Fourier Analysis in Hilbert Space: 4.1 Orthonormal Sequences
Fourier Analysis in Hilbert Space: 4.1 Orthonormal Sequences
Fourier Analysis in Hilbert Space: 4.1 Orthonormal Sequences
In the last section of Chapter 3 we introduced the Lebesgue Lp -spaces for general
measures and discussed their most basic properties. The most important Lp -space,
by far, is L2 . Its importance is its role in applications, especially in Fourier analysis.
The material of this chapter lies at the foundation of the branch of mathematics
called harmonic analysis.
In this chapter we will see that L2 is a Hilbert space (we already really have all the
bits of information we need to see this) and that in some sense the L2 -spaces (with
different µ’s) are the only Hilbert spaces. We will come to see how the problem
that Fourier examined, about decomposing functions as infinite sums of other —
somehow more basic — functions, is a problem best phrased and understood in the
language of abstract Hilbert spaces. One of the triumphs of functional analysis is
to take a very concrete problem — in this case Fourier decomposition — view it in
an abstract setting, and use theoretical tools to obtain powerful results that can be
translated back to the concrete setting. Fourier’s work certainly holds an important
spot at the roots of functional analysis, and it motivated much early work in the
development of the field.
Further Hilbert space theory appears in Section 5.4.
a0 ∞
f (x) + ak cos(kx) + bk sin(kx),
2 k1
then it must be the case that the coefficients ak and bk are given by the formulas
1 π
ak f (x) cos(kx)dx, k 0, 1, 2, . . . ,
π −π
and
π
1
bk f (x) sin(kx)dx, k 1, 2, . . . .
π −π
The big question is this: When is this decomposition actually possible? Even if
the integrals involved make sense, does the series converge? If it does converge,
what type of convergence (pointwise, uniform, etc.) do we get? Even if the series
converges in some sense, does it converge to f ?
The immediate goal is to show you how these questions about Fourier series
can be treated in the abstract setting of an inner product space.
Let us now take stock of what we already know by gathering our information
about L2 . First, recall that L2 L2 (µ), for any abstract measure space (X, R, µ),
denotes the collection of all measurable functions f : X → C such that the integral
|f |2 dµ
X
1
Daniel Bernoulli is the nephew of James Bernoulli, who was mentioned at the beginning
of Section 3.1. The Bernoulli family produced several distinguished mathematicians and
physicists; at least twelve members of the family achieved distinction in at least one of these
fields.
76 4. Fourier Analysis in Hilbert Space
is finite. These functions are often called the “square integrable” functions on X.
With norm
f 2 |f |2 dµ,
X
this collection of functions becomes a Banach space. We can define an inner product
on L2 via
f, g f gdµ.
X
It is easily seen that this is an inner product, and that the norm does indeed come
from this inner product. That is,
f 2 f, f |f |2 dµ.
X
The convergence of this infinite sum is in the norm induced by the inner product.
Further, it would be desirable to be able to do this for all f ∈ V . In general, this
cannot be done. Notice that Fourier was asserting that when {fk }∞ k1 is the trigono-
metric system, the coefficients are of form f, fk (an appropriate indexing of the
trigonometric system has not yet been established) whenever his decomposition
works.
Let {fk }∞
k1 be an orthonormal sequence in V . If it is the case that for each f ∈ V
we can find constants ck (depending on f ) such that
∞
f ck fk ,
k1
4.1 Orthonormal Sequences 77
cos (nx)
0.4
0.2
–3 –2 –1 1 2 3
–0.2
–0.4
(a)
sin (mx)
0.4
0.2
–3 –2 –1 1 2 3
–0.2
–0.4
(b)
2
Note that this is a new usage of the word “complete”; we now have at least two ways
we will use this adjective: a complete metric space, a complete orthonormal system.
78 4. Fourier Analysis in Hilbert Space
100
–1 –0.5 0.5 1
–100
–200
H2,1 H2,2
2 2
1 1
–2 –2
H2,3 H2,4
2 2
1 1
–2 –2
Jean Baptiste Joseph Fourier was born March 21, great talent in many areas by the age of
1768, in Auxerre, France (Figure 4.4). His fourteen. He wanted to join the military,
father had been a tailor, but both of his for some reason was rejected, and instead
parents were dead by the time Fourier was entered a Benedictine abbey to train for
ten. There seems to be some disagreement the priesthood. While there, he was able
among authors as to exactly how many to work on mathematics and submitted
siblings Fourier had, but by all accounts he his first paper in 1789. He never took
had many. According to [51], he was the his vows and returned to his school,
nineteenth (and not the last) child in the teaching math, history, philosophy, and
family. rhetoric. This was the time of the French
Fourier was distinguished in two fields: Revolution, and Fourier became quite
mathematics and Egyptology. He began involved in revolutionary politics. In 1794
both careers when he attended a military he was imprisoned and sentenced to be
school run by the Benedictines. He showed guillotined.
