Complete
Complete
Mathematical Analysis
Handbook
Università degli studi di Roma ”La Sapienza”
Physics and Astrophysics BSc
Matteo Cheri
Mathematical Analysis
Handbook
Written by
Matteo Cheri
Università degli Studi di Roma ”La Sapienza”
Physics and Astrophysics BSc
2 Abstract Spaces 15
2.1 Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.1 Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Convergence and Compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.1 Hölder and Minkowski Inequalities . . . . . . . . . . . . . . . . . . . . . . . . . 24
3 Differential Analysis 27
3.1 Digression on the Notation Used . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.1.1 Differential Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2 Curves in Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3 Differentiability in Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.4 Differentiability in C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.4.1 Holomorphic Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.5 Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.6 Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.6.1 Critical Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.6.2 Convexity and Implicit Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.6.3 Lagrange Multipliers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
1
CONTENTS 2
5 Integral Analysis 61
5.1 Measure Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.1.1 Jordan Measure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.1.2 Lebesgue Measure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.2 Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.2.1 Lebesgue Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.2.2 Lebesgue Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.3 Calculus of Integrals in R2 and R3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.3.1 Double Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.3.2 Triple Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.3.3 Change of Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.3.4 Line Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.3.5 Surface Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.4 Integration in C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.4.1 Integration of Holomorphic Functions . . . . . . . . . . . . . . . . . . . . . . . 81
5.4.2 Integral Representation of Holomorphic Functions . . . . . . . . . . . . . . . . 85
5.5 Integral Theorems in R2 and R3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
8 Distributions 131
8.1 Linear Functionals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
8.1.1 Dual Spaces and Functionals . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
8.2 Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.2.1 Local Integrability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.2.2 Regular and Singular Distributions . . . . . . . . . . . . . . . . . . . . . . . . . 138
8.2.3 Operations with Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
8.2.4 A Physicist’s Trick, The Dirac δ “Function” . . . . . . . . . . . . . . . . . . . . . 154
8.3 Integral Representation of Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . 156
8.3.1 Dirac Delta and Heaviside Theta . . . . . . . . . . . . . . . . . . . . . . . . . . 156
8.3.2 Distributions in Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
8.4 Some Applications in Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
8.4.1 Electrostatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
The idea of writing this came from a mix of pure laziness and despair in preparing two exams, Analisi
Vettoriale (Vector Analysis) and Metodi e Modelli Matematici per la Fisica (Mathematical Models and
Methods for Physics, basically just pure Complex Analysis and Functional Analysis), that I have to sustain
in my BSc course in Physics here at Sapienza University of Rome.
These two courses are deeply intertwined and it’s really difficult to study them apart due to the sheer
volume of things that either are done for half in one course and half in the other one, or they simply
get generalized in the second, breaking up the logical flow one might get from studying these two
seemingly completely different subjects.
There will surely be errors in grammar, typing and probably some mathematical inaccuracies so for
any question, inquiry or you just want to say hi to me, don’t wait in sending me an email here
[email protected].
A huge thanks to anyone of you that’ll read this, I hope it will be useful for you as it is to me.
5
CONTENTS 6
Notation
7
CONTENTS 8
1 Complex Numbers and Functions
Re(z) = x
(1.1)
Im(z) = y
Theorem 1.1. With the previous definitions the set C forms a field.
Definition 1.1.3 (Imaginary Unit). We define the imaginary unit i = (0, 1) ∈ C. From this definition
and the definition of product of two complex numbers, we have that i2 = −1
With this definition, we have
∀z ∈ C z = Re(z) + iIm(z) (1.2)
Definition 1.1.4 (Complex Conjugate). Taken z ∈ C, we call the complex conjugate of z the number
w such that
w = Re(z) − iIm(z) (1.3)
This number is denoted as z
Definition 1.1.5 (Complex Module). We define the module or norm of a complex number, the
following operator.
√ q
kzk = zz = Re2 (z) + Im2 (z) (1.4)
9
CHAPTER 1. COMPLEX NUMBERS AND FUNCTIONS 10
Definition 1.1.6 (Complex Inverse). The inverse of a complex number z ∈ C is defined as z −1 and it’s
calculated as follows
z
z −1 = 2 (1.5)
kzk
Definition 1.1.7 (Polar Form). Taken a complex number z ∈ C one can define it in polar form with its
modulus r and its argument θ. We have that, if z = x + iy
2
p
r = x2 + y 2 = kzk
y (1.6)
tan(θ) =
x
Definition 1.1.8 (Principal Argument). Taken arg(z) = θ we can define two different arguments, due
to the periodicity of the tan function.
As a rule of thumb, using the previous definition of argument of a complex number z = x + iy, we
have
arctan(y/x) − π x < 0, y < 0
Arg(z) = arctan(y/x) x ≥ 0, z 6= 0 (1.7)
arctan(y/x) + π x < 0, y ≥ 0
Definition 1.1.9 (arg+ ). Given z ∈ C we define the arg+ (z) as the only value of arg(z) such that
0 ≤ θ < 2π.
In case we have a polydromic function, in order to specify we’re using this argument, there will be a +
as index. √ +
I.e. log+ (z), [z a ]+ , z , · · · and so on.
Theorem 1.2 (De Moivre Formula). A complex number z ∈ C in polar form can be written with
complex exponential and sine and cosine function as follows.
2 2
z = kzk ei arg z = kzk (cos(arg z) + i sin(arg z)) (1.8)
§ 1.2 Regions in C
Definition 1.2.1 (Line). A line λ in C, from z1 , z2 can be written as follows
If t ∈ R this defines the line lying between z1 , z2 . Its non-parametric representation is the following
z − z1
{λ} := z ∈ C| Im =0 (1.10)
z2 − z1
Where z = λ(t).
Definition 1.2.2 (Circumference). A circumference γ centered in a point z0 ∈ C with radius R is defined
as follows
γ(θ) = z0 + Reiθ θ ∈ [0, 2π] (1.11)
Non parametrically, it can be defined as follows
{γ} := { z ∈ C| kz − z0 k = R} (1.12)
Definition 1.2.3 (Extended Complex Plane). We define the extended complex plane Ĉ as follows
Ĉ = C ∪ {∞} (1.13)
This can be imagined by projecting C into the Riemann sphere centered in the origin.
Definition 1.2.4 (Points in Ĉ). Given a point z ∈ C, z = x + iy we can find its coordinates with the
following transformation
ẑ = (xt, yt, 1 − t) ∈ Ĉ (1.14)
Where the condition kẑk = 1 must hold, defining the value of t ∈ R Inversely, given ẑ = (x1 , x2 , x3 ) ∈ Ĉ
one finds
x1 + ix2
z= (1.15)
1 − x3
This gives
kez k = eRe(z)
(1.17)
arg(ez ) = Im(z) + 2πk k∈Z
We have therefore, for z, w ∈ C
ez ew = ez+w
ez (1.18)
= ez−w
ew
CHAPTER 1. COMPLEX NUMBERS AND FUNCTIONS 12
It’s evident how this function has multiple values for the same z value, and therefore is known as a
polydromic function, like the square root. We also define the principal branch of the logarithm as
Log(z)
Log(z) = log kzk + i Arg(z) (1.20)
Lastly we define the log+ (z) as follows
Definition 1.3.3 (Branch of the Logarithm). A general branch of the log function is defined as the
function f (z) : D ⊂ C −→ C such that
ef (z) = z (1.22)
d s
z = se(s−1) log(z) = sz s−1 (1.24)
dz
Alternatively, we define
sz = ez log(s) (1.25)
Definition 1.3.6 (Hyperbolic Functions). We define the hyperbolic functions as follows, given z = iy
sinh(y) = −i sin(iy)
(1.27)
cosh(y) = cos(iy)
1.3. ELEMENTARY FUNCTIONS 13
1. d(x, y) ≥ 0
2. d(x, x) = 0
3. d(x, y) = d(y, x)
Definition 2.1.2 (Ball). Let (X, d) be a metric space. We then define the open ball of radius r,
centered in x in X (BrX ), and the closed ball of radius r centered in x (BrX ) as follows
When there won’t be doubts on on where the ball is defined, the superscript indicating the set of
reference will be omitted.
Definition 2.1.3 (Open Set). Let (X, d) be a metric space, and A ⊆ X a subset. A is said to be an
open set if and only if
∀x ∈ X ∃BrX (x) ⊂ A (2.3)
15
CHAPTER 2. ABSTRACT SPACES 16
Definition 2.1.4 (Complementary Set). Let A be a generic set, then the set Ac is defined as follows
Ac := {a ∈
/ A} (2.4)
Definition 2.1.5 (Closed Set). Alternatively to the notion of open set, we can say that E ⊆ X is a
closed set, if and only if
∀x ∈ E c ∩ X ∃BrX (x) ⊂ E c ∩ X (2.5)
Proof. Let A = BrX (x). If A is open, we have therefore, applying the definition of open set, that
So
x0 ∈ A =⇒ d(x, x0 ) < r
∴ = r − d(x, x0 ) > 0
Then, by definition of open ball we have
Proof. The first two statements are of easy proof. Let BX ⊂ {}. This means that BX is empty and
therefore BX = {}, which makes it open by definition. Therefore we have that {}c = X, and X must
be closed, but if we reason a bit, we can say that ∀x ∈ X BX (x) ⊂ X, which means that X is open,
thus X c = {} must be closed.
Since we gave a proof for {} and X being open, we have that these two sets are both open and closed.
These two sets are said to be clopen.
For the other statements we use the De Morgan laws on set calculus, therefore we have
n
\
x∈ Ai =⇒ x ∈ Ai
i=1
∴ ∃i : BXi (x) ⊂ Ai
And A is open
If we let C = Ac we have that
n
!c n
\ [
c
C=A = Ai = Aci
i=1 i=1
∴ C is closed
For the last two we proceed as follows
x ∈ Aα =⇒ ∃α0 ∈ I : x ∈ Aα0
[
∴ ∃ > 0 : BX (x) ⊂ Aα0 ⊂ =B
α∈I
For the last one, we use the De Morgan laws and the proposition is demonstrated
Definition 2.1.6 (Internal Points, Closure, Border). Let (X, d) be a metric space and A ⊂ X a subset.
We define the following sets from A
1. A◦ = α∈I Gα is the set of internal points of A, where I is an index set and Gα ⊂ A are open
S
T
2. A = β∈J Fβ is the closure of A, where J is another index set and Fβ ⊂ A are closed
c
3. ∂A = A \ A◦ = A ∪ (A◦ ) is the border of A
2. A is closed iff A = A
c
3. A◦ = (A◦ )
◦ c
4. A = (Ac )
CHAPTER 2. ABSTRACT SPACES 18
◦
5. (A ∩ B) = A◦ ∩ B ◦
6. A ∩ B = A ∪ B
Proof. Let O(A) be a collection of open sets, such that ∀G ∈ O(A) =⇒ G ⊂ A, then
[
A = A◦ =⇒ A = G
G∈O(A)
1. x ∈ A ⇐⇒ ∃ > 0 : B (x) ⊂ A
2. x ∈ A ⇐⇒ ∀ > 0 B (x) ∩ A 6= {}
But
x ∈ A◦ =⇒ ∃G ⊂ X open : x ∈ G =⇒ ∃ > 0 : B (x) ⊂ G ⊂ A
∴ A◦ ⊂ I(A) 3 x, I(A) ⊂ A by definition, ∴ I(A) = A◦
∴ ∀ > 0 B (x) ∩ A 6= {}
3 For the last one, we have, taking into account the first two proofs
x ∈ ∂A ⇐⇒ x ∈ A \ A◦ =⇒ x ∈ A ∧ x ∈
/ A◦
1 ∧ 2 =⇒ x ∈ A ⇐⇒ ∀ > 0 B (x) ∩ A 6= {}
/ A◦ ⇐⇒ ∀ > 0 B (x) ∩ Ac 6= {}
∴x∈
2.2. CONVERGENCE AND COMPACTNESS 19
Definition 2.1.7 (Isometry). Let (X, d), (Y, ρ) be two metric spaces and f an application, defined as
follows
f : (X, d) → (Y, d)
f is said to be an isometry iff
f :[0, 1] → [0, 2]
x → f (x) = x
Definition 2.1.8 (Diameter of a Set). Let A be a set and the couple (A, d) be a metric space. We define
the diameter of A as follows
diam (A) = sup (d(x, y))
x,y∈A
Definition 2.2.1 (Convergence). Let (X, d) be a metric space and x ∈ X. A sequence (xk )k≥0 in X is
said to converge in X and it’s indicated as xk → x ∈ X, iff
Theorem 2.1 (Unicity of the Limit). Let (X, d) be a metric space and (xk )k≥0 a sequence in X. If
xk → x ∧ xk → y, then x = y
Definition 2.2.2 (Adherent point). Let (X, d) be a metric space and A ⊂ X. x ∈ X is said to be an
adherent point of A if ∃(xk )k≥0 ∈ A : xk → x ∈ X. The set of all adherent points of A is called
ad(A)
x ∈ A =⇒ ∀ > 0 B (x) ∩ A 6= {}
∴ ∀n ∈ N B n1 (x) ∩ A 6= {} =⇒ ∀n ∈ N ∃xn ∈ B n1 (x)
Proposition 6. Let (X, d) be a metric space and A ⊂ X. Then A is closed iff ∃(xk )k≥0 ∈ A : xk →
x ∈ A =⇒ ad(A) ⊂ A
Definition 2.2.4 (Dense Set). Let (X, d) be a metric space and A, B ⊂ X. A is said to be dense in
B iff B ⊂ A, therefore ∀ > 0 ∃y ∈ A : d(x, y) < . One example for this is Q ⊂ R, with the usual
euclidean distance defined through the modulus.
Definition 2.2.5. Let (X, d) be a metric space and (xk )k≥0 ∈ X. The sequence xk is said to be a
Cauchy sequence iff
∀ > 0 ∃N > 0 : ∀k, n ≥ N d(xk , xn ) <
Proposition 7. Let (X, d) be a metric space and (xk )k≥0 ∈ X a sequence. Then, if xk → x, xk is a
Cauchy sequence
Definition 2.2.6 (Complete Space). Let (X, d) be a metric space. (X, d) is said to be complete iff
∀(xk )k≥0 ∈ X Cauchy sequences, we have xk → x ∈ X
Theorem 2.2 (Completeness). Let (X, d) be a metric space and Y ⊂ X. (Y, d) is complete iff
Y = Y in X
Proof. Let (Y, d) be a complete space, then
∃(ηk ) ∈ Y : ηk → z =⇒ ∃y ∈ Y : ηk → y ∴ z = y =⇒ ad(Y ) ⊂ Y
Theorem 2.3 (Heine-Borel). Let K ⊂ Rn , then K is compact if and only if K is closed and
bounded
2.2. CONVERGENCE AND COMPACTNESS 21
<++>
Theorem 2.4. Let (X, d) be a compact space. Then (X, d) is also complete
Proof. (X, d) is compact, therefore
xk → x =⇒ xnk → x ∈ X
Definition 2.2.8 (Completely Bounded). Let (X, d) be a metric space. X is totally bounded iff
n
[
∃Y ⊂ X : ∀ > 0, ∀x ∈ Y X = B (x)
i=1
A polygonal chain will be indicated as follows Pz,w and it’s defined as follows
n−1
[
Pz,w = [zk , zk+1 ] = [z, z1 , · · · , zn−1 , w]
k=1
∀z, w ∈ G ∃Pz,w ⊂ G
Definition 2.2.11 (Contraction Mapping). Let (X, d) be a complete metric space. Let T : X −→ X. T
is said to be a contraction mapping or contractor if
xn = T (xn−1 ), ∀n ∈ N
Regrouping, we have
m−n−1 ∞
n
X
k n
X
k n 1
d(xm , xn ) ≤ q d(x1 , x0 ) q ≤ q d(x1 , x0 ) q = q d(x1 , x0 )
1−q
k=0 k=0
Then
q n d(x1 , x0 ) (1 − q)
< =⇒ q n < , ∀n > N
1−q d(x1 , x0 )
Therefore, after taking m > n > N , we have
d(xm , xn ) <
Therefore xn is a Cauchy sequence. Since (X, d) is a complete metric space, this sequence must have a
limit xn → x? ∈ X, but, by definition of convergence and limit, we have that by continuity
x? = lim xn = lim T (xn−1 ) = T lim xn−1 = T (x? )
n→∞ n→∞ n→∞
Therefore
∃!x? ∈ X : T (x? ) = x?
And x? is the fixed point of the contractor T
1. u + v ∈ V sum closure
2.3. VECTOR SPACES 23
2. av ∈ V scalar closure
3. u + v = v + u
4. (u + v) + w = u + (v + w)
5. ∃!0 ∈ V : u + 0 = 0 + u = u
6. ∃!v ∈ V : u + v = 0 =⇒ v = −u
7. ∃!1 ∈ V : 1 · u = u
9. (a + b)u = au + bu
10. a(u + v) = au + av
Definition 2.3.2 (Norm). Let V be a vector space over a field F, then the norm is an application defined
as follows
k·k : V −→ F
Where it satisfies the following properties
1. kuk ≥ 0 ∀u ∈ V
2. kuk = 0 ⇐⇒ u = 0
3. kcuk = |c|kuk ∀u ∈ V c ∈ F
Definition 2.3.3 (Normed Vector Space). A normed vector space is defined as a couple (V, k·k),
where V is a vector space over a field F.
Proposition 8. A normed vector space (NVS), is also a metric vector space (MVS) if we define our
distance as follows
d(u, v) = ku − vk ∀u, v ∈ V
Definition 2.3.4 (Vector Subspace). Let V be a vector space and U ⊂ V. U is a vector subspace of V
iff
1. u, v ∈ U =⇒ u + v ∈ U
2. u ∈ U , a ∈ F =⇒ au ∈ U
Proposition 9. If (V, k·k) is an normed vector space and W ⊂ V is a subspace of V, then (W, k·k) is a
normed vector space
CHAPTER 2. ABSTRACT SPACES 24
Definition 2.3.5 (p-norm). Let (V, k·kp ) be a normed vector space. The norm k·kp is said to be a
p-norm if it’s defined as follows
p1
dim(V)
X
kvkp := (vi )p , ∀v ∈ V, ∀p ∈ N? := N ∪ {±∞} (2.8)
i=1
Definition 2.3.6 (Dual Space). Let V be a vector space over the field F, we define a linear functional
as an application ϕ : V −→ F such that ∀u, v ∈ V and c ∈ F
ϕ(u + v) = γ(u) + ϕ(v)
(2.10)
ϕ(λu) = λϕ(u)
Defining the sum of two linear functionals as (ϕ1 + ϕ2 )(v) = ϕ1 (v) + ϕ2 (v) we immediately see that
the set of all linear functionals forms a vector space over V, which will be called the dual space V ? .
Therefore
n n n
X 1X p 1X q
|sk tk | ≤ |sk | + |tk |
p q
k=1 k=1 k=1
p−1
Letting uk = (|xk | + |yk |) we have, after imposing the condition on q of the p-norm as q(p + 1) = p
and using that the sum is Abelian, we have
! q1
Xn Xn
|xk |uk ≤ kxkp kukq = kxkp (|xk | + |yk |)p
k=1 k=1
X n n
! q1
X
p
|yk |uk ≤ kykp kukq = kykp (|xk | + |yk |)
k=1 k=1
In this chapter (and from now on, mostly), we will use a notation which is called abstract index
notation with the Einstein summation convention. This is usually abbreviated in common literature
as the Einstein index notation. We will give here a brief explanation of how this notation actually
works, and why it’s so useful in shortening mathematical expressions. Let V be a vector space and V ?
be the dual space associated with V. Then we can write the elements v ∈ V, ϕ ∈ V ? with respect to
some basis as follows
v1
..
v = (v1 , v2 , · · · , vn ) = .
(3.1)
vn
ϕ = (ϕ1 , ϕ2 , · · · , ϕn ) = ϕ1 ϕ2 · · · ϕn
The first notation is the ordered tuple notation, meanwhile the second notation is the usual column/row
notation for vectors utilized in linear algebra. In Einstein notation we will have that
v −→ v i
(3.2)
ϕ −→ ϕi
Where the vector in the space will be indicated with a raised index (index, not power!) and the
covector with a lower index, where the index will span all the values i = 1, · · · , dim(V).
Let’s represent the scalar product in Einstein notation. Let’s say that we want to write the scalar product
hv, vi
dim(V)
X
hv, vi = vi vi −→ vi v i (3.3)
i=1
Note how we have omitted the sum over the repeated index. Now one might ask why it’s not written
as vi vi (or v i v i , since v ∈ V), and this is easily explained introducing the matrix gij , which is the matrix
of the scalar product.
Applying this matrix to v i we have gij v i . Note how the low index j is free and i is being summed over,
hence is a dummy index, this means that the result must have a lower index j for consistency. So we
can write vj = gij v i , and due to the lower index we already know that this is a covector, i.e. a linear
27
CHAPTER 3. DIFFERENTIAL ANALYSIS 28
functional V −→ F, hence it will “eat” a vector and “spew” a scalar (with no indices!). Feeding to this
covector the vector v j we have finally
Where, algebraically we have “omitted” the definition of ι(·) = hv, ·i, which is the canonical isomor-
phism between V and V ? .
With this definition we have defined what mathematically are called musical isomorphisms, appli-
cations which raise and lower indexes. Ironically, this operation is called index gymnastics, since
we’re raising and lowering indices. Thanks to these conventions operations with matrices (and ten-
sors) become much much easier. Let aij and bij be two n × n matrices over the ordered field F. The
multiplication of these two matrices will simply be
Note how the k index gets “eaten”. This mathematical cannibalism is called contraction of the index
k. So, the trace for a matrix aij will be
tr(a) = aii (3.6)
And now comes the tricky part. In order to write determinants we need to define a symbol, the so
called Levi-Civita symbol, i1 ...in . In three dimensions it’s ijk , and follows the following rules
1
even permutation of the indices
ijk = −1 uneven permutation of the indices (3.7)
0 i=j∨j =k∨k =i
In n dimensions, it becomes
1
even permutation of indices
i1 ...ik = −1 uneven permutation of indices (3.8)
0 ii = ij for some i, j
It’s obvious by definition that this weird entity is completely antysimmetrical and unitary, and there-
fore it’s perfect for representing permutations (it’s also known as permutation symbol for a reason).
Therefore, remembering the definition of the determinant we can write, for an n × n matrix aij
det(a) = i1 ...in a1i1 a2i2 · · · anin = i1 ...in g ji1 a1j g ki2 a2k · · · g lin anin (3.9)
If dim(V) = 3, we can therefore immediately define the cross product of two vectors as follows
c = v × w −→ ci = g ij jkl v k wl (3.10)
det(A) = A = µ1 ...µn g νµ1 g σµ2 · · · g ζµn A1ν A2σ · · · Anζ (3.11)
3.1. DIGRESSION ON THE NOTATION USED 29
In case we have a vector function f µ (xν ), the following notation will be used
q
kf(x)k −→ fµ f µ (xν ) (3.13)
Or q
kg(x) ± f(y)k −→= gµν (g µ (xγ ) ± f µ (y γ ))(g ν (xγ ) ± f ν (y γ ))
A shorthand notation can be created by directly using the norm symbol, but with the contracted index
in the upper or lower position as follows
For p-norms we have to watch out for a little detail. We have to add a square root in order to “fix” the
squaring of every element. So we get
q
p p
p p
p1
kvkp −→ p (vµ ) 2 (v µ ) 2 = (vµ ) 2 (v µ ) 2 (3.16)
1/p
Theorem 3.1. (vµ v µ )p/2 is wrong
Proof. It’s easy to see why it doesn’t work by expanding the sum on µ
p
(vµ v µ )p = v1 v 1 + v2 v 2 + · · · + vn v n
p p (3.17)
(vµ ) 2 (v µ ) 2 = (v1 v 1 )p + (v2 v 2 )p + · · · + (vn v n )p
Moreover, it’s time to bring down some formal rules for the usage of this notation
Theorem 3.2 (Rules for Index Calculus in Einstein Notation). 1. Free indices must be consistent
in both sides of the equation. I.e. aµν bµγδ = cνγδ . aµν bµγδ 6= cγνδ , aµν bµγδ 6= cνγσ
2. An index can be repeated only two times per factor and must be contracted diagonally. I.e.
aµ bµ fγδ = cδγ is defined correctly, aµ bµ , aµ bµ or aµ bµ fγµ are ill defined
3. Dummy indices can be replaced at will, since they don’t contribute to the “index
equation”
CHAPTER 3. DIFFERENTIAL ANALYSIS 30
2 ∂2f
Hf −→ ∂ν ∂µ f = ∂µν f = ∂µν f = = f,µν (3.19)
∂xµ ∂xν
Derivatives of order > 2 can then be defined recursively.
Now we might ask, what if we have a vector field F µ ? Nothing changes. We simply have to remember
to not repeat indices in order not to represent scalar products.
We have JF as the Jacobian matrix of F µ , basically the derivative matrix which in Einstein notation, as
before, has a quite obvious nature
∂F µ
JF −→ ∂ν F µ = = F,νµ (3.20)
∂xν
And so on, and so on…1
Let’s now define the divergence and curl operators. Take now a vector field g µ . We then have
∂g µ
∇ · g −→ gµν g µδ ∂δ g ν = ∂µ g µ = µ
= g,µ
∂xµ
(3.21)
∂g σ
∇ × g −→ µνσ g νδ ∂δ g σ = µνσ ∂ ν g σ = µνσ = µνσ g σ,ν
∂xµ
1 It’s quite fun to dive into the dumpster of Einstein notation, isn’t it?
3.1. DIGRESSION ON THE NOTATION USED 31
And therefore, defining the Laplacian as ∇2 = ∇ · ∇, we will simply have, for whatever function h
∂2h
∇2 h −→ g µν ∂ν ∂µ h = ∂ µ ∂µ h = = h,µ,µ (3.22)
∂xµ ∂xµ
Note how the operator ∂ µ appears. This can be seen as a derivation along the covector basis (xµ =
gµν xν ).
Now, we can go back to our mathematical rigor.
CHAPTER 3. DIFFERENTIAL ANALYSIS 32
§ 3.2 Curves in Rn
Definition 3.2.1 (Scalar Field). A scalar field is a function f : A ⊆ Rn −→ R where A is an open set
Definition 3.2.2 (Vector Field). A vector field is a function f µ : A ⊂ Rn −→ Rm where A is an open
set
Definition 3.2.3 (Continuity). A scalar field f : A −→ R is said to be continuous in a point pµ ∈ A if
If this function is continuous ∀pµ ∈ A, then the vector field is said to be part of the space C(A), with
A ⊆ Rn
Definition 3.2.4 (Canonical Scalar Product). Let xµ , y µ ∈ Rn , the canonical scalar product is a
bilinear application h·, ·i : Rn × Rn −→ R where, if the components of the two vectors are defined as
xµ , y µ , is defined as
n
X
hx, yi = xi y i −→ xµ y µ (3.25)
i=1
It’s easy to see that the canonical scalar product induces the euclidean norm as follows
p p
kvk = kvk2 = hv, vi = vµ v µ (3.26)
A curve is said to be piecewise regular if it’s not regular in [a, b] but it’s regular in a finite number of
subsets [an , bn ] ⊂ [a, b]
Definition 3.2.7 (Homotopy of Curves). Let γ µ , η µ be two curves from a set [a, b], [c, d] respectively.
These two curves are said to be homotopic to one another, and it’s indicated as γ µ ∼ η µ iff
∼
∃h : [c, d] −→ [a, b], h ∈ C([c, d]), h−1 ∈ C([a, b]), h(s) > h(t) for s > t : η µ = γ µ ◦ h (3.28)
3.2. CURVES IN RN 33
Definition 3.2.8 (Tangent Vector). The tangent vector of a regular curve is defined as the following
vector.
γ̇ µ
T µ (t) = p (3.29)
(γ˙µ γ̇ µ )
Where with γ̇ µ (t) we indicate the derivative of γ µ with respect to the only variable t.
