a nontrivial solution of the diffusion equation<br />

in (??) with zero boundary condition has the form:<br />

u(x, t) = Be −2n2π2t/9 nπ<br />

sin( x), n 2 N<br />

3 (24)<br />

By the superposition principle (which holds for linear differential equations),<br />

the sum of solutions of a homogeneous differential equation is<br />

still a solution; so for any choice of n1, n2, n3 2 N the function<br />

B1e −2n2 1π2t/9 n1π<br />

sin(<br />

3 x)+B2e −2n2 2π2t/9 n2π<br />

sin(<br />

3 x)+B3e −2n2 3π2t/9 n3π<br />

sin(<br />

3 x)<br />

is still a solution of the equation in (??). By comparing this type of<br />

solution with the initial condition in (??) we find:<br />

B1 = 5, n1π/3 = 4π, B2 = ¡3, n2π/3 = 8π, B3 = 2, n3π/3 = 10π<br />

i.e. n1 = 12, n2 = 24, n3 = 30. Thus the unique (bounded) solution<br />

of the diffusion problem (??) is<br />

u(x, t) = 5e −32π2 t sin(4πx) ¡ 3e −128π 2 t sin(8πx) + 2e −200π 2 t sin(10πx)<br />

x


u(x,t)<br />

5<br />

0<br />

−5<br />

0.0<br />

0.5<br />

1.0<br />

1.5<br />

x<br />

2.0<br />

2.5<br />

3.0<br />

0.000<br />

0.002<br />

0.006<br />

t<br />

0.004<br />

0.010<br />

0.008<br />

Figure 3. Graph of the function u(x, t), for (x, t) 2 (0, 3) £ (0, 0.01)<br />

EXAMPLE 4.C - We solve the same diffusion equation with a different<br />

initial value function:<br />

⎧<br />

⎨<br />

⎩<br />

∂u<br />

∂t = 2 ∂2u ∂x2 u(0, t) = u(3, t) = 0, 8t ¸ 0<br />

u(x, 0) = 25x(3 ¡ x), x 2 [0, 3]<br />

Again by separation of variables we obtain a solution of the heat equation,<br />

but it is no more sufficient to make a finite superposition of such<br />

functions to solve the problem. We try by a superposition of infinitely<br />

many functions:<br />

u(x, t) =<br />

∞<br />

n=1<br />

Bne −2n2 π 2 t/9 sin(n π<br />

3 x),<br />

where we require u(x, 0) = 25x(3 ¡ x), that is<br />

25x(3 ¡ x) =<br />

∞<br />

n=1<br />

⎤<br />

⎦<br />

Bn sin(n π<br />

x), x 2 (0, 3).<br />

3<br />

This amounts to expand in a sine Fourier series the function 25x(3¡x).<br />

According to Remarks 2 (b), this is possible if we regard the function<br />

as a restriction of an odd periodic function defined for all x 2 R, with<br />

period 6:<br />

⎧<br />

⎤<br />

⎨ 25x(3 ¡ x), x 2 (0, 3)<br />

f2(x) = 25x(3 + x), x 2 (¡3, 0) ⎦<br />

⎩<br />

... and so on.<br />

13


14<br />

u(x,t)<br />

0 10 20 30 40 50<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0<br />

Figure 4. For initial shape u(x, 0) = 25x(3 ¡ x), graph of<br />

the solution u(x, t) when t = 0, t = 0.01, t = 0.1<br />

Then we have the usual Fourier coefficients for an odd function of<br />

period T :<br />

Bn = 2<br />

T/2<br />

f2(x) sin(n<br />

T −T/2<br />

2π<br />

T/2<br />

4<br />

x) dx = f2(x) sin(n<br />

T T 0<br />

2π<br />

x) dx =<br />

T<br />

= 2<br />

3<br />

25x(3 ¡ x) sin(nπx/3) dx =<br />

3 0<br />

900<br />

n3π3 [1 ¡ (¡1)n ]<br />

Finally<br />

u(x, t) 1800<br />

π3 ∞<br />

(2m + 1) −3 e −2(2m+1)2π2 t/9<br />

sin((2m + 1)πx/3).<br />

m=0<br />

The graph of u(x, t) is first represented for some values of t (Fig. 4),<br />

then for (x, t) 2 (0, 3) £ (0, 1) (Fig. 5).<br />

5. <strong>FOURIER</strong> <strong>TRANSFORM</strong><br />

Mans’s unique grandeur is ultimately based on his capacity to know the truth.<br />

And human beings desire to know the truth.<br />

Yet truth can only be attained in freedom. This is the case for all truth,<br />

as is clear from the history of science; but it is eminently the case with<br />

those truths in which man himself, man as such, is at stake, the truths of<br />

the spirit, the truths about good and evil, about the great goals and horizons<br />