80 4. Fourier Analysis in Hilbert Space
What, then, are the ck ’s to be? We turn to the very simple case of R3 , with its
usual inner product, for inspiration. We take as our orthonormal family the three
Euclidean basis vectors e1 (1, 0, 0), e2 (0, 1, 0), and e3 (0, 0, 1). Then
every vector in R3 can be written in form
3
ck ek .
k1
In this case we know that c1 v, e1 , c2 v, e2 , and c3 v, e3 . This example
illustrates the next theorem.
This is because
sn , fm − f, fm sn − f, fm ,
and
|sn − f, fm | ≤ sn − f · fm sn − f .
Therefore,
∞
∞
∞
sn , fm ck fk , fm ck fk , fm ck δkm cm . (4.2)
k1 k1 k1
Let {fk }∞
be an orthonormal sequence in V , and let f ∈ V . We call
k1
∞ ∞
k1 f, fk k the Fourier series of f with respect to {fk }k1 , and f, fk the
f
∞
Fourier coefficients of f with respect to {fk }k1 . These objects are defined without
any assumptions or knowledge about convergence of the series.
The next theorem tells us something about the size of these coefficients.
Theorem 4.2 (Bessel’s Inequality3 ). Suppose that {fk }∞ k1 is an orthonormal
inner product space V . For every f ∈ V , the series (of nonnegative
sequence in an
∞
|f, fk |2 ≤ f 2 .
k1
Proof. Consider the partial sum sn of the Fourier series for f . Then
f − sn , fk f, fk − sn , fk
n
f, fk − f, fj fj , fk
j 1
n
f, fk − f, fj fj , fk
j 1
n
f, fk − f, fj (fj , fk
j 1
n
f, fk − f, fj δj k
j 1
f, fk − f, fk 0.
This shows that f − sn is orthogonal to each fk . Further,
n
f − sn , sn f − sn , f, fk fk
k1
3
Due to Friedrich Bessel (1784–1846; Westphalia, now Germany).
4.2 Bessel’s Inequality, Parseval’s Theorem, and the Riesz–Fischer Theorem 83
n
f − sn , f, fk fk
k1
n
f, fk f − sn , fk ,
k1
which equals zero by the previous argument. This shows that f − sn is orthogonal
to sn . Then, by Exercise 4.2.1,
f − sn 2 + sn 2 f 2 .
This shows that
sn 2 ≤ f 2 .
Since
n
2
sn 2
f, fk fk
,
k1
we have that
n
n
sn 2 |f, fk |2 · fk 2 |f, fk |2 .
k0 k1
as desired. 2
It is natural to want to determine conditions on {fk }∞
k1 under which equality in
Bessel’s inequality holds.
Theorem 4.3 (Parseval’s Theorem4 ). As in the preceding theorem, suppose that
{fk }∞ ∞
k1 is an orthonormal sequence in an inner product space V . Then {fk }k1 is
a complete orthonormal sequence if and only if for every f ∈ V ,
∞
|f, fk |2 f 2 .
k1
4
Due to Marc-Antoine Parseval des Chênes (1755–1836; France).
84 4. Fourier Analysis in Hilbert Space
(Theorem 4.2), T is linear, one-to-one (Theorem 4.6, together with Theorem 4.5
(c)), onto (Theorems 4.4 and 4.6), and f L2 Tf 2 (Theorem 4.3), for all
f ∈ L2 . This result, that L2 and 2 are isometrically isomorphic, is referred to as
the Riesz–Fischer Theorem (Theorem 4.6 sometimes goes by the same name).
Theorem 4.4. Assume that
∞ 2
(a) {dk }∞
k1 is a sequence of real numbers such that k1 dk converges, and
(b) V is a Hilbert space with complete orthonormal sequence {fk }∞k1 .
Proof. Define
n
sn d k fk .
k1
For m > n,
m
m
m
sn − sm 2 dj dk fj , fk dk2 .
j n+1 kn+1 kn+1
Therefore, {sn }∞
n1 is Cauchy. Because V is a Hilbert space, there is an f ∈ V such
that
lim sn − f 0.
n→∞
That is,
∞
g f, fk fk .
k1
86 4. Fourier Analysis in Hilbert Space
Theorem 4.1 then tells us that the Fourier coefficients of g are the same as the
Fourier coefficients of f with respect to {fk }∞
k1 , i.e., g, fk f, fk . By (b),
f − g must equal 0 almost everywhere. In other words,
∞
f f, fk fk .
k1
By (ii) above, the first and third integrals are bounded in absolute value for each n by
b
2πf (x0 ). The middle integral, however, is greater than or equal to a f (x)t n (x)dx,
4.3 A Return to Classical Fourier Analysis 87
If we add and subtract these two equations, we see that the real and imaginary
parts of f are orthogonal to each of the members of the trigonometric system. By
the first part of the proof, the real and imaginary parts of f are identically 0; hence
f is identically 0.
Finally, we no longer assume that f is continuous. Define the continuous
function
x
F (x) f (t)dt.
−π
We now have shown that F , and hence F − C for every constant C, is orthogonal
to each of the nonconstant members of the trigonometric system. We now take
care of the member √12π . Let
π
1
C0 F (x)dx.
2π −π
Then F − C0 is easily seen to be orthogonal to every member of the trigonometric
system. Since F is continuous, F − C0 is also continuous, and the first part of the
88 4. Fourier Analysis in Hilbert Space
for some sequence {fn }∞n1 ∈ C([a, b]). Actually, L must be considered to be
2
the equivalence classes of such functions, where two functions are equivalent if
and only if they are equal almost everywhere (see the discussion preceding The-
orem 3.17). Therefore, it is not entirely true that this definition avoids discussing
measure. However, we can give this definition with only an understanding of
“measure zero,” and not general measure. (And measure zero can be defined in a
straightforward manner and is much simpler to understand than general measure.)
4.1.2 In this exercise you will actually compute a classical (i.e., with respect
to the orthonormal sequence of Exercise 1) Fourier series, and investigate
its convergence properties. The function given is a basic one; in the next
exercise you are asked to do the same procedure with another very basic
function. You are being asked to do these by hand, and you can no doubt
appreciate that the computations get quite laborious once we depart from
even the most basic functions. There are tricks for doing these computations;
the interested reader can learn more about such techniques in a text devoted
to classical Fourier series.
(a) Let
1 if −π ≤ x < 0,
f (x)
0 if 0 ≤ x < π .
Show that its Fourier series is
1 1 ∞
((−1)n − 1)
+ sin(nx).
2 π n1 n
(b) Explain why this series converges in mean to f .
(c) What can you say about the pointwise and uniform convergence of this
series?
(d) Why are the coefficients of the cosine terms all zero?
4.1.3 In this exercise you will compute another Fourier series and investigate its
convergence properties.
(a) Let f (x) x 2 . Show that its classical Fourier series is
π2 ∞
(−1)n
+4 cos(nx).
3 n1
n2
(b) Explain why this series converges in mean to f .
(c) What can you say about the pointwise and uniform convergence of this
series?
(d) Why are the coefficients of the sine terms all zero?
4.1.4 For a sequence {fn }∞ 2
n1 in L ([−π, π], m), we have seen three ways for
∞
{fn }n1 to converge:
(i) “pointwise,”
(ii) “uniformly,”
(iii) “in mean.”
The point of this exercise is to understand the relation between these three
types of convergence. For the counterexamples asked for below, use what-
ever finite interval [a, b] you find convenient. Please make an effort to
supply “easy” examples.
(a) Prove that uniform convergence implies pointwise convergence.
90 4. Fourier Analysis in Hilbert Space
(b) Give an example to show that pointwise convergence does not imply
uniform convergence.
(c) Prove that uniform convergence implies convergence in mean.
(d) Give an example to show that pointwise convergence does not imply
convergence in mean.
(e) Give an example to show that convergence in mean does not im-
ply pointwise convergence. (Note that the same example shows that
convergence in mean does not imply uniform convergence.)
4.1.5 Apply the Gram–Schmidt process to the functions 1, x, x 2 , x 3 , . . . to obtain
formulas for the first three Legendre polynomials. Then verify that they are
indeed given by the formula
2n + 1 1 d n 2
(x − 1)n , n 1, 2, 3.
2 2n n! dx n
4.1.6 Prove that the Haar family is an orthonormal family in the Hilbert space
L2 ([0, 1]).
4.1.7 (a) Show that the sequence
einx
√ , n 0, ±1 ± 2, . . . ,
2π
is a complete orthonormal sequence in L2 ([−π, π]).
(b) Show that the sequence
2
cos(nx), n 1, 2, 3, . . . ,
π
is a complete orthonormal sequence
in L2 ([0, π ]). (Observe that
2
π
cos(nx) can be replaced by π2 sin(nx).)
Section 4.2
4.2.1 (a) Prove that in any inner product space (V , ·, ·), f and g orthogonal
implies
f 2 + g2 f + g2 .
√
Here, as usual, · ·, ·.
(b) Prove Parseval’s theorem.
4.2.2 Assume that {fn }∞ n1 is a sequence in L and that fn → f in mean. Prove
2
∞
that {fn 2 }n1 is a bounded sequence of real numbers.
4.2.3 Assume that f1 , f2 , . . . , fn is an orthonormal family in an inner product
space. Prove that f1 , f2 , . . . , fn are linearly independent.
4.2.4 For f and g in an inner product space, g 0, the projection of f on g is
the vector
f, g
g.
g2
4.3 A Return to Classical Fourier Analysis 91