0% found this document useful (0 votes)
9 views28 pages

FS Notes 2

The document is a set of notes on Fourier Analysis prepared for an elective course at IIT Madras, covering topics such as trigonometric series, Fourier series, the Riemann Lebesgue lemma, and the Dirichlet kernel. It includes definitions, theorems, and proofs relevant to the convergence and properties of Fourier series. The content is structured into sections with references for further reading.

Uploaded by

Cheng Henry
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
9 views28 pages

FS Notes 2

The document is a set of notes on Fourier Analysis prepared for an elective course at IIT Madras, covering topics such as trigonometric series, Fourier series, the Riemann Lebesgue lemma, and the Dirichlet kernel. It includes definitions, theorems, and proofs relevant to the convergence and properties of Fourier series. The content is structured into sections with references for further reading.

Uploaded by

Cheng Henry
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 28

TOPICS IN FOURIER ANALYSIS-II

M.T. NAIR

Contents

1. Trigonometric series and Fourier series 2


2. Riemann Lebesgue Lemma 4
3. Dirichlet kernel 6
4. Dirichlet-Dini criterion for convergence 8
5. Ces̀aro summablity of Fourier series 12
6. Divergence of Fourier series 18
7. Uniqueness 20
8. Convolution 23
9. L2 -Theory 26
References 28

NOTES PREPARED FOR PART OF AN ELECTIVE COURSE FOURIER ANALYSIS


FOR M.SC. STUDENTS OF IIT MADRAS, JULY-NOVEMBER, 2014.
1
2 M.T. NAIR

1. Trigonometric series and Fourier series

Definition 1.1. A series of the form



X
(1.1) c0 + (an cos nx + bn sin nx)
n=1

is called a trigonometric series, where c0 , an , bn are real numbers.

• If (1.1) converges on [−π, π] to a an integrable function f and if it can be


integrated term by term, then
f (−π) = f (π),
and
Z π Z π Z π
1 1 1
c0 = f (x) cos nxdx, an = f (x) cos nxdx, bn = f (x) sin nxdx.
2π −π π −π π −π

• If the (1.1) converges (pointwise) on [−π, π] to a function f , then f can be


extended as a 2π-periodic function by defining
f (x + 2nπ) = f (x), n ∈ Z.

X
• If the series (|an | + |bn |) converges, then (1.1) converges uniformly on [−π, π]
n=1
and it can be integrated term by term. We know that if f ∈ L1 [−π, π], then
the function f˜ : [−π, π] → C defined by

˜ f (x), x ∈ [−π, π),
f (x) =
f (−π), x = π
satisfies
f˜(−π) = f˜(π) and f˜ = f a.e.
• The series (1.1) can be written as

X
cn einx .
n=−∞

Definition 1.2. Let f ∈ L1 [−π, π]. The Fourier series of f is the series

a0 X
(1.2) + (an cos nx + bn sin nx),
2 n=1

where
Z π Z π
1 1
(1.3) an = f (x) cos nxdx, bn = f (x) sin nxdx.
π −π π −π
TOPICS IN FOURIER ANALYSIS 3

The series
∞ Z π
X 1
(1.4) cn e inx
with cn = f (x)e−inx dx
n=−∞
2π −π

is also called the Fourier series of f . The coefficients cn are called the Fourier
coefficient and are usually denoted by fˆ(n), i.e.,
Z π
ˆ 1
f (n) = f (x)e−inx dx, n ∈ Z.
2π −π
The sum
N
X
SN (f, x) := fˆ(n)einx
n=−N

is called the N -th partial sum of the Fourier series (1.4)).

Notation: In the above and in the following, the integral are w.r.t. the Lebesgue
measure.

The fact that (1.2) is the Fourier series of f is usually written as



a0 X
f (x) ≈ + (an cos nx + bn sin nx).
2 n=1

Equivalently,

X
f (x) ≈ fˆ(n)einx .
n=−∞
inx
Since cos nx, sin x, e are 2π-periodic functions, we can talk about Fourier series of
2π-periodic functions. If (1.2) (resp. (1.4)) converges at a point x ∈ [−π, π], then it
converges at x + 2kπ for every k ∈ Z.

• The Fourier series (1.4)) converges at x ∈ [−π, π] if and only if SN (f, x) → f (x)
as N → ∞.
• If f ∈ L1 [−π, π], then fˆ(n) → 0 as |n| → ∞.
X∞ X∞
• If ˆ
|f (n)| converges, then fˆ(n)einx converges uniformly.
n=−∞ n=−∞

Suppose Fourier series of f ∈ L1 [−π, π] converges uniformly, say to g. Then g is


continuous and
Z π ∞ Z π
1 X
ˆ 1
g(x)e−imx
= f (n) ei(n−m)x dx = fˆ(m),
2π −π n=−∞
2π −π
4 M.T. NAIR

i.e., ĝ(m) = fˆ(m) for all m ∈ Z. A natural question would be whether f = g a.e. We
shall answer this affirmatively.
We know that if the Fourier series of f ∈ L1 [−π, π] converges, then
fˆ(n) → 0 as n → ∞.
Can we assert this for every f ∈ L1 [−π, π]? The answer is in the affirmative as proved
in the next section.

2. Riemann Lebesgue Lemma

Theorem 2.1. (Riemann Lebesgue lemma) Let f ∈ L1 [a, b]. Then


Z b Z b
f (t) cos(λt)dt → 0 and f (t) sin(λt)dt → 0 as λ → ∞.
a a

Corollary 2.2. (Riemann Lebesgue lemma) Let f ∈ L1 [a, b]. Then


Z b Z b
f (t) cos(nt)dt → 0 and f (t) sin(nt)dt → 0 as n → ∞.
a a

For the proof of the Theorem 2.1, we shall make use of

LEMMA 2.3. The span of all step functions1 on [a, b] is dense in L1 [a, b].

Proof of Theorem 2.1. First we observe that if for every ε > 0, there exists a func-
tion g ∈ L1 [a, b] such that kf − gk1 < ε and the the result is true for g, then the result
is true for f also.