of life, about our relationship with God. These truths cannot be attained<br />

without profound consequences for the way we live our lives.<br />

(Benedict XVI)<br />

x


u(x,t)<br />

t<br />

Figure 5. For initial shape u(x, 0) = 25x(3 ¡ x), graph of<br />

the solution u(x, t), (x, t) 2 (0, 3) £ (0, 1)<br />

DEFINITION 5.A (Direct Fourier transform) Let the function f(t) be<br />

given for any t 2 R. If the integral<br />

∞<br />

0<br />

f(t)e −iωt dt<br />

exists 8ω 2 R, then f(t) is said to admit the Fourier transform ˆ f(ω) :<br />

F (f) = ˆ +∞<br />

f(ω) = f(t)e −iωt (25)<br />

dt.<br />

−∞<br />

The operator F , which transforms f(t) into ˆ f(ω), is said ”Fourier<br />

transform”.<br />

Usually the variable t is a time variable, in which case the variable ω is<br />

a frequency. The Fourier transform ˆ f(ω) is a complex valued function<br />

that can be written in exponential notation as<br />

ˆf(ω) = A(ω)e iφ(ω)<br />

where the real function A(ω) is said Fourier spectrum, A 2 (ω) is the<br />

power spectrum, and φ(ω) is the phase angle.<br />

REMARK 5.B The complex conjugate of ˆ f(ω) is equal to the Fourier<br />

transform of the conjugate of f calculated in ¡ω:<br />

+∞<br />

¯ˆf(ω) = ( f(t)e −iωt dt) ¯ +∞<br />

= ¯f(t)e iωt dt = ˆ¯ f(¡ω).<br />

−∞<br />

In particular, if f(t) is a real valued function, then A(ω) = A(¡ω), i.e.<br />

A(ω) is an even function.<br />

−∞<br />

x<br />

15


16<br />

p_a(x)<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

indicatore di [−a,a]<br />

−a a<br />

Figure 6. Graph of pa(t).<br />

EXAMPLE 5.C We want to determine the Fourier transform of the<br />

indicator function of the interval [¡a, a],<br />

<br />

1, t 2 [¡a, a]<br />

(26)<br />

pa(t) =<br />

0, jtj > a<br />

where a > 0 is fixed. In this case the Fourier transform (25) reduces<br />

to the integral<br />

a<br />

ˆpa(ω) = e −iωt · −iωt a<br />

e<br />

dt =<br />

¡iω<br />

= eiωa ¡ e−iωa sin ωa<br />

= 2<br />

iω ω<br />

−a<br />

for ω 6= 0 (while ˆpa(0) = 2a). Fig. 6 and Fig. 7 show pa(t) and ˆpa(ω),<br />

respectively.<br />

EXAMPLE 5.D No sufficient condition for the existence of the Fourier<br />

transform has yet been stated. The expression (25) easily suggests that<br />

if the function f(t) is absolutely integrable on R, that is<br />

(27)<br />

∞<br />

−∞<br />

−a<br />

jf(t)j dt < +1,<br />

then ˆ f(ω) exists for any ω 2 R. In other words, under this assumption<br />

the integral (25) exists and is finite. Such a condition is sufficient but<br />

not necessary. Indeed there exist functions, such as sin(ω◦t)/ω, for<br />

which the Fourier transform exists, but the property (27) does not<br />

hold.<br />

The following is a regularity property.


Trasf. di Fourier di p_a(t)<br />

2a<br />

p ^ (ω)<br />

Figure 7. Graph of ˆpa(ω).<br />

THEOREM 5.E - If f(t) is absolutely integrable, and let ˆ f(ω) be its<br />

Fourier transform. Then<br />

i) limω→±∞ ˆ f(ω) = 0 (Riemann’s lemma)<br />

ii) ˆ f(ω) is continuous on R.<br />

Proof. We omit the proof of the first assertion, which is known as<br />

”Riemann’s lemma”.<br />

As for continuity, let ω◦ 2 R be fixed, and let<br />

∆ ˆ f = ˆ f(ω) ¡ ˆ +∞<br />

f(ω◦) = f(t)[e −iωt ¡ e −iω◦t ] dt =<br />

We need to prove that<br />

−∞<br />

+∞<br />

= f(t)e −iω◦t [e −i(ω−ω◦)t ¡ 1] dt.<br />

−∞<br />

8ǫ > 0, 9δ(ǫ) > 0 : jω ¡ ω◦j < δ =) j∆ ˆ fj < ǫ.<br />

Since f(t) is absolutely integrable, there is R = R(ǫ) > 0 such that<br />

−R +∞<br />

jf(t)jdt + jf(t)jdt · 1<br />

4 ǫ.<br />

−∞<br />

Besides, for such R = R(ǫ) we choose δ(R(ǫ)) such that<br />

max<br />

t∈[−R,R] ¢je−iδt ¡ 1j · 1<br />

2 ǫ<br />

+∞<br />

jf(t)jdt<br />

Now using<br />

and<br />

R<br />

−∞<br />

je −i(ω−ω◦)t ¡ 1j · je −i(ω−ω◦)t j + 1 = 2<br />

−1<br />

je −i(ω−ω◦)t ¡ 1j · je −iδt ¡ 1j, for jω ¡ ω◦j · δ, δt · π/2,<br />

17


18<br />

it follows that<br />

j∆ ˆ · −R +∞ R<br />

fj · 2 jf(t)jdt + jf(t)jdt + jf(t)jje<br />

−∞<br />

R<br />

−R<br />

−i(ω−ω◦)t ¡ 1jdt<br />

· 1 1<br />

ǫ + ǫ = ǫ<br />

2 2<br />

whenever jω ¡ ω◦j · δ. 4<br />

Now let us consider the inversion problem, assuming f(t) absolutely integrable<br />

on R. If ˆ f(ω) = F (f), the problem is how to reconstruct f(t)<br />

starting from ˆ f(ω). The following inversion theorem is stated without<br />

proof.<br />

PROPOSITION 5.F (Inverse Fourier transform) - Let the function f(t)<br />

be absolutely integrable on R and let f(t), f ′ (t) be piecewise continuous<br />

