Sachin Dev
Sachin Dev
Submitted
In partial fulfillment of the requirement for the Degree of
BACHELOR OF SCIENCE
Submitted by
SACHIN DEV
B.Sc. (PCM) - VI Semester
Roll No.: 2101034204016
2024
TOPIC
Submitted
In partial fulfillment of the requirement for the Degree of
BACHELOR OF SCIENCE
Submitted by
SACHIN DEV
B.Sc. (PCM) - VI Semester
Roll No.: 2101034204016
2024
Certificate
Certified that Sachin dev (roll no. 2101034204016) has carried out the
research work presented in this Dissertation entitled topic for the Degree of
Bachelor of Science from Maharishi University of Information Technology, Luc-
know under my supervision. The Dissertation embodies results of original work,
and studies are carried out by the student himself/herself and the contents of
the Dissertation do not form the basis for the award of any other degree to the
candidate or to anybody else from this or any other University/Institution.
i
Acknowledgement
First of all, I am extremely grateful to the almighty GOD for giving me the
opportunity to carry out research work With all due respect, I would like to thank
my supervisor. (Proff) Dr. Chinta-mani Tiwari, for their invaluable guidance and
support throughout my master’s program. Their expertise and encouragement
helped me to complete this research and write this thesis. I would also like
to thank Dr. Birendra Singh for serving on my thesis committee and providing
helpful feedback and suggestions. I would also like to thank my friends and family
for their love and support during this process. Without them, this journey would
not have been possible. Finally, I would like to thank all of the participants in my
study for their time and willingness to share their experiences. This work would
not have been possible without their contribution. I am grateful to everyone who
has supported me throughout this process. Without your help and guidance, this
thesis would not have been possible
Sachin dev
ii
Contents
Declaration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii
Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
1 INTRODUCTION 1
1.0.1 Are Complex Numbers Necessary? . . . . . . . . . . . . . 1
1.0.2 Basic Properties of Complex Numbers . . . . . . . . . . . 3
iii
4 Taylor series 15
4.0.1 Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
5 Laurent series 18
6 Meromorphic Funtions 21
6.0.1 Singularities . . . . . . . . . . . . . . . . . . . . . . . . . . 21
6.0.2 Zeros and Poles . . . . . . . . . . . . . . . . . . . . . . . . 22
9 Conclusion 28
Bibliography 30
Chapter 1
INTRODUCTION
Complex analysis is not an elementary subject, and the author of a book like
this has to make some reasonable assumptions about what his readers know
already. Ideally one would like to assume that the student has some basic
knowledge of complex numbers and has experienced a fairly substantial first
course in real analysis. But while the first of these requirements is realistic
the
second is not, for in many courses with an ”applied” emphasis a course in
com-
plex analysis sits on top of a course on advanced (multi-variable) calculus,
and
many students approach the subject with little experience of f.-0 arguments,
and with no clear idea of the concept of uniform convergence. This chapter
sets
out in summary the equipment necessary to make a start on this book, with
references to suitable texts. It is written as a reminder: if there is anything
you
don’t know at all, then at some point you will need to consult another book,
either the suggested reference or another similar volume.
Given that the following summary might be a little indigestible, you may
find it better to skip it at this stage, returning only when you come across
anything unfamiliar. If you feel reasonably confident about complex numbers,
then you might even prefer to skip Chapter 2 as well.
1
the larger set Q of rational numbers. Once again we get a bonus, for the equation
ax + b = 0 has a solution x = − ab in Q for all a ̸= 0 in Q and all b in Q.
When we come to consider a quadratic equation ax2 + bx + c = 0 (where
a, b, c ∈ Q and a ̸= 0) we encounter our first real difficulty. We may safely
assume that a, b and c are integers: if not, we simply multiply the equation by a
suitable positive integer. The standard solution to the equation is given by the
familiar formula √
−b ± b2 − 4ac
x=
2a
x2 + 4x + 13 = 0,
and decide to write i for i, the formula gives us two solutions x = −2 + 3i and
x = −2 − 3i. If we use normal algebraic rules, replacing i2 by −1 whenever it
appears, we find that
and the validity of the other root can be verified in the same way. We can
certainly agree that if there is a number system containing ”numbers” a + bi,
where a, b ∈ R, then they will add and multiply according to the rules
2
numbers as points on a line, of visualising these new complex numbers.