Indeed,
Z b Z b Z b
f (t) cos(λt)dt ≤ [f (t) − g(t)] cos(λt)dt + g(t) cos(λt)dt
a a a
Z b
≤ ε+ g(t) cos(λt)dt .
a
Rb
Let λ0 > 0 be such that a g(t) cos(λt)dt < ε for all λ ≥ λ0 . Then we
have Z b
f (t) cos(λt)dt < 2ε ∀ λ ≥ λ0
a

1Step
functions are finite linear combinations of characteristic functions. Also, recall that L1 [a, b]
is the vector space of all Lebesgue measurable complex valued functions f such that kf k1 :=
Z b
|f (x)|dx < ∞. Here, dx stands for the Lebesgue measure.
a
TOPICS IN FOURIER ANALYSIS 5
Rb Rb
so that a f (t) cos(λt)dt → 0 as λ → ∞. Similarly, a f (t) sin(λt)dt → 0
as λ → ∞.

Hence, it is enough to prove the result for step functions. Since every step function is a
finite linear combination of characteristic functions on intervals, it is enough to prove
for f of the form f = χ[c,d] , [c, d] ⊆ [a, b]. Note that
Z b Z d
χ[c,d] cos(λt)dt = cos(λt)dt
a c
sin(λd) − sin(λc)
=
λ
2
≤ → 0 as λ → ∞.
|λ|
Z b
Similarly, χ[c,d] sin(λt)dt → 0 as λ → ∞. 
a

Remark 2.4. If f is Riemann integrable on [a, b], then there exists a sequence of (fn )
of step functions such that kf − fn k1 → 0. Thus, conclusion in Theorem 2.1 holds if f
is Riemann integrable.

Proof of Lemma 2.3. If f ∈ L1 [a, b] with f ≥ 0, then there exists an increasing sequence
of non-negative simple measurable functions ϕn , n ∈ N such that ϕn → f pointwise.
Rb
Hence, by DCT, a |f − ϕn | → 0. From this, for any complex valued f ∈ L1 [a, b], there
exists a sequence (ϕn ) of simple complex measurable functions
Z b
|f − ϕn | → 0.
a

We observe (see [3]):

(1) Every simple real valued measurable function is a finite linear combination of
characteristic function of measurable sets.
(2) For every measurable set E ⊆ (a, b) and ε > 0, there exists an open set G ⊇ E
such that m(G \ E) < ε. Hence,
Z b Z b
|χG − χE | = |χ(G\E) | ≤ m(G \ E) < ε.
a a
S∞
(3) If G ⊆ (a, b) is an open set, then G = k=1 In , where {In } is a countable
disjoint family of open intervals in (a, b);
n
X
χG = lim ψn , ψn = χIk ,
n→∞
k=1
6 M.T. NAIR

Since 0 ≤ ψn ≤ χG , by DCT,
Z
|χG − ψn | → 0.

(4) By (1)-(3), if ϕ is a simple measurable function and ε > 0, there exists a step
function ψ such that
Z b
|ϕ − ψ| < ε.
a

Thus, the lemma is proved. 

3. Dirichlet kernel

Note that
N
1 X inx π
Z Z π
−int 1
SN (f, x) := e f (t)e dt = f (t)DN (x − t)dt,
2π n=−N −π 2π −π
where
N
X
DN (t) := eint .
n=−N
Redefining f at the end-points if necessary, and extending it as a 2π-periodic function,
we can also write (verify!),
Z π
1
SN (f, x) = f (x − t)DN (t)dt.
2π −π

Notation: We denote by T the unit circle T := {eit : −π ≤ t < π}. Note that if
f : T → C and if we define f˜ : R → C by f˜(t) = f (eit ), then
f˜(−π) = f˜(π) and f˜(t + 2nπ) = f (t) for all n ∈ Z.
That is, f˜ is a 2π-periodic function. In the due course, we shall identify 2π-periodic
functions with functions on T . We shall denote L1 (T ) for the space of all 2π-periodic
(complex valued) functions on R (with equality replaced equal a.e.) which are inte-
grable on [−π, π] with norm
Z π
1
f 7→ kf k1 := |f (x)|dx.
2π −π
Analogously, for 1 ≤ p < ∞, Lp (T ) denotes the space of all 2π-periodic (complex
valued) functions f on R such that |f |p is integrable on [−π, π] with norm
 Z π 1/p
1 p
f 7→ kf kp := |f (x)| dx
2π −π
TOPICS IN FOURIER ANALYSIS 7

The space L2 (T ) is also a Hilbert space with inner product


Z π
1
(f, g) 7→ hf, gi := f (x)g(x) dx.
2π −π

Definition 3.1. The function DN (·) is called the Dirichlet kernel.

We observe that,

• D
Z Nπ (−t) = DN (t) for all t ∈ [−π, π] and
• DN (t)dt = 1.
−π
N
X N
X N
X
int int −int
• DN (t) = e =1+ [e + e ]=1+2 cos nt.
n=−N n=1 n=1
Z π
Remark 3.2. We shall see that |DN (t)|dt → ∞ as N → ∞.
−π

Theorem 3.3.



 2N + 1, t = 0,

DN (t) =
sin(N + 21 )t
, t 6= 0.


sin( 2t )

Proof. Clearly, DN (0) = 2N + 1. So, let t 6= 0. Note that


N
X
it
(e − 1)DN (t) = [ei(n+1)t − eint ] = ei(N +1)t − e−iN t .
n=−N

But,
(eit − 1)DN (t) = eit/2 (eit/2 − e−it/2 )DN (t) = 2ieit/2 sin(t/2)DN (t).

Thus,

2i sin(t/2)DN (t) = e−it/2 [ei(N +1/2)t − e−i(N +1/2)t ] = 2i sin(N + 1/2)t.

i.e.,
sin(N + 21 )t
DN (t) = , t 6= 2kπ.
sin( 2t )

8 M.T. NAIR

4. Dirichlet-Dini criterion for convergence

We investigate the convergence:


SN (f, x) → f (x).
Z π Rπ
1
Since DN (t)dt = 1 and SN (f, x) = 2π −π
f (x − t)DN (t)dt, we have
−π
Z π
1
f (x) − SN (f, x) = [f (x) − f (x − t)]DN (t)dt.
2π −π

Theorem 4.1. (Dirichlet-Dini criterion) Let f ∈ L1 (T ). If f satisfies


Z π
f (x) − f (x − t)
dt < ∞ (∗)
−π t
at a point x ∈ [−π, π], then
SN (f, x) → f (x).
If (∗) hods uniformly for x ∈ [−π, π], then the convergence {SN (f, x)} to f (x) is
uniform.
f (x) − f (x − t)
Remark 4.2. In the above theorem, by , we mean the function
t
(
f (x)−f (x−t)
t
, t 6= 0
ϕ(t) =
0, t = 0.