in any bounded interval. Then the inversion formula<br />

(28)<br />

f(t) = 1<br />

2π P<br />

+∞<br />

−∞<br />

ˆf(ω)e iωt dω = 1<br />

2π lim<br />

R→+∞<br />

+R<br />

−R<br />

ˆf(ω)e iωt dω<br />

holds in all points t where f(t) is continuous, while the formula<br />

1<br />

2 [f(t◦ + 0) + f(t◦ ¡ 0)] = 1<br />

2π P<br />

+∞<br />

ˆf(ω)e iωt (29)<br />

dω<br />

holds in all discontinuity points t◦. The operator F −1 trasforming ˆ f(ω)<br />

into f(t) (at least where f is continuous) is said the inverse Fourier<br />

transform. 4<br />

REMARK 5.G - The integral in (28)is computed as a Cauchy principal<br />

value. The integral<br />

<br />

is convergent as a Cauchy Principal Value if<br />

+R<br />

the limit limR→+∞ (...) exists and is finite. Recall that the usual<br />

−R<br />

generalized integral of a function g(t) is convergent if the double limit<br />

R<br />

lim g(t) dt<br />

R,S→+∞<br />

−S<br />

exists and is finite with R, S diverging to +1 independently of each<br />

other. Of course if the integral is convergent in the generalized sense,<br />

then it is convergent in the sense of the Cauchy principal value. But<br />

the opposite implication is not true in general. An example is the<br />

function g(t) = t, which is integrable in the Cauchy principal value<br />

sense:<br />

+∞<br />

+R<br />

P<br />

−∞<br />

t dt = lim<br />

R→+∞<br />

−R<br />

t dt = 0.<br />

The same function is not integrable in the generalized sense, since many<br />

different values are attained by the double limit prescription: for example<br />

R<br />

−2R tdt ! ¡1, while 2R<br />

tdt ! +1. A similar situation regards<br />

−R<br />

any odd, piecewise continuous function.<br />

−∞


REMARK 5.H In some textbooks the direct ad the inverse Fourier<br />

transforms are defined in a slightly different way, attributing to both<br />

the same coefficient 1/ p 2π (instead of 1 and 1/2π, respectively). Actually<br />

any other choice is good, provided that the product of the two<br />

coefficients is 1/2π.<br />

REMARK 5.I - (Symmetry property) If ˆ f(ω) is the Fourier transform<br />

of f(t) (and if both satisfy the hypotheses of Theorem 5.3) then the<br />

Fourier transform of ˆ f(t) is 2πf(¡ω), at least in the points in which f<br />

is continuous. Indeed<br />

F ( ˆ +∞<br />

f) = ˆf(t)e −iωt dt = 2π 1<br />

+∞<br />

ˆf(t)e<br />

2π<br />

−iωt dt = 2πf(¡ω).<br />

−∞<br />

−∞<br />

EX. 5.L Compute the Fourier transform of<br />

f(t) = sin(ω◦t)<br />

.<br />

t<br />

By Example 5.C, the indicator function pa(t) of the interval (¡a, a),<br />

admits Fourier transform<br />

ˆpa(ω) = 2 sin(ωa)<br />

.<br />

ω<br />

Hence the given function f(t) can be seen as a Fourier transform:<br />

f(t) = sin(ω◦t)<br />

=<br />

t<br />

1<br />

2 ˆpω◦(t),<br />

where pω◦ is the indicator function of (¡ω◦, ω◦). Finally, by the symmetry<br />

property<br />

ˆf(ω) ´ 1<br />

2 F [ˆpω◦](ω) = 2π 1<br />

2 pω◦(¡ω)<br />

at least in the points where pω◦ is continuous. Explicitly the result is:<br />

(30)<br />

(31)<br />

if f(t) = sin(ω◦t)<br />

,<br />

t<br />

then<br />

⎧<br />

⎨ π,<br />

ˆf(ω)<br />

1<br />

= π,<br />

⎩ 2<br />

0,<br />

⎤<br />

jωj < ω◦<br />

ω = §ω◦ ⎦<br />

jωj > ω◦<br />

EX. 5.M - Compute: a) the Gauss integral +∞<br />

−∞ e−αt2dt<br />

; b) the<br />

Fourier transform of the Gauss function e−αt2 (α > 0).<br />

a) Let I = +∞<br />

−∞ e−αx2dx.<br />

By Fubini’s theorem<br />

2 I 2 · +∞<br />

=<br />

−∞<br />

e −αx2<br />

dx<br />

+∞<br />

=<br />

−∞<br />

e −αx2<br />

dx<br />

+∞<br />

−∞<br />

e −αy2<br />

dy<br />

19


20<br />

+∞ +∞<br />

=<br />

−∞<br />

−∞<br />

e −α(x2 +y 2 ) dxdy<br />

Under a change of variables from cartesian to polar coordinates,<br />

<br />

x = r cos θ<br />

y = r sin θ<br />

, dxdy ! rdrdθ<br />

the integral becomes<br />

I 2 2π +∞<br />

=<br />

So I = π/α.<br />

b)<br />

0<br />

0<br />

e −αr2<br />

rdrdθ = 2π[e −αr2<br />

/(¡2α)] +∞ π<br />

0 =<br />

α .<br />

F [e −αt2<br />

<br />

] = e<br />

R<br />

−iωt e −αt2<br />

<br />

dt = e<br />

R<br />

−(√αt+ iω<br />

2 √ α )2 ω2<br />

−<br />

¢ e 4α dt<br />

Now by Cauchy’s theorem of complex analysis, the integral performed<br />

on the line R + iω<br />

2 √ (parallel to the real line in the complex plane) is<br />

α<br />

equal to the integral performed on the real line, which is π<br />

α :<br />

<br />

= e −(√αt) 2 ω2<br />

−<br />

¢ e 4α dt =<br />

R<br />

<br />

π ω2<br />

e− 4α .<br />

α<br />

Therefore, up to a coefficient, the Fourier transform of a Gauss function<br />

is still a Gauss function, with α replaced by ¡1/(4α).<br />

6. PROPERTIES OF THE <strong>FOURIER</strong> <strong>TRANSFORM</strong><br />

Revelation means that God opens himself, shows himself, and speaks to the<br />

world voluntarily.<br />

”All that is said about God presupposes something said by God” (Edith Stein)<br />

THEOREM 6.A (Linearity property) Let the functions f1, f2 admit<br />

Fourier transform F (f1) = ˆ f1 and F (f2) = ˆ f2, respectively. Then for<br />

arbitrary coefficients c1, c2 2 C there exists the Fourier transform of<br />

c1f1(t) + c2f2(t) and<br />

F (c1f1 + c2f2) = c1 ˆ f1 + c2 ˆ f2.<br />

Proof. A straightforward consequence of the linearity property of the<br />

integral. F and F −1 are linear operators, too. 4.<br />

THEOREM 6.B (Frequency shifting) Let the function f(t) admit<br />

Fourier transform ˆ f(ω). For any constant ω◦ 2 R,<br />

F [e iω◦t f(t)] = ˆ f(ω ¡ ω◦).


