Previous Up Next
Creative Commons License
This work is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License.

5 Fourier Analysis

  All bases are equal, but some are more equal then others.


As we saw already any separable Hilbert space posses an orthonormal basis (infinitely many of them indeed). Are they equally good? This depends from our purposes. For solution of differential equation which arose in mathematical physics (wave, heat, Laplace equations, etc.) there is a proffered choice. The fundamental formula: d/dx eax=aeax reduces the derivative to a multiplication by a. We could benefit from this observation if the orthonormal basis will be constructed out of exponents. This helps to solve differential equations as was demonstrated in Subsection 0.2.

  7.40pm Fourier series: Episode II

 Today’s TV listing


5.1 Fourier series

Now we wish to address questions stated in Remark 9. Let us consider the space L2[−π,π]. As we saw in Example 3 there is an orthonormal sequence en(t)=(2π)−1/2eint in L2[−π,π]. We will show that it is an orthonormal basis, i.e.

  f(t)∈ L2[−π,π]  ⇔   f(t)=
k=−∞
⟨ f,ek  ⟩ek(t),

with convergence in L2 norm. To do this we show that CLin{ek:k∈ℤ}=L2[−π,π].

Let CP[−π,π] denote the continuous functions f on [−π,π] such that f(π)=f(−π). We also define f outside of the interval [−π,π] by periodicity.

Lemma 1 The space CP[−π,π] is dense in L2[−π,π].

Figure 9: A modification of continuous function to periodic

Proof. Let fL2[−π,π]. Given є>0 there exists gC[−π,π] such that ||fg||<є/2. From continuity of g on a compact set follows that there is M such that | g(t) |<M for all t∈[−π,π].

We can now replace g by periodic g′, which coincides with g on [−π,π−δ] for an arbitrary δ>0 and has the same bounds: | g′(t) |<M, see Figure 9. Then

    ⎪⎪
⎪⎪
gg⎪⎪
⎪⎪
22=
π
π−δ

g(t)−g′(t) 
2dt ≤ (2M)2δ.

So if δ<є2/(4M)2 then ||gg′||<є/2 and ||fg′||<є. □

Now if we could show that CLin{ek: k ∈ ℤ} includes CP[−π,π] then it also includes L2[−π,π].

Notation 2 Let fCP[−π,π],write
fn=
n
k=−n
 ⟨ f,ek  ⟩ ek ,   for   n=0,1,2,… (24)
the partial sum of the Fourier series for f.

We want to show that ||ffn||2→ 0. To this end we define nth Fejér sum by the formula

Fn=
f0+f1+⋯+fn
n+1
, (25)

and show that

    ⎪⎪
⎪⎪
Fnf⎪⎪
⎪⎪
 → 0.

Then we conclude

  ⎪⎪
⎪⎪
Fnf⎪⎪
⎪⎪
2=


π
−π

Fn(t)−f
2


1/2



 
 ≤ (2π)1/2 ⎪⎪
⎪⎪
Fnf⎪⎪
⎪⎪
→ 0.

Since FnLin((en)) then fCLin((en)) and hence f=∑−∞f,ekek.

Remark 3 It is not always true that ||fnf||→ 0 even for fCP[−π,π].
Exercise 4 Find an example illustrating the above Remark.

The summation method used in (25) us useful not only in the context of Fourier series but for many other cases as well. In such a wider framework the method is known as .