Proof of Theorem 4.1. We observe that


Z π
1
f (x) − SN (f, x) = [f (x) − f (x − t)]DN (t)dt
2π −π
sin(N + 21 )t
Z π
1
= [f (x) − f (x − t)] dt
2π −π sin( 2t )
1 π f (x) − f (x − t)
Z   
t/2 1
= t sin(N + )tdt
π −π t sin( 2 ) 2

Since (t/2)/[sin(t/2)] is bounded, in view of Riemann Lebesgue lemma, we have the


following. 

The following corollaries are immediate from Theorem 4.1.


Corollary 4.3. Suppose f is Lipschitz at a point2 x ∈ [−π, π]. Then
SN (f, x) → f (x) as N → ∞.
2Afunction ϕ : I → C is said to be Lipschitz at a point x0 ∈ I if there exists K0 > 0 such that
|ϕ(x) − ϕ(x0 )| ≤ K0 |x − x0 | for all x ∈ I.
TOPICS IN FOURIER ANALYSIS 9

Corollary 4.4. Suppose f is Lipschitz3 on [−π, π]. Then


SN (f, x) → f (x) as N →∞
uniformly on [−π, π].

Notation: We denote by C(T ) the space of all 2π-periodic continuous functions on R,


and by C k (T ) for k ∈ N ∪ {0}, the space of all 2π-periodic functions on R which are
k-times continuously differentiable on R.

Corollary 4.5. If f ∈ C 1 (T ), then


SN (f, x) → f (x) as N →∞
uniformly on R.

Now obtain a more general result.

Theorem 4.6. Suppose f is a 2π-periodic function such that the following limits exist
at a point x ∈ R:
f (x+) := lim f (x + t), f (x−) := lim f (x − t),
t→0+ t→0+

f (x + t) − f (x+) f (x−) − f (x − t)
f 0 (x+) := lim , f 0 (x−) := lim .
t→0+ t t→0+ t
Then
f (x+) + f (x−)
SN (f, x) → as N → ∞.
2

Proof. Since DN (t) = DN (−t), we have


Z π
1
SN (f, x) = f (x − t)DN (t)dt
2π −π
Z 0 Z π
1 1
= f (x − t)DN (t)dt + f (x − t)DN (t)dt
2π −π 2π 0
Z π Z π
1 1
= f (x + t)DN (t)dt + f (x − t)DN (t)dt
2π 0 2π 0
Z π
1
= [f (x + t) + f (x − t)]DN (t)dt
2π 0
and Z π Z π
1 2
DN (t)dt = DN (t)dt.
2π −π 2π 0

3A function ϕ : I → C is said to be Lipschitz on I if there exists K > 0 such that |ϕ(x) − ϕ(x0 )| ≤
K0 |x − x0 | for all x ∈ I.
10 M.T. NAIR

Hence, for any β ∈ R,


Z π
1
SN (f, x) − β = [f (x + t) + f (x − t) − 2β]DN (t)dt.
2π 0
f (x+)+f (x−)
Taking β = 2
, we have
f (x + t) + f (x − t) − 2β = [f (x + t) − f (x+)] − [f (x−) − f (x − t)].
Thus,
SN (f, x) − β = AN + BN ,
where
Z π Z π
1 1
AN = [f (x + t) − f (x+)]DN (t)dt, BN = [f (x−) − f (x − t)]DN (t)dt.
2π 0 2π 0

Note that
Z π
1
A = [f (x + t) − f (x+)]DN (t)dt
2π 0
sin(N + 21 )t
Z π
1
= [f (x + t) − f (x+)] dt
2π 0 sin( 2t )
1 π f (x + t) − f (x+)
Z   
t/2 1
= t sin(N + )tdt
π 0 t sin( 2 ) 2

f (x+t)−f (x+)
Since t
→ f 0 (x+) as t → 0+, there exists δ > 0 such that
f (x + t) − f (x+)
0<t<δ =⇒ − f 0 (x+) ≤ 1
t
f (x + t) − f (x+)
=⇒ ≤ 1 + |f 0 (x+)|.
t
Hence, the function
  
f (x + t) − f (x+) t/2
t 7→ , t 6= 0,
t sin( 2t )
is bounded on (0, δ), and hence, belongs to L1 (T ). Therefore,by Riemann Lebesgue
lemma, AN → 0 as N → ∞. Similarly, we see that, BN → 0 as N → ∞. 

An immediate corollary:

Corollary 4.7. If f ∈ C(T ) and has left and right derivative at a point x, then
SN (f, x) → f (x) as N → ∞.

The following result is known as localization lemma.


TOPICS IN FOURIER ANALYSIS 11

LEMMA 4.8. For 0 < r < π and x ∈ [−π, π],


Z
f (x − t)DN (t)dt → 0 as N → ∞.
r≤|t|≤π

Proof. Observe that


Z Z
f (x − t)DN (t)dt = g(x, t) sin(N + 1/2)tdt,
r≤|t|≤π r≤|t|≤π

where 
f (x − t)/ sin(t/2), r ≤ |t| ≤ π,
g(x, t) =
0, |t| ≤ r.
Since g(x, ·) is integrable, by Riemann Lebesgue lemma,
Z
g(x, t) sin(N + 1/2)tdt → 0 as N → ∞.
r≤|t|≤π

Proof of Corollary 4.4 using localization lemma. Suppose f is Lipschitz at a point


x ∈ [−π, π] with Lipschitz constant Kx , i.e., there exists δ > 0 such that
|f (x) − f (x − t)| ≤ Kx |t| whenever |t| < δ.
Now,
Z π
1
f (x) − SN (f, x) = [f (x) − f (x − t)]DN (t)dt
2π −π
Z
1
= [f (x) − f (x − t)]DN (t)dt
2π 0≤|t|<δ
Z
1
+ [f (x) − f (x − t)]DN (t)dt
2π δ≤|t|≤π

By Lemma 4.8,
Z
1
[f (x) − f (x − t)]DN (t)dt → 0 as N → ∞.
2π δ≤|t|≤π

Hence, for a given ε > 0, there exists N0 ∈ N such that for all N ≥ N0 ,
Z
1
[f (x) − f (x − t)]DN (t)dt < ε/2.
2π δ≤|t|≤π
Also, But,
Z Z
1 1
[f (x) − f (x − t)]DN (t)dt ≤ |f (x) − f (x − t)| |DN (t)|dt,
2π 0≤|t|<δ 2π 0≤|t|<δ
12 M.T. NAIR
Z Z
1 1
|f (x) − f (x − t)| |DN (t)|dt ≤ Kx |t| |DN (t)|dt,
2π 0≤|t|<δ 2π 0≤|t|<δ

sin(N + 12 )t t/2 1
|t| |DN (t)| = |t| t =2 t | sin(N + )t| ≤ 2M,
sin( 2 ) sin( 2 ) 2
t/2
where M is a bound for | on 0 < |t| ≤ δ. Hence,
sin( 2t )
Z
1 4M Kx δ 2M Kx δ
[f (x) − f (x − t)]DN (t)dt ≤ = .
2π 0≤|t|<δ 2π π
2M Kx δ
We may take δ such that < ε/2. Hence,
π
Z
1
|f (x) − SN (f, x)| ≤ [f (x) − f (x − t)]DN (t)dt
2π 0≤|t|<δ
Z
1
+ [f (x) − f (x − t)]DN (t)dt
2π δ≤|t|≤π
< ε for all N ≥ N0 .