Proof.<br />

Trasf. di Fourier di p_a(t)cos(omega0 t)<br />

Figure 8. Graph of the Fourier transform of the modulated<br />

indicator function pa(t) cos(ω◦t), with ω◦ = 7, a = 5.<br />

F [e iω◦t +∞<br />

f] = f(t)e −i(ω−ω◦)t dt = ˆ f(ω ¡ ω◦) 4<br />

−∞<br />

EX. 6.C Compute the Fourier transform of the modulated indicator<br />

function<br />

f(t) = pa(t) cos(ω◦t),<br />

where pa(t) is the indicator of (¡a, a) and ω◦ is real and fixed. By Euler’s<br />

formula, linearity, frequency shifting and by the above examples,<br />

ˆf(ω) = F [pa(t) cos(ω◦t)] = 1<br />

2 [F (pa(t)e iω◦t<br />

) + F (pa(t)e −iω◦t<br />

)]<br />

= 1<br />

2 [ˆpa(ω ¡ ω◦) + ˆpa(ω + ω◦)]<br />

= sin[a(ω ¡ ω◦)]<br />

ω ¡ ω◦<br />

+ sin[a(ω + ω◦)]<br />

.<br />

ω + ω◦<br />

THEOREM 6.D (Time shifting) Let the function f(t) admit Fourier<br />

transform ˆ f(ω). Then<br />

Proof.<br />

F [f(t ¡ t◦)] = e −iωt◦ ˆ f(ω).<br />

+∞<br />

F [f(t ¡ t◦)] = f(t ¡ t◦)e −iωt dt<br />

−∞<br />

= e −iωt◦<br />

+∞<br />

f(t ¡ t◦)e −iω(t−t◦) dt<br />

−∞<br />

21


22<br />

= e −iωt◦<br />

+∞<br />

f(t)e −iωt dt = e −iωt◦ ˆ f(ω). 4<br />

−∞<br />

THM. 6.E (Time scaling) Let the function f(t) admit Fourier transform<br />

ˆ f(ω) and let a 2 R ¡ f0g be fixed. Then<br />

F [f(at)] = 1<br />

jaj ˆ f( ω<br />

a ).<br />

Proof. If a < 0 the change of variable τ = at changes the extreme<br />

value ¡1 into +1 and vice-versa, so that<br />

+∞<br />

F [f(at)] = f(at)e<br />

−∞<br />

−iωt +∞<br />

dt = −∞ f(τ)e−iωτ/a dτ/a, a > 0<br />

¡ +∞<br />

−∞ f(τ)e−iωτ/a dτ/a, a < 0<br />

= 1<br />

+∞<br />

f(τ)e<br />

jaj<br />

−iωτ/a dτ = 1<br />

jaj ˆ f( ω<br />

). 4<br />

a<br />

−∞<br />

Let us prove that the Fourier operator transforms derivatives in the<br />

time representation into multiplication by corresponding monomials in<br />

the frequency representation, and vice-versa. This property makes the<br />

Fourier transform able to solve differential equations.<br />

THM. 6.F (Transformation of derivatives into multiplications and viceversa)<br />

Let f 2 C n (R) and let f (r) (t) be absolutely integrable on R for<br />

each r · n. Then<br />

F [f (n) ] = (iω) n ˆ f(ω).<br />

Similarly, assuming that t n f(t) is absolutely integrable,<br />

F [(¡it) n f(t)] = dnf(ω) ˆ<br />

.<br />

dωn Proof. The proof is based on integration by parts and the fact that<br />

(32)<br />

lim<br />

t→±∞ f (r) (t) = 0, r = 0, 1, ..., n ¡ 1<br />

since f (r) (t), r = 0, 1, . . . , n is absolutely integrable on R. We can<br />

realize this fact by writing<br />

f (r) (t) = f (r) t<br />

(0) + f (r+1) (τ)dτ.<br />

Since f (r+1) (τ) is absolutely integrable, then the following limit exists<br />

and is finite:<br />

L = lim<br />

t→+∞ f (r) (t) = f (r) +∞<br />

(0) + f (r+1) (τ)dτ.<br />

Now, either L = 0 or L 6= 0 : we prove that the second case is impossible.<br />

Ab absurdo, let L 6= 0. Then there is an interval [T, +1) in which<br />

0<br />

0


jf (r) (t)j > L/2, so that f (r) (t) cannot be absolutely integrable. Thus<br />

(32) is proved. Consider<br />

F (f (n) +∞<br />

)(ω) = f (n) (t)e −iωt dt<br />

−∞<br />

= f (n−1) (t)e −iωt+∞<br />

+ iω<br />

−∞<br />

+∞<br />

f (n−1) (t)e −iωt dt<br />

−∞<br />

= iω F (f (n−1) )(ω)<br />

since limt→±∞ f (n−1) (t) = 0. By iterating such relation the statement<br />

follows. Finally, by direct inspection the trasform of (¡it)f(t) is equal<br />

to d ˆ f(ω)/dω, and the iteration on n concludes the proof. 4<br />

7. CONVOLUTION PRODUCT<br />

That is one of the reasons why I believe Christianity: it is a religion that no<br />

one could have thought up.<br />

(C.S.Lewis, 1898-1963)<br />

DEFINITION 7.A -(Convolution product) The convolution product of two<br />

functions f1(t), f2(t) defined for t 2 R, is the generalized integral (provided<br />

it exists)<br />

+∞<br />

(33)<br />

(f1 ¤ f2)(t) = f1(x)f2(t ¡ x)dx.<br />

−∞<br />

REMARK 7.B - For the convolution product the usual associative and distributive<br />