Can we find equations that require us to extend our new complex number
system (which we denote by C) still further? No, in fact we cannot: the important
Fundamental Theorem of Algebra, which we shall prove in Chapter 7, states that,
for all n ≥ 1, every polynomial equation with coefficients a0 , a1 , . . . , an in C and
an ̸= 0, has all its roots within C. This is one of many reasons why the number
system C is of the highest importance in the development and application of
mathematical ideas
z = z,
z+w = z + w,
zw = zw,
z+z = 2Re z,
z−z = 2iIm z.
3. |z + w| ≤ |z| + |w|;
3
Chapter 2
2.2 Cauchy-Riemann
The Cauchy-Riemann equations are a pair of partial differential equations that
describe the conditions for a complex-valued function to be analytic. Let f (z) =
u(x, y) + iv(x, y) be a complex-valued function defined on an open set U of the
complex plane. The Cauchy-Riemann equations are given by:
∂u ∂v ∂u ∂v
= and =−
∂x ∂y ∂y ∂x
4
where u(x, y) and v(x, y) are the real and imaginary parts of f (z), respectively,
and z = x + iy.
2.2.1 Question:
Determine whether the following function is analytic, and if so, find its derivative
at z = 1 + i:
2.2.2 Example:
Determine if the following function is analytic, and if so, find its derivative at
z = 2 − i:
z2 + i
f (z) =
z−i
Solution:
2 +i
Given the function f (z) = zz−i , let’s first check if it satisfies the Cauchy-
Riemann equations.
We have z = x + iy, so f (z) = u(x, y) + iv(x, y), where u(x, y) is the real part
and v(x, y) is the imaginary part.
First, let’s express f (z) in terms of its real and imaginary parts:
5
(x + iy)2 + i x2 − y 2 + 2ixy + i
f (z) = =
x + iy − i x + i(y − 1)
2 2
So, we have u(x, y) = x −y
x
and v(x, y) = 2xy+1
x
.
Now, let’s find the partial derivatives of u and v with respect to x and y:
∂u 2x(x) − (x2 − y 2 )
= =2
∂x x2
∂v 2(y − 1)(x) − (2xy + 1)
= =0
∂x x2
∂u −2y(x) − (x2 − y 2 )
= =0
∂y x2
∂v 2(x)(x) + 2(x)(y − 1) 2(x + y − 1)
= 2
=
∂y x x
The Cauchy-Riemann equations are satisfied if and only if:
∂u ∂v ∂u ∂v
= and =−
∂x ∂y ∂y ∂x
In this case, ∂u
∂x
= ∂y∂v
= 2, and ∂u ∂y
∂v
= − ∂x = 0, so the Cauchy-Riemann
equations are satisfied.
2 +i
Therefore, the function f (z) = zz−i is analytic.
To find its derivative at z = 2 − i, we can use the formula for the derivative
of a complex function:
∂u ∂v
f ′ (z) =
+i
∂x ∂x
∂u ∂v
Given that ∂x
= 2 and ∂x
= 0, we have:
f ′ (z) = 2
Therefore, the derivative of f (z) at z = 2 − i is f ′ (2 − i) = 2.
2.2.3 Example:
Determine if the function f (z) = x2 − y 2 + 2ixy is analytic, and if so, find its
derivative.
Solution:
Given the function f (z) = x2 − y 2 + 2ixy, let’s express it in terms of its real
and imaginary parts:
6
∂u
= 2x
∂x
∂v
= 2y
∂x
∂u
= −2y
∂y
∂v
= 2x
∂y
The Cauchy-Riemann equations are satisfied if and only if:
∂u ∂v ∂u ∂v
= and =−
∂x ∂y ∂y ∂x
In this case, ∂u
∂x
= ∂y∂v
= 2x, and ∂u
∂y
∂v
= − ∂x = −2y, so the Cauchy-Riemann
equations are satisfied.
Therefore, the function f (z) = x2 − y 2 + 2ixy is analytic.