Exercise 4.9. Suppose f is 2π-periodic and Hölder continuous at x, i.e., there exist
M > 0 and α > 0 such that |f (x) − f (y)| ≤ M |x − y|α for all y ∈ [−π, π]. Then show
that SN (f, x) → f (x) as N → ∞.

Exercise 4.10. Suppose f is 2π-periodic and Hölder continuous on [−π, π], i.e., there
exist M > 0 and α > 0 such that |f (x) − f (y)| ≤ M |x − y|α for all x, y ∈ [−π, π].
Then show that SN (f, x) → f (x) uniformly.

5. Ces̀aro summablity of Fourier series

Theorem 5.1. (Fejér’s theorem) If f ∈ C(T ), then the Fourier series of f is


uniformly Ces̀aro summable on [−π, π], that is,
N
1 X
σN (f, x) := Sk (f, x) → f (x) as N →∞
N + 1 k=0
uniformly on [−π, π].

Recall that
k Z π
X 1
Sk (f, x) := fˆ(n)einx = f (x − t)Dk (t)dt.
n=−k
2π −π
TOPICS IN FOURIER ANALYSIS 13

Hence,
N
( N
)
Z π
1 X 1 X
σN (f, x) = Sk (f, x) = f (x − t) Dk (t) .
N +1 k=0 −π N +1 k=0

Thus, Z π
σN (f, x) = f (x − t)KN (t)dt,
−π
where
N
1 X
KN (t) := Dk (t).
N + 1 k=0
Definition 5.2. The function KN (t) defined above is called the Fejér kernel.

We observe that Z π
1
KN (t)dt = 1.
2π −π
Hence, Z π
1
f (x) − σN (f, x) = [f (x) − f (x − t)]KN (t)dt.
2π −π
For the proof of Theorem 5.1, we shall make use of the following lemma.
LEMMA 5.3. The following results hold.

(1) For t 6= 0,
1 1 − cos(N + 1)t 1 sin2 [(N + 1)t/2]
KN (t) = = .
N +1 1 − cos t N +1 sin2 (t/2)
(2) KN (t) is an even function and KN (t) ≥ 0 for all t ∈ [−π, π].
(3) For 0 < δ ≤ π,  
1 1
KN (t) ≤ .
N + 1 sin2 (δ/2)
In particular, KN is positive and KN (t) → 0 as N → ∞ uniformly on 0 < δ ≤ |t| ≤ π.

Proof of Theorem 5.1. Sine KN (t) is a non-negative function (see Lemma 5.3), we
have Z π
1
|f (x) − σN (f, x)| ≤ |f (x) − f (x − t)|KN (t)dt.
2π −π
Let ε > 0 be given. Since f is uniformly continuous, there exists δ ∈ (0, π] such that
|f (x) − f (y)| < ε whenever |x − y| < δ.
Hence, Z Z
1 ε
|f (x) − f (x − t)|KN (t)dt < KN (t)dt = ε.
2π |t|<δ 2π |t|<δ
14 M.T. NAIR

Also, since f is uniformly bounded there exists M > 0 such that |f (y)| ≤ M for all
x ∈ [−π, π].
Z Z
1 2M
|f (x) − f (x − t)|KN (t)dt ≤ KN (t)dt.
2π |t|≥δ 2π |t|≥δ
We have observed in Lemma 5.3 that KN (t) is an even function and KN (t) → 0 as
N → ∞ uniformly on [δ, π]. Hence, there exists N0 such that
4M π
Z Z
1
|f (x) − f (x − t)|KN (t)dt ≤ KN (t)dt < ε for all N ≥ N0 .
2π |t|≥δ 2π δ
Hence,
Z π
1
|f (x) − σN (f, x)| ≤ |f (x) − f (x − t)|KN (t)dt < 2ε
2π −π
for all N ≥ N0 . Note that N0 is independent of the point x. Thus, we have proved
that SN (f, x) → f (x) as N → ∞ uniformly for x ∈ [−π, π]. 

Remark 5.4. The proof of Theorem 5.1 reveals more:

If f is peace-wise continuous and 2π-periodic, and continuous at x, then


σN (f, x) → f (x) as N → ∞.

Notation:

• un (x) := einx , n ∈ Z.
• AC(T ) denotes the vector space of all 2π-periodic complex valued functions
defined on R which are absolutely continuous.

• span{un : n ∈ Z} is the space (over C) of all trigonometric polynomials.

Corollary 5.5. The space of all trigonometric polynomials is dense in C(T ) with
respect to the uniform norm, and hence dense in Lp (T ) w.r.t. k · kp for 1 ≤ p < ∞.

Proof. By Theorem 5.1, space of all trigonometric polynomials is dense in C(T ) with
respect to the uniform norm k · k∞ . Hence, for any f ∈ C(T ), there exists a sequence
(fn ) of trigonometric polynomials such that
Z π
p
kf − fn kp = |f (x) − fn (x)|p dx ≤ 2πkf − fn kp∞ → 0
−π

as n → ∞. 

Corollary 5.6. If f ∈ L2 (T ) for some 1 ≤ p < ∞ and fˆ(n) = 0 for all n ∈ Z, then
f = 0 a.e.
TOPICS IN FOURIER ANALYSIS 15

Proof. Suppose f ∈ L2 (T ) for some 1 ≤ p < ∞ and fˆ(n) = 0 for all n ∈ Z, i.e.,
hf, un i = 0 for all n ∈ Z. By Corollary 5.5, it follows that kf kL2 = 0. Hence, f = 0
a.e. 