  It took 19 years of his life to prove this theorem


5.2 Fejér’s theorem

Proposition 5 (Fejér, age 19) Let fCP[−π,π]. Then
     
    Fn(x)=
1
π
−π
f(t) Kn(xt) dt,     where   
(26)
    Kn(t)=
1
n+1
n
k=0
k
m=−k
eimt,  
(27)
is the Fejér kernel.
Proof. From notation (24):
    fk(x)=
k
m=−k
 ⟨ f,em  ⟩ em(x)
 =
k
m=−k
π
−π
f(t)
eimt
dt  
eimx
 =
1
π
−π
f(t)
k
m=−k
eim(xt)dt.
Then from (25):
    Fn(x)=
1
n+1
n
k=0
fk(x)
 =
1
n+1
1
n
k=0
π
−π
f(t)
k
m=−k
eim(xt)dt
 =
1
π
−π
f(t) 
1
n+1
n
k=0
k
m=−k
eim(xt)dt,
which finishes the proof. □
Lemma 6 The Fejér kernel is -periodic, Kn(0)=n+1 and can be expressed as:
Kn(t)=
1
n+1
sin2
(n+1)t
2
sin2
t
2
,    for  t∉2πℤ. (28)

   1   
  z−11z  
 z−2z−11zz2 
Table 1: Counting powers in rows and columns

Proof. Let z=eit, then:
    Kn(t)=
1
n+1
n
k=0
(zk+⋯+1+z+⋯+zk)
 =
1
n+1
n
j=−n
(n+1−
j
)
zj, 
by switch from counting in rows to counting in columns in Table 1. Let w=eit/2, i.e. z=w2, then
     
    Kn(t)=
1
n+1
(w−2n+2w−2n+2+⋯+(n+1)+nw2+⋯+w2n)
 
 =
1
n+1
(wn+wn+2+⋯+wn−2+wn)2 
(29)
 =
1
n+1



wn−1wn+1
w−1w



2



 
   Could you sum a geometric progression?
 
 =
1
n+1






2isin
(n+1)t
2
2isin
t
2






2






 
, 
 
if w≠ ± 1. For the value of Kn(0) we substitute w=1 into (29). □

Figure 10: A family of Fejér kernels with the parameter m running from 0 to 9 is on the left picture. For a comparison unregularised Fourier kernels are on the right picture.

The first eleven Fejér kernels are shown on Figure 10, we could observe that:

Lemma 7 Fejér’s kernel has the following properties:
  1. Kn(t)≥0 for all t∈ ℝ and n∈ℕ.
  2. −ππKn(t) d t=2π.
  3. For any δ∈ (0,π)
          
    −δ
    −π
    +
    π
    δ
    Kn(t) dt → 0    as    n→ ∞.
Proof. The first property immediately follows from the explicit formula (28). In contrast the second property is easier to deduce from expression with double sum (27):
    
π
−π
Kn(t) dt
=
π
−π
1
n+1
n
k=0
k
m=−k
eimtdt
 =
     
1
n+1
n
k=0
k
m=−k
π
−π
eimtdt
 =
     
1
n+1
n
k=0
 2π
 =2π,
since the formula (19).

Finally if | t |>δ then sin2(t/2)≥ sin2(δ/2)>0 by monotonicity of sinus on [0,π/2], so:

    0≤ Kn(t) ≤ 
1
(n+1) sin2(δ/2)

implying:

  0≤ 
 
δ≤
t
≤ π
Kn(t)  dt ≤ 
1(π−δ)
(n+1) sin2(δ/2)
→ 0   as  n→ 0.

Therefore the third property follows from the squeeze rule. □

Theorem 8 (Fejér Theorem) Let fCP[−π,π]. Then its Fejér sums Fn (25) converges in supremum norm to f on [−π,π] and hence in L2 norm as well.
Proof. Idea of the proof: if in the formula (26)
      Fn(x)=
1
π
−π
f(t) Kn(xt) dt, 
t is long way from x, Kn is small (see Lemma 7 and Figure 10), for t near x, Kn is big with total “weight” 2π, so the weighted average of f(t) is near f(x).

Here are details. Using property 2 and periodicity of f and Kn we could express trivially

      f(x)= f(x)
1
x
x−π
Kn(xt)  dt  =
1
x
x−π
f(x) Kn(xt) dt. 

Similarly we rewrite (26) as

      Fn(x)=
1
x
x−π
f(t) Kn(xt) dt, 

then

    
f(x)−Fn(x) 
=
    
1



x
x−π
 (f(x)−f(t)) Kn(xt) dt


 
1
x
x−π

f(x)−f(t) 
Kn(xt) dt.