To find its derivative, we can use the formula for the derivative of a complex
function:
∂u ∂v
f ′ (z) = +i
∂x ∂x
∂u ∂v
Given that ∂x
= 2x and ∂x
= 2y, we have:
f ′ (z) = 2x + 2iy
So, the derivative of f (z) is f ′ (z) = 2x + 2iy.
7
Z b
f (x)g(x) dx = 0
a
Orthogonal functions are widely used in various areas of mathematics, includ-
ing functional analysis, differential equations, and signal processing.
∂ 2u ∂ 2u
∇2 u = + =0
∂x2 ∂y 2
Harmonic functions arise naturally in the study of potential theory, where
they represent steady-state solutions to the heat equation and the wave equation
in regions with no sources or sinks.
2.2.7 Example:
Let’s consider the function f (z) = sin(x) cosh(y), where z = x + iy and x, y ∈ R.
We’ll verify whether f (z) is a harmonic function.
Solution:
Given f (z) = sin(x) cosh(y), we need to check if u(x, y) = sin(x) cosh(y)
satisfies Laplace’s equation:
∂ 2u ∂ 2u
∇2 u = + =0
∂x2 ∂y 2
Let’s compute the second partial derivatives:
∂ 2u
= − sin(x) cosh(y)
∂x2
∂ 2u
= sin(x) cosh(y)
∂y 2
Summing these, we have:
∂ 2u ∂ 2u
+ = − sin(x) cosh(y) + sin(x) cosh(y) = 0
∂x2 ∂y 2
Since the Laplacian of u(x, y) is zero, f (z) = sin(x) cosh(y) is a harmonic
function.
2.2.8 Example:
Let f (z) = z1 where z = x + iy, and x, y ∈ R. We’ll determine whether f (z) is a
harmonic function.
Solution:
Given f (z) = z1 , we can express it as f (z) = u(x, y) + iv(x, y), where u(x, y)
and v(x, y) are the real and imaginary parts of f (z), respectively.
8
Since f (z) is analytic on its domain except at z = 0, we can use the Cauchy-
Riemann equations to find the real and imaginary parts of f (z).
The Cauchy-Riemann equations are:
∂u ∂v ∂u ∂v
= and =−
∂x ∂y ∂y ∂x
1 1
For f (z) = z
= x+iy
, we have:
x y
u(x, y) = and v(x, y) = −
x2 + y2 x2 + y2
Now, let’s compute the Laplacian of u(x, y) to determine if it satisfies Laplace’s
equation:
∂ 2u ∂ 2u
∇2 u =
+
∂x2 ∂y 2
∂ 2x(x2 + y 2 ) − 2x2
∂ −2xy
= +
∂x (x2 + y 2 )2 ∂y (x2 + y 2 )2
(x4 − 6x2 y 2 + y 4 + 2x2 − 2y 2 )
=
(x2 + y 2 )3
1
Since ∇2 u is not zero, f (z) = z
is not a harmonic function.
2.2.9 Example:
Let f (x) = sin(nx) and g(x) = cos(mx), where n and m are positive integers.
We’ll determine if f (x) and g(x) are orthogonal over the interval [0, π].
Solution:
To determine if f (x) and g(x) are orthogonal over the interval [0, π], we need
to compute their inner product and check if it equals zero:
Z π
f (x)g(x) dx
0
Example :
Show that the function f (z) = |z|2 is not analytic even though it is continuous
everywhere and differentiable at zero.
9
Solution:
Let f (z) = 2 + 2i.
= lim z + z∆z
∆z→0
r · e−iθ
= lim z + z · (where ∆z = r · eiθ )
r→0 r · e−iθ
= z + z · e−2iθ
which does not tend to a unique limit as this limit depends on θ, the argument
of ∆z, which is arbitrary.
f (z) is not differentiable at any point z = 0 and hence f (z) is not analytic at
any point z = 0. Thus, continuity and differentiability of f (z) at z = 0 are as
follows:
f (z) − f (0) = 2 − 0 = r2 , the modulus of z, for which 0 < 2 − 0, where r is
for |w| for all w for which 0 < 2 − 0, and hence f (z) has the limit f (0) as z → 0
and hence f (z) is continuous at z = 0. Also,
f (0 + ∆z) − f (0) ∆z · ∆z
f ′ (0) = lim = lim = lim ∆z = 1
∆z→0 ∆z ∆z→0 ∆z ∆z→0
10
Chapter 3
The Cauchy Integral Theorem, Basic Version, states that if D is a domain and
f (z) is analytic in D with f ′ (z) continuous, then
I
f (z) dz = 0
C
for any closed contour C lying entirely in D having the property that C is con-
tinuously deformable to a point.