Corollary 5.7. If f ∈ C(T ) such that fˆ(n) = 0 for all n ∈ Z, then f = 0. In


particular, if f, g ∈ C(T ) such that fˆ(n) = ĝ(n) for all n ∈ Z, then f = g.

Proof. Suppose f ∈ C(T ) such that fˆ(n) = 0 for all n ∈ Z. Thus, hf, un iL2 = 0 for all
n ∈ Z. Since C(T ) ⊆ L2 [−π, π], f ∈ L2 [−π, π]. Hence by Corollary, f = 0 a.e. Since
f is continuous, f = 0. 

The above corollary shows:

The Fourier coefficients of f ∈ C(T ) determines f uniquely.

Corollary 5.8. If f ∈ C 2 (T ), then

fc00 (n) = (in)2 fˆ(n) for all n ∈ Z.

In particular, fˆ(n) = o( n12 ), and the Fourier series of f converges uniformly to f .

Proof. Let f ∈ C 2 (T ). Then, using integration by parts, we obtain,


Z π
ˆ
2π f (n) = f (x)e−inx dx
−π
Z π
h e−inx iπ h e−inx i
= f (x) − f 0 (x) dx
−in −π −π −in
1 π 0
Z
= f (x)e−inx dx
in −π
e−inx iπ 1 π 00 h e−inx i
Z
1h 0
= f (x) − f (x) dx
in −in −π in −π −in
Z π
1
= 2
f 00 (x)e−inx dx.
(in) −π

Hence, fc00 (n) = (in)2 fˆ(n) for all n ∈ Z. In particular, fˆ(n) = o(1/n2 ). Therefore,
P ˆ
n∈Z |f (n)| converges, and hence the Fourier series converges uniformly. Suppose
SN (f, x) → g(x) uniformly. Then it follows that g ∈ C(T ) and ĝ(n) = fˆ(n) for all
n ∈ Z. Therefore, by Corollary 5.7, g = f. 

Following the same arguments as in the proof of Corollary 5.8, we obtain:


16 M.T. NAIR

Corollary 5.9. If f ∈ C 1 (T ) and f 0 is absolutely continuous, then f 00 exists almost


everywhere, f 00 ∈ L1 [−π, π] and

fc00 (n) = (in)2 fˆ(n) for all n ∈ Z,


and the Fourier series of f converges uniformly to f .

More generally,

Theorem 5.10. If f ∈ C k−1 (T ) and f (k−1) is absolutely continuous for some k ∈ N,


then f (k) exists almost everywhere f (k) ∈ L1 (T ) and
(k) (n) = (in)k fˆ(n) for all
fd n ∈ Z.

Proof of Lemma 5.3. We have


N
1 X sin(k + 1/2)t
KN (t) := Dk (t) where Dk (t) =
N + 1 k=0 sin t/2

Hence,
N N
X sin(k + 1/2)t X ei(k+1/2)t − e−i(k+1/2)t
(N + 1)KN (t) = =
k=0
sin t/2 k=0
eit/2 − e−it/2
But,
ei(k+1/2)t − e−i(k+1/2)t ei(k+1)t − e−ikt
= ,
eit/2 − e−it/2 eit − 1
ei(k+1/2)t − e−i(k+1/2)t eikt − e−i(k+1)t
= ,
eit/2 − e−it/2 1 − e−it
Therefore,
N
X
[eit − 1](N + 1)KN (t) = [ei(k+1)t − e−ikt ], (1)
k=0

N
X
−it
[1 − e ](N + 1)KN (t) = [eikt − e−i(k+1)t ] (2)
k=0

Subtracting the (2) from (1),


N
X
[2 cos t − 2](N + 1)KN (t) = 2 [cos(k + 1)t − cos kt] = 2[cos(N + 1)t − 1]
k=0

Thus,
1 cos(N + 1)t − 1 1 sin2 [(N + 1)t/2]
KN (t) = = .
N +1 cos t − 1 N +1 sin2 (t/2)
TOPICS IN FOURIER ANALYSIS 17

Thus, we have proved (1). It is clear that KN (t) is even and non-negative. Now, for
0 < δ ≤ π, sin2 (t/2) ≥ sin δ/2, so that
Z π Z π Z π
1 sin2 [(N + 1)t/2] 1 1
KN (t)dt = 2 dt ≤ 2 dt.
δ N +1 δ sin (t/2) N + 1 δ sin (δ/2)
Thus,
π
π−δ
Z
KN (t)dt ≤ → 0 as N → ∞.
δ (N + 1) sin2 (δ/2)


Exercise 5.11. Suppose f is piecewise continuous and 2π-periodic. If fˆ(n) = 0 for all
n ∈ Z, then f (x) = 0 for all x at which f is continuous.

Exercise 5.12. If f ∈ C 1 (T ), then fˆ(n) = O(1/n). More generally, f ∈ C k (T ) implies


fˆ(n) = O(1/nk ).

Example 5.13. Let f (x) = x2 , |x| ≤ π. Note that


Z π
ˆ π3
2π f (0) = x2 dx = 2
−π 3
so that fˆ(0) = π 2 /3, and for n 6= 0,
Z π
ˆ
2π f (n) = x2 e−inx dx
−π
−inx iπ π
e−inx
Z
2e
h
= x − 2x dx
−in −π −π −in
−inx iπ
h e h e−inx iπ
= x2 − 2x
−in −π (−in)2 −π
h e−inx iπ h e−inx iπ einx
= − 2x = 2x = 4π
(−in)2 −π n2 −π n2
(−1)n
= 4π 2
n
Hence, for n 6= 0,
n
(−1)
fˆ(n) = 2 2 .
n
Thus,
∞ ∞
π2 X (−1)n inx π 2 X (−1)n
x2 ≈ +2 2
e = + 4 2
cos nx.
3 n6=0
n 3 n=1
n
Since the series of coefficients converges absolutely, we have

π3 X (−1)n
f (x) = +4 2
cos nx.
3 n=1
n
18 M.T. NAIR

Taking x = 0,

π3 X (−1)n
0= +4 2
.
3 n=1
n
Thus,

X (−1)n+1 π2
= .
n=1
n2 12
Taking x = π,
∞ ∞
2π3 X (−1)n n π3 X 1
π = +4 2
(−1) = +4 .
3 n=1
n 3 n=1
n2
Thus,