Given є>0 split into three intervals: I1=[x−π,x−δ], I2=[x−δ,x+δ], I3=[x+δ,x+π], where δ is chosen such that | f(t)−f(x) |<є/2 for tI2, which is possible by continuity of f. So

     
1
 


I2

f(x)−f(t) 
Kn(xt) dt
є
2
1
 


I2
  Kn(xt) dt <
є
2
.

And

     
1
 


I1⋃ I3

f(x)−f(t) 
Kn(xt)  dt
2⎪⎪
⎪⎪
f⎪⎪
⎪⎪
1
 


I1⋃ I3
  Kn(xt)  dt
 =
⎪⎪
⎪⎪
f⎪⎪
⎪⎪
π
 
δ<
u
  Kn(u)  du
 <
є
2
, 

if n is sufficiently large due to property 3 of Kn. Hence | f(x)−Fn(x) |<є for a large n independent of x. □

Remark 9 The above properties 13 and their usage in the last proof can be generalised to the concept of approximation of the identity. See § 15.4 for a further example.

We almost finished the demonstration that en(t)=(2π)−1/2eint is an orthonormal basis of L2[−π,π]:

Corollary 10 (Fourier series) Let fL2[−π,π], with Fourier series
    
n=−∞
⟨ f,en  ⟩en=
n=−∞
cneint   where  cn=
⟨ f,en  ⟩
=
1
π
−π
f(t)eintdt.
Then the series −∞f,enen=∑−∞cneint converges in L2[−π,π] to f, i.e
    
 
lim
k→ ∞
⎪⎪
⎪⎪
⎪⎪
⎪⎪
f
k
n=−k
cneint⎪⎪
⎪⎪
⎪⎪
⎪⎪
2=0.
Proof. This follows from the previous Theorem, Lemma 1 about density of CP in L2, and Theorem 15 on orthonormal basis. □
Remark 11 There is a reason why we had used the Fejér kernel and the Cezàro summation Fn (25) instead of plain partial sums fn (24) of the Fourier series. It can be shown that point-wise convergence fnf does not hold for every continuous function f, cf. Cor. 41.

5.3 Parseval’s formula

The following result first appeared in the framework of L2[−π,π] and only later was understood to be a general property of inner product spaces.

Theorem 12 (Parseval’s formula) If f, gL2[−π,π] have Fourier series f=∑n=−∞cneint and g=∑n=−∞dneint, then
⟨ f,g  ⟩=
π
−π
f(t)
g(t)
dt=2π
−∞
cn
dn
. (30)

More generally if f and g are two vectors of a Hilbert space H with an orthonormal basis (en)−∞ then

    ⟨ f,g  ⟩=
k=−∞
cn
dn
,    where  cn=⟨ f,en  ⟩, dn=⟨ g,en  ⟩,

are the Fourier coefficients of f and g.

Proof. In fact we could just prove the second, more general, statement—the first one is its particular realisation. Let fn=∑k=−nn ckek and gn=∑k=−nn dkek will be partial sums of the corresponding Fourier series. Then from orthonormality of (en) and linearity of the inner product:
    ⟨ fn,gn  ⟩=⟨ 
n
k=−n
ckek,
n
k=−n
dkek  ⟩=
n
k=−n
ck
dk
.
This formula together with the facts that fkf and gkg (following from Corollary 10) and Lemma about continuity of the inner product implies the assertion. □
Corollary 13 A integrable function f belongs to L2[−π,π] if and only if its Fourier series is convergent and then ||f||2=2π∑−∞| ck |2.
Proof. The necessity, i.e. implication fL2 ⇒ ⟨ f,f ⟩=||f||2=2π∑| ck |2 , follows from the previous Theorem. The sufficiency follows by Riesz–Fisher Theorem. □
Remark 14 The actual rôle of the Parseval’s formula is shadowed by the orthonormality and is rarely recognised until we meet the wavelets or coherent states. Indeed the equality (30) should be read as follows:
Theorem 15 (Modified Parseval) The map W: Hl2 given by the formula [Wf](n)=⟨ f,en ⟩ is an isometry for any orthonormal basis (en).
We could find many other systems of vectors (ex), xX (very different from orthonormal bases) such that the map W: HL2(X) given by the simple universal formula
[Wf](x)=⟨ f,ex  ⟩ (31)
will be an isometry of Hilbert spaces. The map (31) is oftenly called wavelet transform and most famous is the Cauchy integral formula in complex analysis. The majority of wavelets transforms are linked with group representations, see our postgraduate course Wavelets in Applied and Pure Maths.