Definition 3.2.9 (Tangent Line). A curve γ µ : [a, b] −→ Rn is said to have tangent line at a point
t0 ∈ [a, b] if it’s regular, therefore the line will have parametric equations
Definition 3.2.10 (Length of a Curve). The length of a curve γ : [a, b] −→ Rn is defined as follows
ˆ b q
Lγ := γ̇µ (t)γ̇ µ (t) dt (3.31)
a
The graph of a function f : [a, b] −→ R, f ∈ C 1 (a, b) can also be parametrized from a curve ϕµ (t),
where
ϕµ (t) → (t, f (t)) (3.34)
Its length will be then calculated with the following integral
ˆ b q
Lϕ := 1 + (ḟ (x))2 dx (3.35)
a
Theorem 3.3 (Length Invariance under Homotopy of Curves). Let {γ1 }, {γ2 } ⊂ Rn be two curves,
such that γ1 ∼ γ2 , then
Lγ1 = Lγ2
Proof. By definition we have that since the two curves are homotopic, the two domains of definition
of both are diffeomorphic to each other, and therefore
∼
Where ϕ : I −→ J is our C 1 diffeomorphism. Therefore
ˆ ˆ ˆ
Lγ1 = kγ̇1µ kµ dt = ˙
γ2µ (ϕ(t)) dt = k(γ2µ )0 kµ |ϕ̇(t)|dt
I I µ I
Definition 3.2.11 (Curviline Coordinate). Let ϕ : [a, b] −→ Rn , we can define a function s(t) as follows
ˆ tq
s(t) = ϕ̇µ (τ )ϕ̇µ (τ ) dτ (3.37)
a
Then q
ds = ϕ̇µ (t)ϕ̇µ (t) dt (3.38)
And the length of a curve can also be indicated as follows
ˆ
Lϕ = ds (3.39)
ϕ
Definition 3.2.12 (Curvature, Normal Vector). The curvature of a curve is defined as follows
q q
κ(s) = Tµ (s)T µ (s) = ϕ̈µ (s)ϕ̈µ (s) (3.40)
Ṫ µ (s)
N µ (s) = (3.41)
κ(s)
Definition 3.2.13 (Simple Curve, Closed Curve). A simple curve is an injective application γ : [a, b] −→
Rn . A curve is said to be closed iff γ µ (a) = γ µ (b)
Theorem 3.4 (Jordan Curve). Let γ µ be a simple and closed curve in R2 or C (note that C ' R2 ),
then the set {γ}c is defined as follows
{γ}c = {γ}◦ ∪ extr ({γ}) (3.42)
Note that {γ} ⊂ R2 is the image of the application γ and extr ({γ}) is the set of points that
lay outside of the closed curve.
In C everything that was said about curves holds, however one must watch out for the definition of
modulus, for a curve γ µ ∈ C we will have
q
|γ̇(t)| = (Re0 (γ))2 + (Im0 (γ))2 = γ(t)γ(t)
p
(3.43)
3.3. DIFFERENTIABILITY IN RN 35
§ 3.3 Differentiability in Rn
Definition 3.3.1 (Directional Derivative). Let A ⊆ Rn be an open set, and f : A −→ R. The function is
said to be derivable with respect to the direction v µ ∈ A at a point pµ ∈ A, if the following limit is
finite
f (pµ + hv µ ) − f (pµ )
∂vµ f (pν ) = lim (3.44)
h→0 h
If v µ = xµ then this is called a partial derivative, and it will be indicated in the following ways
∂f
= ∂µ f = ∂xµ f (3.45)
∂xµ
1. f ∈ C(A)
5. There exist a tangent hyperplane to the graphic of the function at the point (pµ , f (pµ )) ∈
Rn+1 and has the following equation
3. ∂vµ f = ∂µ f v µ
Theorem 3.6 (Total Differential). Let f : A ⊂ Rn −→ R, with A open. If f ∈ C 1 (A) (i.e. the
derivatives of f are continuous), then f is differentiable in A, the vice versa is also true
Therefore
n
|f (pµ + hµ ) − f (aµ ) − ∂µ f hµ | X
µ hi
lim p ≤ lim |∂i f (c i ) − ∂i f (p )| p =0 (3.51)
hµ hµ h→0 hµ hµ
p
hµ hµ →0
i=1
Since A is open, both A1 , A2 must be open. Define a path γ from xµ to xµ0 and define the composite
function
ϕ(t) = f (xµ0 + t(xµ − xµ0 )) , t ∈ [0, 1]
Since f is differentiable and ∂µ f = 0 by definition ϕ is constant and therefore continuous, which
confirms that A1 is open.
Since A1 ∩ A2
Theorem 3.8 (Differentiability of Vector Fields, Jacobian Matrix). Let f µ : A ⊆ Rn −→ Rm be a vector
field and A an open set, then the function f µ is differentiable iff exists a matrix Jνµ ∈ Mnm (R)
such that
kf µ (pν + hν ) − f µ (pν ) − Jνµ hν kµ
p lim p (3.52)
hµ hµ →0 hµ hµ
Or, equivalently p
fµ (pν + hν ) = f µ (pν ) + Jνµ hν + O hµ hµ (3.53)
The then Jνµ is the matrix of partial derivatives of the vector field, called the Jacobian matrix
of the vector field f µ , and can be calculated as follows
g ν (f µ (pσ + sσ )) = g ν (f µ (pσ )) + ∂σ g ν ◦ f µ̃ ∂σ f µ +
+ ∂σ (g ν ◦ f µ ) O sµ sµ + O kf µ (pσ + sσ ) − f µ (pσ )kµ
p
But √
∂σ (g ν ◦ f µ ) O( sµ sµ )
√ µ →0
sµ s
And
kf µ (pσ + sσ ) − f µ (pσ )kµ ≤
p p p p
∂ σ fµ ∂σ f µ sµ sµ + O sµ sµ ≤ C sµ sµ
Therefore
O kf µ (pσ + sσ ) − f µ (pσ )kµ O kf µ (pσ + sσ ) − f µ (pσ )kµ kf µ (pσ + sσ ) − f µ (pσ )k
µ
√ µ = √ µ →0
sµ s kf µ (pσ + sσ ) − f µ (pσ )kµ sµ s
§ 3.4 Differentiability in C
Definition 3.4.1 (Differentiability). A function f : G ⊂ C −→ C with G open, is said to be differen-
tiable or derivable at a point a ∈ G if exists finite the following limit
df f (z) − f (a)
= f 0 (a) = lim (3.56)
dz a
z→a z−a
Proof.
f (z) − f (a)
lim (f (z) − f (a)) = lim (z − a) lim =0
z→a z→a z→a z−a
4. f (z) = c =⇒ f 0 (z) = 0
Proof. Since G is open, ∃Br (a) ⊂ G. Therefore, taking a sequence (z)n ∈ Br (a) : limn→∞ (z)n = a.
Letting f (zn ) 6= a, we can directly write in the definition of derivative
(g ◦ f )(zn ) − (g ◦ f )(a)
lim = (g ◦ f )0 (a) = g 0 (f (a))f 0 (a)
n→∞ zn − a
Thus, rewriting the function inside the limit
(g ◦ f )(zn ) − (g ◦ f )(a) (g ◦ f )(zn ) − (g ◦ f )(a) f (zn ) − f (a)
= →0
zn − a f (zn ) − f (a) zn − a
Since f is continuous in a ∈ G
∼
Theorem 3.13 (Inverse Function Derivation). Let f : G ⊂ C −→ C be a bijective continuous map,
with f −1 (w) = z. If f (a) 6= 0 and it’s derivable at that same point, we have
df −1 1
= (3.58)
dw f (a) f 0 (a)
Definition 3.4.3 (Wirtinger Derivatives). Before demonstrating the previous theorem, we define the
Wirtinger derivatives as follows.
Let z ∈ C, z = x + iy and f : G ⊂ C −→ C.
∂f 1 ∂ ∂
∂z = ∂f (z) = − i f (z)
2 ∂x ∂y
(3.63)
∂f 1 ∂ ∂
= ∂f (z) = +i f (z)
∂z 2 ∂x ∂y
Then, the Cauchy-Riemann equations will be equivalent to the following equation
∂f
= ∂f (z) = 0 (3.64)
∂z
And, therefore, along the imaginary axis and the real axis, we have
f (z0 + Re(h)) − f (z0 ) ∂f
lim =
Re(h)→0 Re(h) ∂x z0
f (z0 + iIm(h)) − f (z0 ) 1 ∂f ∂f
lim = = −i
Im(h)→0 iIm(h) i ∂y z0 ∂y z0
Due to the continuity of the derivative (f ∈ H(B (z0 ))) we must have an equality between these limits
∂f ∂f ∂f ∂f
+i =2 = 0, ∴ f ∈ H(B (z0 )) =⇒ =0
∂x ∂y ∂z ∂z
But, since f (z) = u(x, y) + iv(x, y), we will have that
∂f ∂ 1 ∂ ∂
= (u(x, y) + iv(x, y)) = +i (u(x, y) + iv(x, y))
∂z ∂z 2 ∂x ∂y
1 ∂u ∂u ∂v ∂v
= +i +i − =0
2 ∂x ∂y ∂x ∂y
∂u ∂v ∂v ∂u
∴ − =i +i
∂x ∂y ∂x ∂y
Rewriting the previous equation in a system, we immediately get back the Cauchy-Riemann equations
∂u ∂v
− =0
∂x ∂x
∂v ∂u
+ =0
∂x ∂y
3.5. SURFACES 41
Is written as
f (t) = Re(f (t)) + iIm(f (t)) ∈ C
§ 3.5 Surfaces
Definition 3.5.1 (Regular Surface). Let K ⊂ R2 , K = E where E is an open and connected subset. A
regular surface in R3 is an application
rµ : K −→ R3
Such that
1. rµ ∈ C 1 (K), i.e. ∃∂ν rµ ∈ C(K)
2. rµ is injective in K
3. rank (∂ν rµ ) = 2
The image Im(rµ ) = Σ ⊂ R2 is then defined by the following parametric equations
1
x(u, v) = r (u, v)
rµ (u, v) = y(u, v) = r2 (u, v) (3.65)
z(u, v) = r3 (u, v)
Remark. A function f ∈ C 1 (K) defines automatically a surface with parametric equations rµ (u, v) =
(u, v, f (u, v)). This surface is always regular since µνσ ∂1 rν ∂2 rσ = (−2u, −2v, 1) 6= 0 ∀(u, v) ∈ K
Definition 3.5.2 (Coordinate Lines). The curves obtained fixing one of the two variables are called
coordinate lines in the surface Σ. We have therefore, for a parametric surface rµ (u, v) and two fixed
values ũ, ṽ ∈ I ⊂ R
xµ1 (t) = rµ (t, ṽ)
(3.67)
xµ2 (t) = rµ (ũ, t)
CHAPTER 3. DIFFERENTIAL ANALYSIS 42
Example 3.5.1. The sphere centered in a point pµ0 ∈ R3 , pµ0 = (x0 , y0 , z0 ) with radius R ≥ 0 has the
following parametric equations
x = x0 + R sin(u) cos(v)
y = y0 + R sin(u) sin(v) (3.68)
z = z0 + R cos(v)
With (u, v) ∈ [0, π] × [0, 2π]. It’s a regular surface, since
kµνσ ∂1 rν ∂2 rσ kµ = R2 sin(u) > 0 ∀(u, v) ∈ [0, π] × [0, 2π]
The regular curve pµ (t) = rµ (u(t), v(t)) has Im pµ ⊂ Σ. If it passes for a point pµ0 = (u0 , v0 ) it has
tangent line
pµ (t) = pµ0 + ṙµ (t)(t − t0 ) = pµ0 + ∂u rµ (u(t), v(t))u̇(t) + ∂v rµ (u(t), v(t))v̇(t) (3.70)
The line is contained inside the following plane
(x − x0 ) (y − y0 ) (z − z0 )
det ∂1 r1 ∂1 r2 ∂3 r 1 (3.71)
1 2
∂2 r ∂2 r ∂3 r 3
For a cartesian surface, i.e. the surface generated from the graph of a function f (x, y), the tangent
plane will be
z = f (xµ0 ) + ∂µ f (xν0 )(xµ − xµ0 ) (3.72)
Definition 3.5.4 (Normal Vector). The normal vector to a surface Σ, nµ (u, v) is the vector
1
nµ (u, v) = q µνσ ∂u rν ∂v rσ (3.73)
µνσ µδγ ∂1 rν ∂2 rσ ∂1 rδ ∂v rγ
Definition 3.5.5 (Implicit Surface). Let F : A ⊂ R3 −→ R be a function such that F ∈ C 1 (A), letting
Σ := { xµ ∈ R3 F (xµ ) = 0}. If xν0 ∈ Σ and ∂µ F (xν0 ) 6= 0, Σ coincides locally to a cartesian surface,
and the equation of the tangent plane at the point xν0 is the following
∂µ F (xν0 )(xµ − xµ0 ) = 0 (3.75)
3.6. OPTIMIZATION 43
Definition 3.5.6 (Metric Tensor). Let ds be the curviline coordinate of some curve γ µ inside a regular
surface Σ. Then we have that
sµ (t) = rµ (u(t), v(t)) (3.76)
And therefore
2 2
ds2 = drµ drµ = ∂1 rµ ∂1 rµ dx1 + 2∂1 rµ ∂2 rµ dx1 dx2 + ∂2 rµ ∂2 rµ dx2 (3.77)
gµν = ∂µ rσ ∂ν rσ (3.79)
And it’s called the first fundamental quadratic form in the language of differential geometry. Then,
we can write
2 2
ds2 = E dx1 + 2F dx1 dx2 + G dx2 (3.82)
§ 3.6 Optimization
∂µ f (pν ) = 0 (3.83)
The point pν satisfying this condition is then called a stationary point or a critical point for
the function f
Proof. Let v µ ∈ A be a direction. The function g(t) = f (pµ + tv µ ) has a point of local maximal or
minimal for t = 0. Then
Definition 3.6.1 (Hessian Matrix). Let f : A ⊆ Rn −→ R, and let f ∈ C 1 (A), then we define the
Hessian matrix as the matrix of the second partial derivatives of the function f
∂11 f · · · ∂1n f
∂µ ∂ν f (xγ ) = ∂µν f (xγ ) ... .. .. (xγ ) (3.85)
. .
∂n1 f ··· ∂nn f µν
Definition 3.6.2 (Nature of Critical Points). Let pγ be a critical point for a function f ∈ C 1 (A).
Then
1. ∂µν f (pγ ) is definite positive, then pγ is a local minimum
2. ∂µν f (pγ ) is definite negative, then pγ is a local maximum
3. ∂µν f (pγ ) is indefinite, then pγ is a saddle point
Theorem 3.17. Here is a list of some rules in order to determine the definition of the matrix
∂µν f .
Let v µ ∈ A be a direction, and pγ ∈ A a critial point of the function f : A ⊂ Rn −→ R then
1. If ∂µν f (pγ )v µ v ν > 0 ∀v µ ∈ A =⇒ ∂µν f (pγ ) positive definite
2. If ∂µν f (pγ )v µ v ν < 0 ∀v µ ∈ A =⇒ ∂µν f (pγ ) negative definite
3. If ∂µν f (pγ )v µ v ν ≥ 0 ∀v µ ∈ A =⇒ ∂µν f (pγ ) semi-positive definite
4. If ∂µν f (pγ )v µ v ν ≤ 0 ∀v µ ∈ A =⇒ ∂µν f (pγ ) semi-negative definite
5. If v µ 6= wµ are two directions, and ∂µν f (pγ )v µ v ν > 0 ∧ ∂µν f (pγ )wµ wν < 0 =⇒ ∂µν f (pγ )
indefinite
Theorem 3.18 (Sylvester’s Criteria). Let Aµν ∈ Mnn (R), and (Ak )µν be the reduced matrix with
order k ≤ n, then
1. detµν ((Ak )µν ) > 0 =⇒ Aµν positive definite
2. (−1)k detµν ((Ak )µν ) > 0 =⇒ Aµν negative definite
3. If detµν ((A2k )µν ) < 0 or if detµν ((A2k+1 )µν ) < 0 ∧ detµν ((A2n+1 )µν ) > 0 for k 6= n, then Aµν is
indefinite
Theorem 3.19 (Compact Weierstrass). Let f : K ⊆ Rn −→ R, f ∈ C(K), with K a compact set,
then
∃pµ , q µ ∈ K : min(f ) = f (pµ ) ≤ f (xµ ) ≤ max(f ) = f (q µ ) ∀xµ ∈ K (3.87)
K K
3.6. OPTIMIZATION 45
Proof. Being K a compact set, we have that every sequence (pµ )n converges inside the set, therefore,
letting (pµ )n being a minimizing sequence for f . Then there exist a converging subsequence (pµ )nk
such that
f (pµnk ) → f (pµ )
But, since (pµ )n is a minimizing sequence, we have
f (pµ ) = min(f )
K
By definition of minimizing sequence. Analogously, one can define a maximizing sequence and obtain
the same result for the maximum of the function in K
Theorem 3.20 (Heine-Cantor). Given f : K → R with K a compact set, then if f ∈ C(K), f is
uniformly continuous
<++>
Theorem 3.21 (Existence of Intermediate Values). Let f : A ⊂ Rn −→ R, f ∈ C(A), then, given
u ∈ R such that
min(F ) ≤ u ≤ max(F )
A A
Then
∃cµ ∈ A : f (cµ ) = u
Theorem 3.22 (Squeeze). Let g(xµ ), f (xµ ), h(xµ ) be three functions such that g ≤ f ≤ h for
xµ ∈ Bδ (xµ0 ), then
Proof. By definition of limit we can say, supposing that the affirmation is true, that
Since we’re working in Rn with the standard topology, the last condition on the right tells us that
we’re working in Bδ (xµ0 ), for which g ≤ f ≤ h for some δ3 > 0.
Taken δ = min {δ1 , δ2 , δ3 } then
Therefore
|f (xµ ) − L| ≤
Which proves the theorem
<++>
CHAPTER 3. DIFFERENTIAL ANALYSIS 46
Then
∃xµ ∈ L : min(f ) = f (xµ ) (3.89)
L
Proof. µ
p Let (p )n be a minimizing sequence for f in L. If this sequence wasn’t limited, we would have
that (pµ )n (pµ )n → ∞, and therefore
Therefore (pµ )n must be limited, and the proof is the same as in the case of a compact set.
Theorem 3.24 (Topology and Functions). Let f : Rn −→ R, f ∈ C(Rn ). Then
{ xµ ∈ Rn | f (xµ ) < a ∈ R}
(3.90)
{ xµ ∈ Rn | f (xµ ) > b ∈ R}
{ xµ ∈ Rn | f (xµ ) ≤ a ∈ R}
(3.91)
{ xµ ∈ Rn | f (xµ ) ≥ b ∈ R}
Definition 3.6.4 (Matrix Infinite Norm). Let Aµν (xγ ) ∈ V −→ Mmn (F), where dim(V) = n. We can
define a norm for this space as follows
√
r
(µ)
kAµν k∞ = m max sup Aν Aν(µ) (xγ ) (3.95)
µ xγ ∈V
Theorem 3.26 (Average Value). Let f µ : A ⊆ Rn −→ Rm , with f ∈ C 1 (A), A an open set and
K ⊂ A a compact convex subset, then
Proof. Let rν (t) = (1 − t)y ν + txν be a smooth parametrization of a segment connecting the two
points xν , y ν , then
2 2
kf µ (rν (1) − f µ (rν (0)kµ ≤ ∂ ν fµ ∂ν f µ (rν (t)) ≤ sup(∂ ν fµ ∂ν f µ (rγ ))kxν − y ν kν
rγ
2
≤ m max sup ∂ ν f(µ) ∂ν f (µ) (xγ ) kxν − y ν kν
µ γ
Therefore
√ r
kf µ (xν ) − f µ (y ν )kµ ≤ m max sup ∂ ν f(µ) ∂ν f (µ) kxν − y ν kν =
µ γ
µ ν ν ν ν
= k∂ν f k∞ kx − y kν ∀x , y ∈ K
−1
Proof. Let Bµγ = ∂y0γ f µ , then we know that
We have therefore
Gγ (xν , g σ (xν )) = g γ (xν ) − Bµγ f µ (xν , g σ (xν )) = g γ (xν ) ∀xν ∈ B r (xν0 ) = I
∂Gγ ∂f µ
σ
= δσγ − Bµγ σ
∂y ∂y
∂Gγ γ γ ∂f µ
= δ σ − B µ = δσγ − δσγ = 0
∂y0σ ∂y σ
Now take (X, d) = (C(I, J), k·k∞ ), with J = B (y0γ ), and define an application H : X −→ X such
that
H γ (wσ (xν )) = Gγ (xν , wσ (xν ))
We need to demonstrate that this application is a contraction, i.e. that ∃!g γ (xν ) : f µ (xν , g γ (xν )) =
0 ∀(xν , y γ ) ∈ I × J
Definition 3.6.6 (Argmax, Argmin). Let f µ : A −→ Rn a function which reach its maximum in
xνi ∈ A i = 1, · · · , m and its minimum at yjν ∈ A j = 1, · · · , k Then we can define
Proof. Let ∂n g 6= 0, then we can see M as a graph of a regular implicit function of g, h : Rn−1 −→ R,
where
g(xµ , h(xµ )) = 0 ∀xµ ∈ Br (x˜µ 0 ) ⊂ Rn−1
Letting ϕ : (−, ) −→ Br (xµ0 ) a smooth curve, such that ϕµ (0) = xµ0 , we have that ψ ν (t) =
(ϕµ (t), h(t)) ∈ M is the parameterization of a smooth curve that passes through xµ0 ∈ M. We
have
d
f (ψ ν (0)) = ∂µ f φ̇µ (0) + ∂n f ḣ(φµ (0)) = ∂ν f (xµ0 )sν
dt
d
g(ψ ν (0)) = ∂ν g(xµ0 )sν
dt
With sν = ψ˙ν (0), therefore ∂ν f kψν (0) ∂ν g
Theorem 3.29 (Generalized Lagrange Multiplier Method). Let f, gi : A ⊆ Rn −→ R, 0 < i < n,
f, gi ∈ C 1 (A) with A an open set, let M := { xν ∈ A| g(xν ) = 0}. Take xν0 ∈ M such that
rank ∂ν g µ (xγ0 ) = k
CHAPTER 3. DIFFERENTIAL ANALYSIS 50
Then xν0 is a critical point for f |M , and it’s a free critical point for the Lagrangian L
∂µ f (xγ0 )
rank(A) = =k
∂ν g µ (xγ )
4 Tensors and Differential Forms
Definition 4.1.3 (Dual Space). We define the dual space of a real vector space V as the space of all
linear functionals from the space to the field over it’s defined, and it’s indicated with V ? . I.e. let
ϕµ ∈ V ? , then ϕµ : V −→ R.
It’s easy to see how V ? = T 1 (V).
51
CHAPTER 4. TENSORS AND DIFFERENTIAL FORMS 52
Theorem 4.1. Let B = {vµ1 , · · · , vµn } be a basis for the space V, and let B ? := {ϕµ1 , · · · , ϕµn } be
the basis of the dual space, i.e. ϕµ vν = δνµ ∀ϕµ ∈ B ? , vµ ∈ B, then the set of all k-fold tensor
products has basis BT , where
Definition 4.1.4 (Alternating Tensor). Let V be a real vector space, and ω ∈ T k (V). ω is said to be
alternating if
Or, compactly
ωµ...ν...γ...σ = −ωµ...γ...ν...σ
(4.10)
ωµ...ν...ν...γ = 0
The space of all alternating k−tensors on V is indicated as Λk (V), and we obviously have that Λk (V) ⊂
T k (V).
We can define an application Alt : T k (V) −→ Λk (V) as follows
1 X
Alt(T )(v1µ , · · · , vkµ ) = µ
sgn(σ)T (vσ(1) µ
, · · · , vσ(k) ) (4.11)
k!
σ∈Σk
With σ = (i, j) a permutation and Σk the set of all permutations of natural numbers 1, · · · , k Compactly,
we define an operation on the indices, indicated in square brackets, called the antisymmetrization of
the indices inside the brackets.
This definition is much more general, since it lets us define a partially antisymmetric tensor, i.e. anti-
symmetric on only some indices.
1
Alt(Tµ1 ...µk ) = T[µ1 ...µk ] (4.12)
k!
4.1. TENSORS AND K-FORMS 53
(k + l)!
ωµ1 ...µk ∧ ην1 ...νk = ω[µ1 ...µk ην1 ...νl ] (4.15)
k!l!
With the following properties
∀ω, ω1 , ω2 ∈ Λk (V), ∀η, η1 , η2 ∈ Λl (V), ∀a ∈ R, ∀f ? ∈ L : T k (V) −→ T l (V) ∀θ ∈ Λm (V)
(ω1 + ω2 ) ∧ η = ω1 ∧ η + ω2 ∧ η
ω ∧ (η1 + η2 ) = ω ∧ η1 + ω ∧ η2
(ω ∧ η) ∧ θ = ω ∧ (η ∧ θ)
(4.16)
aω ∧ η = ω ∧ aη = a(ω ∧ η)
ω ∧ η = (−1)kl η ∧ ω
f ? (ω ∧ η) = f ? (ω) ∧ f ? (η)
CHAPTER 4. TENSORS AND DIFFERENTIAL FORMS 54
Where dim(V) = n.
Therefore, dim(Λn (V)) = 1
Theorem 4.6. Let vµ1 , · · · , vµn be a basis for V, and take ω ∈ Λn (V), then, if wµ = aνµ vν
ωµ1 ...µn tµ1 · · · tµn = det(aµν )ων1 ...νn tν1 · · · tνn (4.19)
µν
[vµ ]
−[wµ ]
Definition 4.1.7 (Volume Element). Take a vector space V such that dim(V) = n and it’s equipped
with an inner product g, such that there are two bases (v µ1 , · · · , v µn ), (wµ1 , · · · , wµn ) that satisfy the
orthonormality condition with respect to this scalar product
Then
ωµ1 ...µn v µ1 · · · v µn = ωµ1 ...µn wµ1 · · · wµn = det(aµν ) = ±1
µν
Where
wµ = aµν v ν
Therefore
∃!ω ∈ Λn (V) : ∃![wµ1 , · · · , wµn ] = O
Where O is the orientation of the vector space.
Definition 4.1.8 (Cross Product). Let v1µ , · · · , vnµ ∈ Rn+1 and define ϕν wν as follows
µ
v 1
..
ϕν wν = det .
µ
v n
wν
z µ = v ν1 × · · · × v νn = µν1 ...νn v ν1 · · · v νn
Remark. If a vector v µ ∈ Rn can be seen as an arrow from 0 to the point v, a vector (p, v µ ) ∈ Tp Rn
can be seen as an arrow from the point p to the point p + v. In concordance with the usual notation
for vectors in physics, we will write (p, v µ ) = v µ directly, or vpµ when necessary to specify that we’re
referring to the vector v µ ∈ Tp Rn . The point p + v is called the end point of the vector vpµ .
Definition 4.2.2 (Inner Product in Tp Rn ). The usual inner product of two vectors vpµ , wpµ ∈ Tp Rn is
defined as follows
h·, ·ip :Tp Rn × Tp Rn −→ R
(4.21)
vpµ wµp = v µ wµ = k
Analogously, one can define the usual orientation of Tp Rn as follows
[(eµ1 )p , · · · , (eµn )p ]
CHAPTER 4. TENSORS AND DIFFERENTIAL FORMS 56
Definition 4.2.3 (Vector Fields, Again). Although we already stated a definition for a vector field, we’re
gonna now state the actual precise definition of vector field
Let p ∈ Rn be a point, then a function f µ (p) : Rn −→ Tp Rn is called a vector field, if ∀p ∈ A ⊆ Rn we
can define
f µ (p) = f µ (p)(eµ )p (4.22)
Where (eµ )p is the canonical basis of Tp Rn
All the previous (and already stated) considerations on vector fields hold with this definition.
Definition 4.2.4 (Differential Form). Analogously to vector fields, one can define k−forms on the
tangent space. These are called differential (k-)forms and “live” on the space Λk (Tp Rn ).
?