X 1 π2
2
= .
n=1
n 6

Example 5.14. Let f (x) = x, x ∈ [−π, π]. Note that fˆ(0) = 0 and for n 6= 0,
Z π h e−inx iπ Z π −inx h e−inx iπ
e
2π fˆ(n) = xe−inx
dx = x − dx = x .
−π −in −π −π −in −in −π
Thus,
h e−inx iπ 1 einπ
2π fˆ(n) = x = [πe−inπ + πeinπ ] = 2π
−in −π −in −in
so that
(−1)n (−1)n+1
fˆ(n) = = .
−inπ inπ
Hence,

X (−1)n+1 X (−1)n+1 X (−1)n+1
x= einx = [einx − e−inx ] = 2 sin nx
n6=0
in n=1
in n=1
n

Taking x = π/2 we obtain the Madhava-Nīlakaṅtha series


∞ ∞
π X (−1)n+1 nπ X (−1)n
= sin = . ♦
4 n=1
n 2 n=0
2n + 1

6. Divergence of Fourier series

Theorem 6.1. There exists f ∈ C(T ) such that {SN (f, 0)} is unbounded; in particular,
the Fourier series of f does not converge to f at 0.

For this we shall make use of the Uniform Boundedness Principle from Functional
Analysis:
TOPICS IN FOURIER ANALYSIS 19

Theorem 6.2. (Uniform Boundedness Principle) Let (Tn ) be a sequence of con-


tinuous linear transformations from a Banach space X to a normed linear space Y . If
for each u ∈ X, the set {kTn uk : n ∈ N} is bounded, then there exists M > 0 such that

sup kTn uk ≤ M ∀ n ∈ N.
kuk≤1

Let
ϕN (f ) := SN (f, 0), f ∈ C(T ).
We see that ϕN : C(T ) → C is a linear functional for each N ∈ N and
Z π  Z π 
1 1
|ϕN (f )| = |SN (f, 0)| = f (−t)DN (t)dt ≤ kf k∞ |DN (t)|dt .
2π −π 2π −π
Hence, each ϕN is a continuous linear functional on C(T ) and
Z π
1
sup |ϕN (f )| ≤ |DN (t)|dt.
kuk∞ ≤1 2π −π
In fact,

Theorem 6.3. Z π
1
sup |ϕN (f )| = |DN (t)|dt
kuk∞ ≤1 2π −π

and
Z π N
8X1
|DN (t)|dt ≥ .
−π π k=1 k

Proof of Theorem 6.1. By Theorem 6.3, there does not exist M > 0 such that
supkuk∞ ≤1 |ϕN (f )| ≤ M for all n ∈ N. Hence, by Theorem 6.2, there exists f ∈ C(T )
such that {|ϕn (f )| : n ∈ N} is unbounded. Hence, there exists f ∈ C(T ) such that
Fourier series of f diverges at 0. 

Remark 6.4. Let D := {f ∈ C(T ) : {SN (f, 0)} does not converge}. Then C(T ) \ D
is a subspace of C(T ), and by Theorem 6.1, C(T ) \ D is a proper subspace. Hence,
C(T ) \ D is nowhere dense, and hence D is dense in C(T ). Thus, we have proved the
following:

There exists a dense subset D of C(T ) such that for each f ∈ D, the
Fourier series of f diverges at 0.

In place of 0, we can take any point in [−π, π] and obtain similar divergence result at
that point.
20 M.T. NAIR

7. Uniqueness

Theorem 7.1. (Uniqueness of Fourier series) Let f ∈ L1 (T ). If fˆ(n) = 0 for all


n ∈ N, then f = 0 a.e.

Proof. Let
Z t
g(t) = f (x)dx, t ∈ [−π, π].
−π

Then, by Fundamental Theorem of Lebesgue Integration (FTLI), g is absolutely con-


tinuous, g 0 exists a.e. and g 0 = f a.e. Note that
Z t+2π Z π
g(t + 2π) − g(t) = f (x)dx = f (x)dx = 2π fˆ(0) = 0.
t −π

Hence g is 2π-periodic. Let


Z t
h(t) = g(x)dx, t ∈ [−π, π].
−π

Then we see that


Z t+2π Z π
h(t + 2π) − h(t) = g(x)dx = g(x)dx = 2πĝ(0).
t −π

Taking
Z t
G(t) = [g(x) − ĝ(0)]dx, t ∈ [−π, π],
−π

we have
Z π
G(t + 2π) − G(t) = [g(x) − ĝ(0)]dx = 2π[ĝ(0) − ĝ(0)] = 0.
−π

Thus, G is 2π-periodic, and G00 = f a.e.Hence,

fˆ(n) = G
c00 (n) = (in)2 Ĝ(n) for all n 6= 0.

Therefore, Ĝ(n) = 0 for all n 6= 0. Hence, by Corollary 5.9, G(x) = Ĝ(0), and hence
G00 = 0, so that f = 0 a.e. 

Recall that for each f ∈ L1 (T ),

fˆ(n) → 0 as |n| → ∞.

Thus, (fˆ(n)) ∈ c0 (Z) for every f ∈ L1 [−π, π].


Notation: c0 (Z) is the set of all sequences ϕ : Z → C such that ϕ(n) → 0 as |n| → ∞.
TOPICS IN FOURIER ANALYSIS 21

Theorem 7.2. The map F : L1 (T ) → c0 (Z) be defined by


F(f ) = (fˆ(n)), f ∈ L1 (T )
is an injective continuous linear operator which is not onto.

Proof. For f, g ∈ L1 (T ) and α ∈ C, we have

+ g)(n)) = fˆ(n) + ĝ(n) for all n ∈ Z,


(f\
c (n)) = αfˆ(n). for all n ∈ Z,
αf
Thus, F is a linear operator. Note that
Z π Z π
ˆ 1 −inx 1
|f (n)| = f (x)e dx ≤ |f (x)|dx.
2π −π 2π −π
Thus, if we endow L1 [−π, π] with the norm
Z π
1
kf kL1 := |f (x)|dx, f ∈ L1 (T ),
2π −π
then we see that F is a continuous linear operator. By Theorem 7.1, F is injective. So,
it remains to show that F is not onto. If it is onto, then my Bounded Inverse Theorem,
its inverse is also continuous. Note that
F(DN ) = {D
dN (n)}

and
b N (n) = 1 for |n| ≥ N
D
so that
k(F(DN ))k∞ = 1 for all N ∈ N.
If F is onto, then, by Bounded Inverse Theorem4 its inverse F −1 is continuous so that
(kDN k) = {kF −1 (F(DN )k} is bounded, which is not true. 