  Heat and noise but not a fire?

 Answer:


5.4 Some Application of Fourier Series

We are going to provide now few examples which demonstrate the importance of the Fourier series in many questions. The first two (Example 16 and Theorem 17) belong to pure mathematics and last two are of more applicable nature.

Example 16 Let f(t)=t on [−π,π]. Then
    ⟨ f,en  ⟩=
π
−π
teintdt=





        (−1)n
2π i
n
,
n≠ 0
        0,n=0
   (check!),
so f(t)∼ ∑−∞(−1)n (i/n) eint. By a direct integration:
    ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
22=
π
−π
t2dt=
3
3
.
On the other hand by the previous Corollary:
  ⎪⎪
⎪⎪
f⎪⎪
⎪⎪
22=2π
 
n≠ 0



(−1)ni
n



2=4π
n=1
1
n2
.
Thus we get a beautiful formula
  
1
1
n2
=
π2
6
.

Here is another important result.

Theorem 17 (Weierstrass Approximation Theorem) For any function fC[a,b] and any є>0 there exists a polynomial p such that ||fp||.
Proof. Change variable: t=2π(xa+b/2)/(ba) this maps x∈[a,b] onto t∈[−π,π]. Let P denote the subspace of polynomials in C[−π,π]. Then eint$P_^$ for any n∈ℤ since Taylor series converges uniformly in [−π,π]. Consequently P contains the closed linear span in (supremum norm) of eint, any n∈ℤ, which is CP[−π,π] by the Fejér theorem. Thus $P_^$CP[−π,π] and we extend that to non-periodic function as follows (why we could not make use of Lemma 1 here, by the way?).

For any fC[−π,π] let λ=(f(π)−f(−π))/(2π) then f1(t)=f(t)−λ tCP[−π,π] and could be approximated by a polynomial p1(t) from the above discussion. Then f(t) is approximated by the polynomial p(t)=p1(t)+λ t. □

It is easy to see, that the rôle of exponents eint in the above prove is rather modest: they can be replaced by any functions which has a Taylor expansion. The real glory of the Fourier analysis is demonstrated in the two following examples.


Figure 11: The dynamics of a heat equation:
x—coordinate on the rod,
t—time,
T—temperature.

Example 18 The modern history of the Fourier analysis starts from the works of Fourier on the heat equation. As was mentioned in the introduction to this part, the exceptional role of Fourier coefficients for differential equations is explained by the simple formula x einx= ineinx. We shortly review a solution of the heat equation to illustrate this.

Let we have a rod of the length . The temperature at its point x∈[−π,π] and a moment t∈[0,∞) is described by a function u(t,x) on [0,∞)×[−π,π]. The mathematical equation describing a dynamics of the temperature distribution is:

∂ u(t,x)
∂ t
=
2 u(t,x)
∂ x2
  or, equivalently,  
t−∂x2
u(t,x)=0.  (32)

For any fixed moment t0 the function u(t0,x) depends only from x∈[−π,π] and according to Corollary 10 could be represented by its Fourier series:

    u(t0,x)=
n=−∞
⟨ u,en  ⟩en=
n=−∞
cn(t0)einx, 

where

    cn(t0)=
⟨ u,en  ⟩
=
1
π
−π
u(t0,x)einxdx, 

with Fourier coefficients cn(t0) depending from t0. We substitute that decomposition into the heat equation (32) to receive:

     
     
t−∂x2
u(t,x)
=
t−∂x2
n=−∞
cn(t)einx
         
 
=    
n=−∞

t−∂x2
cn(t)einx
         
 
=    
n=−∞
(cn(t)+n2cn(t))einx=0 . 
        (33)

Since function einx form a basis the last equation (33) holds if and only if

cn(t)+n2cn(t)=0   for all  n  and  t.  (34)

Equations from the system (34) have general solutions of the form:

cn(t)=cn(0)en2t    for all  t∈[0,∞), (35)

producing a general solution of the heat equation (32) in the form:

u(t,x)=
n=−∞
cn(0)en2teinx =
n=−∞
cn(0)en2t+inx, (36)

where constant cn(0) could be defined from boundary condition. For example, if it is known that the initial distribution of temperature was u(0,x)=g(x) for a function g(x)∈L2[−π,π] then cn(0) is the n-th Fourier coefficient of g(x).

The general solution (36) helps produce both the analytical study of the heat equation (32) and numerical simulation. For example, from (36) obviously follows that

The example of numerical simulation for the initial value problem with g(x)=2cos(2*u) + 1.5sin(u). It is clearly illustrate our above conclusions.


Figure 12: Two oscillation with unharmonious frequencies and the appearing dissonance. Click to listen the blue and green pure harmonics and red dissonance.


  
  
  
Figure 13: Graphics of G5 performed on different musical instruments (click on picture to hear the sound). Samples are taken from Sound Library.


Figure 14: Fourier series for G5 performed on different musical instruments (same order and colour as on the previous Figure)


(a)   (b)
(c)
Figure 15: Limits of the Fourier analysis: different frequencies separated in time

Example 19 Among the oldest periodic functions in human culture are acoustic waves of musical tones. The mathematical theory of musics (including rudiments of the Fourier analysis!) is as old as mathematics itself and was highly respected already in Pythagoras’ school more 2500 years ago.

The earliest observations are that

  1. The musical sounds are made of pure harmonics (see the blue and green graphs on the Figure 12), in our language cos and sin functions form a basis;
  2. Not every two pure harmonics are compatible, to be their frequencies should make a simple ratio. Otherwise the dissonance (red graph on Figure 12) appears.

The musical tone, say G5, performed on different instruments clearly has something in common and different, see Figure 13 for comparisons. The decomposition into the pure harmonics, i.e. finding Fourier coefficient for the signal, could provide the complete characterisation, see Figure 14.

The Fourier analysis tells that:

  1. All sound have the same base (i.e. the lowest) frequencies which corresponds to the G5 tone, i.e. 788 Gz.
  2. The higher frequencies, which are necessarily are multiples of 788 Gz to avoid dissonance, appears with different weights for different instruments.

The Fourier analysis is very useful in the signal processing and is indeed the fundamental tool. However it is not universal and has very serious limitations. Consider the simple case of the signals plotted on the Figure 15(a) and (b). They are both made out of same two pure harmonics:

  1. On the first signal the two harmonics (drawn in blue and green) follow one after another in time on Figure 15(a);
  2. They just blended in equal proportions over the whole interval on Figure 15(b).

This appear to be two very different signals. However the Fourier performed over the whole interval does not seems to be very different, see Figure 15(c). Both transforms (drawn in blue-green and pink) have two major pikes corresponding to the pure frequencies. It is not very easy to extract differences between signals from their Fourier transform (yet this should be possible according to our study).

Even a better picture could be obtained if we use windowed Fourier transform, namely use a sliding “window” of the constant width instead of the entire interval for the Fourier transform. Yet even better analysis could be obtained by means of wavelets already mentioned in Remark 14 in connection with Plancherel’s formula. Roughly, wavelets correspond to a sliding window of a variable size—narrow for high frequencies and wide for low.

site search by freefind advanced

Last modified: November 6, 2024.
Previous Up Next