(with respect to sum) properties hold. Moreover, the commutative<br />

property can be easily proved: by a change of variable t ¡ x = y,<br />

+∞<br />

+∞<br />

f1 ¤ f2(t) = f1(x)f2(t ¡ x)dx = f1(t ¡ y)f2(y)dy = f2 ¤ f1(t).<br />

−∞<br />

THM. 7.C - (Convolution theorem in time domain, convolution theorem in<br />

frequency domain)<br />

If the functions f1, f2 are sufficiently regular 1 then<br />

(34)<br />

(35)<br />

Scheme of proof.<br />

−∞<br />

F (f1 ¤ f2)(ω) = ˆ f1(ω) ¢ ˆ f2(ω)<br />

F −1 ( ˆ f1 ¤ ˆ f2)(t) = (f1 ¤ f2)(t).<br />

1 A generic assumption of regularity so as to ensure the existence of the Fourier<br />

transforms, the change in the order of integration, etc.<br />

23


24<br />

Let us consider the fourier transfor of a convolution product:<br />

+∞<br />

F (f1 ¤ f2)(ω) = e −iωt +∞<br />

dt f1(x)f2(t ¡ x)dx<br />

−∞<br />

−∞<br />

+∞ +∞<br />

=<br />

e −iωt f1(x)f2(t ¡ x)dt dx<br />

−∞<br />

−∞<br />

By the change of variables (t, x) ! (y, x), where x = x and y = t ¡ x, we<br />

obtain<br />

+∞ +∞<br />

F (f1 ¤ f2)(ω) =<br />

e −iω(x+y) f1(x)f2(y)dy dx<br />

−∞<br />

and vice-versa<br />

−∞<br />

−∞<br />

+∞<br />

= e −iωx +∞<br />

f1(x) dx e −iωy f2(y) dy = ˆ f1(ω) ¢ ˆ f2(ω)<br />

F −1 ˆ f1 ¢ ˆ f2<br />

−∞<br />

<br />

(t) = (f1 ¤ f2) (t) 4<br />

EXAMPLE 7.D - (The diffusion equation on a line) Consider the heat<br />

equation with initial condidion f(x) for 2 R:<br />

⎧<br />

⎨ ∂u<br />

∂t = k<br />

⎩<br />

∂2u ∂x2 ⎤<br />

u(x, 0) = f(x), ¡1 < x < +1 ⎦<br />

ju(x, t)j < M, ¡1 < x < +1, t > 0<br />

In this problem it is useful to consider the Fourier transform of both sides<br />

with respect to x, so that the second order derivative (with respect to x)<br />

is transformed into a multiplication by the square of the conjugate variable<br />

(Thm. 6.5). Denoting p the Fourier variable conjugate with x, we obtain:<br />

dF [u]<br />

dt = ¡kp2 ¢ F [u]<br />

where the Fourier transform of u is denoted by F [u]. Here we deal with a<br />

first order ordinary differential equation, since the unknown function F [u]<br />

no more depends on x. The general solution of such ordinary equation is:<br />

(36)<br />

Setting t = 0 in (36) we see that<br />

F [u(x, t)] = Ce −kp2 t<br />

F [u(x, 0)] = F [f(x)] = C<br />

so that<br />

F [u] = ˆ f(p) ¢ e −kp2t .<br />

Now we can apply the convolution theorem (Thm. 7.1):<br />

<br />

u(x, t) = f(x) ¤ F −1 [e −kp2 <br />

t<br />

]<br />

From (??) we know that<br />

F [e −αx2<br />

] =<br />

With α = 1/4kt, it implies<br />

π<br />

α e−p2 /4α , or F −1 [e −p 2 /(4α) ] =<br />

F −1 [e −kp2 t ] =<br />

1<br />

4kπt e−x2 /4kt .<br />

α<br />

π e−αx2


u(x,t)<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

u(x,t) per t=0.01,t=1,t=10<br />

t=0.01<br />

t=1<br />

t=10<br />

−10 −5 0 5 10 15 20<br />

Figure 9. Graph of the solution of the heat equation, starting<br />

from an initial function u(x, 0) with compact support,<br />

the indicator of (0, 10). Its diffusion is represented at times<br />

t = 0.01, t = 1, t = 10.<br />

Then the solution is the convolution product<br />

<br />

1<br />

u(x, t) = f(x) ¤ p4πkt e −x2 +∞<br />

/4kt<br />

1<br />

= f(z) p e<br />

4πkt −(x−z)2 /4kt<br />

dz.<br />

Notice that in this formula the well known ”heat kernel” appears:<br />

G(z; t) =<br />

x<br />

−∞<br />

1<br />

p 4πkt e −z2 /4kt<br />

which coincides with a Gaussian probability density with mean zero and<br />

variance 2kt. So we have three facts: a) the diffusion equation on the line<br />

with initial condition u(x, 0) = f(x) is solvable with a known explicit kernel:<br />

+∞<br />

u(x, t) = f(z)G(x ¡ z; t)dz.<br />

−∞<br />

b) The physical interpretation of u(x, t) as the temperature at time t suggests<br />

that a choice of a compact support initial temperature f(x) gives rise to a<br />

temperature which is at once everywhere nonzero (!): u(x, t) 6= 0 8x 2 R,<br />

8t > 0. An instantaneous diffusion to all points of space takes place. Setting<br />

σ2 = 2kt, the convolution of any function f(x) with the probability density<br />

is convergent to f as σ ! 0.<br />

1<br />

σ √ 2π e−z2 /2σ2 25


26<br />

u(x,t)<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

u(x,t) per t=0.01,t=1,t=10<br />

t=0.01<br />

t=1<br />

t=10<br />

−10 −5 0 5 10 15 20<br />

Figure 10. Graph of the solution of the heat equation,<br />

starting from an initial function u(x, 0) with compact support,<br />

the indicator of (0, 10). Evidence of its ”diffusion” is<br />

given at times t = 0.01, t = 1, t = 10.<br />

8. THE SPACE OF DISTRIBUTIONS AND DIRAC’S DELTA<br />

We read that man cannot exist ”alone” (Gen 2:18); he can exist only as<br />

”unity of the two”, and therefore in relation to anther human person. It is a<br />

question of mutual relationship: man to woman and woman to man. Being<br />

a person in the image and likeness of God thus also involves existing in a<br />

relationship, in relation to the other ”I”. This is a prelude to the definitive<br />

self-revelation of the Triune God: a living unity in the communion of the<br />

Father, Son and Holy Spirit.<br />

(John Paul II, 1920-2005)<br />

Let φ be a function of one real variable. We recall that the support of a<br />

function is the closure of the set where the function is nonzero: supp(ϕ) =<br />

fx 2 R : ϕ(x) 6= 0g. Also, we recall that a compact subset of R is a closed<br />

and bounded set.<br />

Let Ω be any open subset of R; then the set of infinitely differentiable<br />

functions with compact support in Ω is denoted C∞ ◦ (Ω). A compact support<br />

is bounded and has a positive distance from the boundary of Ω.<br />

DEFINITION 8.A - (The space of test functions) Let Ω be any open subset<br />

of R. In the set C∞ ◦ (Ω), of infinitely differentiable functions with compact<br />

support, a sequence fϕng is said to converge to ϕ 2 C∞ ◦ if<br />

(i) there is a compact set K such that supp(ϕn) ½ K, 8n 2 N<br />

(ii) the sequence fϕ (k)<br />

n gn∈N of the k¡th order derivatives is uniformly convergent<br />

to ϕ (k) as n ! 1, for k = 0, 1, 2... The vector space C∞ ◦ (Ω), when<br />

x


endowed with such notion of convergence, is said ”the space of test functions”’<br />

and is denoted by D(Ω).<br />

Notice that the space D(Ω) is not normalizable, i.e. it is impossible to<br />

introduce a norm compatible with the above type of convergence.<br />

DEF. 8.B - Such functions do exist: an example is<br />

<br />

exp(¡1/(1 ¡ jxj2 ), jxj < 1<br />

ϕ(x) =<br />

0, jxj ¸ 1.<br />

It is easily seen that ϕ 2 D(R) and that supp(ϕ) = fx 2 Rn : jxj · 1g =<br />

B1(0), i.e. the support is the unitary ball centered at the origin. Setting<br />

ρ(x) = ϕ(x)/ <br />

R ϕ(x)dx, again ρ 2 D(R) and supp(ρ) = B1(0). Now for any<br />

ǫ > 0 and x◦ 2 R we define<br />

ρǫ,x◦(x) = ǫ −n x ¡ x◦<br />

ρ( ).<br />

ǫ<br />

Then<br />

(i) ρǫ,x◦ 2 D(R)<br />

(ii) supp(ρǫ,x◦) = fx 2 R : jx ¡ x◦j · ǫg = Bǫ(x◦)<br />

(iii) ρǫ,x◦(x) ¸ 0, 8x 2 R,<br />

<br />

and ρǫ,x◦(x) > 0 () jx ¡ x◦j < ǫ<br />

(iv) ρǫ,x◦(x)dx = 1.<br />

R<br />

Therefore for any open set Ω ½ R, there always exist functions of the type<br />

ρǫ,x◦(x) with the property ρǫ,x◦ 2 D(Ω): indeed it is enough to choose<br />

x◦ 2 Ω and ǫ sufficiently small.<br />

DEF. 8.C - (The space of distributions) Let Ω be an open set and let D(Ω) be<br />

the space of test functions on Ω. T : D(Ω) ! R is said a linear functional<br />

if<br />

< T, (λ1ϕ1 + λ2ϕ2) > = λ1 < T, ϕ1 > +λ2 < T, ϕ2 > .<br />

T is continuous if<br />

ϕn ! ϕ in D(Ω), =) < T, ϕn >!< T, ϕ > in R.<br />

The space D ∗ (Ω) of linear and continuous functionals on D(Ω) is said the<br />

”space of distributions” and each element T 2 D ∗ (Ω) is said a ”distribution”.<br />

D ∗ (Ω) is endowed with a topology by means of the notion of weak convergence:<br />

DEF. 8.D - (The convergence of distributions)<br />

A sequence of distributions fTngn∈N ½ D ∗ (Ω) is convergent to a distribution<br />

T 2 D ∗ (Ω) if and only if<br />

lim<br />

n→∞ < Tn, ϕ > = < T, ϕ >, 8ϕ 2 D(Ω).<br />

27


28<br />

EXAMPLE 8.E - Let Ω be an open set of R and let f be a ”locally integrable”’<br />

real function on Ω:<br />

f 2 L 1 <br />

loc (Ω), i.e. 9 jf(x)j dx < +1, 8 compact K ½ Ω.<br />

The map Tf : D(Ω) ! R is defined by<br />

<br />

< Tf, ϕ >= f ¢ ϕ dx, ϕ 2 D(Ω).<br />

K<br />

Ω<br />

Such a map is well defined since ϕ has compact support in Ω. Moreover Tf<br />

is linear and continuous: indeed, if ϕn ! ϕ in D(Ω, there exists a compact<br />

K ½ Ω such that all ϕn vanish out of K. Since the convergence is uniform<br />

in K,<br />

<br />

<br />

j < Tf , ϕn > ¡ < Tf, ϕ > j = j (fϕn ¡ fϕ)dxj · ǫ jfjdx<br />

Ω<br />

K<br />

if only n is larger than some nǫ. Therefore < Tf, ϕn >!< Tf, ϕ >, i.e. Tf<br />

is continuous. Therefore Tf is a distribution, Tf 2 D∗ (Ω). Distributions of<br />

this kind are said ”regular”. All the remaining distributions, which do not<br />

admit a representation by means of L1 loc functions, are said ”singular”.<br />

DEFINITION 8.F - Dirac’s delta distribution -<br />

Let Ω be an open set of R and a 2 Ω. For any ϕ 2 D(Ω), Dirac’s delta in<br />

a acts the following way:<br />

δa(ϕ) = ϕ(a).<br />

Such a linear functional on D(Ω) is a distribution in the sense of Def. 8.C.<br />

Indeed it is continuous:<br />

δa(ϕn) = ϕn(a)toϕ(a) = δa(ϕ) as n ! 1.<br />

In case a = 0, it is simply said ”Dirac’s delta”: < δ, ϕ >= ϕ(0), 8ϕ 2 D(Ω).<br />

It is a singular distribution and, more precisely, it belongs to the class of<br />

distributions called ”measures”. Dirac’s delta is usually written<br />

∞<br />

−∞<br />

δ(x)ϕ(x)dx = ϕ(0)<br />

just as if a ”function” would exist such that<br />

Similarly,<br />

δ(x) = 0, 8x 6= 0, δ(0) = +1,<br />

∞<br />

−∞<br />

<br />

δ(x ¡ a)ϕ(x)dx = ϕ(a).<br />

δ(x)dx = 1.<br />

This notation can be understood in terms of sequences of ”approximants”:<br />

THM. 8.G - Sequence of approximants of Dirac’s delta -<br />

Let fn be a sequence of probability densities, i.e. fn 2 L1 loc , with fn ¸ 0,<br />

and <br />

R fn(x)dx = 1, 8n 2 N. Let, moreover, fn(x) ! 0 8x 6= 0. Then<br />

<br />

lim<br />

n→+∞<br />

R<br />

fn(x)ϕ(x)dx = ϕ(0), as n ! 1.


The same is true for other sequences, such as:<br />

.<br />

Proof.<br />

fn = i e−inx<br />

πx<br />

For the proof, we restrict to a simple example of approximants:<br />

fn(x) = n<br />

2 I <br />

n2 , x 2 [¡<br />

[−1/n,1/n](x) =<br />

1 n , 1 n ]<br />

<br />

0, elsewhere.<br />

Indeed<br />

min<br />

[− 1<br />

n<br />

,+ 1<br />

n ]<br />

where, by continuity in 0,<br />

ϕ(x) = mn · n<br />

1<br />

+ n<br />

2 − 1<br />

n<br />

whence the formula. 4<br />

ϕ(x)dx · Mn = max<br />

lim<br />

n→∞ mn = ϕ(0), lim<br />

n→+∞ Mn = ϕ(0),<br />

[− 1 1<br />

,+ n n ]<br />

ϕ(x)<br />

In order to understand how to generate a derivative of a distribution, let us<br />

consider the simple case of a differentiable function f, at least with continuous<br />

derivative in an open connected set Ω. For any ϕ 2 D(Ω) we consider<br />

the distribution Tf ′ associated with f ′ (x). Integrating by parts we obtain:<br />

< Tf ′, ϕ >=<br />

<br />

Ω<br />

f ′ (x)ϕ(x) dx = [f(x)ϕ(x)] b a ¡<br />

b<br />

f(x)ϕ ′ (x) dx<br />

Taking into account that ϕ ′ 2 D(Ω), it follows that<br />

< Tf ′, ϕ > = ¡ < f, ϕ′ ><br />

is the distribution which is determined by f ′ .<br />

DEFINITION 8.H- The derivative of a distribution. -<br />

Let Ω be an open set of R and let r 2 N. The linear map from D ∗ (Ω) to<br />

D ∗ (Ω) defined by<br />

T ! D r T, with < D r T, ϕ, >= (¡1) r < T, D r ϕ ><br />

is said derivative, and D r T is said the r-th order derivative of the distribution<br />

T.<br />

EXAMPLE 8.I- The derivatives of the Heaviside function, of δ and δ ′ . -<br />

Let us consider the Heaviside function<br />

<br />

1, x > 0<br />

H(x) =<br />

0, elsewhere.<br />

The corrspoinding distribution<br />

has the derivative:<br />

< Tf, ϕ >=<br />

< DTf, ϕ >= ¡ < Tf, ϕ ′ >= ¡<br />

∞<br />

0<br />

∞<br />

0<br />

ϕ(x) dx<br />

a<br />

ϕ ′ (x) dx = ϕ(0) =< δ, ϕ ><br />

29


30<br />

Therefore the derivative of the Heaviside function is the Dirac’s delta distribution.<br />

The derivative of δ is given by:<br />

The second derivative of delta is:<br />

< Dδ, ϕ >= ¡ < δ, ϕ ′ >= ¡ϕ ′ (0).<br />

< D 2 δ, ϕ >= (¡1) 2 < δ, ϕ ′′ >= ϕ ′′ (0).<br />

and so on.<br />

Notice that Dirac’s δ is still a ”measure”, and in particular, a discrete measure<br />

2 All derivatives of δ are more singular distributions.<br />

An example of distributional, non-classical derivative which is still a function:<br />

the absolute value function<br />

<br />

x, x > 0<br />

f(x) = jxj =<br />

¡x, elsewhere.<br />

admits a distributional derivative given by:<br />

<br />

¡1, x < 0<br />

sign(x) =<br />

1 x > 0.<br />

EX. 8.L- Let f : R ! R be continuous except in the points a1 < a2 < ... <<br />

an. For sake of simplicity, fix ideas with n = 2. Let f admit a continuous<br />

derivative in the intervals (¡1, a1), (a1, a2), (a2, +1). Let, moreover,<br />

exists the finite limits f(a − 1 ), f(a+ 1 ), f(a−2 ), f(a+ 2 ). f is locally integrable,<br />

so the distribution Tf makes sense:<br />

<br />

s<br />

ak+1<br />

< Tf, ϕ >= f(x)ϕ(x) dx = f(x)ϕ(x) dx<br />

The derivative is:<br />

< DTf , ϕ >= ¡ < Tf, ϕ ′ a1<br />

>= ¡<br />

= ¡[fϕ] a1<br />

−∞ +<br />

a1<br />

<br />

R−{a1,a2}<br />

−∞<br />

R<br />

−∞<br />

k=0<br />

ak<br />

fϕ ′ a2<br />

¡ fϕ<br />

a1<br />

′ ¡<br />

ϕf ′ ¡ [fϕ] a2<br />

a1 +<br />

a2<br />

ϕf<br />

a1<br />

′ ¡ [fϕ] +∞<br />

a2 +<br />

∞<br />

fϕ ′<br />

a2<br />

+∞<br />

ϕf ′ =<br />

ϕf ′ dx + ϕ(a1)[f(a + 1 ) ¡ f(a− 1 )] + ϕ(a2)[f(a + 2 ) ¡ f(a− 2 )].<br />

2 In the general framework of measure theory and integration theory - which goes<br />

far beyond our aims - discrete measures correspond with discrete distributions (and<br />

variables) of probability theory. Indeed, setting<br />

we have <br />

µ(A) =<br />

<br />

1, if 0 2 A<br />

0, if 0 /2 A.<br />

ϕ(x) µ(dx) = ϕ(0)µ(f0g) = ϕ(0) ´< δ, ϕ ><br />

On the contrary, there is no measure µ such that ϕ dµ = ¡ϕ ′ (0).<br />

a2


Therefore<br />

d<br />

dx Tf = Tf ′ +<br />

n<br />

[f(a + k ) ¡ f(a−)]δak k .<br />

k=1<br />

In probability, for example, this means that the derivative of a discrete<br />

distribution function F (x) is a ”probability density” in the distributional<br />

sense:<br />

d<br />

dx F (x) = F ′ (x) ¢ I R−{a1,...,an} +<br />

n<br />

P (X = ak) ¢ δ(x ¡ ak).<br />

To conclude, let us see some Fourier transforms involving Dirac’s Delta.<br />

k=1<br />

REMARK 8.M - As an immediate cosequence of the definition<br />

<br />

e −iωt δ(t)dt = 1, F −1 (1) = δ.<br />

R<br />

By virtue of the symmetry property<br />

<br />

e −iωt dt ´ F [1] = 2πδ(ω), F −1 [2πδ(ω)] = 1.<br />

R<br />

Similarly,<br />

F [e iω◦t <br />

] =<br />

and<br />

R<br />

e −i(ω−ω◦)t dt ´ F [1](ω ¡ ω◦) = 2πδ(ω ¡ ω◦)<br />

F [δa] = e −iωa and F −1 (e −iωa = δa(t) = δ(t ¡ a).<br />

REMARK 8.N - Let us compute the Fourier transform of the function<br />

cos(ω◦t). From Euler’s formulas, from linearity property, and from the<br />

above transforms we obtain<br />

F [cos(ω◦t)] = 1<br />

2 F (eiω◦t 1<br />

) +<br />

2 F (e−iω◦t )<br />

= π[δ(ω ¡ ω◦) + δ(ω + ω◦)].<br />

REMARK 8.O - Compute the Fourier transform of the Heaviside function<br />

H(t)<br />

<br />

0, t < 0<br />

H(t) = I [0,∞)(t) =<br />

+1 t ¸ 0.<br />

Notice that<br />

+∞<br />

F (H) =<br />

−∞<br />

= lim<br />

R→+∞<br />

H(t)e −iωt +R<br />

dt = lim<br />

R→+∞<br />

R<br />

0<br />

−R<br />

e −iωt ·<br />

e−iωt dt = lim<br />

R→+∞ ¡iω<br />

= 1<br />

+ lim<br />

iω R→+∞ fR(ω)<br />

H(t)e −iωt dt<br />

R<br />

0<br />

31


32<br />

where fR(ω) = i<br />

ω e−iωR . By Thm. 8.G, since i e−int<br />

πt converges to δ(t) as<br />

n ! 1 in the distributional sense,<br />

F (H) = πδ(ω) + 1<br />

iω .<br />

f(t)<br />

ˆ f(ω)<br />

c1f1(t) + c2f2(t) c1 ˆ f1(ω) + c2 ˆ f2(ω)<br />

f ′ (t) iω ˆ f(ω)<br />

f (n) (t) (iω) n ˆ f(ω)<br />

t n f(t) i n d n ˆ f(ω)<br />

dω n<br />

f(t ¡ t◦) e −iωt◦ ˆ f(ω)<br />

e iω◦t f(t)<br />

(f1 ¤ f2)(t)<br />

f1 ¢ f2<br />

ˆ f(ω ¡ ω◦)<br />

ˆ f1(ω) ¢ ˆ f2(ω)<br />

1<br />

2π ( ˆ f1 ¤ ˆ f2)(ω)<br />

pa(t) = I (−a,a)(t) 2 sin(aω)/ω<br />

sin(ω◦t)/t<br />

e −αt2<br />

δ(t) 1<br />

ˆ f(ω) = π ¢ I(−ω◦,ω◦)(ω)<br />

π<br />

α e−ω2 /(4α)<br />

1 2πδ(ω)<br />

δ(t ¡ a) e −iωa<br />

e iω◦t 2πδ(ω ¡ ω◦)<br />

e −iω◦t 2πδ(ω + ω◦)<br />

cos(ω◦t) π[δ(ω ¡ ω◦) + δ(ω + ω◦)]<br />

H(t) = I (0,∞)(t) πδ(t) + 1/(iω)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!