By the above theorem there exists (cn ) ∈ c0 (Z) such that there is no f ∈ L( T )
satisfying cn = fˆ(n) for all n ∈ N. It is a natural urge to have an example of such a
sequence cn ). We shall show that cn ) with

1/ log(n), n ≥ 2,
cn =
0, n ≤ 1,
is such a sequence. This is a consequence of the first part of the following theorem.
4If X and Y are Banach spaces and T : X → Y is a continuous bijective linear operator, then T −1
is also continuous.
22 M.T. NAIR

X fˆ(n)
Theorem 7.3. Let f ∈ L1 (T ). Then einx converges at every x ∈ R and
n6=0
n
Z b XZ b
f (x)dx = fˆ(n)einx dx.
a n∈Z a

For proving the above theorem we shall make use of the following theorem:

Theorem 7.4. (Jordan) If f ∈ L1 (T ) is of bounded variation5, then for every x ∈ R,


1
SN (f, x) → (f (x+) + f (x−) as N → ∞.
2
In particular, if f ∈ AC(T ), then
SN (f, x) → f (x) as N →∞
for every x ∈ R.

It can be easily shown that:

Every absolutely continuous function is of bounded variation.

Proof of Theorem 7.3. Let


Z t
g(t) = [f (x) − fˆ(0)]dx.
−π

Then g is absolutely continuous and g is 2π-periodic, i.e., g ∈ AC(T ), g 0 ∈ L1 (T ) and


g 0 = f − fˆ(0) a.e. Therefore, ĝ 0 (n) = inĝ(n) for all n 6= 0 so that
fˆ(n)
ĝ(n) = , n 6= 0.
in
By Jordan’s theorem,
X X fˆ(n)
g(x) = ĝ(0) + ĝ(n)einx = ĝ(0) + einx .
n6=0 n6=0
in

X fˆ(n)
In particular, einx converges. Also,
n6=0
n

X fˆ(n) X Z x
g(x) − g(y) = [e inx
−e inx
]= fˆ(n) eint dt.
n6=0
in n6=0 y

5A
function f : [a, b] → C is of bounded variation if there exits κ > 0 such that for every partition
Pn
x0 < x1 < · · · < xn = b, k=1 |f (xk+1 ) − f (xk | ≤ κ.
TOPICS IN FOURIER ANALYSIS 23

But,
Z x Z x Z x
g(x) − g(y) = 0
g (t)dt = [f (t) − fˆ(0)]dt = f (t)dt − fˆ(0)(x − y)dt.
y y y

Hence,
Z x XZ b
f (t)dt = fˆ(n)eint dt.
y n∈Z a

This competes the theorem. 

Corollary 7.5. Let (cn ) be with



1/ log(n), n ≥ 2,
cn =
0, n ≤ 1,

Then there is no f ∈ L1 (T ) satisfying cn = fˆ(n) for all n ∈ N.

Proof. Suppose f ∈ L1 (T ) satisfying cn = fˆ(n) for all n ∈ N. Then by the first part of
Theorem 7.3, the series ∞
P einx
P∞ 1
n=2 n log n converges. In particular, taking x = 0, n=2 n log n
converges, which is not true (e.g., by integral test).


8. Convolution

Given f, g ∈ L1 (T ), it can be shown that

(x, y) 7→ f (x − y)g(y)

is measurable on R × R, and hence, for each x ∈ [−π, π], the integral


Z π
f (x − y)g(y)dy
−π

converges.

Definition 8.1. The convolution of f, g ∈ L1 (T ) is defined by


Z π
1
(f ∗ g)(x) = f (x − y)g(y)dy, x ∈ [−π, π].
2π −π

We observe the following:


24 M.T. NAIR

(1) f ∗ g ∈ L1 (T ) and kf ∗ gk1 ≤ kf k1 kgk1 :

Z π Z π Z π Z π 
|f (x − y)| |g(y)|dydx = |f (x − y)| dx |g(y)|dy
−π −π −π −π
Z π
= 2πkf k1 |g(y)|dy
−π
= (2π)2 kf k1 kgk1 .

(2) f ∗ g = g ∗ f :
Z π Z π
f (x − y)g(y)dy = f (τ )g(x − τ )dy
−π −π
Z x+π
= f (τ )g(x − τ )dy
x−π
Z π
= f (τ )g(x − τ )dy.
−π

∗ g(n) = ĝ(nfˆ(n) for all n ∈ Z:


(3) f[
Z π
1
∗ g(n) =
f[ (f ∗ g)(x)e−inx dx,
2π −π

Z π
−inx 1
(f ∗ g)(x)e = f (x − y)g(y)e−inx dy
2π −π
Z π
1
= f (x − y)g(y)e−in(x−y) e−iny dy,
2π −π

Z π Z π Z π 
−inx −in(x−y) −iny
(f ∗ g)(x)e dx = f (x − y)g(y)e e dy dx
−π −π −π
Z π Z π 
−in(x−y)
= f (x − y)e dx g(y)e−iny dy
−π −π
Z π
= 2π fˆ(n)g(y)e−iny dy
−π

= (2π)2 fˆ(n)ĝ(n).
TOPICS IN FOURIER ANALYSIS 25

(4) (f ∗ g) ∗ h = f ∗ (g ∗ h):
Z π Z π Z π 
(f ∗ g)(x − y)h(y)dy = f (x − y − t)g(t)dt h(y)dy
−π −π −π
Z π Z π 
= f (x − τ )g(τ − y)dτ h(y)dy
−π −π
Z π Z π 
= f (x − τ ) g(τ − y)h(y)dy dτ
−π −π
Z π
= 2π f (x − τ )(g ∗ h)(τ )dτ
−π
2
= (2π) [f ∗ (g ∗ h)](x).

Theorem 8.2. With respect to convolution as multiplication, L1 (T ) is a Banach alge-


bra.

• The Banach algebra L1 (T ) does not have a multiplicative identity:


Suppose there exists ϕ ∈ L( T ) such that f ∗ ϕ = f for all f ∈ L1 (T ). Then
fˆ(n)ϕ̂(n) = fˆ(n) for all f ∈ L1 (T ). Hence, ϕ̂(n) = 1 whenever ϕ̂(n) 6= 0. But,
ϕ̂(n) → 0 as |n| → ∞. Hence, there exists N ∈ N such that ϕ̂(n) = 0 for all
n ≥ N . Let f ∈ L1 (T ) be such that fˆ(n) 6= 0 for some n ≥ N . Then for such
n, we obtain

0 = fˆ(n)ϕ̂(n) = fˆ(n) 6= 0,

which is a contradiction.