Let ϕµp 1 , · · · , ϕµp k ∈ (Tp Rn ) be a basis on such space, then the differential form ω ∈ Λk (Tp Rn ) is
defined as follows
X
ωµ1 ...µk (p) = ωµ1 ...µk ϕ[µ1 µk ]
p · · · ϕp → ωi1 ...ik (p)ϕi1 (p) ∧ · · · ∧ ϕik (p) (4.23)
i1 <...<ik
Basically, the vectors dxµ are the dual basis with respect to the canonical basis (eµ )p
Theorem 4.7. Since df (vpν ) = ∂ν f (v ν ) we have, expressing the differential of a function with
the basis vectors,
∂f
df = dxµ = ∂µ f dxµ (4.27)
∂xµ
Definition 4.2.6. Having defined a smooth linear transformation fνµ : Rn −→ Rm , it induces another
linear transformation ∂γ fνµ : Rn −→ Rm , which with some modifications becomes the application
(f? )µν : Tp Rn −→ Tf (p) Rm defined such that
µ
(f? )µν (v ν ) = df |f (p) (v ν ) (4.28)
ν
4.2. TANGENT SPACE AND DIFFERENTIAL FORMS 57
Which, in turn, also induces a linear transformation f ? : Λk (Tf (p) Rm ) −→ Λk (Tp Rn ), defined as
follows. Let ωp ∈ Λk (Rm ), then we can define f ? ω ∈ Λk (Tf (p) Rn ) as follows
(Just remember that in this way we are writing explicitly the chosen base, watch out for the indexes!)
2. f ? (ω1 + ω2 ) = f ? ω1 + f ? ω2
3. f ? (gω) = (g ◦ f )f ? ω
4. f ? (ω ∧ η) = f ? ω ∧ f ? η
d
Definition 4.2.7 (Exterior Derivative). We define the operator das an operator Λk (Tp V) −→ Λk+1 (Tp V)
for some vector space V. For a differential form ω it’s defined as follows
4. f ? (dω) = d(f ? ω)
Definition 4.2.8 (Closed and Exact Forms). A form ω is called closed iff
dω = 0 (4.32)
Proof. The proof is quite straightforward. Since ω is exact we can write ω = dρ for some differential
form ρ, therefore
2
dω = ddρ = d ρ = 0
Hence dω = 0 and ω is closed.
ωµ = p dx + q dy (4.34)
The external derivative will be of easy calculus by remembering the mnemonic rule d → ∂µ ∧ dxµ , or
also as ∂[ν then we have
dωµν = ∂[ν ωµ]
But
∂1 ω1 ∂1 ω2
∂ν ωµ =
∂2 ω1 ∂2 ω2 µν
And
1 1
∂[ν ωµ] = (∂ν ωµ − ∂µ ων ) = (∂ω − ∂ω T )
2 2
Therefore
1 0 ∂x q − ∂y p
dωµν =
2 ∂y p − ∂x q 0 µν
2 2
Which, expressed in terms of the basis vectors of Λ (R ), dx ∧ dy, we get
1 1
dω = (∂x q − ∂y p) dx ∧ dy + (∂y p − ∂x q) dy ∧ dx = (∂x q − ∂y p) dx ∧ dy (4.35)
2 2
Therefore
dω = 0 ⇐⇒ ∂x q − ∂y p = 0 (4.36)
Definition 4.2.9 (Star Shaped Set). A set A is said to be star shaped with respect to a point a iff
∀x ∈ A the line segment [a, x] ⊂ A
Lemma 4.2.1 (Poincaré’s). Let A ⊂ Rn be an open star shaped set, with respect to 0. Then every closed
form on A is exact
Definition 4.3.2 (Standard n−cube). We define a standard n-cube as a function I n : [0, 1]n −→ Rn
such that I n (xµ ) = xµ .
4.3. CHAIN COMPLEXES AND MANIFOLDS 59
Definition 4.3.3 (Face). Given a standard n−cube I n we define the (i, α)−face of the cube as
n
I(i,α) = (x1 , · · · , xi−1 , α, xi , · · · , xn−1 ) α = 0, 1 (4.37)
Definition 4.3.5 (Boundary). Given an n−cube ci we define the boundary as ∂ci . For a standard
n−cube we have
n X
X
∂I n = (−1)i+α I(i,α)
n
(4.39)
i=1 α=0,1
Where ∂s is a (k − 1)-chain
Theorem 4.11. For a chain c, we have that ∂∂c = ∂ 2 c = 0
§§ 4.3.2 Manifolds
Definition 4.3.6 (Manifold). Given a set M ⊂ Rn , it is said to be a k-dimensional manifold if
∀xµ ∈ M we have that
1. ∃U ⊂ Rk open set xµ ∈ U and V ⊂ Rn and ϕ a diffeomorphism such that U ' V and
ϕ (U ∩ M ) = V ∩ Rk × {0} , i.e. U ∩ M ' Rk ∩ {0}
2. ∃U ⊂ Rk open and W ⊂ Rk open, xµ ∈ U and f : W −→ Rn a diffeomorphism
(a) f (W ) = M ∩ U
(b) rank (f ) = k ∀xµ ∈ W
(c) f −1 ∈ C(f (W ))
The function f is said to be a coordinate system in M
Definition 4.3.7 (Half Space). We define the k-dimensional half space Hk ⊂ Rk as
Hk := xµ ∈ Rk xi ≥ 0
(4.41)
Definition 4.3.8 (Manifold with Boundary). A manifold with boundary (MWB) is a manifold M such
that, given a diffeomorphism h, an open set U ⊃ M and an open set V ⊂ Rn
h (U ∩ V ) = V ∩ Hk ∩ {0}
(4.42)
The set of all points that satisfy this forms the set ∂M called the boundary of M
CHAPTER 4. TENSORS AND DIFFERENTIAL FORMS 60
Definition 4.3.9 (Tangent Space). Given a manifold M and a coordinate set f around xµ ∈ M , we
define the tangent space of M at xµ ∈ M as follows
f : W ⊂ Rk −→ Rn =⇒ f? Tx Rk = Tx M
(4.43)
Definition 4.3.10 (Vector Field on a Manifold). Given a vector field f µ we identify it as a vector field
on a manifold M if f µ (xν ) ∈ Tx M . Analogously we define a k−differential form
5 Integral Analysis
Theorem 5.1. The function 1Q is not integrable over the set [0, 1] with the usual definition of
the integral (Riemann sums)
Proof. Indicating the integral I as usual
ˆ 1
I= 1Q (x) dx
0
61
CHAPTER 5. INTEGRAL ANALYSIS 62
Basically, what we did before, was demonstrating that the set Q ∩ [0, 1] is not measurable in the
Riemann integration theory. This is commonly indicated with saying that the set Q ∩ [0, 1] is not Jordan
measurable.
For clarity, let K be some measure theory. We will say that a set is K-measurable if the following
calculation exists ˆ
µK (A) = 1A (x) dx (5.4)
X
1. {} ∈ A
2. X ∈ A
3. A ∈ A =⇒ Ac ∈ A
Sn Tn
4. A1 , · · · , An ∈ A =⇒ i=1 Ai , i1 Ai ∈ A
Example 5.1.1 (Simple Set Algebra). Let X = R2 and call R the set of all rectangles Ii ⊂ R? × R? ,
where R? = R ∪ {±∞}. It’s easy to see that this is not an algebra, since, by taking [0, 1] ∈ R, we have
that [0, 1]c ∈
/ R, hence it cannot be an algebra.
But, taken S as follows ( )
[n
2
S := A ⊂ R A = Ii Ii ∈ R
i=1
Definition 5.1.6 (Simple Set). A set A is a simple set iff, for some Ri ∈ S, we have
n
G
A= Ri
i=1
Definition 5.1.7 (Measure of a Simple Set). Let A be a simple set, the Jordan measure of a simple set
is given by the sum of the measure of the rectangles, i.e. the “area” of A is given by the sum of the
area of each rectangle Ri
Xn
µJ (A) = µJ (Ri ) (5.6)
i=1
5.1. MEASURE THEORY 63
Definition 5.1.8 (External and Internal Measure). We define the external measure µJ and the internal
measure µJ as follows.
Taken a limited set B and a simple set A we have
µJ (B) = inf{ µJ (A)| B ⊂ A}
(5.7)
µJ (B) = sup{ µJ (A)| A ⊂ B}
Remark (A Non Measurable Set). A good example for showing that the Jordan measure is the set we
were trying to measure, the set Q ∩ [0, 1]. We can easily see that
µJ (Q ∩ [0, 1]) = 1
µJ (Q ∩ [0, 1]) = 0
We have that
∞
X
µ(A) ≤ = 2
n=1
2n−1
But µ(Q ∩ [0, 1]) ≤ µ(A) ≤ 2 → 0, therefore Q ∩ [0, 1] is measurable with µ(Q ∩ [0, 1]) = 0
S∞
2. If Yj ⊂ X, with j ∈ J ⊆ N, µ(Yj ) < ∞ then X = j=1 Yj
Definition 5.1.11 (Measure Space). A measure space is a triplet (X, F, µ) with F a σ−algebra and µ
a measure.
Remark. The empty set has null measure.
Proof. Due to σ−additivity we have that
µ({}) = µ({} ∪ {}) = µ({}) + µ({})
Therefore, µ({}) = 0 necessarily.
Definition 5.1.12 (Lebesgue Measure). Consider again X = R2 and S the algebra of simple sets.
The external Lebesgue measure of a set B ⊂ R2 is then defined as follows
( ∞ ∞
)
X [
µL (B) := inf Area(Ri ) Ri ∈ S, B ⊂ Ri (5.9)
i=1 i=1
µL (C) = µL (C ∩ B) + µL (C \ B) (5.10)
If it’s measurable, then, µL (B) = µL (B) and it’s called the Lebesgue measure of the set.
In other words ∃ > 0 : ∃A, C ⊂ R2 , with A = A◦ , C = C such that
C ⊂ B ⊂ A ∨ µL (A \ C) < (5.11)
Definition 5.1.13 (Borel Algebra). Let R be the set of all rectangles. The smallest σ−algebra containing
R is called the Borel algebra and it’s indicated as B
Definition 5.1.14 (Lebesgue Algebra). The set of (Lebesgue) measurable sets is a σ−algebra, which we
will indicate as L. In particular, we have that, if I is a rectangle, I ∈ L.
If we add the fact that in B there are null measure sets which have subsets which aren’t part of B, we
end up with the conclusion that B ⊂ L
Definition 5.1.15 (Null Measure Sets). A set with null measure is a set X ⊂ F such that
µ(X) = 0 (5.12)
Where µ is a measure function.
It’s obvious that sets formed by a single point have null measure.
I.e take a set A = {a}, then it can be seen as a rectangle with 0 area, and therefore
µ ({a}) = 0 (5.13)
5.2. INTEGRATION 65
Theorem 5.2. Every set such that |A| = ℵ0 has null measure
For the corollary, it’s obvious if the line is thought as a rectangle in R2 with null area
§ 5.2 Integration
Definition 5.2.1 (Measurable Function). Given a Borel space (X, F) a measurable function is a
function f : X −→ F such that, ∀k ∈ F the following set is measurable
Or, in other words If ∈ F , with F the given σ−algebra of the Borel space.
The space of all measurable functions on X will be identified as M(X)
Theorem 5.3. Given a set A ∈ F with F a σ−algebra, the function 1A (x) is measurable
Definition 5.2.2 (Simple Measurable Function). Given a Borel space (X, F), a simple measurable
function is a function f : X −→ F which can be written as follows
n
X
f (x) = ck 1Ak (x) (5.16)
k=1
Where Ak ∈ F , ck ∈ F 0≤k≤n
Definition 5.2.3 (Integral). Given a measure space (X, F, µ) and a simple function f (x), we can define
the integral of the function f with respect to the measure µ over the set X as follows
ˆ n
X
f (x)µ (dx) = ck µ(Ak ) (5.17)
X k=1
f = f+ − f− (5.19)
Where (
f + = max {f, 0}
(5.20)
f − = max {−f, 0}
With the only constraint that the function f (x) must be misurable in the σ-algebra F
Defining the integral as an operator K̂µ [f ] we can see easily that this is a vector spaces due to the
properties of K̂µ .
It’s easy to note that if the chosen σ−algebra and measure are the Lebesgue ones, then this integral is
simply an extension of the usual Riemann integral.
It’s important to note that a norm in Lp (µ) can’t be defined as an usual integral p−norm, since there
are nonzero functions which have actually measure zero.
Definition 5.2.5 (Almost Everywhere Equality). Taken two functions f, g ∈ Lp (µ) we say that these two
function are almost everywhere equal if, given a set A := { x ∈ X| f (x) 6= g(x)} has null measure.
Therefore
f ∼ g ⇐⇒ µ(A) = 0 (5.23)
This equivalence relation creates equivalence classes of functions compatible with the vector space
properties of Lp (µ).
Definition 5.2.6 (Lp -Spaces). With the definition of the almost everywhere equality we can then define
a quotient space as follows
Lp (µ) = Lp (µ)\ ∼ (5.24)
This is a vector space, obviously, where the elements are the equivalence classes of functions f ∈ Lp (µ),
indicated as [f ].
If we define our σ−algebra and measure as the Lebesgue ones, this space is called the Lebesgue
space Lp (X), where an integral p−norm can be defined.
5.2. INTEGRATION 67
µ (F+∞ ) = µ (F−∞ ) = 0
Letting r > 0 we will indicate with 1r (x) the set function of the set F+∞ ∩ Br (0), therefore we have
that
f + (x) ≥ k1r (x) ∀k ∈ N
Therefore ˆ ˆ
1
µ (F+∞ ∩ Br (0)) = 1r (x) dx ≤ f + (x) dx −→ 0
k E
Theorem 5.5. Let (X, L, µ) be a measure space, where L is the Lebesgue σ−algebra and µ is
the Lebesgue measure. Given a function f ∈ L1 (X) we have that
ˆ
f (x) dx = 0 ⇐⇒ f ∼ 0 (5.25)
X
Theorem 5.6 (Monotone Convergence (B. Levi)). Let (f )k be a sequence of measurable functions
over a Borel space E, such that
Proof. Let F0k = 0 < y < fk (x) and F0 = 0 < y < f (x) be two sets defined as seen. They are all
measurable since fk (x), f (x) are measurable, and due to the monotony of fk (x) we have that
∞
G
F01 ⊂ F02 ⊂ · · · ⊂ F0k ⊂ · · · ∧ F0 = F0k
k=1
Due to σ−additivity of the measure function, we have that F0 is measurable, and that
∞
X
µ (F0 ) = µ (F0k ) ∴ µ (F0 ) = lim µ (F0k )
k−→∞
k=1
Notation (For Almost All). We now introduce a new (unconventional) symbol in order to avoid writing
too much, which would complicate the already difficult to understand theorems.
In order to indicate that we’re picking almost all elements of a set we will use a new quantifier, which
means that we’re picking all elements of a null measure subset of the set in question. The quantifier in
question will be the following
∀† (5.28)
Corollary 5.2.1. Let fk (x) be a sequence of non-negative measurable functions over a measurable set
E, then ∀† x ∈ E ˆ X Xˆ
fk (x) dx = fk (x) dx (5.29)
E k≥0 k≥0 E
Theorem 5.7 (Fatou). Let fk (x) be a sequence of measurable functions over a measurable set
E, such that ∀† x ∈ E ∃Φ(x) measurable : fk (x) > Φ(x), then
ˆ ˆ
lim inf fk (x) dx ≤ lim inf fk (x) dx
E k→∞ k→∞ E
It’s also (obviously) true taking the lim sup of the RHS, and for the theorem on the monotone conver-
gence, we have that
ˆ ˆ ˆ
lim gj (x) dx = lim gj (x) dx ≤ hk (x) dx
E j→∞ j→∞ E E
∴ lim gj (x) = sup gj (x) = sup inf hk (x) = lim inf hk (x)
j→∞ j j k≥j k→∞
And
f (x) = lim fk (x) ∀† x ∈ E
k→∞
Then ˆ ˆ
f (x) dx = lim fk (x) dx
E k→∞ E
Proof. By definition we have that −h(x) ≤ fk (x) ≤ h(x) ∀† x ∈ E, and we can apply Fatou’s theorem
ˆ ˆ ˆ ˆ
f (x) dx ≤ lim inf fk (x) dx ≤ lim sup fk (x) dx ≤ f (x) dx
E k→∞ E k→∞ E E
Corollary 5.2.2. Let E be a measurable set such that µ (E) < ∞ and let fk (x) be a sequence of
functions in E such that |fk (x)| ≤ M ∀† x ∈ E and fk (x) → f (x), ∀† x ∈ E. Then the theorem (5.8)
is valid.
Example 5.2.1. Take the sequence of functions fk (x) = kxe−kx over E = [0, 1]. We already know
that fk (x) −→ f (x) = 0 for x ∈ E, but fk (x) 6⇒ f (x) in E.
We have that
sup fk (x) = e−1 = h(x) 6= f (x)
E
We have that h(x) is measurable in E and we can apply the theorem (5.8)
Note that if µ is a finite measure, then f bounded =⇒ f locally uniformly integrably bounded or LUIB.
CHAPTER 5. INTEGRAL ANALYSIS 70
Theorem 5.9 (Leibniz’s Derivation Rule). Let (X, F, µ) be a measure space and A ⊂ Rn an open
set. If f : X × A −→ R is a LUIB Carathéodory function we can define
ˆ
µ
g(a ) = f (xν , aµ ) dµ (xσ ) ∈ C(A)
X
Then
∂xµ f (xν , aσ ) ∈ C(A)
Is LUIB, and therefore
g(aµ ) ∈ C 1 (A)
And ˆ
∂µ g = ∂aµ f (aν , xσ ) dµ (xγ )
X
In other terms ˆ ˆ
∂aµ f (aν , xσ ) dµ(xγ ) = ∂aµ f (aν , xσ ) dµ(xγ ) (5.30)
X X
Proof. Since f is a LUIB Carathéodory function we have that ∃haµ (xν ) : X −→ R and B (aµ ) ⊂ A :
∀y µ ∈ B (aν )
|f (y µ , xν )| ≤ haµ (xν )
Therefore ˆ
|g(aµ )| ≤ haµ (xν ) dµ(xσ ) < ∞
X
Since f is differentiable and its derivative is measurable, we have for the mean value theorem
And therefore ˆ
g(aµ + teµ ) − g(aµ ) 1
= t∂µ f (ξ ν (t, xσ ), xγ ) dµ(xδ )
t t X
For t → 0 ∂µ f (ξ ν , xσ ) → ∂µ f (aν , xσ ), and the LHS is simply the gradient of g. Therefore for theorem
(5.8) ˆ ˆ
ν ∂ ν σ γ
∂µ g(a ) = f (a , x ) dµ(x ) = ∂µ f (aν , xσ ) dµ(xγ )
∂aµ X X
5.3. CALCULUS OF INTEGRALS IN R2 AND R3 71
But
Y1x ⊂ Y2x ⊂ · · ·
G∞
Ax = Ykx
k=1
Let E ⊂ R2 be a measurable set. Define a sequence of compact sets Ki and a sequence of open sets
Aj such that
K1 ⊂ · · · ⊂ Kj ⊂ E ⊂ Aj ⊂ · · · ⊂ A1
We have that limj→∞ µ2 (Aj ) = limj→∞ µ2 (Kj ) = µ2 (E) and that Kjx ⊂ E ⊂ Ajx .
From the previous derivation we can write that
ˆ
lim (µ1 (Ajx ) − µ1 (Kjx )) dx = 0
j→∞ R
CHAPTER 5. INTEGRAL ANALYSIS 72
Building a sequence of non-negative functions fj (x) = µ1 (Ajx )−µ1 (Kjx ) we have that fj (x) ≤ fj−1 (x)
and due to Beppo-Levi we have that
ˆ ˆ
lim fj (x) dx = lim fj (x) dx
j→∞ R R j→∞
1. ∀† x ∈ R y 7→ f (x, y) is measurable in R
´
2. g(x) = R f (x, y) dy is measurable in R
˜ ´ ´
3. R2 f (x, y) dx dy = R R f (x, y) dx dy
Proof. Let f (x, y) ≥ 0. Defining F0 := { (x, y) ∈ E × R| 0 < z < f (x, y)} ⊂ R3 , we have that F0 is
measurable, and ¨
µ3 (F0 ) = f (x, y) dx dy
R2
Theorem 5.12 (Tonelli). Let f (x, y) be a measurable function and E ⊂ R2 be a measurable set.
If one of these integrals exists, the others also exist and have the same value
¨ ˆ ˆ ˆ ˆ
f (x, y) dx dy f (x, y) dx dy f (x, y) dy dx
R2 R R R R
Theorem 5.13 (Integration Over Rectangles). Let R = [a, b] × [c, d] ⊂ R2 be a rectangle, and f (x, y)
a measurable function over R. Then
´d
1. If ∀† x ∈ [a, b] ∃G(x) = c
f (x, y) dy, the function G(x) is measurable in [a, b] and
¨ ˆ b ˆ b ˆ d
f (x, y) dx dy = G(x) dx = f (x, y) dy dx
R a a c
5.3. CALCULUS OF INTEGRALS IN R2 AND R3 73
´b
2. If ∀† y ∈ [c, d] ∃F (y) = a
f (x, y) dx, the function F (y) is measurable in [c, d] and
ˆ ˆ d ˆ d ˆ b
f (x, y) dx dy = F (y) dy = f (x, y) dx dy
R2 c c a
Definition 5.3.1 (Normal Set). A set E ⊂ R2 is said to be normal with respect to the x axis if
It can be considered normal with respect to both axes, and we can use the inversion formula
¨ ˆ b ˆ x ˆ b ˆ b
f (x, y) dx dy = dx f (x, y) dy = dy f (x, y) dx (5.36)
T a a a y
Theorem 5.18 (Center of Mass). Take a plane E ⊆ R2 with surface density ρ(x, y) > 0. We define
the total mass M as follows ¨
M= ρ(x, y) dx dy (5.39)
E
The coordinates of the center of mass will be the following
¨
1
xG = ρ(x, y)x dx dy
M E
¨ (5.40)
1
yG = ρ(x, y)y dx dy
M E
Theorem 5.19 (Moment of Inertia). Taken the same plane E, we define the moment of inertia
with respect to a line r as the following integral
¨
2
Ir = ρ(x, y) (d(pµ , r)) dx dy (5.41)
E
Where d(pµ , r) is the distance function between the point (x, y) and the rotation axis r.
Both formulas are easily generalizable in R3
M 'N
∼
Theorem 5.20. Let A, B ⊂ Rn be two open sets and ϕµ : A −→ B a diffeomorphism, such that
ϕµ (E) = F
∼
Theorem 5.21 (Change of Variables). Let ϕµ : Rn −→ Rn be a diffeomorphism such that
∂1 g(y 1 , y 2 ) = f (y 1 , y 2 )
Taken c ∈ R, c > 1 : K ⊂ Q = [−c, c] × [−c, c], we have that ϕµ (xν ) = δνµ ∀kxµ kµ > 1 ∧ f (xµ ) =
0 ∀xµ ∈
/ Q.
Therefore f (ϕµ ) = 0 also and we have
ˆ ˆ ˆ
f (ϕµ ) det ∂µ ϕν dxγ = f (ϕµ ) det ∂µ ϕν dxγ = ∂1 g(ϕµ ) det ∂µ ϕν dxγ
Rn µν Q µν Q µν
ν cos θ −ρ sin θ
∂µ ϕ =
sin θ ρ cos θ (5.43b)
det ∂µ ϕν = ρ
µν
2. Spherical Coordinates
ρ ∈ R+
x(ρ, θ, φ) = ρ sin φ cos θ
µ ν
ϕ (x ) = y(ρ, θ, φ) = ρ sin φ sin θ θ ∈ [0, 2π) (5.44a)
z(ρ, θ, φ) = ρ cos φ φ ∈ [0, π]
sin φ cos θ −ρ sin φ sin θ ρ cos φ cos θ
∂µ ϕν = sin φ sin θ ρ sin φ cos θ ρ cos φ sin θ
cos φ 0 −ρ sin φ (5.44b)
det ∂µ ϕν = ρ2 sin φ
µν
3. Cylindrical Coordinates
ρ ∈ R+
x(ρ, θ, z) = ρ cos θ
µ ν
ϕ (x ) = y(ρ, θ, z) = ρ sin θ θ ∈ [0, 2π) (5.45a)
z(ρ, θ, z) = z z∈R
det ∂µ ϕν = ρ (5.45b)
µν
Definition 5.3.3 (Rotation Solids). Let D ⊂ R2 be a bounded measurable set contained in the half-plane
y = 0, x > 0. Suppose we let D “pop up” into R3 through a rotation by an angle θ0 around the z axis.
What has been obtained is a rotation solid E ⊂ R3 . We have that
˚ ¨ ˆ θ0 ¨ ¨
µ(E) = dx dy dz = ρ dρ dθ dz = θ0 ρ dρ dz = θ0 x dx dy (5.46)
E D 0 D D
Or
µ(E) = θ0 xG µ2 (D)
Theorem 5.23 (Guldino). The measure of a rotation solid is given by the measure of the rotated
figure times the circumference described by the center of mass of the solid.
This is exactly the previous formula.
5.3. CALCULUS OF INTEGRALS IN R2 AND R3 77
Theorem 5.24 (Center of Mass of a Curve). Given a curve γ µ : [a, b] −→ R3 with linear mass
density m : {γ} −→ R, we define the total mass of γ as follows
ˆ ˆ b
dγ µ
M= m ds = m(γ µ ) dt (5.48)
γ a dt µ
Definition 5.3.5 (Line Integral of the Second Kind). Given a vector field f µ : A −→ R3 and a smooth
curve γ µ : [a, b] −→ A ⊂ R3 we define the line integral of the second kind as follows
ˆ ˆ b
µ dγµ
f Tµ ds = f µ (γ ν ) dt (5.50)
γ a dt
Defining a differential form ω = f µ dxµ we can also see this integral as follows
ˆ ˆ
ω= f µ Tµ ds (5.51)
γ γ
Definition 5.3.6 (Conservative Field). Let f µ : A −→ R3 be a vector field such that f µ ∈ C 1 (A) and A
is open and connected. This field is said to be conservative, if ∀xµ ∈ A
Theorem 5.25 (Line Integral of a Conservative Field). Given a conservative field f µ : A −→ R3 and
a smooth curve {γ} ⊂ A, γ µ : [a, b] −→ R3 with A open and connected, we have that
ˆ
f µ Tµ ds = U (γ(a)) − U (γ(b)) (5.53)
γ
Definition 5.3.7 (Rotor). Given a vector field f µ : A −→ R3 with f µ ∈ C 1 (A), we define the rotor of
the vector field as follows
rot(f µ ) = µνγ ∂ ν f γ (5.54)
Theorem 5.26. Given f µ a conservative vector field on an open connected set A, we have that
µνγ ∂ ν f γ = 0 (5.55)
Alternatively, if γ µ : [a, b] −→ R3 is the parameterization of a smooth closed curve, we have
that ˛
f µ Tµ ds = 0 (5.56)
γ
Definition 5.3.9 (Rotation Surface). Given a smooth curve γ µ : [a, b] −→ R3 , the rotation of this curve
around the z−axis generates a smooth surface Σ with the following parameterization
1
γ (t) cos θ
µ
r (t, θ) = γ 2 (t) sin θ (t, θ) ∈ [a, b] × [0, θ0 ] (5.59)
3
γ (t)
The area of a rotation surface is calculated as follows
ˆ b
s 2 2 2
1 dγ 1 dγ
µ (Σ) = θ0 γ (t) + dt (5.60)
a dt dt
Theorem 5.27 (Guldino II). Given Σ a smooth rotation surface defined as before, we have that
its area will be ˆ
µ (Σ) = θ0 x1 ds = θ0 x1G Lγ (5.61)
γ
Where x1G is the first coordinate of the center of mass of the curve, calculated as follows
ˆ
1 1
xG = x1 ds
Lγ γ
5.3. CALCULUS OF INTEGRALS IN R2 AND R3 79
Definition 5.3.11 (Center of Mass of a Surface). Given a smooth surface Σ with parameterization
rµ (u, v) and mass density δ, we define its total mass as follows
ˆ
M= δ dσ (5.64)
Σ
Definition 5.3.12 (Moment of Inertia of a Surface). Given a smooth surface Σ with parameterization
rµ (u, v) and mass density δ we define its moment of inertia around an axis r, I, as the following
integral ˆ
2
I= δ(xµ ) (d(pµ , r)) dσ pµ ∈ Σ (5.66)
Σ
Definition 5.3.14 (Boundary of a Surface). Given a smooth surface as before, we define the boundary
∂Σ as follows
∂Σ = Σ \ Σ (5.69)
Note how, given the parameterization rµ , we have rµ (∂K) = ∂Σ
Definition 5.3.15 (Closed Surface). A surface Σ ⊂ R3 is said to be closed iff ∂Σ = {}
Definition 5.3.16 (Flux). Given a vector field f µ : A ⊂ R3 −→ R3 and a smooth orientable surface
Σ ⊂ A, we define the flux of the vector field f µ on the surface as follows
ˆ ¨
µ µ
ΦΣ (f ) = f nµ dσ = f µ (rν )µγσ ∂1 rγ ∂2 rσ du dv (5.70)
Σ K
CHAPTER 5. INTEGRAL ANALYSIS 80
§ 5.4 Integration in C
Definition 5.4.1 (Piecewise Continuous Function). Let γ : [a, b] −→ C be a piecewise continuous curve
such that {γ} ⊂ D ⊂ C, and f : D −→ C, f ∈ C(D). Then the function (f ◦ γ) γ 0 (t) : [a, b] −→ C
is a piecewise continuous function
Definition 5.4.2 (Line Integral in C). Let γ : [a, b] −→ D ⊂ C be a piecewise continuous curve and
f : D −→ C a measurable function f ∈ C(D).
We define the line integral over γ the result of the application of the integral operator K̂γ [f ], where
ˆ ˆ b
K̂γ [f ] = f (z) dz = (f ◦ γ) γ 0 (t) dt (5.71)
γ a
2. γ ∼ η =⇒ K̂γ [f ] = K̂η [f ]
Proof. The proof is quite straightforward using the definition given for the line integral
ˆ ˆ b ˆ b
0
f (z) dz = (f ◦ γ) γ (t) dt ≤ k(f ◦ γ) γ 0 (t)k dt ≤
γ a a
ˆ b
≤ sup kf (z)k kγ 0 (t)k dt = Lγ sup kf (z)k
z∈{γ} a z∈{γ}
5.4. INTEGRATION IN C 81
dF
= f (z) ∀z ∈ D (5.72)
dz
Notation. Given a closed curve γ and a measurable function f (z) we define the following notation
ˆ ˛
f (z) dz = f (z) dz
γ γ
Theorem 5.30 (Existence of the Primitive Function). Let f : D −→ C f ∈ C(D) with D ⊂ C open
and connected. Then these statements are equivalent
1. ∃F : D −→ C : F 0 (z) = f (z)
´z ´
2. ∀z1 , z2 ∈ D, ∀{γ} ⊂ D piecewise smooth f (z) dz = z12 f (z) dz
γ
¸
3. ∀γ : [a, b] −→ {γ} ⊂ D closed piecewise smooth γ f (z) dz = 0
The result of the integral is obviously F (z2 ) − F (z1 ), therefore we can immediately write that, if
ˆ ˆ z2
∃F : D −→ C : F 0 (z) = f (z) =⇒ f (z) dz = f (z) dz
γ z1
Since D is open we can define δz ∈ R, δz > 0 ∧ ∃Bδ1 (z) ⊂ D. Taken ∆z ∈ C : k∆zk < δ1 we have
that ˆ z+∆z
F (z + ∆z) − F (z) = f (w) dw
z
Dividing by ∆z and taking the limit as ∆z → 0 we have that using the Darboux inequality we get that
ˆ z+∆z
F (z + ∆z) − F (z) 1
− f (z) = f (w) dw ≤
∆z k∆zk z
CHAPTER 5. INTEGRAL ANALYSIS 82
2 =⇒ 3. Taken an arbitrary piecewise smooth curve γ and z1 6= z2 ∈ {γ}. We can now find two
curves such that γ(t) = γ1 (t) − γ2 (t). Since the integral of f is path independent, we get
ˆ ˆ ˆ
f (z) dz = f (z) dz − f (z) dz = 0
γ γ1 γ2
For n = 1 we still have that f ∈ C(D) but @F (z) : D −→ C primitive of f1 (z), but there exists one in
the domain G of holomorphy of the logarithm.
Although we have that G ⊂ D, and we can take a curve γ : 0 ∈ extr γ, and therefore {γ} ⊂ G and
we have that ˛
1
dz = 0
γ z
If we otherwise have 0 ∈ γ ◦ the integral is non-zero.
Take a branch of the logarithm σ and a curve η has only one point of intersection with such branch
at zi = u0 eiα . Taken η(a) = η(b) = u0 eiα , we define η : [a + , b + ] −→ C with > 0 : η (t) =
η(t) ∀t ∈ [a + , b + ], then ˛ ˛
1 1
dz = lim dz
η z →0 η z
√
Example 5.4.2. Let’s calculate the integral of f (z) = z along a closed simple piecewise smooth
curve γ : [a, b] −→ C : 0 ∈ γ ◦ and it intersects the line σα = u0 eiα , where
√ √ θ
z= rei 2 r ∈ R+ , θ ∈ (α, α + 2π], α ∈ R
5.4. INTEGRATION IN C 83
Taken a parametrization γ(t) : γ(a) = γ(b) = u0 eiα we have that f (z) ∈ H(D) where D = C \ {σα }.
Proceding as before, we have ˛ ˛
√ √
z dz = lim z dz
γ →0 γ
Lemma 5.4.1. Taken a closed simple pointwise smooth curve γ : [a, b] −→ C and taken D = {γ}◦ ∪γ =
{γ}◦ and a function f ∈ H(D), for a finite cover of D, Q composed by squares Qj ∈ Q ∀j ∈ [1, N ] ⊂ N,
we have that
f (z) − f (zj ) df
∃zj ∈ Qj ∩ {γ}◦ : − < ∀z ∈ Qj ∩ {γ}◦ \ {zj }
z − zj dz zj
We have that An+1 ⊂ An , and taking a sequence (w)n ∈ {γ}◦ we have due to the compactness of
{γ}◦ that ∃(w)nj → w ∈ {γ}◦ . Since f ∈ H({γ}◦ ) we have that f is holomorphic in w, therefore
f (z) − f (w) df
∀ > 0 ∃δ > 0 : − < ∀z ∈ Bδ (w) \ {w}
z−w dz w
√
2
Taken an ñ such that diam(Qñj ) = 2ñ d < δ we have that still w ∈ An ∀n ∈ N, and due to its closedness
we can also say
∃Nj ∈ N : ∀nj > Nj (w)nj ∈ An
Therefore
∃k0 ∈ N : w ∈ Qñk0 ∩ {γ}◦ ⊂ Añ
Theorem 5.31 (Cauchy-Goursat). Taken γ : [a, b] −→ C a closed simple piecewise smooth curve
and D = {γ} ∪ {γ}◦ and a function f ∈ H(D), we have
˛
f (z) dz = 0 (5.73)
γ
CHAPTER 5. INTEGRAL ANALYSIS 84
Proof. Using the previous lemma we can say that for a finite cover {γ}, Qj ∈ Q ∃zj ∈ Qj ∩ {γ}◦ and
a function
f (z) − f (zj ) − f 0 (zj ) z 6= zj
δj (z) = z − zj
0 z=z j
The first two integrals on the second line are null, and we have therefore
˛ ˛
f (z) dz = δj (z)(z − zj ) dz
ηj ηj
SN
By definition {γ} = j=1 {ηj } and therefore
˛ N ˛
X
f (z) dz = δj (z)(z − zj ) dz
γ j=1 ηj
Using the theorem on the Jordan curve, we have that ∃Qn ∈ Q such that {γ} ⊂ Qn . Taken diam(Qn ) =
D
˛ XN
√
f (z) dz ≤ 2D(4D + L) → 0
γ j=1
Definition 5.4.4 (Simple Connected Set). An open set G ⊂ X with X some metric space, is said to be
simply connected iff ∀{γj } ⊂ G simple curves we have that γj ∼ 0.
γ ∼ 0 implies that the curve is homotopic to a point
Theorem 5.32 (Cauchy-Goursat II). Let G ⊂ C open and simply connected. Then, ∀f ∈ H(G), {γ} ⊂
G with γ simple closed and smooth
˛
f (z) dz = 0
γ
5.4. INTEGRATION IN C 85
Theorem 5.33. Let G ⊂ C be a simply connected open set. If f ∈ H(G), then there exists a
primitive for f (z)
Theorem 5.34 (Cauchy Integral Representation). Taken a positively oriented closed simple piece-
wise smooth curve γ : [a, b] −→ C and a function f : G ⊂ C −→ C such that if D = {γ} ∪ {γ}◦ ⊂
G, f ∈ H(D), we have that
ffi
1 f (w)
f (z) = dw ∀w ∈ {γ}◦ (5.74)
2πi γ w − z
Proof. Taken γρ (θ) = z + ρeiθ such that γρ ∼ γ, {γρ } ⊂ {γ}◦ is a simple curve, we have
ffi ffi
f (w) f (w)
dw = dw
γ w−z γρ w − z
Where γ is a closed simple piecewise smooth curve such that z ∈ {γ}◦ and {γ} ⊂ D
Corollary 5.4.1. Let f ∈ H(D), then
dn f
∀n ∈ N ∈ H(D)
dz n
Theorem 5.36 (Morera). Let D ⊂ C be an open and connected set. Take f : D −→ C : f ∈ C(D).
Then, if ∀{γ} ⊂ D closed piecewise smooth
˛
f (z) dz = 0 =⇒ f ∈ H(D) (5.76)
γ
Proof. Since f ∈ C(D) ∃F (z) ∈ C 1 (D) : f (z) = F 0 (z). Since C 1 (C) ' H(C) we have that, due to
the previous corollary
dF
= f (z) ∈ H(D)
dz
Theorem 5.37 (Cauchy Inequality). Let f ∈ H(BR (z0 )) with z0 ∈ C. If kf (z)k ≤ M ∀z ∈ BR (z0 )
df n!M
≤ (5.77)
dz z0 Rn
dn f
Proof. Take γr (θ) = z0 + reiθ with θ ∈ [0, 2π], r > R, then the derivative dz n can be written using
z0
the Cauchy integral representation, since f ∈ H(Br (z0 ))
ffi
dn f n! f (w)
= dw
dz n z0 2πi γr (w − z0 )n+1
5.5. INTEGRAL THEOREMS IN R2 AND R3 87
dn f n! n!M
≤ sup kf (z)k ≤ n
dz n z0
n
r z∈{γr } r
Theorem 5.38 (Liouville). Let f : C −→ C a function such that f ∈ H(C), i.e. whole. If
∃M > 0 : kf (z)k ≤ M ∀z ∈ C the function f (z) is constant
Proof. f ∈ H(C), kf (z)k ≤ M and we can write, taken γR (θ) = z + Reiθ with θ ∈ [0, 2π]
ffi
0 1 f (w)
f (z) = dz
2πi γR (w − z)2
Theorem 5.39 (Fundamental Theorem of Algebra). Take a polynomial Pn (z) ∈ Cn [z], where Cn [z]
is the space of complex polynomials with variable z and degree n. If we have
n
X
Pn (z) = ak z k , z, ak ∈ C, an 6= 0
k=0
Proof. As an absurd, say that ∀z ∈ C, Pn (z) 6= 0. Then f (z) = 1/Pn (z) ∈ H(C).
Since limz→∞ Pn (z) = ∞, we have that kf (z)k ≤ M ∀z ∈ C, and limz→∞ f (z) = 0.
Therefore ∃R > 0 : ∀kzk > R, kf (z)k < 1. Since f ∈ H(C), we have that f ∈ C(B R (z)). Due to the
Liouville theorem we have that f (z) is constant
Theorem 5.41 (Stokes). Given D ⊂ R2 an open set with ∂D piecewise smooth and a vector
field f µ : A −→ R2 with D ⊂ A
ˆ ˆ
3µν ∂ µ f ν dx dy = f µ tµ ds (5.79)
D ∂+D
We can analogously write these theorems in the language of differential forms and manifolds, after
giving a couple of definitions
Definition 5.5.1 (Volume Element). Given a k−dimensional compact oriented manifold M with bound-
ary and ω ∈ Λk (M ) a k−differential form on M , we define the volume of M as follows
ˆ ˆ
V (M ) = dV = ω (5.83)
M M
Where dV is the volume element of the manifold, given by the unique ω ∈ Λk (M ), defined as follows
ωµν v µ wν = nµ µνγ v ν wγ = dA
5.5. INTEGRAL THEOREMS IN R2 AND R3 89
Therefore
µ
dA = kµνγ v ν wγ k (5.85)
Which is the already known formula.
For a 2−manifold we can write immediately the following formulas
dA = n1 dy ∧ dz + n2 dz ∧ dx + n3 dx ∧ dy (5.86)
And, on M 1
n dA = dy ∧ dz
n2 dA = dz ∧ dx (5.87)
3
n dA = dx ∧ dy
ω = α dx + β dy (5.90)
We have ˆ ˆ ˆ ¨
∂β ∂α
α dx + β dy = ω= dω = − dx ∧ dy (5.91)
∂M ∂M M M ∂x ∂y
Theorem 5.46 (Gauss, Differential Forms). Given M a 3−manifold smooth with boundary and
compact with outer normal nµ and a vector field f µ ∈ C 1 (M ), we have
ˆ ˆ
µ
∂µ f dV = f µ nµ dA (5.92)
M ∂M
CHAPTER 5. INTEGRAL ANALYSIS 90
ω = f 1 dy ∧ dz + f 2 dz ∧ dx + f 3 dx ∧ dy
Theorem 5.47 (Stokes, Differential Forms). Given M ⊂ R3 a compact oriented smooth 2−manifold
with boundary with nµ as outer normal and tµ as tangent vector in ∂M , given a vector field
f µ ∈ C 1 (A) where M ⊂ A, we have
ˆ ˆ
nµ µνγ ∂ ν f γ dA = f µ tµ ds (5.93)
M ∂M
ω = f µ dxµ
We have that
f µ tµ ds = f µ dxµ = ω
And therefore ˆ ˆ ˆ ˆ
µ ν γ
n µνγ ∂ f dA = dω = ω= f µ tµ ds
M M ∂M ∂M
These last formulas are a good example on how they can be generalized through the use of
differential forms, bringing an easy way of calculus in Rn of the various integral theorems, all condensed
in one formula, the Gauss-Green-Stokes-Ostogradskij theorem
6 Sequences, Series and Residues
Definition 6.1.3 (Uniform Convergence). Defining an k·k∞ = supi≤n |·| we have that the convergence
of a sequence of functions is uniform, and it’s indicated as fn ⇒ f , iff
∀ > 0 ∃N ∈ N : d(fn (x), f (x)) < ∀n ≥ N ∀x ∈ S (6.4)
Or, using the norm k·k∞
∀ > 0 ∃N ∈ N : kfn − f k∞ < (6.5)
Theorem 6.1 (Continuity of Uniformly Convergent Sequences). Let (fn )n≥0 : (S, dS ) −→ (X, d) be a
sequence of continuous functions. Then if fn ⇒ f , we have that f ∈ C(S), where C(S) is the
space of continuous functions
Proof.
∀x ∈ S, ∃ > 0 : fn ⇒ f, ∴ ∀n ≥ N ∈ N : d(fn (x), f (x)) <
3
(6.6)
fn ∈ C(S) =⇒ ∃δ > 0 : d(fn (x), fn (y)) < , ∀x, y ∈ S : dS (x, y) < δ
3
∴ d(f (x), f (y)) ≤ d(f (x), fn (x)) + d(fn (x), fn (y)) + d(fn (y), f (y)) < ⇐⇒ dS (x, y) < δ
91
CHAPTER 6. SEQUENCES, SERIES AND RESIDUES 92
Theorem 6.2 (Integration of Sequences of Functions). Let (fn )n≥0 be a sequence of functions such
that fn ⇒ f Then we can define the following equality
ˆ b ˆ b ˆ b
lim fn (x) dx = lim fn (x) dx = f (x) dx (6.7)
n→∞ a a n→∞ a
Proof. We already know that in the closed set [a, b] we can say, since fn ⇒ f , that
∀ > 0 ∃N ∈ N : ∀n ≥ N kfn − f k∞ < (6.8)
b−a
1. ∃x0 ∈ I : fn (x0 ) → l
2. fn0 ⇒ g ∀x ∈ I
Then
fn (x) ⇒ f =⇒ ∀x ∈ I, f 0 (x) = lim fn0 (x) = g(x) (6.10)
n→∞
Proof. For the fundamental theorem of integral calculus, we can write, using the regularity of the
fn (x) that
ˆ x
fn (x) = fn (x0 ) + fn (t) dt
x0
Definition 6.2.1 (Series of Functions). Let (fn )n≥0 ∈ C be a sequence of functions, such that fn : S → C.
We can define the series of functions as follows
n
X
sn (x) = fk (x) (6.11)
k=1
Proof.
∀x ∈ S lim fk (x) = lim (sn (x) − sn+1 (x)) = 0
k→∞ n→∞
P∞ A
Theorem 6.5. Let k=0 fk (x) −→ s(x), then
∞ ∞
A
X X
fk (x) −→ s(x) =⇒ fk (x) → s(x) (6.16)
k=0 k=0
Proof. Let
n
X ∞
X
sn (x) = fk (x) ∴ ∃g(x) : (S, d) −→ C, ∃N (x) ∈ N : g(x) − fk (x) =
k=0 k=0
X∞
= |fk (x)| < ∀n ≥ N (x)
k=n+1
∴ ∀n, m ∈ N, m > n
m
X ∞
X
|sm (x) − sn (x)| = fk (x) ≤ |fk (x)| < ∀x ∈ S
k=n+1 k=n+1
Definition 6.2.5 (Total Convergence). A series of functions sk (x) is said to be totally convergent if
T
The total convergence is then indicated as sk (x) −→ s(x)
Then
A
3. sk (x) −→ s(x) =⇒ sk (x) → s(x)
A
4. sk (x) ⇒ s(x) =⇒ sk (x) −→ s(x)
T
5. sk (x) −→ s(x) =⇒ sk (x) ⇒ s(x)
6.2. SERIES OF FUNCTIONS 95
Definition 6.2.6 (Power Series). Let z, z0 , (an ) ∈ C. A power series centered in z0 is defined as
follows
X∞
ak (z − z0 )k (6.17)
k=0
Example 6.2.1. Take the geometric series. This is the best example of a power series centered in
z0 = 0, and it has the following form
∞
X
zk (6.18)
k=0
P∞
Theorem 6.7 (Cauchy-Hadamard Criteria). Let k=0 ak (z−z0 )k be a power series, with an , z, z0 ∈ C.
We define the Radius of convergence R ∈ R? = R ∪ {±∞}, with the Cauchy-Hadamard criteria
1
+∞ =0
R
1 1
1
= lim sup |an | n = l 0< =l<∞ (6.21)
R n→∞
R
1
0
= +∞
R
Then sk (z) ⇒ s(z) ∀|z| ∈ (−R, R)
Theorem 6.8 (D’Alambert Criteria). From the power series we have defined before, we can write
the D’Alambert criteria for convergence as follows
1 ak+1 ak
= lim =⇒ R = lim (6.22)
R k→∞ ak k→∞ ak+1
Where R is the previously defined radius of convergence
CHAPTER 6. SEQUENCES, SERIES AND RESIDUES 96
Theorem 6.9 (Abel). Let R > 0, then if a power series converges for |z| = R, it converges
uniformly ∀|z| ∈ [r, R] ⊂ (−R, R]. It is valid analogously for x = −R
Remark (Power Series Integration). If the series has R > 0 and it converges in |z| = R, calling s(x) the
sum of the series, with x = |z| we can say that
ˆ R ∞ ˆ R ˆ ∞
RX ∞
X X Rk+1
s(x) dx = ak xk dx = ak xk dz = ak (6.23)
0 0 0 k+1
k=0 k=1 k=0
Remark (Power Series Derivation). If Abel’s theorem holds, we have also that, if we have s(x) our
power series sum, we can define the n−th derivative of this series as follows
∞
dn s X
= k(k − 1) · · · (k − n + 1)ak xk−n (6.24)
dxn
k=n
Proof. Taken z ∈ Br (z0 ) and γ(t) = z0 + reit t ∈ [0, 2π] and kz − z0 k < r < R we can write, using
the integral representation of f
ffi ffi
1 f (w) 1 f (w)
f (z) = dw = dw
2πi γ (w − z) 2πi γ (w − z0 ) − (z − z0 )
1 1 − (z/w)n
z n
1 1
= + =
w−z w 1 − z/w 1 − z/w w
n−1
1 z n X 1 z n
= +
w−z w w w
k=0
On the RHS as first term we have the k−th derivative of f and on the right there is the so called
remainder Rn (z). Therefore
n
X 1 dk f
f (z) = (z − z0 )k + Rn (z)
k! dz k z0
k=0
n→∞
It’s easy to demonstrate that Rn (z) −→ 0, and therefore
∞
X 1 dk f
f (z) = (z − z0 )k
k! dz k z0
k=0
Definition 6.3.1 (Taylor Series for Scalar Fields). Given a function f : A ⊂ Rn −→ R f ∈ C m (A), given
a multi-index α one can define the Taylor series of the scalar field as follows
X 1
f (x) = ∂ α f (x0 )(x − x0 )α + Rm (x)
α!
|α|≤m
Definition 6.3.2 (MacLaurin Series). Taken a Taylor series, such that z0 = 0, we obtain a MacLaurin
series.
∞
X 1 dk f
f (z) = zk (6.26)
k! dz k z=0
k=0
Definition 6.3.3 (Remainders). We can have two kinds of remainder functions while calculating series:
n
1. Peano Remainders, Rn (z) = O(kz − z0 k )
2. Lagrange Remainders, Rn (x) = (n + 1)!−1 f (n+1) (ξ)(x − x0 )n+1 , x, x0 ∈ R ξ ∈ (x, x0 )
What we saw before as Rn (z) is the remainder function for functions f : D ⊂ C −→ C.
A particularity of remainder function is that Rn (z) → 0 always, if f is holomorphic
Theorem 6.11 (Integration of Power Series II). Let f, g : BR (z0 ) −→ C and {γ} ⊂ BR (z0 ) a piecewise
smooth path. Taken
X∞
f (z) = an (z − z0 )n g ∈ C({γ})
n=0
We have that ˆ ˆ
∞
X
n
an g(z)(z − z0 ) dz = g(z)f (z) dz (6.27)
n=0 γ γ
CHAPTER 6. SEQUENCES, SERIES AND RESIDUES 98
Proof. Since f, g ∈ C({γ}) by definition, and f ∈ H(Br (z0 )) with r < R, we have that ∃K̂γ [f g].
Firstly we can write that ∀z ∈ BR (z0 )
n−1
X n−1
X ∞
X
g(z)f (z) = ak g(z)(z − z0 )k + g(z)Rn (z) = ak g(z)(z − z0 )k + g(z) ak (z − z0 )k
k=0 k=0 k=n
Letting M = supz∈{γ} kg(z)k, and noting that kRn (z)k < for ∀ > 0 and for some n ≥ N ∈ N, z ∈
{γ} we have, using the Darboux inequality
ˆ
g(z)Rn (z) dz ≤ M Lγ → 0
γ
Theorem 6.12 (Holomorphy of Power Series). If a function f (z) is expressable as a power series
P∞
f (z) = k=0 ak (z − z0 )k , kz − z0 k < R we have that f ∈ H(BR (z0 ))
Proof. Take the previous theorem on the integration of power series, and choose g(z) = 1. Since
g(z) ∈ H(C) we also have that it’ll be continuous on all paths {γ} ⊂ C piecewise smooth.
Take now a closed piecewise smooth path {γ}, then we can write
ffi ffi X
∞ ∞
X ffi
f (z)g(z) dz = ak (z − z0 )k = ak (z − z0 )k dz
γ γ k=0 k=0 γ
Since the function h(z) = (z − z0 )k ∈ H(C) ∀k = 6 1, we have, thanks to the Morera and Cauchy
theorems ffi
f (z) dz = 0 =⇒ f (z) ∈ H(BR ({γ}))
γ
P∞
Corollary 6.3.1 (Derivative of a Power Series II). Take f (z) = k=0 ak (z − z0 )k kz − z0 k < R. Then,
∀z ∈ BR (z0 ) we have that
∞
df X
= ak k(z − z0 )k−1 (6.28)
dz
k=1
Proof. Taken z ∈ BR (z0 ) and a continuous function g(w) ∈ C({γ}), with {γ} ⊂ BR (z0 ) a closed
simple piecewise smooth path. If z ∈ {γ}◦ and
1 1
g(w) =
2πi (w − z)2
6.3. SERIES REPRESENTATION OF FUNCTIONS 99
Since h(w) = (w − z0 )k ∈ H(C) ∀k 6= 1 we have, using again the integral representation for
holomorphic functions ffi
1 f (w) df
dw =
2πi γ (w − z)2 dz
ffi k
1 (w − z0 )
dw = k(z − z0 )k−1
2πi γ (w − z)2
Therefore
ffi ∞
1 f (w) X df
2
dw = ak k(z − z0 )k =
2πi γ (w − z) dz
k=0
Corollary 6.3.2 (Uniqueness of the Taylor Series). Taken an holomorphic function f ∈ H(D) with
D ⊂ C some connected open set, we have that
∞
X 1 dk f
f (z) = ak (z − z0 )k ak = ∀kz − z0 k < R
k! dz k z0
k=0
Proof. Taken g(z) a continuous function over a closed piecewise simple path {γ} ⊂ C, where
1 1
g(z) =
2πi (z − z0 )k+1
We have that
ffi ∞ ffi
1 f (z) X ak
n+1
dz = (z − z0 )k−n−1 dz
2πi γ (z − z0 ) 2πi γ
k=1
The integral on the RHS evaluates to δnk , and thanks to the integral representation of f (z) we can write
ffi
1 f (z) 1 dn f
dz = = n!an
2πi γ (z − z0 )n+1 n! dz n z0
Theorem 6.13 (Laurent Series Expansion). Let f : AR1 R2 (z0 ) −→ C be a function such that
f ∈ H (AR1 R2 (z0 )), and {γ} ⊂ AR1 R2 (z0 ) a closed simple piecewise smooth curve.
Then f is expandable in a generalized power series or a Laurent series as follows
∞ ∞ ∞
X X c−
n
X
f (z) = c+ n
n (z − z0 ) + n
= ck (z − z0 )k (6.30)
n=0 n=1
(z − z 0 )
k=−∞
Proof. Taken a random point z ∈ AR1 R2 (z0 ), a closed simple piecewise smooth curve {γ} ⊂ AR1 R2 (z0 )
and two circular smooth paths {γ2 }, {γ3 } : {γ2 } ∪ {γ3 } = ∂Ar1 r2 (z0 ) ⊂ AR1 R2 (z0 ) ∧ {γ} ⊂ Ar1 r2 (z0 )
and a third circular path {γ3 } ⊂ Ar1 r2 (z0 ), we can write immediately, using the omotopy between all
the paths ffi ffi ffi
f (w) f (w) f (w)
dw = dw + dw
γ2 w − z γ1 w − z γ3 w −z
Using the Cauchy integral representation we have that the integral on γ3 yields immediately 2πif (z),
hence we can write
ffi ffi
1 f (w) 1 f (w)
f (z) = dw + dw
2πi γ2 (w − z0 ) − (z − z0 ) 2πi γ1 (z0 − z) − (w − z0 )
Using the two following identities for z 6= w
n−1
1 1 z n X 1 z k
= +
w−z w−z w w w
k=0
w n n
1 1 X 1 w k
= +
z−w z−w z w z
k=1
We obtain that
n−1 ffi n ffi
X (z − z0 )k f (w) X 1
f (z) = k+1
dw + ρn (z) + f (w)(w − z0 )k−1 dw + σn (z)
2πi γ2 (w − z0 ) 2πi(z − z0 )k γ1
k=0 k=1
6.3. SERIES REPRESENTATION OF FUNCTIONS 101
−
Where, after choosing appropriate substitutions with some coefficients c+
k , ck we have
n−1 n
X X c−
f (z) = c+ k
k (z − z0 ) + ρn (z) +
k
+ σn (z)
(z − z0 )k
k=0 k=1
Where ρn , σn are the two remainders of the series expansion, and are
ffi
(z − z0 )n f (w)
ρn (z) = dw
2πi γ2 [(w − z 0 ) − (z − z0 )] (w − z0 )n
ffi
1 f (w)
σn (z) = dw
2πi(z − z0 )n γ1 (w − z0 ) − (z − z0 )
n→∞
In order to prove the theorem we now need to demonstrate that ρn , σn −→ 0. Taken M1 =
supz∈{γ1 } kf (z)k, M2 = supz∈{γ2 } kf (z)k, we have, using the fact that both γ1 , γ2 are circular
n
M2 kz − z0 k n→∞
kρn (z)k ≤ −→ 0 kz − z0 k < r2
1 − r12 kz − z0 k r2
n
M1 r1 n→∞
kσn (z)k ≤ 1 −→ 0 r1 < kz − z0 k
r1 kz − z0 k − 1
kz − z 0 k
Theorem 6.15 (Integral of a Laurent Series). Let f (z) ∈ H (AR1 R2 (z0 )) and take {γ} ⊂ AR1 R2 (z0 ) a
piecewise smooth curve, and g ∈ C({γ}), then we have
∞
X ˛ ˛
cn g(z)(z − z0 )n dz = g(z)f (z) dz
n=−∞ γ γ
Proof. We begin by separating the sum in two parts, ending up with the following
X∞ ˛ ˛
c+n g(z)(z − z 0 )n
dz = g(z)f+ (z) dz
n=0 γ γ
∞ ˛ ˛
X g(z)
c−
n dz = g(z)f− (z) dz
n=1 γ (z − z0 )n γ
CHAPTER 6. SEQUENCES, SERIES AND RESIDUES 102
Which is analogous to the integration of Taylor series. The same could be obtained keeping the bounds
of the sum in all Z
As for Taylor series, in a completely analogous fashion, a Laurent series is holomorphic and unique.
The derivative of a Laurent series, is then obviously the following
∞
df X
= cn n(z − z0 )n−1
dz n=−∞
Proof. Due to the holomorphy of both f and g, we have that the function f g has a Taylor series
expansion
∞
X
f (z)g(z) = cn (z − z0 ) kz − z0 k < R
n=0
Theorem 6.17 (Division of Power Series). Taken the two functions as before, with the added
necessity that g(z) 6= 0, we have that
∞ n−1
!
f (z) X n 1 X
= dn (z − z0 ) dn = an − dk bn−k
g(z) n=0 b0
k=0
Proof. Everything hold as in the previous proof. Remembering that (f /g)g = f and using the previous
theorem, we obtain
Xn
an = dk bk−n
k=0
6.3. SERIES REPRESENTATION OF FUNCTIONS 103
∞
X zn
ez = kzk < ∞ (6.33)
n=0
n!
∞
1 iz −iz
X (−1)n 2n+1
sin(z) = e −e = z kzk < ∞ (6.34)
2i n=0
(2n + 1)!
∞
d X (−1)n 2n
cos(z) = sin(z) = z kzk < ∞ (6.35)
dz n=0
(2n)!
∞
X z 2n
cosh(z) = cos(iz) = kzk < ∞ (6.36)
n=0
(2n)!
∞
d X z 2n+1
sinh(z) = cosh(z) = kzk < ∞ (6.37)
dz n=0
(2n + 1)!
∞
1 X
= zn kzk < 1 (6.38)
1 − z n=0
∞
1 X
= (−1)n z n kzk < 1 (6.39)
1 + z n=0
∞
1 X
= (−1)n (z − 1)n kz − 1k < 1 (6.40)
z n=0
∞
X s
(1 + z)s = zn s ∈ C, kzk < 1 (6.41)
n
n=0
∞
1
X 1
ez = n
0 < kzk < ∞ (6.42)
n=0
n!z
CHAPTER 6. SEQUENCES, SERIES AND RESIDUES 104
§ 6.4 Residues
Definition 6.4.4 (Winding Number). Given a closed curve {γ} we define the winding number or
index of the curve around a point z0 the following integral
˛
1 dz
n(γ, z0 ) = (6.46)
2πi γ z − z0
Theorem 6.18 (Residue Theorem). Given a function f : G −→ C such that f ∈ H(D) where
D = G \ {z1 , · · · , zn } and zk are isolated singularities, we have, taken a closed piecewise
smooth curve {γ}, such that {z1 , · · · , zn } ⊂ {γ}◦
˛ ∞
X
f (z) dz = 2πi n(γ, zk ) Res f (z) (6.47)
γ z=zk
k=0
6.4. RESIDUES 105
Proof. Firstly we can say that γ ∼ k γk where γk are simple curves around each zk , then since the
P
function is holomorphic in the regions A0r (zk ) with k = 1, · · · , n we can write
∞
X
f (z) = cn (z − zk )n
n=−∞
Therefore, we have
˛ n ˛
X n
X ∞
X ˛
f (z) dz = f (z) dz = cj (z − zk )j dz
γ k=0 γk k=0 j=−∞ γk
We can then use the linearity of the integral operator and write
˛ n −2 ˛ ˛ ∞ ˛
X X dz X
f (z) dz = cj (z − zk )j dz + c−1 + cj (z − zk )j dz
γ j=−∞ γ k γk
z − z k j=0 γk
k=0
Thanks to the Cauchy theorem we already know that the first and last integrals on the RHS must be
null, therefore
˛ n ˛
X dz
f (z) dz = c−1
γ γk z − zk
k=0
Recognizing the definition of residue and the winding number of the curve, we have the assert
˛ X
f (z) dz = 2πi n(γ, zk ) Res f (z)
γ z=zk
k=0
Definition 6.4.5 (Residue at Infinity). Given a function f : G −→ C and a piecewise smooth closed
curve γ. If f ∈ H({γ} ∪ extr {γ}) we have
ffi
1 1
f (z) dz = −2πi Res f (z) = 2πi Res 2 f (6.48)
γ z=∞ z=0 z z
Theorem 6.19. Given a function f : G −→ C as before, if the function has zk singularities with
k = 1, · · · , n
n
X
Res f (z) = Res f (z) (6.49)
z=∞ z=zk
k=1
Definition 6.4.7 (Removable Singularity). Given f (z), z0 as before, we have that z0 is a removable
singularity if bk = 0 ∀k ≥ 1
Definition 6.4.8 (Essential Singularity). Given f (z), z0 as before, we have that z0 is an essential
singularity if bk 6= 0 for infinite values of k
Definition 6.4.9 (Meromorphic Function). Let f : G ⊂ C −→ C be a function. f is said to be
meromorphic if f ∈ H(G̃) where G̃ = G \ {z1 , · · · , zn } where zk ∈ G are poles of the function
Theorem 6.20. Let z0 be an isolated singularity of a function f (z). z0 is a pole of order m if
and only if
1 dm−1 g
f (z) = g ∈ H(B (z0 )) > 0 (6.50)
(m − 1)! dz m−1 z0
Proof. Let f : G −→ C be a meromorphic function and g : G −→ C, g ∈ H(G) where f (z) has a pole
in z0 ∈ G and g(z0 ) 6= 0
g(z)
f (z) =
(z − z0 )m
Since g(z) is holomorphic in z0 we have that, for some r
∞
X 1 dk g
g(z) = (z − z0 )k z ∈ Br (z0 )
k! dz0k
k=0
g(z) is holomorphic in the previous domain of expansion, and therefore we have, since the Taylor
expansion is unique
1 dm−1 g
c−1 = = Res f (z)
(m − 1)! dz0m−1 z=z0
6.4. RESIDUES 107
g(z)
f (z) = g(z0 ) 6= 0, g ∈ H(G)
(z − z0 )m
Proof. Taken f (z) = (z − z0 )m g(z) such that g(z0 ) 6= 0 we can expand g(z) with Taylor and at the
end obtain
∞
X 1 dk g
f (z) = (z − z0 )k+m
k! dz0k
k=0
Since this is a Taylor expansion also for f (z) we have that, for j = 1, · · · , m − 1
dj f dm f
=0 = m!g(z0 ) 6= 0
dz0j dz0m
The same is obtainable with the vice versa demonstrating the theorem
Notation. Let f be a meromorphic function. We will define the following sets of points accordingly
Theorem 6.22. Let f : D −→ C be a function such that f ∈ H(D), with D an open set, then
1. f (z) = 0 ∀z ∈ D
Proof. 3) =⇒ 2)
Take z0 ∈ D as the limit point of Zf . Since f ∈ C(D) we have that z0 ∈ Zfm . therefore
Therefore
f (z) 6= 0 ∀z ∈ A0δ (z0 )
2) =⇒ 1)
Suppose that Zf (k) := z ∈ D| f (k) (z) = 0 6= {}. We have to demonstrate that this set is clopen in
D.
Take z ∈ Zf (k) and a sequence (z)k ∈ Zf (k) such that zk → z. We have then
∞
(
X
k z=w
∀w ∈ Br (z), z 6= w f (w) = ak (w − z) = 0 =⇒
k=0
ak = 0 ∀k ≥ 0
Since w 6= z we have that Br (z) ⊂ Zf (k) and the set is open. Taking both results we have that the set
is clopen and D = Zf (k)
Corollary 6.4.1. Let f, g : D −→ C and f, g ∈ H(D). We have that f = g iff the set {f (z) = g(z)}
has a limit point in D
∀z ∈ Zfm m<∞
Proof. Take z0 ∈ Zf , then since f is non-constant we have that Zf has no limit points in D, therefore
dk f dm f
∃δ > 0 : f (z) 6= 0 ∀z ∈ A0δ (z0 ) ∧ ∃m ≥ 1 : = 0 k ∈ [0, m), 6= 0
dz0k dz0m
Therefore z0 ∈ Zfm
p(z)
f (z) = p, q ∈ H(D)
q(z)
Proof.
g(z)
z0 ∈ Pfm =⇒ f (z) = , z0 ∈
/ Zg
(z − z0 )m
Then
1 (z − z0 )
lim = lim =0
z→z0 f (z) z→z0 g(z)
Theorem 6.26. If z0 ∈ Pf1 , ∃ > 0 such that f ∈ A0 (z0 ) and kf (z)k ≤ M , ∀z ∈ A0 (z0 )
Proof. By definition we have that
∃r > 0 : f ∈ H(A0 (z0 ))
And therefore the function is Laurent representable in this set as follows
∞
X
f (z) = ck (z − z0 )k 0 < kz − z0 k <
k=0
Lemma 6.4.1 (Riemann). Take a function f ∈ H (A0 (z0 )) for some > 0, then if kf (z)k ≤ M ∀z ∈
A0 (z0 )
The point z0 is a removable singularity for f
Proof. In the set of holomorphy the function is representable with Laurent, therefore
∞ ∞
X X c−
f (z) = c+ k
k (z − z0 ) +
k
(z − z0 )k
k=0 k=1
ffi
1
c−
k = f (z)(z − z0 )k−1 dz
2πi γ
c− k
k ≤ρ M →0 ∀k ≥ 1
Therefore z0 ∈ Pf1
Theorem 6.27 (Quick Calculus Methods for Residues). Let f be a meromorphic function, then
1. z0 ∈ Pfn then
1 dn−1 n
Res f (z) = lim (z − z0 ) f (z) (6.51)
z=z0 (n − 1)! z→z0 dz n−1
2. z0 ∈ Pfm and f (z) = p(z)/(z − z0 )m , where p ∈ Ck [z] with k ≤ m − 2 and p(z0 ) 6= 0, then
p(z)
Res f (z) = Res =0
z=z0 z=z0 (z − z0 )m
Definition 6.5.1 (Improper Integral). An improper integral is defined as the integral of a function in
a domain where such function has a divergence, or where the interval is infinite. Some examples of
6.5. APPLICATIONS OF RESIDUE CALCULUS 111
such integrals, given a function f (x) with divergences at a, b ∈ R are the following
ˆ ∞ ˆ R
f (x) dx = lim f (x) dx
c R→∞ c
ˆ d ˆ d
f (x) dx = lim f (x) dx
−∞ R→∞ −R
ˆ ∞ ˆ ˆ R
f (x) dx = f (x) dx = lim f (x) dx
−∞ R R→∞ −R
ˆ b ˆ b−
f (x) dx = lim+ f (x) dx
a →0 a+
ˆ h ˆ a− ˆ h
!
f (x) dx = lim f (x) dx + f (x) dx a ∈ (e, h)
e →0+ e a+
Definition 6.5.2 (Cauchy Principal Value). The previous definitions give rise to the following definition,
the Cauchy principal value. Given an improper integral we define the Cauchy principal value as
follows
Let f (x) be a function with a singularity c ∈ (a, b), and g(x) another function then
ˆ ∞ ˆ ˆ R
PV g(x) dx = PV g(x) dx = lim g(x) dx
−∞ R R→∞ −R
ˆ b ˆ ˆ !
c− b
PV f (x) dx = lim f (x) dx + f (x) dx
a →0+ a c+
Notation (Circumferences and Parts of Circumference). For a quick writing of the integrals in this
section, we will use this notation for the following circumferences
CR (t) = Reit t ∈ [0, 2π]
it
CRαβ = Re t ∈ [α, β]
+ it
CR (t) = Re t ∈ [0, π]
− −it
CR (t) = Re t ∈ [0, π]
C˜± R = ±
CR × [−R, R]
lim f (z) = 0
z→∞
Hypothesis 3. Let f (z) = g(z)h(z) with g(z) a meromorphic function such that Sg 6⊂ R+ and
1. h ∈ H(C \ R+ )
2. limz→∞ zf (z) = 0
6.5. APPLICATIONS OF RESIDUE CALCULUS 113
3. limz→0 zf (z) = 0
Definition 6.5.3 (Pacman Path). Let ΓRr be what we will call as the pacman path, this path is formed
by 4 different paths
Where
∆h(x) = lim+ (h(x + i) − h(x − i)) (6.60)
→0
h(z) ∆h(x)
1
− 2πi log+ (z) 1
2
log+ (z) −2πi log(x) + 4π 2
(6.61)
log+ (z) − 2πi log+ (z) −4πi log(x)
i 2 1
4π log+ (z) + 2 log+ (z) log(x)
[z α ]+ xα 1 − e2πiα
All the previous integrals are solved through a direct application of the residue theorem.
Theorem 6.34 (Integrals of Rational Functions). Let f (x) = pn (x)/qm (x) with m ≥ n + 2 and
qm (x) 6= 0 ∀x ∈ R, then ˆ
pn (x) X pn (z)
dx = 2πi Res (6.63)
q
R m (x) z=zk qm (z)
k
Lemma 6.5.1 (Jordan’s Lemma). Let f (z) be a holomorphic function in A := { z ∈ C| kzk > R0 , Im(z) ≥ 0}.
Taken γ(t) = Reit 0 ≤ t ≤ π with R > R0 .
If ∃MR > 0 : kf (z)k ≤ MR ∀z ∈ {γ} and MR → 0, we have that
ˆ
PV f (z)eiaz dz = 0 a > 0 (6.64)
γ
Theorem 6.35. Let f (x) = pn (x)/qm (x) and m ≥ n + 1 with qm (x) 6= 0 ∀x ∈ R, then ∀a > 0 we
have that ˆ
pn (x) iax X pn (z) iaz
e dx = 2πi Res e (6.65)
R mq (x) z=zk q m (z)
k
Lemma 6.5.2. Let f (z) be a meromorphic function such that z0 ∈ Pf1 and γr± are semi circumferences
parametrized as follows
γr± (t) = z0 + re±iθ θ ∈ [−π, 0]
Then ˆ
PV f (z) dz = ±πi Res f (z) (6.66)
γr± z=z0
Theorem 6.36. Let f (x) = pn (x)/qm (x) with m ≥ n + 2 and qm (x) has xj ∈ Zg1 R
then
ˆ
pn (x) X pn (z) X pn (z)
dx = 2πi Res + πi Res (6.67)
R qm (x) z=zk qm (z)
j
z=xj qm (z)
k
We can imagine a sequence as a point in a space. We will call this space FN , and an element of this
space will be indicated as follows
Therefore, every point in FN is a sequence. Note that the infinite sequence of 0s and 1s will be indicated
as 0 = (0)n , 1 = (1)n
Definition 7.1.3 (Sequence of Sequences). We can see a sequence of sequences as a mapping from N
to the space FN , as follows
x :N −→ FN
n → ((x)k )n
It’s important to note how there are two indexes, since every element of the sequence is a sequence in
itself (i.e. ((x)k )n ∈ FN for any fixed n ∈ N)
Definition 7.1.4 (Convergence of a Sequence of Sequences). A sequence of sequences is said to
converge to a sequence in FN if and only if
115
CHAPTER 7. HILBERT AND BANACH SPACES 116
2) kc(x)n k∞ = |c|k(x)n k∞
sup |(x)n + (y)n | ≤ sup (|(x)n | + |(y)n |) = sup |(x)n | + sup |(y)n | = k(x)n k∞ + k(y)n k∞
n∈N n∈N n∈N n∈N
Since `∞ (F) is a vector space, the couple (`∞ (F), k·k∞ ) is a normed vector space
Remark. Let V be a vector space over some field F. If dim(V) = ∞, a closed and bounded subset
W ⊂ V isn’t necessarily compact, whereas, a compact subset Z ⊂ V is necessarily closed and bounded.
7.1. BANACH SPACES 117
We have that diam(B1 ) = 2, therefore this set is bounded and closed by definition.
Take the canonical sequence of sequences ((e)k )n , defined as follows:
Therefore, ∀n 6= m
((e)k )n − ((e)k )m ∞
= k(1)k k∞ = 1
Therefore there aren’t converging subsequences, and therefore B1 can’t be compact.
Proposition 11. `0 (F) ⊂ `∞ (F), and the couple (`0 (F), k·k∞ ) is a normed vector space, where the
norm k·k∞ gets induced from the space `∞ (F)
Proof.
lim (x)k = 0 =⇒ ∀ > 0 ∃N ∈ N : |(x)n | < ∀n ≥ N
k→∞
∴ sup |(x)n | = ≤ M ∈ F =⇒ (x)n ∈ `∞ (F), ∴ `0 (F) ⊂ `∞ (F)
n∈N
`p (F) Spaces
Definition 7.1.8. The sequence space `p (F) is defined as follows
p
`p (F) := { (x)n ∈ FN k(x)n kp = M ∈ F} (7.6)
Proposition 12. The application k·kp : `p (F) −→ F is a norm in `p (F), and the couple (`p (F), k·kp ) is a
normed vector space
Proof. We begin by proving that `p (F) is actually a vector space, therefore 1) ∀(x)n , (y)n ∈ `p (F), (x)n +
(y)n = (x + y)n ∈ `p (F)
∞
X p p
(x + y)n ∈ `p (F) =⇒ |(x)n + (y)n | = k(x)n + (y)n kp < M ∈ F
n=0
p p p
k(x)n + (y)n kp ≤ k(x)n kp + k(y)n kp < M ∈ F
CHAPTER 7. HILBERT AND BANACH SPACES 118
Definition 7.1.10 (Some Function Spaces). We are already familiar from the basic courses in one
dimensional real analysis, about the space of continuous functions C(A), where A ⊂ R. We can define
three other spaces directly, adding some restrictions.
Due to the properties of continuous functions, these spaces are obviously vector spaces.
Proof. The inclusion of these spaces is obvious, due to the definition of these. For the proof that the
application k·ku is a norm, it’s immediately given from the proof that the application k·k∞ is a norm in
`∞ (F), and that k·ku = k·k∞
Remark. Take fn ∈ Cb (F) a sequence of functions. The uniform convergence of this sequence means
that fn → f in the norm k·ku = k·k∞
Proposition 14. If f ∈ C0 (F), then f is uniformly continuous
Proof. Let f ∈ C0 (F), then
∀ > 0 ∃l : |x| ≥ l =⇒ |f (x)| <
2
Since every continuous function is uniformly continuous in a closed set, then
Hence we can have two cases. We either have |x − y| < δ or x, y ∈ [−L − 1, L + 1]. Hence we have,
in the first case
|f (x) − f (y)| ≤ |f (x)| + |f (y)| <
Or, in the second case
Cp (F) spaces
Definition 7.1.11. We can define a set of function spaces analogous to the `p (F) spaces. These spaces
are the Cp (F) spaces. We define analogously the p−norm for functions as follows
sˆ
p p
kf kp := |f (x)| dx (7.10)
F
Thanks to what said about `p (F) spaces and p−norms, it’s already obvious that these spaces are normed
vector spaces
Remark. Watch out! Cp (F) 6⊂ C0 (F), and Cp (F) 6⊂ Cq (F) for 1 ≤ p ≤ q. It’s easy to find counterex-
amples
Proposition 15. If 1 ≤ p ≤ q, then
kf kα,β := xα ∂ β f ∞
= sup xα ∂ β f (x) (7.11)
x∈F
Definition 7.1.13 (Schwartz Space). The space S(Rn ) is called the Schwartz space, and it’s defined as
follows n o
S(Rn ) := f ∈ C ∞ (Rn )| kf kα,β < ∞, α, β multiindices (7.12)
Example 7.1.3. Taken p(x) ∈ R[x] a polynomial, a common example of functions f (x) ∈ S(R) is the
following.
2n
f (x) = p(x)e−a|x| (7.13)
With a > 0, n ∈ N
Theorem 7.2. A function f ∈ C ∞ (Rn ) is in S(Rn ) if ∀β multiindex, ∀a > 0 ∃Cα,β such that
Cα,β
∂ β f (x) ≤ a2 ∀x ∈ Rn (7.14)
2
1 + kxk
Therefore
kf kj,k ≤ Cj,k < ∞ =⇒ f ∈ S(R)
Taken f ∈ S(R) we have that, if |x| ≥ 1
a a a
(1 + x2 ) 2 ≤ 2 2 |x|
Taken j = dae
a a
Definition 7.1.14 (Space of Compact Support Function). Given a function with compact support f , we
define the space of compact functions Cc∞ (Rn ) as the space of all such functions.
We have obviously that Cc∞ (Rn ) ⊂ S(Rn )
Theorem 7.3. Both Cc∞ (Rn ) and S(Rn ) are dense in (Cp (Rn ), k·kp )
7.2. HILBERT SPACES 121
Definition 7.2.2 (Hilbert Space). Given (V, h·, ·i) an euclidean space. It’s said to be a Hilbert space if
it’s complete
Theorem 7.5 (Cauchy-Schwartz Inequality). Given (V, h·, ·i) a complex euclidean space, then
∀u, v ∈ V
2
khu, vik ≤ hu, uihv, vi (7.16)
CHAPTER 7. HILBERT AND BANACH SPACES 122
Proof. Taken t ∈ C, we define p(t) = htu + v, tu + vi. Then by definition of the Hermitian product,
we have
2
p(t) = ktk hu, ui + thu, vi + thv, ui + hv, vi
Writing hu, vi = ρeiθ , t = se−iθ we have
Theorem 7.6 (Induced Norm). Given a Hermitian product h·, ·i we can define an induced norm
k·k by the definition p
k·k = h·, ·i (7.17)
Theorem 7.7. Addition, multiplication by a scalar and the scalar product are continuous in an
euclidean space V, then, given two sequences un −→ u ∈ V, vn → v ∈ V and
un + vn −→ u + v
c ∈ C =⇒ cun −→ u
hun , vn i −→ hu, vi
|hu, vi − hun , vn i| = |hu, vi − hu, vn i + hu, vn i − hun , vn i| ≤ |hu, vi − hu, vn i| + |hu, vn i − hun , vn i| =
= kukkv − vn k + kvn kku − un k
Since the successions are convergent, we have that ∃M > 0 : kvn k ≤ M ∀n ∈ N, therefore
The spaces C2 aren’t complete therefore they’re not Hilbert spaces. The spaces L2 (C) and the
weighted alternative are the completion of such spaces and are therefore Hilbert spaces
Theorem 7.8 (Polarization Identity). Given a complex euclidean space (V, h·, ·i) we have, ∀u, v ∈ V
1 2 2
2 2
hu, vi = ku + vk − ku − vk + i ku + ivk ku − ivk (7.18)
4
Theorem 7.9 (Parallelogram Rule). Let (V, k·k) be a normed vector space. A necessary and
sufficient condition that the norm is induced by a scalar product is that
2 2 2 2
ku + vk + ku − vk = 2 kuk + kvk ∀u, v ∈ V (7.19)
§ 7.3 Projections
§§ 7.3.1 Orthogonality
Definition 7.3.1 (Angle). Given a real euclidean space (V, h·, ·i) we define the angle θ = u∠v as follows
hu, vi
θ = arccos (7.20)
kukkvk
Definition 7.3.2 (Orthogonal Complement). Given an euclidean vector space V and two vectors u, v,
we say that the two vectors are orthogonal u ⊥ v if
hu, vi = 0 (7.21)
CHAPTER 7. HILBERT AND BANACH SPACES 124
X ⊥ := { v ∈ V| hv, wi = 0 ∀w ∈ X} (7.22)
Theorem 7.10. Given X ⊂ V with V an euclidean space, the set X ⊥ is a closed subspace of V
Proof. X ⊥ is a subspace, hence ∀v1 , v2 ∈ X ⊥ and c1 , c2 ∈ C we have
Hence c1 v1 + c2 v2 ∈ X ⊥ .
Given a sequence (v)n ∈ X ⊥ : (v)n −→ v ∈ V we have, given w ∈ X
hvn , wi = 0 ∀n ∈ N
X ⊂ Y =⇒ Y ⊥ ⊂ X ⊥
⊥
X ⊥ = (X)⊥ = span(X)
hv, yi = 0 ∀y ∈ Y
X ⊥ ⊃ (X)⊥ ⊃ (span(X))⊥
Now take w ∈ span(X). Take a sequence (w)n ∈ span(X) such that (w)n −→ w. Thanks to the
continuity of the scalar product we have
hv, wi = hv, lim wn i = lim hv, wn i = 0
n→∞ n→∞
Proof. Take d = infw∈W kv − wk. By definition of infimum we have that ∃(w)n ∈ W such that
lim kv − wn k = d
n→∞
And therefore
1 2 2
2 1 2
d2 = kv − w1 k + kv − w2 k = kv − w3 k + kw1 − w2 k
2 4
I.e. if w1 6= w2 , w3 is the infimum between v ∈ V and W
CHAPTER 7. HILBERT AND BANACH SPACES 126
Proof. Given v ∈ V, due to the previous lemma we have that ∃!w ∈ W such that
d = kv − wk ≤ kv − w0 k ∀w0 ∈ W
Taken z = v − w, and an element x ∈ W, define the vector w + tx with t ∈ C. Since W is a subspace
w + tx ∈ W and ∀t ∈ C we have
2 2 2 2
d2 ≤ kv − (w + tx)k = kv − wk − thv − w, x, −ithx, v − wi + ktk kxk
Writing hx, v − wi = khx, v − wikeiθ and t = se−iθ with s ∈ R we have
2
−2skhv − w, xik + s2 kxk ≥ 0 ∀s ∈ R
Which implies
hv − w, xi = 0 =⇒ z = v − w ∈ W ⊥
Therefore there exists a representation v = w + z with w ∈ W, z ∈ W ⊥ Now, we suppose that
v = w0 + z 0 , then
0 = (w − w0 ) + (z − z 0 )
Therefore
2 2
0 = k(w − w0 ) + (z − z 0 )k = kw − w0 k + kz − z 0 k
Therefore the representation is unique.
Theorem 7.13. If W ⊂ V with V a Hilbert space, we have
⊥
W⊥ = W
If W is closed ⊥
W⊥ =W
Definition 7.3.3 (Orthogonal Projection). Given a closed subspace W ⊂ V we can define an operator
π̂W : V −→ W such that
π̂W v = w ⇐⇒ w ∈ W, v − w ∈ W ⊥ (7.23)
π̂W is linear and called a orthogonal projection
Theorem 7.14. Given W ⊂ V a closed subspace of the Hilbert space V, then given an orthogonal
projection π̂W : V −→ W we have, ∀v, z ∈ V and another closed subspace Z ⊂ V
1. π̂W
2
= π̂W
Definition 7.3.4 (Direct Sum). An euclidean space V is called the direct sum of closed subspaces
Vi ⊂ V and it’s indicated as follows
∞
M
V = V1 ⊕ V2 ⊕ · · · = Vk (7.24)
k=1
If
V = W ⊕ W⊥ (7.25)
Theorem 7.15. Given an orthogonal system X ⊂ V with (V, h·, ·i) an euclidean space, then we
have that X is a system of linearly independent vectors
Definition 7.3.7 (Basis). Given an orthogonal and complete set of vectors (v)α in an euclidean space V
it’s said to be an orthogonal basis of V. If it’s an orthonormal and complete set of vectors it’s said to
be an orthonormal basis of V
CHAPTER 7. HILBERT AND BANACH SPACES 128
Lemma 7.3.2. Given an orthogonal system (v)nk=1 ∈ V and let u ∈ V an arbitrary vector. Then given
z ∈ V as follows
n
X hu, vk i
z := u − 2 vk
k=1
kvk k
We have that z ⊥ vi ∀1 ≤ i ≤ n and therefore z ⊥ span {v1 , · · · , vn }
Proof. ∀i = 1, · · · , n it’s obvious that hz, vi i = 0, therefore
⊥
z ∈ {v1 , · · · , vn }
Therefore
⊥ ⊥ ⊥
z ∈ {v1 , · · · , vn } = span {v1 , · · · , vn } = span {v1 , · · · , vn }
Proof. Defining
n−1
X hvn , vk i wn
wn = vn − 2 wk un =
kwk k kwn k
k=1
⊥
We can say immediately that ∀n ≥ 1 wn ∈ {w1 , · · · , wn−1 } . By induction we can say that it holds
∀(w)n∈N , therefore (w)n is an orthogonal system and (u)n∈N is an orthonormal system
We can also say that
k−1
X
vn = wn + βnj wj
j=1
The final result will be called the Legendre Polynomials We begin by taking the canonical scalar
product in C2 [−1, 1] as follows
ˆ 1
hf, gi = f (x)g(x) dx
−1
And using that
ˆ 2
1 n = 2k ∈ N
xn dx = n+1
−1
0 n = 2k ∈ N
Therefore, we have that
ˆ 1
2
w0 = 1 kw0 k = dx = 2
−1
ˆ 1
1 2 −1 2 2
w1 = x − x dx = x kw1 k = x dx =
2 −1 1 3
ˆ 1 2
1 2 1 8
w2 = x2 − kw2 k = x2 dx =
3 −1 3 45
..
.
Normalizing, we find
1
p0 = √
2
r
3
p1 (x) = x
2
r r
45 2 1 1 5
3x2 − 1
p2 (x) = x − =
8 3 2 2
..
.
And so on. The set (p)n is called the set of Legendre polynomials 2) Hermite Polynomials
2
Using the same procedure, we can find the Hermite polynomials Hn (x) in the space C2 (R, e−x dx).
The first 5 are the following
H1 = 1
H2 (x) = x
1
H3 (x) = x2 −
2
3 3
H4 (x) = x − x
2
3
H5 (x) = x4 − 3x2 +
4
CHAPTER 7. HILBERT AND BANACH SPACES 130
span {(w)∞ ∞
n=1 } = span {(v)n=1 }
span {(w)∞ ∞
n=1 } = span {(v)n=1 } = V
Orthonormalizing the system (w)n∈N with the Gram-Schmidt procedure I obtain then a new set (u)n∈N
such that
span {(u)n∈N } = span {(w)n∈N } = V
(u)n∈N is complete and therefore a basis.
Remark. If V is a complete euclidean space but not separable, we can find thanks to Zorn’s lemma a
maximal orthonormal basis, but
1. There isn’t a standard procedure for finding this basis
2. Taken the basis (u)α∈I it can’t be numerable. If it was then V = span {(u)α }. Taken X =
spanQ {(u)k } as the set of finite linear combinations with rational coefficients, we have that
X = span {(u)k } and therefore X = V contradicting the fact that V is not separable
8 Distributions
Definition 8.1.2 (Bounded Linear Functional). Given f ∈ V ? , we call f a bounded linear functional
if, ∀x ∈ V
sup |f (x)| < ∞ (8.2)
kxk≤1
Definition 8.1.3 (Dual Norm). Given a normed vector space (V, k·k) we can give a structure of normed
vector space to its dual with the couple (V ? , k·k? ), where the application k·k? is called the dual norm
k·k? : V ? −→ F
|f (v)|
kf k?1 = sup (8.3a)
v6=0 kvk
131
CHAPTER 8. DISTRIBUTIONS 132
Theorem 8.1. For a bounded linear functional f ∈ V ? the two definitions of the dual norm
coincide, i.e.
kf k?1 = kf k?2 = kf k?
And we can define the following inequality
Proof. Since f ∈ V ? is bounded we have that both norms exist and must be finite. We can therefore
write, thanks to the homogeneity of f , taken v ∈ V
v
kf k?1 = sup f = sup |f (v)| ≤ sup |f (v)| = kf k?2
v6=0 kvk kvk=1 kvk≤1
|f (v)| |f (v)|
kf k?2 = sup |f (v)| ≤ sup ≤ sup = kf k?2
kvk≤1 0<kvk≤1 kvk v6=0 kvk
kf k?1 = kf k?2 = kf k?
Theorem 8.2. Given f ∈ V ? with V normed vector space, we have that the following assump-
tions are equivalent
1. f is continuous
3. f is bounded
Proof. 1) =⇒ 2)
Since f ∈ V ? is linear by definition, we have that it’s also continuous and injective, therefore
2) =⇒ 3)
Since f is continuous at the origin, we have by definition of continuity and limit
∃x ∈ V : |f (x)| ≥ (C − kf k? )kxk
Definition 8.1.4 (Kernel). Given a linear functional f ∈ V ? we define the kernel of f as the set of
zeros of f , i.e.
ker f = { v ∈ V| f (v) = 0} (8.5)
Theorem 8.3 (Riesz Representation Theorem). Given (V, h·, ·i) a Hilbert space, we can uniquely
∼
define its dual V ? through the isomorphism φv : V −→ F defined as follows
1) ∀v ∈ V φv ∈ V ? , kφv k? = kvk
2) ∀c1 , c2 ∈ F, v1 , v2 ∈ V φc1 v1 +c2 v2 = c1 φv1 + c2 φv2 (8.6)
?
3) f ∈ V =⇒ ∃v ∈ V : f = φv
CHAPTER 8. DISTRIBUTIONS 134
f (v) = 0 ∀v ∈ V ∴f =0
By definition of φv , then
Then
f (w) f (w) f (u)
φv (u) = 2 hw, u1 i + 2 hw, wi = f (u)
kwk kwk f (w)
Therefore
φv (u) = f (u) ∀f ∈ V ? , ∀u ∈ V
8.2. DISTRIBUTIONS 135
§ 8.2 Distributions
§§ 8.2.1 Local Integrability
Definition 8.2.1 (Weak Convergence). Given a normed vector space (V, k·k), a sequence (v)k∈N ∈ V
is said to be weakly convergent to v ∈ V and it’s indicated as vk *w v if
Theorem 8.4. Given a normed vector space V and a sequence (v)k∈N ∈ V such that vk → v ∈ V
(i.e. converges strongly), we have that
vk → v =⇒ vk *w v
vk → v ⇐⇒ lim kvk − vk = 0
k→∞
Now, writing the definition of weak convergence and applying the linearity of the limit
Therefore
∀f ∈ V ? lim (f (vk − v)) ≤ kf k? lim kvk − vk
k→∞ k→∞
And therefore
vk → v =⇒ vk *w v
Proof. Take (V, k·k) = (`2 , k·k2 ) and take the standard sequence (ei )j where
Now, considering that (`2 , k·k2 ) is a Hilbert space, we have that, due to Riesz representation theorem
∀f ∈ `2? f = h·, ai = φa a ∈ `2
Therefore
∞
X
f (eij ) = heij , ai i = ai eij = aj → 0
i=1
Therefore
∀(a)j ∈ `2 ∀f ∈ `2? lim f (eij ) = 0
j→∞
3.
∃ lim+ f (x) = f (a+ )
x→a
(8.8)
∃ lim− f (x) = f (b− )
x→b
Definition 8.2.4 (Jump). Given a piecewise continuous function f we define the jump of the function
at a discontinuity x ∈ Sf as follows
The set of locally integrable functions forms a subspace of L1 (R) and it’s indicated as L1loc (R)
Theorem 8.5. Let f : R −→ C be a piecewise continuous function, then f ∈ L1loc (R)
Proof. Take f (x) a piecewise continuous function, then ∀a, b ∈ R ∃(u)0≤k≤n ∈ R : u0 = a, un = b
for which we have f ∈ C((uk−1 , uk )) ∀0 ≤ k ≤ n. I can therefore define f̃k : [uk−1 , uk ] −→ C, where
x ∈ (uk−1 , uk )
f (x)
+
f̃k (x) = f (uk−1 ) x = uk−1
f (u−
k) x = uk
We have that
∞
X
f (x) = f̃k (x)
k=0
And therefore ˆ b n ˆ
X uk n ˆ
X uk
|f (x)| dx = |f (x)| dx = f̃k (x) dx
a k=0 uk−1 k=0 uk−1
Therefore
ˆ b X
(f g)(b− ) − (f g)(a+ ) = (f g)0 (x) dx + f (x− )∆g(x) + g(x− )∆f (x) + ∆f (x)∆g(x)
a x∈Sf ∪Sg
1. ∃I ⊂ R : ∀x ∈ I c fn (x) = 0
(k)
2. ∀k ∈ N fn (x) ⇒ f (n) (x)
Then ϕg ∈ K?
Proof. Using the fact that f ∈ K we can immediately say that |f (x)| ≤ M ∀x ∈ [−a, a] ⊂ R, therefore
ˆ ˆ a ˆ a
|g(x)f (x|) dx = |g(x)f (x)| dx ≤ M |g(x)| dx < ∞
R −a −a
Definition 8.2.11 (Ceiling and Floor Functions). We define the ceiling function d·e : R −→ Z and the
floor functions b·c : R −→ Z as follows
bxc = max { m ∈ Z| m ≤ x}
(8.10)
dxe = min { n ∈ Z| n ≥ x}
Example 8.2.1 (Floor and Ceiling Distributions). Since both the floor and ceiling distributions are locally
integrable, we can build a regular distribution ϕ ∈ K? as follows
ˆ Xˆ k+1
ϕbxc (f ) = f (x)bxc dx = kf (x) dx
R k∈Z k
ˆ Xˆ k+1
ϕdxe (f ) = f (x)dxe dx = (k + 1)f (x) dx
R k∈Z k
Example 8.2.2 (Theta Distribution). Given H ∈ L1loc (R) we can define the associated theta distribu-
tion ϕH = ϑ ∈ K, as follows
ˆ ˆ
ϑ(f ) = f (x)H(x) dx = f (x) dx
R R+
It’s already obvious that ϕH̃ = ϕH , but it’s better to formalize it in the following theorem
Theorem 8.8. Let f, g ∈ L1loc (R) such that { x ∈ R| f (x) 6= g(x)} = {u1 , · · · , un }. Then we have
ϕg = ϕf ∈ K?
CHAPTER 8. DISTRIBUTIONS 140
Therefore
ϕg = ϕh
Definition 8.2.14 (Dirac Delta Distribution). An example of singular distribution is the Dirac delta
distribution δa ∈ K? . This distribution is defined as follows
δa (f ) = f (a) ∀f ∈ K, a ∈ R
Theorem 8.9. Given the Dirac delta distribution δa , @δ(x) ∈ L1loc (R) such that (taken a = 0
without loss of generality) ˆ
δ0 (f ) = f (x)δ(x) dx
R
Proof. Let’s say that ∃δ(x) ∈ L1loc (R), therefore, we could define δ0 = h·, δ(x)i. This function therefore
must have these properties
δ(x) 6= 0 ⇐⇒ x = 0, x ∈ Sδ
But, since for b ∈ R, b 6= 0, we have that δ(x) → δ(b) continuously, i.e. δ(b) = A > 0, therefore
A
∃ > 0 : δ(x) ≥ ∀x ∈ [b − , b + ]
2
But, then ∃f ∈ K such that f (x) → f (b) continuously, f > 0, f (b) = 1. Taken < b we can say
1
∃˜ ∈ (0, ) : f (x) ≥ ∀x ∈ [b − 0 , b + 0 ]
2
8.2. DISTRIBUTIONS 141
Note that ˆ
δ0 (1) = δ(x) dx = 1
R
Theorem 8.10. Given gn ∈ L1loc (R) such that gn ∈ C[−an , an ] with an → 0, an > 0 ∀n ∈ N, we
have that, if ˆ
gn (x) dx = 1
R
Then
lim ϕgn (f ) = δ0 (f )
n→∞
Therefore
lim ϕgn (f ) = δ0 (f )
n→∞
CHAPTER 8. DISTRIBUTIONS 142
And therefore
ˆ u ˆ L ˆ
u u
f − f (0) g(u) du = f − f (0) g(u) du + f − f (0) g(u) du
R n −L n |u|≥L n
By using the properties of the integral, we know that
ˆ u u
f − f (0) g(u) du ≤ sup |f (x) − f (0)| + sup f − |f (0)|
R n x≤ L u∈R n
n
Since f (x) → f (0) continuously, due to the fact that f ∈ K, we have that
ˆ
|f (x) − f (0)|g(nx)n dx ≤ 2kf ku
R
Therefore ˆ
lim f (x)gn (x) dx − δ0 (f ) = 0
n→∞ R
Definition 8.2.15 (PV (x−n ) −Distribution). Taken n > 0, n ∈ N, we have that x−n ∈ / L1loc (x) and
?
therefore there is no associated distribution ϕx−n ∈ K .
A useful thing we could do is utilizing the definition of the Cauchy principal value of the function.
Therefore, we define the following singular distribution
ˆ a n−1
X f (k) (0)
!
1 1 k
PV [f ] = lim f (x) − x dx (8.12)
xn a→∞ −a xn k!
k=0
8.2. DISTRIBUTIONS 143
Using the Lagrange formulation for the remainder of the McLaurin expansion, we have
n−1
X f (k) (0) k
Rn−1 (x) = f (x) − x = xn rn−1 (x)
k!
k=0
Since rn−1 (x) ∈ C ∞ (R) (it’s a polynomial) and supp f ⊂ [−b, b] for some b ∈ R, b > 0, we have that
for R > b
X f (k) (0) ˆ
n−1
1
n n
IR (f ) = Ib (f ) − n−k
dx
k! b≤|x|≤R x
k=0
Therefore
n 2R
|IR (f )| ≤ 2 sup |rn−1 (x)| ≤ kf ku
x∈[−R,R] n!
Which means
n−1
f (j) (0)
1 2b (n) X 2
PV (f ) ≤ f +
xn n! u
j=0
j! (j − 1)bj−1
Taken b1 = max {b, 1} we have that supp f ⊂ [−b1 , b1 ] and whenever j − 1 > 1, b1 ≥ 1
2
≤2
(j − 1)bj−1
1
Therefore, finally
n−1
f (j) (0)
1 2 X
PV (f ) ≤ max {b, 1} f (n) +2
xn n! u
j=0
j!
Now, defined fj ∈ K and hj (x) = f (x) − fj (x) with fj → f we have that supp hj ⊂ [−b, b], ∀j ∈ N.
(n)
By definition we have that hj ⇒ 0 and hj ⇒ 0 ∀n ∈ N
n−1 (j)
1
2 hj (0)
(n)
X
PV (hj ) ≤ max {b, 1} hj +2
xn n! u
j=0
j!
This means
(k)
∀k ∈ N, ∀ > 0, ∃NR ∈ N : ∀k ≥ Nj hj (x) ≤
2(b + 1 + )
Chosen N = max0≤k≤n−1 Nk we have that ∀j ≥ N
n−1
1 2 X 1
PV (hj ) ≤ max {b, 1} + ≤→0
xn 2(b + 1 + ) j=0
j!
Therefore, PV(x−n ) ∈ K?
v(x) = D u(x)
8.2. DISTRIBUTIONS 145
+ : K? × K? −→ K?
· : C × K? −→ K?
(8.13)
◦ : C ∞ × K? −→ K?
D : K? −→ K?
Theorem 8.13. Given g(x) ∈ L1loc (R) and ϕg ∈ K? its associated distribution. Then, ∀h ∈
C ∞, f ∈ K
(hϕg )(f ) = ϕg (hf ) = ϕhg (f ) (8.14)
Theorem 8.14. Taken g ∈ L1loc (R) and g 7→ ϕg ∈ K? its associated distribution, we have that
D ϕg = ϕD g
Theorem 8.15. ∀g, f ∈ L1loc (R) and given ϕg , γf ∈ K? , c ∈ C their associated distributions, we
have that
ϕg + ϕf = ϕg+f
ϕcf = cϕf
CHAPTER 8. DISTRIBUTIONS 146
Proof. The proof is obvious using the linearity of the integral operator
Theorem 8.16 (Useful Identities). Here is a list of some useful identities
xδ0 = 0
1
xPV = ϕ1 (8.15)
x
x D δ0 = −δ0
Proof. 1) xδ0 = 0.
Taken f ∈ K
xδ0 (f ) = δ0 (xf ) = 0f (0) = 0
1
2) xPV xn = ϕ1
Again, taken g ∈ K
ˆ
1 1
x PV (g) = PV (xg) = lim f (x) dx = ϕ1
x x →0 |x|≥0
Notation (Abuse of Notation). Given g ∈ L1loc (R) and its associated distribution g 7→ ϕg , is quite
common to use the original function to indicate actually the distribution. Together with this, the
distributional derivative D gets indicated as an usual derivative, therefore
ϕg → g(x)
dg
D ϕg → g 0 (x) =
dx
Therefore it isn’t uncommon to see identities like this
1
xPV =1
x
Where actually 1 → ϕ1 is the identity distribution.
Or
δ00 = −δ0
This makes an easy notation for calculating distributional derivatives and have some calculations, but
one should watch out to this common abuse of notation
Definition 8.2.17 (K? Convergence). Given (σ)n∈N ∈ K? a sequence of distributions, it’s say to
K? −converge if
(σ)n (f ) → σ(f ) ∈ K?
It’s indicated as follows
(σ)n *? σ
8.2. DISTRIBUTIONS 147
Theorem 8.17. Given κ ∈ K? a distribution, we have that f ∈ C ∞ (K) in the sense of distribu-
tional derivatives, where the n−th derivative is defined as follows
Dn κ(f ) = (−1)n κ(f (n) ) ∀f ∈ K
Proof. Taken f ∈ K we have that f ∈ Cc∞ (R) ⊂ C ∞ (R) hence f is smooth, therefore we have, for
κ ∈ K?
D κ(f ) = −κ(f 0 )
Iterating, we get for n = 2
D2 κ(f ) = − D κ(f 0 ) = κ(f 00 )
Iterating till n we get finally
Dn κ(f ) = (−1)n κ(f (n) )
Hence the derivability depends only on the test function f , and since it’s smooth, the distributional
derivatives can be defined ∀n ∈ N
Example 8.2.3. Take for example the delta distribution δ0 ∈ K? , given f ∈ K we want to calculate
the following value
Dn δ0 (f )
Using the previous identity, we have that
Dn δ0 (f ) = (−1)n δ0 [f (n) ] = (−1)n f (n) (0)
Example 8.2.4 (The Absolute Value Distribution). Taken g(x) = |x| we have that (obviously) |x| ∈
L1loc (R), therefore there exists a distribution ϕ|x| ∈ K? defined as follows
ˆ
ϕ|x| (f ) = |x|f (x) dx
R
We have that the distributional derivative it’s actually the function sgn (x), (in this case, since it’s a locally
integrable function it coincides with its weak derivative), therefore it’s not unusual to see expressions
like this
d
|x| = sgn(x) weakly/distributionally
dx
The proof of this is quite easy. Taken f ∈ K
ˆ
D ϕ|x| (f ) = −ϕ|x| (f 0 ) = − |x|f 0 (x) dx =
R
ˆ ˆ
= [xf (x)]R− − [xf (x)]R+ + f (x) dx − f (x) dx =
R−
ˆ R+
Theorem 8.21. In general, given g ∈ C 1 (R), we can say that g(|x|) has the following weak
derivative
D g(|x|) = D g(|x|) sgn(x)
Proof. Since g(|x|) ∈ L1loc (R) we can say that exists g 7→ ϕg ∈ K? such that
D ϕg(|x|) = ϕD g(|x|)
Where D g is the weak derivative of g(|x|). Since it’s a derivative (and it’s also demonstrable) we can
use the rules for composite derivation, and knowing that D |x| = sgn(x) we have
D ϕg(|x|) = ϕD g(|x|) = ϕD g D |x| = ϕsgn D g
Theorem 8.22 (Derivative of the ϑ Distribution). Given ϑ ∈ K? the theta distribution, we define
ˆ ˆ
ϑ(f ) = f (x)H(x) dx = f (x) dx
R R+
Proof. We have ∀f ∈ K
ˆ ˆ
0 0
D ϑ(f ) = −ϑ(f ) = − f (x)H(x) dx = − f 0 (x) dx = f (0) = δ0 (f )
R R+
Therefore
D ϑ = δ0
Or, using a common abuse of notation
ϑ = ϑ(x) = H(x), D H(x) = δ(x)
And, ∀g ∈ K, taken f 7→ ϕf ∈ K?
ˆ u ˆ ∞
0 0
D ϕf (g) = −ϕf (g ) = − g (x)f (x) dx − g(x)f (x) dx
−∞ u
And therefore
D f (x) = D◦ f + ∆f (u)δ(u)
Example 8.2.5 (Deriving the Sign Function). Take a ∈ R and the function sgn(x − a), using the
previous formula we have that Ssgn(x−a) = {a} and ∆ sgn(a) = 2, therefore
D sgn(x − a) = 2δ(x − a)
And equivalently
D sgn(a − x) = −2δ(x − a)
Example 8.2.6 (A General Piecewise Differentiable Function). Take the function g : R −→ R defined
as follows (
1 x<0
g(x) =
x−2 x≥0
We have that (
0 x≤0
D◦ g(x) = H(x) =
1 x>0
Since the discontinuity is in the origin and we have ∆g(0) = −3 we have
Example 8.2.7 (Derivative of the Floor Function). Given the floor function bxc : R −→ Z we have that
Sbxc = Z, ∆bxc = 1 and D◦ bxc = 0 therefore
X ∞
X
Dbxc = δ(x − k) = δ(x − k)
k∈Z k=−∞
n √ o
Given instead a different function, bx2 c we have that D◦ bx2 c = 0 and Sbx2 c = k ∈ Z| ± k ⊂R
√
with ∆b± kc = ±1, therefore
√ ∞ √
X √ X √
Dbx2 c = δ(x − k) − δ(x − −k) = δ(x − k) − δ(x − −k)
k≥1 k=1
D7 xex D2 δ0
d2 d x
xex D2 δ0 = (xex )δ0 = (e + xex )δ0 = (2 + x)ex δ0
dx2 dx
CHAPTER 8. DISTRIBUTIONS 152
D7 xex D2 δ0 = 2 D7 (δ0 − D δ0 ) = 2 D7 δ0 − D8 δ0
And
Di (xm )(0) = m!δm
i
Theorem 8.26 (Properties of the PV(x−n ) Distribution). Here there will be listed some properties
of the PV(x−n ) distribution
1. D log |x| = PV(x−n )
2. Given n, m ∈ N, then
m−n
x
m≥n
1
xm PV =
1
xn PV
m<n
xn−m
3. Dm PV(x−n ) = (−m)!PV(x−n−m )
Proof. 1) Taken log |x| ∈ L1loc (R) we know that ∃ϕlog |x| ∈ K? such that log |x| 7→ ϕlog |x| and therefore
we can write as follows the previous derivative, that ∀f ∈ K
ˆ ˆ ∞
D ϕlog |x| (f ) = −ϕlog |x| (f 0 ) = − lim f 0 (x) log(−x) dx − lim f (x) log(x) dx
→0− −∞ →0+
8.2. DISTRIBUTIONS 153
Or
1
D log |x| = PV
xn
2) Taken f ∈ K
ˆ R n−1
!
Dk (xm f )(0) k
1 1 1 X
xm PV (f ) = PV (x m
f ) = lim m
x f (x) − x dx
xn xn R→∞ −R xn k!
k=0
Taken m ≥ n we have that the sum is null ∀k ≤ m, therefore since k ≤ n − 1 < m it sums to 0
Therefore ˆ
n m 1 m
lim IR (x f ) = PV (x f ) = xm−n f (x) dx = ϕxm−n (f )
R→∞ xn R
Taken m < n we have that
n−1 n−1 n−m−1
X xk k m X k m! X f (k) (0)
D (x f )(0) = f (k−m) (0) = xk
k! m k! k!
k=0 k=m k=0
Therefore
ˆ R n−m−1
!
f k (0) k
n m 1 X 1
lim IR (x f ) = lim lim f (x) − x dx = PV (f )
R→∞ R→∞ →0 −R xn−m k! xn−m
k=0
Definition 8.2.19 (Giving it Some Meaning). Taken gn ∈ L1loc (R) a sequence of functions such that
gn ∈ C[−an , an ] with an > 0, an → 0, we have that ∃!gn 7→ ϕgn such that ϕgn *? δ0
A good choice for this sequence would be the Gaussian function γn defined as follows
n 2 2
γn (x) = √ e−n x (8.17)
π
Then, we have that ϕγn *? δ0 , or written in a clear abuse of notation with non-existing functions
Theorem 8.27. Let g ∈ C 1 (R) be a function with isolated and simple zeros Zg := {x1 , x2 , x3 , · · · },
then ∀f ∈ K we have
ˆ ˆ X δx (f )
k
lim γn (g(x))f (x) dx = f (x)δ(g(x)) dx = 0 (x )|
(8.18)
n→∞ R R |g k
xk ∈Zg
8.2. DISTRIBUTIONS 155
Proof. Taken g(x) ∈ C 1 (R) such that g(x) ≥ g(y) ∀x ≥ y, supposing that Zg = {x0 } we have, since
it’s a simple zero, that g 0 (x0 ) 6= 0
Taking in the previous integral the substitution y = g(x), we have that x = g −1 (y) and dx =
D g −1 (y) dy, therefore
ˆ ˆ sup(g)
f (g −1 (y))
I = lim γn (g(x))f (x) dx = lim γn (y) dy
n→∞ R n→∞ inf(g) g 0 (g(y))
f (g −1 (0)) f (x0 ) 1
I= = 0 = 0 δx
g 0 (g −1 (0)) g (x0 ) |g (x0 )| 0
Example 8.2.11 (Composition with a Polynomial). Take g(x) = (x2 − a2 ), then Zg := {−a, a} and
g 0 (±a) = ±2a, therefore, if a 6= 0
1
δ(x2 − a2 ) = (δ(x − a) + δ(x + a))
2|a|
Example 8.2.12 (Composition with an Exponential). Taken g(x) = ex we have that Zg = {} therefore
δ(ex ) = 0
Example 8.2.13 (Composition with a Trigonometric Function). Taken now g(x) = cos(x) we have
that Zg = { k ∈ Z| kπ/2} and g 0 (x) = − sin(x), therefore
0 kπ
g (xk ) = − sin = (−1)k+1
2
Therefore
X ∞
kπ X kπ
δ(cos(x)) = δ x− = δ x−
2 2
k∈Z k=−∞
CHAPTER 8. DISTRIBUTIONS 156
Example 8.2.14 (Calculating an Integral with the Composite Delta). Take the following integral
ˆ
e−|x| δ(sin(2x)) dx
R
In order for solving it we expand the composite delta. We have that Zg = { k ∈ Z| kπ/2} and
g 0 (xk ) = (−1)k 2, therefore
ˆ ∞ ∞
−|x| 1 X −|x| kπ 1 X − kπ
e δ(sin(2x)) dx = e δ x− = e 2
R 2 2 2
k=−∞ k=−∞
Which means ¨
1
δy (f ) = eik(x−y) f (x) dk dx
2π R2
Theorem 8.29 (Integral Representation of the Theta Distribution). Taken ϑ(x) the Heaviside regular
distribution we can write ∀x ∈ R \ {0}
ˆ
1 eikx
ϑ(x) = PV dk (8.20)
2πi R k − i
Proof. This equation has two cases, one with x > 0 and one with x < 0. Therefore taking x > 0 we
have ˆ
eikx
I(x) = PV dk
R k − i
This integral can be evaluated using the residue theorem. We have defining the function f (k) as follows
1
f (k) =
k − i
8.3. INTEGRAL REPRESENTATION OF DISTRIBUTIONS 157
Due to the Jordan lemma we know for sure that JR = 0 therefore we have that, writing Re(z) = x,
thanks to the residue theorem
X eizx
J(x) = I(x) = 2πi Res
z=zk z − i
zk ∈Sf
The set of singularities of f is formed by a single point Sf = {i} which is a simple pole, and therefore,
since i ∈ {γR }◦ we have
eizx eizx
J(x) = 2πi Res = 2πi lim (z − i) = e−x
z=i z − i z→i z − i
Therefore, we get
J(x) = 2πie−x
For getting back to the first integral, we have
1
I(x) = lim+ e−x = 1 = ϑ(x) x > 0
2πi →0
In the second case, for x < 0 we have that the new curve will be
− −
{γR } = CR ∪ [−R, R]
− ◦
Since Sf 6⊂ {γR } we have thanks to Cauchy-Goursat that
1 1 −
I(x) = J (z) = 0 = ϑ(x) x < 0
2πi 2πi
Proving our assumption
CHAPTER 8. DISTRIBUTIONS 158
§§ 8.3.2 Distributions in Rn
Definition 8.3.1 (Scalar Test Fields). Given a scalar field f : Rn −→ C one can define the space of
scalar test fields K(Rn ) as follows
K(Rn ) = Cc∞ (Rn ) (8.21)
A function f ∈ Cc∞ (Rn ) is a function such that supp f ⊂ A ⊂ Rn with A a compact set and
∃∂ α f ∀|α| ∈ N
Definition 8.3.2 (K(Rn )−Convergence). Given fn ∈ K(Rn ) a sequence of scalar fields. fn converges
to f ∈ K(Rn ) if
1. ∃A ⊂ Rn compact set such that fn (Ac ) = {0}
2. ∀α ∈ Nn multi-index ∂ α fn ⇒ ∂ α f , ∀x ∈ A
It’s indicated as
fn *Kn f
Then ϕ ∈ K? (Rn )
Theorem 8.30 (Operations in K? (Rn )). We can define the usual distributional operations in
K(Rn ). The usual operations defined for K? are defined in the same way, the only difference
is given by the definition of the derivative Given α ∈ Nn , f ∈ K(Rn ), ϕ ∈ K? (Rn )
∂µ ∂ µ ϕ(f ) = ϕ(∂µ ∂ µ f )
Remark (Local Integrability). The local integrability of some functions is different in Rn . In fact taken
the spherical (n − 1)-dimensional coordinate transformation we have
dn x = rn−1 dr dSn−1
−a
And taken the function g(xν ) = kxµ kµ we get
ˆ ˆ ˆ
f (x) n rn−1
a d x= f (x) dSn−1 dr
Rn kxµ kµ R Sn−1 ra
−a
Which means that kxµ kµ ∈ L1loc (Rn ) ∀a < n
8.4. SOME APPLICATIONS IN PHYSICS 159
∂µ ∂ µ V (xν ) = −4πρ(x)
ρ(x) = qδ 3 (xν )
Therefore we get
˚ ¨
1 3 1 1
∂µ ∂ µ f d x = − ∂µ f − f ∂µ rµ dS2
AR kxµ kµ kxµ kµ = r r
Since → 0+ and since f ∈ K(R3 ) we have that the first and second integral converge to 0, and
therefore ˚ ¨
1 µ 3
µ
∂µ ∂ f d x = − lim f (x) dS2 = −4πf (0)
R3 kx kµ →0 kxµ kµ =
9 Ordinary Differential Equations
dm y
= f (x, y, · · · , y (m−1) )
dxm
It’s called in normal form.
If only total derivatives appear, the differential equation is called ordinary, whereas if also partial
derivatives appear, the differential equation is called partial.
Theorem 9.1 (Reduction of Order). Given an ODE (ordinary differential equation) of order m,
one can reduce the order of the equation through a mapping to Rm , where
y(x)
y 0 (x)
(y, y 0 , · · · , y (m) ) 7→ y µ (x) = ..
.
y (m) (x)
dy µ
= f µ (x, y µ )
dx
1 Captain Picard, Star Trek: The Next Generation
161
CHAPTER 9. ORDINARY DIFFERENTIAL EQUATIONS 162
Definition 9.1.2 (Cauchy Problem). A Cauchy or initial value problem is defined as the following
system (
y 0 (x) = f (x, y(x))
(9.1)
y(x0 ) = y0
Where y0 , t0 ∈ R are known values
Theorem 9.2. Given the Cauchy problem (9.1) we say that y(x) ∈ C 1 is a solution of the system
if and only if is a continuous solution of the following integral equation
ˆ x
y(x) = y0 + f (s, y(s)) ds (9.2)
x0
Proof. Supposing y ∈ C 1 we have that for the fundamental theorem of integral calculus
ˆ x ˆ x
y(x) = y(x0 ) + y 0 (s) ds = y0 + f (s, y(s)) ds
x0 x0
Instead, considering f ◦ y we have that the new composed function must be continuous, therefore we
have for the fundamental theorem of integral calculus that y(x) ∈ C 1 and that
ˆ x0
y(x0 ) = y0 + f (s, y(s)) ds = y0
x0
And therefore ˆ x
dy d
= f (s, y(s)) ds = f (x, y(x))
dx dx x0
Proof. We firstly define a Banach space (X, k·ku ) where X := { y ∈ C (I )| ky(x) − y0 ku ≤ b}.
Define the Picard operator as follows
T̂ : X −→ X
Where ˆ x
T̂ (w) = y0 + f (s, w(s)) ds = z(x)
x0
We have
ˆ x
T̂ (y) − T̂ (z) = T̂ (y − z) = f (s, u(s)) − f (s, v(s)) ds ≤ Lku − vku ∀u, v, y, z ∈ X
u u x0 u
Definition 9.1.3 (Maximal and Global Solutions). Given y(x) a solution to the Cauchy problem (9.1)
we define a maximal solution ym (x) a solution for which ym (x) = y(x) ∀x ∈ I and which still
solves the problem in a set (a, b) ⊃ I , and (a, b) is the biggest set for which ym (x) solves the ODE.
If ym is defined ∀x ∈ R, the solution is called global
Theorem 9.4 (Prolungability of Solutions). Given the ODE (9.1) and let f : (a, b) × R −→ R. Let
c1 , c2 ∈ K ⊂ (a, b) with K a compact set, such that
Then the solution can be extended in the whole set (a, b). The previous statement means
that f is sublinear in the second variable
Theorem 9.5. Let ym (x) be a maximal solution to (9.1) defined on (a, b). Then ∀K ⊂ A ∃δ >
0 : ∀x 6∈ (a + δ, b − δ), (x, ym (x)) 6∈ K
Theorem 9.6. Taken y(x) a solution to the ODE (9.1). If ∃c > 0 such that
|y(x)| ≤ c ∀t ∈ A
Method 1 (Separation of Variables). Suppose that we have the previous ODE, then, thanks to the
Picard-Lindelöf-Cauchy theorem we can immediately say that
1. There is at least one local solution in B (t0 ) if f ∈ C, i.e. if g, h ∈ C
2. There is a unique local solution in B (t0 ) if ∂y f ∈ C, i.e. if h ∈ C 1
The first thing to do is finding all constant solution to the ODE, called the equilibrium solutions.
Then, a general solution can be found by direct integration, where we’re integrating the following
differential form
1
dy = g(t) dt
h(y)
Integrating, and putting H(t), G(t) as the two primitives, we can say that
H(y) = G(t) + H(y0 ) − G(t0 )
Where c = H(y0 ) − G(t0 ) is the integration constant in terms of the initial values, if given
Example 9.2.1. Take the following initial value problem
(
w0 (x) = w2 (x)
w(0) = 1
Definition 9.2.2 (Autonomous Differential Equation). Given a differential equation in normal form
y 0 (x) = f (x, y(x)), it’s said to be an autonomous differential equation if
Therefore
y 0 (x) = g(y(x)) (9.3)
Note that it’s a separable equation, and therefore, writing the primitive of 1/g(y) = G(y) we have that
the general solution of this problem would be
G(y) = x + c
Where c ∈ R is a constant
|f (t, y) − f (t, z)| = |t|y| − t|z|| = |t|||y| − |z|| ≤ max {|α|, |β|} |y − z|
Taken L = max {|α|, |β|} we have that f is Lipschitz continuous in the second variable in every compact
set K ⊂ R and therefore there exists a unique solution y : K −→ R.
Successively we move on to integrate directly the differential equation. We have
ˆ y ˆ t
dy
= t dt
y0 |y| t0
I.e.
y 1 2
t − t20
log =
y0 2
y 1 2 1 2
= e 2 t0 e 2 t
y0
Since t0 = 0 we get
1 2
|y|(t) = |y0 |e 2 t
The solutions depend on the sign of y0 and we finally get
( 1 2
y0 e 2 t y0 > 0
y(t) =
− 12 t2
y0 e y0 < 0
CHAPTER 9. ORDINARY DIFFERENTIAL EQUATIONS 166
This equation has a unique solution in all of R, since f (t, y) is Lipschitz continuous and sublinear in R.
We immediately know that the solution y ∈ C 1 exists since f ∈ C(R2 ) and by Cauchy-Picard-Lindelöf
we have
|f (t, y) − f (t, z)| = |p(t)y(t) − p(t)z(t)| = |p(t)||y − z| ≤ kpku |y − z|
Taken L = kpku we see immediately that f satisfies the theorem and the solution is uniquely defined
in every compact set K ⊂ R.
We also see that f is sublinear, since
Which means that the solution can be extended in R, and ∃!y : R −→ R solution to the ODE.
We now proceed to integrate the ODE, for which we will proceed with a different method than usual.
Let µ(t) be what we will call the integrating factor, defined as follows
´
µ(t) = e− p(t) dt
Method 2 (Integrating Factor). Let y 0 (t) = p(t)y(t) + q(t) be a linear ODE, we solve by finding the
integrating factor µ(t) and then multiplying both sides of the equayion by the integrating factor and
moving the p(t)y(t) term from the RHS to the LHS
By definition of the integrating factor we have that p(t)µ(t) = −µ0 (t) and therefore
t
y(−1) = 2
We immediately see the previous definition, therefore we can immediately say that since f is Lipschitz
continuous and sublinear the solution to the problem will be uniquely defined on all of R.
For integrating this ODE we use the integrating factor method, taking p(t) = t−1 and q(t) = 3t3 .
By definition, we have µ(t) = |t|, and therefore, since t < 0, µ(t) = −t−1 , which gives
ˆ
1
− y(t) = −3 t2 dt + c
t
Therefore
y(t) = t4 − ct
Applying the initial condition
y(−1) = 1 + c = 2 =⇒ c = 1
And therefore the final, unique solution of the initial value problem is the following
y(t) = t4 − t
Theorem 9.7 (General Solutions). Taken y 0 (x) + p(x)y(x) = q(x) a linear non homogeneous ODE,
we define the associated homogeneous ODE as the equation
z 0 (x) + p(x)z(x) = 0
The solution of the complete ODE will be given by the sum of the particular solution y(x)
and the homogeneous solution z(x)
Proof. Take y(x) a solution of the equation and y(x) the particular solution of the ODE, we then have,
since both are solutions (
y 0 (x) + p(x)y(x) = q(x)
y 0 (x) + p(x)y(x) = q(x)
Subtracting term a term we have
z 0 (x) + p(x)z(x) = 0
Therefore z(x) is a solution, but it’s also the solution to the associated homogeneous ODE, and therefore,
inverting for the general solution y(x), we have
y 0 (x) + 2 y(x) = 1
x x2
y(−1) = 2
We want to find the general integral of the equation, y(x). We begin by solving the associated
homogeneous equation
2
z 0 (x) + z(x) = 0
x
In order to integrate this we use the integrating factor µ(x), where
ˆ x
1
dx = exp (2 log(|x|)) = exp log x2 = x2
µ(x) = exp 2
−1 x
d 2
x z(x) = 0
dx
Therefore
c
x2 z(x) = c =⇒ z(x) =
x2
The particular solution y(x) will be
ˆ x
1 1
y(x) = µ(x) dx
µ(x) x0 x2
Therefore ˆ x
1 1
y(x) = dx =
x2 −1 x
The general solution will be y(x) = y(x) + z(x), and we finally get
c 1
y(x) = 2
+
x x
Imposing the initial condition we have
y(−1) = c − 1 = 2 =⇒ c = 3
Therefore
3 1
y(x) = +
x2 x
Note how we could have found the general integral directly using the formula
ˆ x ˆ x
c 1 3 1
y(x) = 2 + 2 dx = 2 + 2 dx
x x −1 x x −1
9.2. COMMON SOLVING METHODS 169
Method 3 (Variation of Constants). Given a linear ODE y 0 (t) + a(t)y = f (t), with a, f ∈ C(I), I ⊆ R
suppose that z(t) is the solution of the homogeneous equation, we have therefore
ˆ
c
z(t) = = c exp a(t) dt
µ(t)
We can find a particular solution through variation of constants, i.e. we suppose c = c(t), and we
have ˆ
y(t) = c(t) exp a(t) dt
And therefore
1 a(t)c(t)
(c0 (t) − a(t)c(t)) + = f (t)
µ(t) µ(t)
Therefore ˆ
c0 (t)
= f (t) =⇒ c(t) = f (t)µ(t) dt
µ(t)
Which implies ˆ
1
y(t) = f (t)µ(t) dt
µ(t)
We then recover the previous formula by adding the particular and homogeneous solutions
ˆ
c 1
y(t) = + f (t)µ(t) dt
µ(t) µ(t)
This system is solvable only if the determinant of the matrix Wµν is non zero, and therefore the two
functions must be linearly independent and span the vector space of solutions S
Method 4 (Characteristic Polynomial). Take the following linear homogeneous ODE of order 2
We begin by “guessing” a solution in exponential form z(x) = eλx , where λ ∈ C. Therefore, reinserting
into the ODE, we have
eλx λ2 + aλ + b = 0
Since the exponential is never zero ∀x ∈ R, ∀λ ∈ C we have that the polynomial must be zero, and
therefore we get two roots
a 1p 2
λ1,2 = − ± a − 4b
2 2
We can now discern 3 cases.
Case 1), a2 − 4b > 0
There are two real roots λ1 , λ2 , and therefore we have
Or
z(x) = AeRe(λ)x cos(Im(λ) + ϕ)
With A, ϕ ∈ R Case 3) a2 − 4b = 0
There is only one root λ = a/2 with multiplicity 2. In order to find the general integral z(x) we impose
the variation of constants, and we have
z(x) = c(x)eλx
And
z 0 (x) = c0 (x)eλx + λc(x)eλx = eλx (c0 (x) + λc(x))
z 00 (x) = c00 (x)eλx + λc0 (x)eλx + λ2 c(x)eλx = eλx (c00 (x) + λc0 (x) + λ2 c(x))
9.2. COMMON SOLVING METHODS 171
Method 5 (Similarity Method). Another method for finding a particular solution is the similarity method.
Given the following linear ODE
y 00 (x) + ay 0 (x) + by(x) = f (x)
The particular solution of this problem can be found using the shape of f (x).
Case 1) f (x) ∈ Rn [x]
In this case we will suppose that the solution will be a polynomial, and we can discern 3 cases
q(x)
b 6= 0
y(x) = xq(x) b = 0, a 6= 0
2
x q(x) b = 0, a = 0
Where q(x) ∈ Rn [x]
Case 2) f (x) = Aeλx , λ ∈ C
In this case we will suppose the particular solution as the real part of a complex exponential, where
y(x) = Re (ỹ(x)) = γ(x)Re eλx
Where γ(x) ∈ C 2 (I), I ⊂ R is some unknown function. From this we will have, deriving twice
ỹ 0 (x) = eλx (γ 0 (x) + λγ(x))
ỹ 00 (x) = eλx λ2 γ(x) + 2λγ 0 (x) + γ 00 (x)
Example 9.2.5. We will use the similarity method to solve the following differential equation
We see immediately that f (x) = 2e3x so we choose a particular solution y(x) with the following form
y(x) = ke3x
We therefore have
y 0 (x) = 3ke3x
y 00 (x) = 9ke3x
Substituting
1
18k = 2 =⇒ k =
9
Therefore our searched solution will be the following
1 3x
y(x) = e
9
Following the same approach, we see that the characteristic polynomial is null, and we end up with a
contradiction, therefore we find a solution of the following kind
y(x) = kxe−3x
Substituting
3k(3x − 2) + 2k(1 − 3x) − 3kx = 2
−4k = 2
1
k=−
2
Therefore the particular solution we’re searching for is, finally
x
y(x) = − e−3x
2
9.2. COMMON SOLVING METHODS 173
Method 6 (Variation of Constants). Another method that might be used to solve a second order linear
differential equation is the method of variation of constants.
Given the following differential equation and the associated homogeneous equation
Suppose that z1 (x), z2 (x) are two linearly independent homogeneous solutions, and
We search for a particular solution supposing that the two constants are actually functions, and we
have
y(x) = c1 (x)z1 (x) + c2 (x)z2 (x)
Due to the necessity of a non-singular Wronskian determinant, we impose the following condition
Deriving two times the particular solution and then substituting in the original equation we end up
with the following system (
c01 (x)z1 (x) + c02 (x)z2 (x) = 0
c01 (x)z10 (x) + c02 (x)z20 (x) = f (x)
The necessity for the linear independence of the solutions is clear here, since the system is solvable if
and only if the Wronskian determinant is zero.
We solve the system by substitution, and we have
z (x)
c01 (x) = −c02 (x) 2
z1 (x)
c0 (x) =
f (x) − c01 (x)z10 (x)
2
z20 (x)
Therefore
z1 (x)
0
c1 (x) = − f (x)
z1 (x)z20 (x) − z2 (x)z10 (x)
z2 (x)
c02 (x) =
f (x)
z1 (x)z20 (x) − z2 (x)z10 (x)
Written in terms of the Wronskian determinant detµν Wµν (x) we finally have as particular solution
ˆ ˆ
z2 (x) z1 (x)
y(x) = z2 (x) f (x) dx − z1 (x) f (x) dx + c
detµν Wµν (x) detµν Wµν (x)
And the general solution of the differential equation will be the following
ˆ x ˆ x
z2 (x) z1 (x)
y(x) = z2 (x) f (x) dx − z1 (x) f (x) dx + c1 z1 (x) + c2 z2 (x)
x0 detµν Wµν (x) x0 detµν Wµν (x)
CHAPTER 9. ORDINARY DIFFERENTIAL EQUATIONS 174
10 Fourier Calculus
Then
kv − Sn k ≤ kv − Snα k
∞
X 2 2
kck k ≤ kvk
k=1
The last inequality is known as Bessel’s inequality
Lastly we also have Parseval’s equality or Parseval’s theorem, which states
∞
X ∞
X ∞
X 2 2
v= hv, uk iuk = ck uk ⇐⇒ kck k = kvk
k=1 k=1 k=1
Due to this the operator F̂U actually acts into ` (C), i.e. 2
F̂U : V −→ `2 (C)
175
CHAPTER 10. FOURIER CALCULUS 176
Proof. By definition of euclidean norm and using the bilinearity of the scalar product we have
2 2
0 ≤ kv − Snα k = kvk − 2Re (hv, Snα i) + kSnα k =
n
! n
2
X X 2
= kvk − 2Re hv, uk iαk + kαk k
k=1 k=1
Therefore
n
X n
X
2 2 2
0 ≤ kvk − kck k + kαk − ck k
k=1 k=1
The minimum on the left is given for αk = ck and therefore, since Snc = Sn we have
kv − Sn k ≤ kv − Snα k
Therefore
∞
X 2 2
khv, uk ik ≤ kvk
k=1
Which means that the sum on the left converges uniformly, and therefore ck ∈ `2 (C). This demonstrates
that F̂U : V −→ `2 (C) and Bessel’s inequality.
This also gives Parseval’s equality, since, for n → ∞
∞ 2 ∞
X 2
X 2
v− hv, uk iuk = 0 ⇐⇒ kvk = khv, uk ik
k=1 k=1
Definition 10.1.2 (Closed System). An system (uk )k∈N ∈ V is said to be closed iff ∀v ∈ V
∞
X
2 2
kvk = khv, uk ik
k=1
X∞
v= hv, uk iuk
k=1
10.1. BESSEL INEQUALITY AND FOURIER COEFFICIENTS 177
Theorem 10.2 (Closeness and Completeness). Given an orthonormal system U = (uk )k∈N ∈ V
with V an euclidean space, we have that U is a complete set if and only if U is a closed system.
If U is complete or closed, V is separable
Proof. Defined Sn the partial sums of the Fourier representation of v (ndr the series that represents v
with respect to the system (uk )), we have that for the theorem to be true the following two things
must hold
lim Sn = v span(uk ) = V
n→∞
v ∈ span(U) =⇒ V = span(U)
The last implication is given by the fact that v ∈ V is arbitrary, and it implies the completeness of U and
the separability of V
Theorem 10.3 (Riesz-Fisher). Given V a hilbert space and U = (uk )k∈N ∈ V an orthonormal
system, therefore ∀c ∈ `2 ∃v ∈ V : F̂U [v] = c and
ck = hv, uk i
∞
X
2 2 2
kvk = kck2 = kck k
k=1
∞
X
v= hv, uk iuk
k=1
Where ˆ
1 π
ak = f (x) cos(kx) dx
π −π
ˆ (10.2)
1 π
bk =
f (x) sin(kx) dx
π −π
The notation ∼ indicates that the Fourier series of the function converges to the function f (x). Usually
an abuse of notation is used, where the function is actually set as equal to the Fourier expansion.
Definition 10.2.2 (Trigonometric Polynomial). A function p ∈ L2 [−π, π] is said to be a trigonometric
polynomial if, for some coefficients αk , βk we have
∞
α0 X
p(x) = + αk cos(kx) + βk sin(kx) (10.3)
2
k=1
10.2. FOURIER SERIES 179
Theorem 10.4 (Completeness of Trigonometric Functions). Given (uk ), (vk ) ∈ L2 [−π, π] two se-
quences of functions, where (
uk (x) = cos(kx)
vk (x) = sin(kx)
1
cos(nx) cos(kx) = (cos[(n + k)x] + cos[(n − k)x])
2
1
sin(nx) sin(kx) = (cos[(n − k)x] − cos[(n + k)x]) (10.4)
2
1
cos(nx) sin(kx) = (sin[(n + k)x] − sin[(n − k)x])
2
Proof. We begin by demonstrating that the two function sequences uk , vk are orthogonal in L2 [−π, π].
Therefore, by explicitly writing the scalar product, we have, for k 6= n
ˆ π ˆ
1 π
hun , uk i = cos(nx) sin(kx) dx = cos[(n + k)x] + cos[(n − k)x] dx
−π 2 −π
Therefore π
1 sin[(n + k)x] sin[(n − k)x]
hun , uk i = + =0
2 n+k n−k −π
Analogously π
1 sin[(n − k)x] sin[(n + k)x]
hvn , vk i = − =0
2 n−k n+k −π
And, finally π
1 cos[(n + k)x] cos[(n − k)x] 1 1
hun , vk i = − = − =0
2 n+k n−k −π 2 2
Which demonstrates that, for k 6= n uk ⊥ un , vk ⊥ vn , uk ⊥ vk .
Now, taken a trigonometric polynomial p(x) ∈ L2 [−π, π] we need to prove that span {uk , vk } =
L2 [−π, π], i.e.
∀ > 0 ∀f ∈ L2 [−π, π] kp − f k2 <
We have already that C[−π, π] = L2 [−π, π] and that for a Weierstrass theorem (without proof), every
periodic function with period 2π is the uniform limit of a trigonometric polynomial.
Using these two results, given f ∈ L2 [−π, π], ∃g ∈ C[−π, π] : kf − gk2 < /3. Taken ĝ(x) as the
periodic extension of g(x), for Weierstrass we have
kg − ĝk2 < kp − ĝk2 < =⇒ kp − ĝku < √
3 3 3 2π
Proof. The proof is quite straightforward, since trigonometric polynomials form a basis for L2 [−π, π]
we have that this is simply the already known Parseval identity, since
∞
X
2 2
kf k2 = kck k
k=0
Writing ck = ak + bk we have
2 ∞
2 ka0 k X 2 2
kf k2 = + kak k + kbk k
2
k=1
Where ˆ b
1 kπ
ã = f (x) cos (2x − b − a) dx
k
b−a a b−a
ˆ b (10.6)
1 kπ
b̃k = f (x) sin (2x − b − a) dx
b−a a b−a
10.2. FOURIER SERIES 181
§§ 10.2.3 Fourier Series in Symmetric Intervals, Expansion in Only Sines and Cosines
Definition 10.2.5. We firstly begin finding the Fourier series of a function in L2 [−l, l]. Using the
previous general case in L2 [a, b] and setting a = −l, b = l we have ∀f ∈ L2 [−l, l]
∞
a0 X kπx kπx
f (x) ∼ + ak cos + bk sin (10.7)
2 l l
k=1
With coefficients ˆ
1 l
kπx
ak = l f (x) cos dx
−l l
ˆ (10.8)
1 l
kπx
bk = f (x) sin dx
l −l l
Theorem 10.6. Taken the space L2 [0, π] we have that both trigonometric sequences (uk (x))
and (vk (x)) are orthogonal bases in this space, and the following equalities hold.
∀f ∈ L2 [0, π]
∞
a0 X
f (x) ∼ 0 + a0k cos(kx)
2
k=1
∞
X
f (x) ∼ b0k sin(kx)
k=1
Where ˆ
2 π
a0k = f (x) cos(kx) dx
π 0
ˆ π
2
b0k = f (x) sin(kx) dx
π 0
Proof. The proof of this theorem is straightforward, we firstly define the even and uneven extensions
of the function f (x) in L2 [−π, π] as follows
(
f (x) x ∈ [0, π]
f e (x) =
f (−x) x ∈ [−π, 0)
And (
u f (x) x ∈ [0, π]
f (x) =
−f (−x) x ∈ [−π, 0)
Expanding both these functions in [−π, π] we have that, indicating the coefficients of each as aek , bek , auk , buk
ˆ ˆ
1 π e 2 π
aek = f (x) cos(kx) dx = f (x) cos(kx) dx = a0k
π −π π 0
bek = 0
auk = 0
ˆ ˆ
1 π u 2 π
buk = f (x) sin(kx) dx = f (x) sin(kx) dx = b0k
π −π π 0
CHAPTER 10. FOURIER CALCULUS 182
Therefore
∞
a0 X
f (x) ∼ 0 +
e
a0k cos(kx)
2
k=1
∞
X
f u (x) ∼ b0k sin(kx)
k=1
The coefficients ak can be calculated using the fact that x2 is even, and therefore we have
ˆ l ˆ l
1 l 2
kπx 1 2 kπx l 4 kπx
ak = x cos dx = x sin − x sin dx =
l −l l l l kπ −l kπ 0 l
l ˆ l
4l kπx 4l kπx
= x cos − sin dx
(kπ)2 l −l (kπ)2 0 l
Example 10.2.2 (Parseval’s Equality). Having now found the Fourier expansion for x2 , we can use
Parseval’s equality in order to calculate the sum
∞
X 1
k2
k=1
10.2. FOURIER SERIES 183
The integral on the left is obvious, and moving the terms around we finally have
∞ ˆ l
π4 π4 π4 2 2
X 1 4
= x dx − = −
k4 16l5 −l 36 16 5 9
k=1
Therefore
∞
X 1 π4
=
k4 90
k=1
Therefore, due to the linearity of the scalar product, these functions are orthogonal to each other, and
due to linearity we also have
Which, implies
span{eikx } = L2 [−π, π]
Note that
2
eikx 2
= 2π
Definition 10.2.6 (Complex Fourier Series). Given f ∈ L2 [−π, π] we can now define a Fourier expansion
in complex exponentials as follows
∞
X hf (x), eikx i X
f (x) ∼ 2 eikx = ck eikx (10.9)
k=−∞ keikx k2 k∈Z
Example 10.2.3. Taken the function f (x) = ex , x ∈ [−π, π], we want to find the Fourier series in
terms of complex exponentials. Since f (x) is a real valued function, we have
∞
X
e x ∼ c0 + 2 Re ck eikx
k=1
1 (−1)k
eikπ eπ − e−ikπ e−π =
ck = sinh(π)
2π(1 − ik) π(1 − ik)
The final expansion will then be given from finding the real part of this coefficient times the basis vector,
i.e.
(−1)k sinh(π) ikx (−1)k sinh(π)
Re e = Re ((1 + ik)(cos(kx) + i sin(kx)))
π(1 − ik) 1 + k2
0 f (x + ) − f (x+ )
f+ (x) = lim
→0+
0 f (x + ) − f (x+ )
f− (x) = lim− −
→0
Where
f (x± ) = lim± f (y)
y→x
0 x<0
1
f (x) = x=0
2
1−x x>0
We have
0 f () − f (0+ ) 1−−1
f+ (0) = lim+ = = −1
→0
0 f () − f (0+ ) 0
f− (0) = lim+ = =0
→0 −
It’s important to see how the right and left derivatives might not coincide with the right and left limits
of the derivative, as explained in the following theorem
Theorem 10.8. Let f (x) : [a, b] −→ C be a piecewise differentiable function, then given x ∈ [a, b)
we have
0
f+ (x) = f 0 (x+ )
0
f− (x) = f 0 (x− )
Proof. Since f (x) is piecewise differentiable, we have that ∃γ > 0 such that f (x) is differentiable
∀x ∈ (x, x + γ) and we can define f 0 (x+ )
Therefore ∀α > 0, ∃1 > 0 such that
Thanks to the Lagrange theorem, and the definition of one sided limit, we have, given 0 < δ < ≤ 1
f (x + ) − f (x + δ)
− f 0 (x+ ) < α
−δ
Which implies therefore
f (x + ) − f (x+ )
lim lim − f 0 (x+ ) = 0
→0+ δ→0+
CHAPTER 10. FOURIER CALCULUS 186
0 f (x + ) − f (x+ )
f+ (x) = lim = f 0 (x+ )
→0+
Definition 10.2.8 (Periodic Extension). Given a function f : [−π, π] −→ C, non periodic, we define the
periodic extension f̃ : R −→ C as
Note that it coincides with the same function, given that x ∈ (−π, π] and therefore the periodic
extension has discontinuities of the first kind at the points xk = (2k + 1)π, k ∈ Z
Lemma 10.2.1 (Riemann-Lebesgue). Let f ∈ C[a, b] be a function such that f 0 is piecewise continuous
(also holds ∀f ∈ L1 [a, b]), then
ˆ b
lim f (x) sin(sx) dx = 0
s→∞ a
Proof. The proof comes directly from the calculus of the integral
ˆ b ˆ b
1 b 1
f (x) sin(sx) dx = − [f (x) cos(sx)]a + f 0 (x) sin(sx) dx
a s s a
Definition 10.2.9 (Dirichlet Kernel). We define the Dirichlet kernel as the following function
! n
1 sin 2n+1
2 z 1 1X
Dn (z) = = + cos(kz)
2π sin z2 2π π
k=1
0
Lemma 10.2.2. Given f : R −→ C a piecewise continuous function, and x ∈ R such that exists f+ (x)
we have that ˆ π
1
lim f (x + z)Dn (z) dz = f (x+ )
n→∞ 0 2
ˆ 0
1
f (x + z)Dn (z) dz = f (x− )
lim
n→∞ −π 2
10.2. FOURIER SERIES 187
Proof. By definition of the Fourier expansion, we have that, in the trigonometric basis
n
a0 X
Sn (x) = + ak cos(kx) + bk sin(kx)
2
k=1
Inserting the usual definitions of the Fourier coefficients and using the fact that the sum is finite, hence
it converges, we have
"ˆ n
! #
π
1 1 X
Sn (x) = f (s) + cos(kx) cos(ks) + sin(kx) sin(ks) ds
π −π 2
k=1
Simplifying
ˆ π n
!
1 1 X
Sn (x) = f (s) + cos (k(s − x)) ds
π −π 2
k=1
Rearranging the second term we see that it’s the Dirichlet kernel, and using a transformation z = s − x,
we have ˆ π
Sn (x) = f (x + z)Dn (z) dz
−π
Note that the extremes of integration are the same since both these functions are 2π−periodic.
Using the definition of the integral between f and the Dirichlet kernel, we have finally the statement
of the theorem, in the case that the function has a discontinuity at the point x
1
f (x+ ) + f (x− )
lim Sn (x) =
n→∞ 2
Theorem 10.10 (Pointwise Convergence of the Fourier Series). Given a piecewise continous 2π−pe-
riodic function f : R −→ C, we have that the Fourier series converges pointwise to the
following two cases, in case the function is continous or not in the point x ∈ R
Theorem 10.11 (Uniform Convergence of the Fourier Series). Given a 2π−periodic function f ∈
C(R), such that its derivative is piecewise continuous, we have that
Sn ⇒ f
CHAPTER 10. FOURIER CALCULUS 188
Proof. Since f (x) ∈ C(R) for the previous theorem we have that Sn (x) → f (x) ∀x ∈ R, therefore
∞
a0 X
f (x) = + ak cos(kx) + bk sin(kx) x ∈ R
2
k=1
Using the Weierstrass M-test we have that, taken the following sequence
ck = ak cos(kx) + bk sin(kx)
In order to check that the sum of this extremum is convergent we find the Fourier expansion of the
derivative of f
∞
a0 X
f 0 (x) ∼ 0 + a0k cos(kx) + b0k sin(kx)
2
k=1
Example 10.2.5 (Heat Equation with Von Neumann Boundary Conditions). We firstly write the heat
equation with Von Neumann boundary conditions
∂t u = λ∂x2 u
∂x u(0, t) = ∂x u(l, t) = 0 (10.13)
u(x, 0) = f (x)
Since we want the derivative on the x to vanish in order to satisfy immediately the boundary conditions,
we suppose the following expansion
∞
a0 (t) X kπx
u(x, t) = + ak (t) cos
2 l
k=1
Deriving, we have
∞
a00 (t) X 0
kπx
∂t u = + ak (t) cos
2 l
k=1
∞
πX kπx
∂x u = − kak (t) sin
l l
k=1
∞
π2 X 2
2 kπx
∂x u = − 2 k ak (t) cos
l l
k=1
It’s immediate to see that the boundary conditions are immediately satisfied, and therefore, reinserting
it back into the differential equation, we get
∞ ∞
π2 λ X 2 a0 (t) X 0
kπx kπx
− 2 k ak (t) cos = 0 + ak (t) cos
l l 2 l
k=1 k=1
Reinserting into the second boundary condition u(x, 0) = f (x) we end up determining the coefficients
as the cosine-Fourier coefficients of the function f (x)
ˆ l
2 kπx
ak = f (x) cos dx
l 0 l
ˆ l
1
a0 = f (x) dx
l 0
Example 10.2.6 (Dirichlet Boundary Conditions). Let’s now take the heat equation with different
boundary conditions, namely
2
∂t u = λ∂x u
u(x, 0) = f (x)
u(0, t) = u(l, t) = 0
+
Where (x, t) ∈ [0, l] × R \ {0} Since the first derivative doesn’t appear in the boundary conditions
we choose a particular form of u(x, t) in terms of an only sine Fourier expansion (assuming uniform
convergence). We have therefore
∞
X kπx
u(x, t) = bk (t) sin
l
k=1
Sending l → ∞ we have that the sum of coefficients behaves like a Riemann sum and converges to the
following integral ˆ
g(λ)eikx dx = f (x)
R
Where the last equality is given by the fact that fl (x) → f (x).
We define the integral used for finding these “coefficients” the Fourier Integral Transform of f
ˆ
1
g(λ) = F̂[f ](λ) = f̂ (λ) f (x)e−iλx dx
2π R
Definition 10.3.1 (Parity, Translation and Dilation Operators). Let f : R −→ C be some function. We
define the following operators
Definition 10.3.2 (Fourier Operator). Given a function f ∈ L1 (R) we define the Fourier operator
F̂[f ] as follows ˆ
F̂[f ] = f (x)e−iλx dx λ ∈ R
R
Theorem 10.12 (Properties of the Fourier Transform). Given f, g ∈ L1 (R) and a, b ∈ C we hafe that
1. F̂[af + bg] = aF̂[f ] + bF̂[g]
Theorem 10.13 (Action of the Dilation, Parity and Translation Operators on the Fourier Operator).
Given f ∈ L1 (R) and a ∈ R \ {0} we have
1. F̂ P̂ = P̂ F̂
2. F̂ T̂a = e−iλa F̂
3. T̂a F̂ = F̂[eiax f (x)]
−1
4. F̂ Φ̂a = |a| Φ̂a−1 F̂
−1 −iλb
5. F̂Φa T̂b = |a| e Φ̂a−1 F̂
−1 −iλb/a
6. F̂ T̂b Φ̂a = |a| e Φ̂a−1 F̂
7. F̂ D̂ = iλF̂
Example 10.3.1 (Fourier Transform of the Set Index Function). Let f (x) = 1[−a,a] (x) be the index
function of [−a, a], we have
ˆ ˆ a
i 2
F̂[1[−a,a] ](λ) = 1[−a,a] (x)e−iλx dx = e−iλx dx = [e−iλx ]a−a = sin(λa)
R −a λ λ
Example 10.3.2 (Fourier Transform of the Triangle Function). We define the triangle function tri(x) =
max {1 − |x|, 0} = rect(x/2)(1 − |x|). We then have
ˆ ˆ 1
F̂[max {1 − |x|, 0}] = max {1 − |x|, 0} e−iλx dx = (1 − |x|)e−iλx dx
R −1
Using the properties of the absolute value and using a change in coordinayes we have
ˆ 1 ˆ 1
(1 − x) eiλx + e−iλx dx = 2
F̂[max {1 − |x|, 0}] = (1 − x) cos(λx) dx
0 0
Example 10.3.3 (A Couple Fourier Transforms More). 1) f (x) = H(x)e−ax where a ∈ R, a > 0 This
one is quite straightforward. We have
ˆ ˆ +∞
e−(a+iλ)x
1
F̂[H(x)e−ax ] = H(x)e−(a+iλ)x dx = e−(a+iλ)x = − =
R R+ a + iλ 0 a + iλ
10.3. FOURIER TRANSFORM 193
2) f (x) = 1/(1 + x2 ) For this we immediately choose to use the residue theorem, so, we firstly suppose
that λ < 0 and we choose a path enclosing the region Im(z) < 0, for which the only singularity is
given by z̃ = −i ˆ
e−iλz
−iλz
1 e
F̂ = lim dz = −2πi Res
1 + x2 R→∞ γ − 1 + z 2
R
z=−i 1 + z 2
Therefore
e−iλz
1
F̂ = −2πi lim (z + i) = πeλ
1 + x2 z→−i (z + i)(z − i)
Analogously, for λ > 0 we choose a curve enclosing the region Im(z) > 0, and therefore, noting that
the only encompassed singularity is z̃ = i
ˆ
e−iλz
−iλz
1 e
F̂ = lim dz = 2πi Res
1 + x2 R→∞ γ + 1 + z 2
R
z=i 1 + z2
Therefore
e−iλz
1
F̂ = 2πi lim (z − i) = πe−λ
1 + x2 z→i (z + i)(z − i)
Uniting both cases, i.e. for λ ∈ R, we have finally
1
F̂ = πe−|λ|
1 + x2
3) f (x) = e−a|x| Last but not least, we can calculate this Fourier transform using the properties of the
Fourier operator. Firstly
e−a|x| = H(x)e−ax + H(−x)eax
We can also write this as follows
Therefore, using the linearity of the Fourier transform and the behavior of it under parity transformations,
we have
1 1 1 1 2a
F̂[e−a|x| ] = + P̂ = + = 2
a + iλ a + iλ a + iλ a − iλ a + λ2
2
Example 10.3.4 (A Particular Way of Solving a Fourier Integral). Take now the function f (x) = e−x ,
using the properties of this function under derivation we can build easily a differential equation
df
= −2xf (x)
dx
Applying the Fourier operator on both sides we get
df
F̂ = −2F̂ [xf (x)]
dx
d
iλF̂[f (x)] = −2i F̂[f (x)]
dλ
CHAPTER 10. FOURIER CALCULUS 194
Therefore, we’ve built a differential equation in the Fourier domain, where the searched function is
actually the Fourier transform of the initial equation. Solving, we get
d log λ
F̂[f (x)] = −
dλ 2
´ 2 √
Therefore, imposing the condition of integration on all R and remembering that R
e−x dx = π, we
have
√ λ2 2
F̂[f (x)] = πe− 4 = F̂[e−x ]
Theorem 10.14. Let f ∈ L1 (R) be some function. Taken f̂ (λ) = F̂[f ](λ), we have that f̂ ∈ C0 (R)
Proof. Firstly, we need to demonstrate that f̂ ∈ C(R). Therefore, by definition of continuity we have
ˆ
f̂ (λ + ) − f̂ (λ) = f (x)e−iλx (e−ix − 1) dx
R
f̂ (λ + ) − f̂ (λ) ≤ |a|kf k1
We have then
ˆ
f̂ (λ) ≤ f (x)e−iλx dx +
|x|≤a
Theorem 10.15. Given f ∈ C p−1 (R) a function, such that ∂ p f (x) is piecewise continuous, and
∂ k f (x) ∈ L1 (R) for k = 1, · · · , p. If this holds, we have that
1. F̂[∂ k f ](λ) = (iλ)k F̂[f ], for k = 1, · · · , p
2. limλ→±∞ λp F̂[f ](λ) = 0
Proof. Through integration by parts of the definition of the Fourier transform we have
If the evaluation of f (x)e−iλx on all R gives back 0 we have that the first part of the theorem is
demonstrated through iteration.
Using that f 0 ∈ L1 (R) tho, we can define using the fundamental theorem of calculus
ˆ x
f (x) = f (0) + f 0 (s) ds
0
Also, since f 0 ∈ L1 (R) we must have that the limits at ±∞ of f (x) must be finite, therefore
lim f (x)e−iλx = 0
x→±∞
And
F̂[f 0 ](λ) = iλF̂[f ](λ)
Through this, we therefore have by iteration that
1
λp F̂[f ](λ) = F̂[f (p) ](λ)
ip
Which, thanks to Riemann-Lebesgue, gives
1
lim λp F̂[f ](λ) = lim F̂[f ](λ) = 0
λ→∞ ip λ→∞
Theorem 10.16. Given f ∈ L1 (R) such that xk f ∈ L1 (R) for k = 1, · · · , p, we have that F̂[f ] ∈
C p (R), and
∂ k F̂[f ](λ) = F̂[(−ix)k f (x)](λ)
CHAPTER 10. FOURIER CALCULUS 196
Proof. In order to see if it’s true, we start for the first derivative and apply the definition. Therefore,
given f̂ (λ) = F̂[f ](λ) ˆ
1
(f̂ (λ + ) − f̂ (λ)) − (−ix)f (x)e−iλx dx
R
Dividing the improper integral around a point a ∈ R we have that everything is lesser or equal to the
following quantity
ˆ ˆ
e−ix − 1 e−ix − 1
|xf (x)| + i dx + 2 |xf (x)| + i dx
|x|≤a x |x|>a x
Using the Taylor expansions for the cosine and the sine we get, approximating, that
1 2
|cos θ − 1| ≤ θ
2
1 3
|sin(θ) − θ| = |θ|
3!
Therefore
e−ix − 1 1 2 1 3
+ i ≤ |x| + |x|
x 2 6
Putting that back into the integral, we have that it must be smaller or equal to the following quantity
ˆ 2
! ˆ 2
!
|x| |x| x |x|
|xf (x)| + dx + 2 |xf (x)| + dx
|x|≤a 2 6 |x|>a 2 6
Using in the first integral that |x| ≤ |a| we get, that all the quantity will be surely less than the
supremum of such, and therefore
ˆ
|a| |a|
+ kxf (x)k1 + 2 |xf (x)| dx
2 6 |x|>a
Theorem 10.17 (Invariance of the Schwartz Space under Fourier Transforms). Given f ∈ S(R) ⊂ L1 (R),
then F̂[f ] ∈ S(R).
Proof. We will prove a weaker assumption. Since S(R) = C ∞ (R), k·kj,k where the seminorm
k·kj,k is defined as follows
kf kj,k = xj ∂ k f u
<∞ j, k ∈ N
We have, ∀f ∈ S(R) that for a given constant Ca,k ∈ R
Ca,k
∂kf ≤ a a, k ∈ N, x ∈ R
(x2 + 1) 2
Therefore, taken a = j + 2
ˆ ˆ j ˆ
j k |x| 1
x ∂ f (x) dx ≤ Cj+2,k j dx ≤ Cj+2,k dx < ∞
R R (x2 + 1) 2 +1 R x2 + 1
Therefore, we have that xj ∂ k f ∈ L1 (R) ∀j, k ∈ N. But we can also write the following result using
Leibniz’s rule
j
j k
X j
∂ x f (x) = ∂ m xj ∂ j−m f (x) ∈ L1 (R)
m=0
m
Therefore, applying the Fourier transform and using the previous property, we have
(iλ)j (∂ k f̂ (λ)) = (iλ)j (−i)k F̂[xk f ](λ) = (−i)k F̂[∂ j (xk f )](λ) ∈ C0 (R)
Theorem 10.18 (Inversion Formula). Given f ∈ S(R), we have that F̂ a F̂[f ] = F̂ F̂ a [f ] = 1̂[f ] = f ,
therefore F̂ a = F̂ −1 and the Fourier transform is a bijective map in S(R)
∼
F̂ : S(R) −→ S(R)
Proof. Weakening the statement, we can say that taken f ∈ K such that f (x) = 0 for |x| > a ∈ R,
and taken ∈ (0, 1/a), we have that, Fourier expanding the function in [−−1 , ], we get
ˆ 1
X 2
f (x) = ck e ikπx
, ck = f (x)e−ikπx dx → F̂[f ](kπ)
− 1 2
k∈Z
Where we let → 0+ in the Fourier series, which written in that way gives a Riemann-Lebesgue sum
converging to the integral of the antitransform, therefore proving that in S(R) F̂ a = F̂ −1
Theorem 10.19 (Plancherel). Given f, g ∈ S(R), and let h·, ·i be the usual scalar product defined
as follows ˆ
hf, gi = f (x)g(x) dx
R
Then, we have
√
F̂[f ] = 2πkf k2
2 (10.15)
hF̂[f ], F̂[g]i = 2πhf, gi
Taking the limit → 0+ the sum on the right converges to the following value
ˆ ˆ 2 √
2 1
|f (x)| dx = F̂[f ] dλ =⇒ kf k2 = 2πkf k2
R 2π R
Then for the polarization identity, and this result, we get the final proof of the theorem
10.3. FOURIER TRANSFORM 199
Theorem 10.20 (Continuity Expansion). Let p, q ∈ C(R), then if p(x) = q(x) ∀x ∈ Q we have that
p(x) = q(x) ∀x ∈ R
Proof. Since Q is dense in R we can take a sequence (xn )n∈N ∈ Q such that xn → x ∈ R. Therefore,
using the continuity of p, q we have
p(x) = p lim xn = lim p(xn ) = lim q(xn ) = q lim xn = q(x)
n→∞ n→∞ n→∞ n→∞
Theorem 10.21 (Extension of the Inversion Formula). Let f, g ∈ S(R). Taken a metric dS (·, ·) :
S(R) × S(R) → R defined as follows
∞ X
∞
X 1 kf − gkj,k
dS (f, g) = j, k ∈ N
j=0 k=0
2j+k 1 + kf − gkj,k
F̂ a F̂ = F̂ F̂ a = 1̂ =⇒ F̂ a = F̂ −1 ∀f ∈ S(R)
Theorem 10.22. Given f ∈ L1 (R), then, since we might have that F̂[f ] ∈ / L1 (R), using the
Cauchy principal value
ˆ
1 1
PV F̂[f ](λ)eiλx dλ = f (x− ) + f (x+ )
∀x ∈ R (10.16)
2π R 2
Theorem 10.23 (New Calculation Rules). With what we added so far, in operatorial form, we can
write down two new calculation rules. Supposing the inversion formula holds, and therefore
F̂ a = F̂ −1
1
F̂ −1 = F̂ P̂
2π (10.17)
F̂ −1 F̂ = 2π 1̂
CHAPTER 10. FOURIER CALCULUS 200
Example 10.3.5. Let’s calculate the Fourier transform of the following function f : R −→ C
1
f (x) =
(x + i)3
Theorem 10.24. Defined the convolution product of two bounded functions, we have that
∗ : L1 (R) × L1 (R) −→ L1 (R)
And
kf ∗ gku ≤ kf ku kgk1
kf ∗ gk1 ≤ kf k1 kgk1
For the second proof, we begin taking a compact set [a, b] ⊂ R and we move forward sending a → −∞
and b → ∞. Therefore
ˆ b ˆ ˆ b−y ¨
|(f ∗ g)(x)| dx ≤ |f (y)| |g(u)| du ≤ |f (y)||g(u)| dy du = kf k1 kgk1
a R a−y R2
Theorem 10.25 (Properties of the Convolution Product). Given f, g, h ∈ L1 (R) three bounded
functions, then
f ∗g =g∗f
(10.19)
(f ∗ g) ∗ h = f ∗ (g ∗ h) = f ∗ g ∗ h
These properties are easily demonstrable using the properties of the integral
df ∗ g dg
=f∗ = f ∗ g0 (10.20)
dx dx
And therefore ˆ
df ∗ g ∂
= A(x, t) dt
dx ∂x R
∂A
= |f (t)g 0 (x − t)| ≤ M |f (t)|
∂x
Due to the fact that f (t) ∈ L1 (R) the integral is well defined, and using Leibniz’s derivation rule, we
have ˆ ˆ
df ∗ g ∂A
= dt = f (t)g 0 (x − t) dt = (f ∗ g 0 )(x)
dx R ∂x R
Theorem 10.27 (Fourier Transform of a Convolution). Given f, g ∈ L1 (R) two bounded functions,
we have that
F̂[f ∗ g] = F̂[f ]F̂[g] (10.21)
Example 10.3.6 (Convolution of Two Set Functions). Given a set A = [−a, a] and a set B = [−b, b],
we know that
1A (x)1B (x) = 1A∩B (x)
1[−a,a] (x + y) = 1[−a+y,a+y] (x)
Suppose we need to calculate the convolution product 1A ∗ 1A . The calculation is quite easy with
those properties ˆ
(1A ∗ 1A )(x) = 1A (y)1A (x − y) dy
R
Taken C = [−a + x, a + x], we have
ˆ
(1A ∗ 1A )(x) = 1A∩C (y) dy = µ (A ∩ C)
R
Summarized
µ (A ∩ C) = max {2a − |x|, 0} = (1A ∗ 1A )(x)
Remembering the definition of the rect(x), tri(x) functions, we get that
Example 10.3.7 (Two Gaussians). Taken α, β ∈ R we might want to calculate the following convolution
product ˆ
−αx2 −βx2 2 2
e ∗e = e−αy e−β(x−y) dx
R
The direct calculation of the integral is immediate, but we might want to test here the power of the
last theorem that was stated. Hence we have
h 2 2
i h 2 2
i
F̂ −1 F̂[e−αx ∗ e−βx ] = F̂ −1 F̂[e−αx ]F̂[e−βx ]
Using that r
2 π − λ2
F̂[eax ](λ) = e 4a
a
We have immediately that the searched convolution will be
4
π
α+β
2 2 −λ
e−αx ∗ e−βx = √ F̂ −1 e 4 αβ (x)
αβ
Which gives s
π λ2 π 1 αβ − α+β
α+β αβ
−1 − x2
√ F̂ e 4 αβ
(x) = √ √ e
αβ αβ π α + β
10.3. FOURIER TRANSFORM 203
With some rearrangement of the constant terms, we get finally the expected result
r
π
2 2 − αβ x2
e−αx ∗ e−βx = e α+β
α+β
We begin by applying the Fourier transform on the solution as follows. The transformed variable will
be indicated as an index
ˆ
v(λ, t) = F̂x [u(x, t)](λ) = u(x, t)e−iλx dx (10.22)
R
Note that
2 1 h x2 i
e−kλ = √ t
F̂x e− 4kt (λ)
2 πkt
Therefore, remembering that the product of transforms gives the transform of the convolution
1 h x2
i
v(λ, t) = F̂x [u(x, t)] = √ F̂x u0 ∗ e− 4kt (λ)
2 πkt
And, the searched solution will therefore be the following
1 x2
u(x, t) = √ u0 ∗ e− 4kt
2 πkt
CHAPTER 10. FOURIER CALCULUS 204
Appendices
205
A Useful Concepts
In order to ease various calculations one can utilize more abstract index constructions. One of these is
the multi-index notation, where instead of having an index i ∈ N or j ∈ Z, one constructs a “vector”
of indices, like α = (a1 , · · · , an ) ∈ Nn or β = (b1 , · · · , bn ) ∈ Zn .
This notation includes a set of operations on such multi-indexes, defined as follows
Theorem A.1 (Operations on Multi-indexes). Given a multi-index α ∈ Nn , one can define the
following operations on them
n
X
|α| = ai
i=1
n
(A.1)
Y
α! = ai !
i=1
n
Y
xα = xai i
i=1
n
(A.2)
Y ∂ |α| ∂ |α|
α
∂ = ∂iai = a1 an =
i=1
∂x1 · · · ∂xn ∂xα
207
APPENDIX A. USEFUL CONCEPTS 208
1
F̂ Φ̂a = Φ̂ 1 F̂ a ∈ R \ {0}
|a| a
e−iλb
F̂ Φ̂a T̂b = Φ̂ a1 F̂ a ∈ R \ {0}, b ∈ R
|a|
b
(A.3)
e−iλ a
F̂ T̂b Φ̂a = Φ̂ a1 F̂ a ∈ R \ {0}, b ∈ R
|a|
F̂ ∂ˆ = iλF̂
∂ˆ F̂ = F̂ [−ixf (x)] f ∈ L1 (R)
1 1
F̂ −1 = F̂ P̂ = P̂ F̂
2π 2π
F̂ −1 F̂ = F̂ F̂ −1 = 2π 1̂
B Common Fourier Transforms
In this appendix, you’ll find a table of some particular Fourier transforms that might pop up in Fourier
calculus, and might also be helpful for calculating more complex transforms.
Note that
ˆ
F̂[f ](λ) = f (x)e−iλx dx (B.1)
R
2 sin(λa)
1[−a,a] (x)
λ
rect(x) sinc(λ/2π)
sinc(x) rect(λ/2π)
209
APPENDIX B. COMMON FOURIER TRANSFORMS 210
2a
e−|x|
λ2 + a2
1
H(x)e−ax a>0
a + iλ
sgn(a)
H(ax)e−ax
a + iλ
1 π −a|λ|
e
a2 + x2 a
√
1 4ac − b2 π −β|λ|+i λ
β= >0 e 2a
a2 x2 + bx + c 2|a| βa
1 π −|λ|
e (|λ| + 1)
(x2 + 1)2 2
1 |λ|
− √2 λ π
πe sin 2 + 4
1 + x4
r
2 π − λ2
e−ax e 4a
a
r
2 π iλ2
eix (1 + i)e− 4
2
r
−ix2 π iλ2
e (1 − i)e 4
2
√ h
λ2 −π
i
cos x2 π cos 4
√ h 2 i
− π sin λ 4−π
sin x2
πλ
sech(x) π sech
2
x2 √ λ2
e− 2 (−i)n 2πe 2
p√ Hn (x) p√ Hn (λ)
π2n n! π2n n!
B.1. COMMON FOURIER TRANSFORMS 211
[Ces18] F. Cesi. Rudimenti di Analisi Infinito Dimensionale, Versione 0.3 (Armadillo). 2018.
[Fla18] Eugenio Montefusco Flavia Lanzara. Esercizi Svolti di Analisi Vettoriale con Elementi di
Teoria. Edizioni LaDotta, 2018. isbn: 9788898648429.
[Pre13] C. Presilla. Elementi di Analisi Complessa, Funzioni di una Variabile. Springer, 2013.
isbn: 9788847055001.
[Spi65] Michael Spivak. Calculus on Manifolds, a Modern Approach to Classical Theorems of
Advanced Calculus. Addison-Wesley Publishing Company, 1965. isbn: 0805390219.
213