We also showed that if C is any closed contour oriented counterclockwise in
C and a is inside C, then I
1
dz = 2πi.
C z −a
Our goal now is to derive the celebrated Cauchy Integral Formula which can
be viewed as a generalization of (∗).
11
and
f (z) − f (a) f (z) − f (a)
I I
dz = dz
C z−a Ca z−a
since the integrand f (z)−f
z−a
(a)
is analytic everywhere except at z = a and its deriva-
tive is continuous everywhere except at z = a so that integration over C can be
continuously deformed to integration over Ca .
However, if we write
f (z) − f (a)
I I
f (z)
dz − 2πif (a) = dz.
C z −a Ca z−a
f (z) − f (a)
Z
lim dz
r→0 C
a
z−a
Z
f (z)
= lim dz − 2πf (a)i
r→0 Ca z − a
f (z) − f (a)
Z
= lim dz.
r→0 C
a
z−a
Hence, the proof will be complete if we can show that
f (z) − f (a)
Z
lim dz = 0.
r→0 C
a
z−a
To this end, suppose that Mr = max{|f (z) − f (a)| : z on Ca }. Therefore, if z
is on Ca = {|z − a| = r}, then
f (z) − f (a)
Z
dz
Ca z−a
f (z) − f (a)
Z
≤ dz
Ca z−a
Z
Mr
≤ dz
Ca r
1
= Mr · length(Ca )
r
1
= Mr · 2πr
r
= 2πMr .
However, since f (z) is analytic in D, we know that f (z) is necessarily contin-
uous in D so that limz→a |f (z) − f (a)| = 0 or, equivalently, limr→0 Mr = 0.
Therefore,
f (z) − f (a)
Z
lim dz ≤ lim(2πMr ) = 0
r→0 Ca z−a r→0
12
as required.
1
R z·ez
Example:. Compute 2πi C z−i
dz where C = {|z| = 2} is the circle of radius
2 centered at 0 oriented counterclockwise.
Solution. Observe that f (z) = z · ez is entire, f ′ (z) = z · ez is continuous, and
i is inside C. Therefore, by the Cauchy Integral Formula,
z · ez
Z Z
1 1 f (z)
dz = dz = f (i) = i · ei .
2πi C z − i 2πi C z − i
z
Example : Compute C zz·e
R
2 +1 dz where C = {|z| = 2} is the circle of radius 2
and so
z · ez zez zez
Z Z Z
i i
2
dz = dz − dz
C z +1 2 C z+i 2 C z−i
Let f (z) = z ·ez . Note that f (z) is entire and f ′ (z) = z ·ez +z ·ez is continuous.
Since both i and −i are inside C, the Cauchy Integral Formula implies...
Z
zez dz = 2πif (−i) = 2πi · (−i)e−i = 2πe−i
ZC
zez dz = 2πif (i) = 2πi · iei = −2πei
C
so that
zez
Z
i −i i
2
dz = · 2π · e − · (−2π) · ei
C z +1 2 2
−i
= πi · e + πi · ei
ei + e−i
= 2πi ·
2
= 2πi · cos 1.
13
3.0.2 Theorem (The derivative of an analytic function)
Let f(z) be analytic within and on a closed contour C and let z0 be any point
inside C, then
Z
′ 1 f (z)
f (z0 ) = dz
2πi C (z − z0 )2
f (z0 + h) − f (z0 )
Z
1 f (z)
lim = dz
h→0 h 2πi C (z − z0 )2
R f (z)
Hence, f (z) is differentiable at z0 and f ′ (z0 ) = 2πi 1
C (z−z0 )2
dz, which is
′
Cauchy’s integral formula for f (z) at points within C.
H e2x −2
Example: Using Cauchy’s Integral Formula, show that C (z+1) 4 dz = 8πe 3i
where C is the circle |z| = 3.
Solution: jBy Cauchy’s Integralk Formula for derivatives, we have
(n) n!
H f (z)dz
f (z0 ) = 2πi C (z−z0 )n+1 where f (z) is analytic inside and on C. In the
present case, C is |z| = 3, f (z) = e2z , z0 = −1, n = 3 and f (z) is analytic
H einside
2x
(3) 2 −2 13
and on circle |z| = 3. Also, f (−1) = 8e ... (1) becomes 8e = 2πi C (z+1)4 dz
H e2z −2
C (z+1)4
dz = 8πe3 .i. Hence the result.
14
Chapter 4
Taylor series
In real analysis there are distinctions between functions that are dif-
ferentiable, infinitely differentiable (having derivatives of all orders),
and analytic (having a Taylor4 series expansion). We have already
seen that a holomorphic function is infinitely differentiable. In fact it
also has a Taylor series expansion. Precisely, we have the following
theorem:
4.0.1 Theorem
Let c ∈ C and suppose that the function f is holomorphic in some neighborhood
N (c, R) of c. Then, within N (c, R), we have
∞
X f (n) (c)
f (z) = an (z − c)n , an = , where n = 0, 1, 2, . . . .
n=0
n!
Proof:
It is helpful first to record the sum of the following finite geometric series:
z − c (z − c)2 hn 1 hn+1
1+ + +. . .+ = − . (7.6)
h h2 (z − c)n+1 z − c − h (z − c)n+1 (z − c − h)
Let 0 < R1 < R2 < R. Then, f is holomorphic throughout the closed disc
N (c, R2 ). Let C be the circle (c, R2 ), and let c+h ∈ N (c, R1 ). Then, by Theorems
7.1 and 7.5 and Equation (7.6), we have
Z
1 f (z)
f (c + h) = dz
2πi C z − c − h
Z Z Z
1 2 n f (z)
= f (z)(z − c) dz + h f (z)(z − c) dz + . . . + h dz
2πi C C C (z − c)n+1
15
hn (n)
f (c) + hf ′ (c) + . . . + f (c) + En
n!
where,
n
|h|(n+1)
M M |h| |h|
⇒ |En | ≤ · 2πR2 · (n+1) = ·
2π R2 · (R2 − R1 ) (R2 − R1 ) R2
|h|
Since < 1 we deduce that En → 0 as n → ∞.
R2
Thus, f(c + h) = ∞
P hn (n)
P∞
n=0 n!
f (c) and substituting z = c+h gives f (z) = n=0 an (z−
c)n , an = n!1 f (n) (c)
Z
1 f (z)
= dz.
2πi C (z − c)n+1
Example:
Show that, for all real a,
∞
X a(a − 1) · · · (a − n + 1)
(1 + z)a = 1 + zn (|z| < 1).
n=1
n!
16
Solution
The function is holomorphic in the open set C \ {−1}, so certainly in the
neighborhood N (0,1). Moreover, the principal argument of f1 + 2 lies safely in
the interval − π2 , π2 and so there is no ambiguity in the meaning of (1 + z)a = ea .
Hence, log(1 + 2). A routine calculation gives f (n) (z) = a(a − 1) · · · (a − n + 1)(1 +
(n)
z)a−n for n = 1, 2, . . .. Thus, f n!(0) = a(a−1)···(a−n+1)
n!
and
∞
a
X a(a − 1) · · · (a − n + 1)
(1 + z) = 1 + zn (|z| < 1)
n=1
n!
as required. □
article amsmath, amssymb
We sometimes want to say that the Taylor series f (z) = ∞ n
P
n=0 an (z −c) is the
Taylor series of f at c, or that the series is centered on c, or that c is the center of
the series. To qualify Remark 7.17 above, the Taylor series of a function is unique
once we choose the center, but a function has many different Taylor series, with
1
different centers. For example, the function (1+z) 2 is holomorphic in C \ {−1} and
17
Chapter 5
Laurent series
18
CR+ and CR− , respectively, as also indicated in figure 16.1. Finally, keeping track
of the orientations of the subcurves, let Γ1 and Γ2 be the closed curves given by
f (ζ) f (ζ)
IΓ+ dζ and IΓ− dζ.
ζ −z ζ −z
By the Cauchy integral theorem, we know that the integral over Γ− (the curve
not encircling z) is 0, while the basic Cauchy integral formula tells us that the
other integral is i2πf (z). Thus,
Z Z Z Z
f (ζ) f (ζ) f (ζ) f (ζ) f (ζ) f (ζ)
i2πf (z)+0 = IΓ+ dζ+IΓ− dζ = dζ+ dζ− dζ+ dζ
ζ −z ζ −z + ζ − z
CR l1 ζ − z Cr+ ζ − z l2 ζ − z
That is, Z Z
1 f (ζ) f (ζ)
f (z) = dζ − dζ .
i2π CR ζ −z Cr ζ −z
Consider the first integral on the right side of equation (16.1). More precisely,
let ζ ∈ CR . Then |z − z0 | < R = |ζ − z0 |, which means that
z − z0
< 1,
ζ − z0
and the formula for the sum of the geometric series can be applied as follows:
∞ k X ∞
1 1 1 1 1 X z − z0 (z − z0 )k
= = · = · = .
ζ −z ζ − z0 − (z − z0 ) ζ − z0 1 − z−z
ζ−z0
0
ζ − z0 k=0 ζ − z0 k=0
(ζ − z0 )k+1
Consequently,
∞ ∞ Z ∞
(z − z0 )k (z − z0 )k (z −
Z Z
1 f (ζ) 1 X 1 X X
dζ = k+1
f (ζ) dζ = k+1
f (ζ) dζ =
i2π CR ζ −z i2π CR k=0 (ζ − z0 ) i2π k=0 CR (ζ − z0 ) k=0
i2
19
where Z
1 f (ζ)
ak = dζ
i2π C (ζ − z0 )k+1
and C is any simple, counterclockwise oriented loop in A about z0 . An analogous
derivation can be done for the other integral in equation (16.1).
m
Example :
Calculate the first few terms of the Laurent series for sin1 z at 0.
Solution
The function sin1 z has a singularity at 0 but is otherwise holomorphic in the
3
neighborhood N (0, π). We know that, as z → 0, sin z = z − z6 + O(z 5 ).
Hence, for z near 0,
2 −1
z2
1 1 z 4 1 1 z
= 1− + O(z ) = 1+ + O(z ) = + + O(z 3 ).
4
sin z z 6 z 6 z 6
Example :
Show that cot z = z1 − z3 + O(z 3 ).
Solution
cos z z2 1 z
4
From what we know already, cot z = sin z = 1 − 2
+ O(z ) z
+ 6
+ O(z 3 ) =
1
− z 12 − 61 + O(z 3 ) = z1 − z3 + O(z 3 ).
z
article amsmath, amssymb lipsum
20
Chapter 6
Meromorphic Funtions
6.0.1 Singularities
There is a general principle in the theory, already implicit in Riemann’s work,
which states that analytic functions are in an essential way charac- terized by
their singularities. That is to say, globally analytic functions are ”effectively”
determined by their zeros, and meromorphic functions by their zeros and poles.
While these assertions cannot be formulated as precise general theorems, there
are nevertheless significant instances
where this principle applies. We begin this chapter by considering singulari-
ties, in particular the different kind of point singularities (”isolated” singularities)
that a holo- morphic function can have.
• removable singularities
• poles
• essential singularitiesIn order of increasing severity, these are:
• removable singularities
• poles
• essential singularities
The question that occurs is: what happens if f has a pole in the
interior of the curve? To try to answer this question consider the
21
example f(z) = 1/z, and recall that if C is a (positively oriented) circle
centered at 0, then
Z
dz
= 2πi
C z
vmm
This turns out to be the key ingredient in the calculus of residues.
A new aspect appears when we consider indefinite integrals of holomor-
phic functions that have singularities. As the basic example f(z) = 1/2
shows, the resulting ”function” (in this case the logarithm) may not be
single-valued, and understanding this phenomenon is of importance for
a number of subjects. Exploiting this multi-valuedness leads in effect
to the ”argument principle.” We can use this principle to count the
number of zeros of a holomorphic function inside a suitable curve. As
a simple consequence of this result, we obtain a significant geometric
property of holomorphic functions: they are open mappings. From
this, the maxi- mum principle, another important feature of holomor-
phic functions, is an easy step.
22
to 0 on the negative real axis. Finally, h oscillates rapidly, yet remains
bounded, as z approaches the origin on the imaginary axis.
Since singularities often appear because the denominator of a frac-
tion vanishes, we begin with a local study of the zeros of a holomorphic
function.
A complex number z0 is a zero for the holomorphic function f if
f (z0 ) = 0. In particular, analytic continuation shows that the zeros of
a non-trivial holomorphic function are isolated. In other words, if f is
holomorphic in C and f (z0 ) = 0 for some z0 ∈ C, then there exists an
open neighborhood U of z0 such that f (z) ̸= 0 for all z ∈ U \ {z0 }, unless
f is identically zero. We start with a local description of a holomorphic
function near a zero.
23
Chapter 7
Theorem:
Suppose that f is holomorphic in an open set containing a circle C and
its interior, except for a pole at z0 inside C. Then
I
f (z) dz = 2πi resz0 f.
C
where Cϵ is the small circle centered at the pole zc and of radius ϵ. Now
we observe that I
1 a1
dz = a1
2πi C z − z0
is an immediate consequence of the Cauchy integral formula (Theorem
4.1 of the previous chapter), applied to the constant function f = a−1 .
Similarly,
I
1 ak
dz = 0
2πi C (z − z0 )k
When k > 1, by using the corresponding formulae for the derivatives
(Corollary 4.2 also in the previous chapter). But we know that in a
neighborhood of z0 we can write
24
R R
C
G(z) dz = 0, hence C f (z) dz = a−1 . This implies the desired result.
This theorem can be generalized to the case of finitely many poles
in the circle, as well as to the case of toy contours.
**Corollary 2.2** Suppose that f is holomorphic in an open set
containing a circle C and its interior, except for poles at the points
z1 , . . . , zN inside C. Then
Z N
X
f (z) dz = 2πi reszk f.
C k=1
For the proof, consider a multiple keyhole which has a loop avoiding
each one of the poles. Let the width of the corridors go to zero. In the
limit, the integral over the large circle equals a sum of integrals over
small circles to which Theorem 2.1 applies.
**Corollary 2.3** Suppose that f is holomorphic in an open set
containing a toy contour γ and its interior, except for poles at the
points z1 , . . . , z2N inside. Then
Z N
X
f (z) dz = 2πi reszk f.
γ k=1
25
Chapter 8
26
In conclusion, complex analysis is a versatile and powerful tool with
applications across various fields. Its ability to simplify complex prob-
lems, provide elegant solutions, and offer insights into the behavior of
systems makes it an indispensable tool in mathematics and its appli-
cations.
27
Chapter 9
Conclusion
28
I n
X
f (z)dz = 2πi Res(f, zk ),
C k=1
29
Bibliography
[7] Ahuja, G., 1992, The product of an arbitrary functions and the
fractional derivative of direct delta distribution, Jour. of the In-
dian Math Soc., 58(4), 231–36.
[12] Fisher, B., 1976, Neutrices and the product of distributions, Stu-
dia Mathematica T.L., VII, 263–274.
30
[15] Fisher, B., 1982, A theorem on the products of distributions, Pub-
lications Mathematicae T, 27, 243–249.
[22] Osler, T., 1970, Leibniz rule for fractional derivatives Generalized
and an application to infinite series, SIAM J. Appl., 18, .658–74.
[23] Osler, T., 1971, LeTaylor’s series generalized for fractional deriva-
tives and applications, SIAM J. Appl., 2, .658–74.
[24] Osler, T., 1972, A further extension of the Leibniz rule to frac-
tional derivatives and its relation to Parseval’s formula, SIAM J.
Appl., 3, 1–16.
31