However,

• There exists (ϕn ) in L1 (T ) such that kf ∗ ϕn − f k1 → 0.

In fact, we have the following.

Theorem 8.3. Let Kn be the Fejér kernel. Then, for every f ∈ L1 (T ),

kf ∗ Kn − f k1 → 0 as N → ∞.

Proof. Recall that if g ∈ C(T ), then kg ∗ ϕn − gk1 → 0. Let f ∈ L1 (T ) and ε > 0


be given. Let g ∈ C(T ) be such that kf − gk1 < ε, and let N ∈ N be such that
26 M.T. NAIR

kg ∗ ϕn − gk1 < ε for all n ≥ N . Then, for n ≥ N , we have


kf ∗ Kn − f k1 ≤ kf ∗ Kn − g ∗ Kn k1 + kg ∗ Kn − gk1 + kg − f k1
≤ k(f − g) ∗ Kn k1 + ε + ε
≤ k(f − g) ∗ Kn k1 + 2ε
≤ kf − gk1 kKn k1 + 2ε
≤ 3ε.
1
Rπ 1

The last inequality is due the fact that 2π −π
|Kn (t)|dt = 2π −π
Kn (t)dt = 1. 

9. L2 -Theory

The norm on L2 (T ) is given by


 Z π 1/2
1 2
kf k2 = |f (x)| dx .
2π −π

Observe:

(1) If un (x) := einx , n ∈ Z, then the set {un : n ∈ Z} is an orthonormal set and
span{un : n ∈ Z}, the space of all trigonometric polynomials, is dense in L2 (T ).
(2) Let f ∈ L2 (T ) and SN (f ) := SN (f, ·). Then
X N
(a) SN (f ) = hf, un iun .
n=−N
N
X
(b) kSN (f )k22 = |fˆ(n)|2 .
n=−N
N
X
(c) kf − SN (f )k22 = kf k22 − kSN (f )|22 = kf k22 − |fˆ(n)|2 .
n=−N
N
X
(d) (Bessel’s inequality): |fˆ(n)|2 ≤ kf k22 ∀ N ∈ N. In particular,
n=−N
fˆ(n) → 0 as |n| → ∞.
(e) hf − SN (f ), un i = 0 ∀ |n| ≤ N }.
(f) kf − SN (f )k2 ≤ kf − gk ∀ g ∈ span{un : n ∈ Z, |n| ≤ N }.

Only (2)(f) requires some explanation.


Note that for every g ∈ span{un : n ∈ Z, |n| ≤ N },
kf − gk22 = kf − SN (f )k22 + kSN (f ) − gk22 ,
because, in view of (2)(e), hf − SN (f ), SN (f ) − gi = 0.
TOPICS IN FOURIER ANALYSIS 27

• The result in (2)(d) gives another proof for the Riemann Lebesgue lemma,
because L2 (T ) is dense in L1 (T ).
• In view of (2)(f),
kf − SN (f )k2 = inf{kf − gk2 : g ∈ XN },
where XN := span{un : n ∈ Z, |n| ≤ N }. In other words, SN (f ) is the (unique!)
best approximation of f from XN . Uniqueness is due to the following: Suppose
ϕ be in XN such that
kf − ϕk2 = inf{kf − gk2 : g ∈ XN }.
Then,
kf − ϕk22 = kf − SN (f )k22 + kSN (f ) − ϕk22
since hf − SN (f ), SN (f ) − ϕi = 0 so that we obtain kSN (f ) − ϕk2 = 0.

Theorem 9.1. Let f ∈ L2 (T ). Then we have the following:

(1) span{un : n ∈ Z} is an orthonormal basis of L2 (T ), i.e., a maximal orthonormal


set in L2 (T ). X
(2) (Fourier expansion) f = fˆ(n)un in L2 (T ).
n∈Z X
(3) (Parseval’s formula) 2
kf k2 = |fˆ(n)|2 .
n∈Z

Proof. (1) It can be seen that hf, un i = 0 for all n ∈ Z implies f = 0 in L2 (T ). Hence,
span{un : n ∈ Z} is a maximal orthonormal set in L2 (T ).
(3) We observe that, for n > m,
X
kSn (f ) − Sm (f )k2 ≤ |fˆ(n)|2 .
n≤|k|≤m

Hence, {Sn (f )} is a cauchy sequence in L2 (T ). Therefore, it converges to some g ∈


L2 (T ). It can be seen that ĝ(n) = fˆ(n) for all n ∈ Z. Therefore, g = f in L2 (T ).
(3) Follows from (2). 

Now, we give another proof for the following theorem:

Theorem 9.2. If f ∈ C 1 (T ), then the Fourier series of f converges absolutely, and


uniformly to f . Further,
 
1
kf − SN (f, ·)k∞ = O √ .
N
28 M.T. NAIR

Proof.
f ∈ C 1 (T ) =⇒ fˆ0 (n) = infˆ(n).
Hence,
!1/2
X X1 X1 X 1 π
|fˆ(n)| = | |infˆ(n)| = | |fˆ0 (n)| ≤ kfˆ0 k2 = √ kfˆ0 k2 .
n6=0 n6=0
n n6=0
n n6=0
n2 3
Hence the Fourier series of f converges absolutely, and uniformly to a continuous
function, say g ∈ C(T ). Since ĝ(n) = fˆ(n) for all n ∈ Z, we obtain g = f . We also
observe that, for all x ∈ R,
 1/2
X X1 X 1 fˆ0 k2
 kfˆ0 k2 ≤ k√
|f (x) − SN (f, x)| ≤ |fˆ(n)| = | |fˆ0 (n)| ≤  2
.
n n N
|n|>N n6=0 |n|>N

This completes the proof. 

References
[1] R. Bhatia, Fourier Series, TRIM, Hindustan Book Agency, 1993 (Second Edition: 2003) .
[2] S. Kesavan, Lectures on Fourier Series, Notes: Third Annual Foundation School, December,
2006.
[3] M.T. Nair, Measure and Integration, Notes for the MSc. Course, Jan-may, 2014.
[4] R. Radha & S. Thangavelu, Fourier Series, Web-Course, NPTEL, IIT Madras, 2013.
[5] B. O. Turesson, Fourier Analysis, Distribution Thoery, and Wavelets, Lecture Notes, March,
2012.

Department of Mathematics, I.I.T. Madras, Chennai-600 036, INDIA

E-mail address: mtnair@iitm.ac.in

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy