Sei sulla pagina 1di 130

Complex Analysis

Istvn Mez, PhD


Contents
I Complex numbers, sequences, series, topology 1
1 Complex numbers 3
1.1 The denition of the complex numbers and the basic operations 3
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Polar form of complex numbers . . . . . . . . . . . . . . . . . 6
1.2.1 The argument function . . . . . . . . . . . . . . . . . . 8
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2 Complex sequences and series 10
2.1 Complex sequences . . . . . . . . . . . . . . . . . . . . . . . . 10
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Complex series . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.1 Complex power series . . . . . . . . . . . . . . . . . . . 15
2.2.2 The Polylogarithm functions . . . . . . . . . . . . . . . 18
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3 Topological basics 20
3.1 Metric spaces and their properties . . . . . . . . . . . . . . . . 20
3.2 Continuous functions on metric spaces . . . . . . . . . . . . . 23
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
II Complex Dierentiability and the Integral For-
mula of Cauchy 25
4 Dierentiability 27
4.1 The Cauchy-Riemann equations . . . . . . . . . . . . . . . . . 28
5 The exponential and logarithm function and their deriva-
tives 31
5.1 The sine and cosine functions . . . . . . . . . . . . . . . . . . 35
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.2 The logarithm function . . . . . . . . . . . . . . . . . . . . . . 38
i
5.3 General powers . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
6 Path integrals 43
6.1 Curves in the complex plane . . . . . . . . . . . . . . . . . . . 43
6.2 The path integral . . . . . . . . . . . . . . . . . . . . . . . . . 45
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
7 The theorem of Cauchy 49
7.1 The theorem of Cauchy for triangle paths . . . . . . . . . . . . 49
7.2 Cauchys theorem for starlike domains . . . . . . . . . . . . . 51
7.3 Cauchys theorem on simply connected domains . . . . . . . . 53
8 Cauchy Integral formula 56
8.1 A special version of the Cauchy Integral Formula . . . . . . . 56
8.2 Example circle integrals . . . . . . . . . . . . . . . . . . . . . 58
8.3 The general version of Cauchys integral formula . . . . . . . . 61
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
III Local and global analysis 65
9 The Laurent series 67
9.1 The Laurent series. Main part of a function . . . . . . . . . . 67
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
10 Singularities and zeroes of analytic functions 73
10.1 Classication of singularities . . . . . . . . . . . . . . . . . . . 73
10.1.1 The denition of zeros . . . . . . . . . . . . . . . . . . 77
10.2 The connection between zeros and poles . . . . . . . . . . . . 77
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
10.3 The Theorem of Casorati and Weierstrass . . . . . . . . . . . 80
10.4 Further properties of zeros and poles . . . . . . . . . . . . . . 81
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
11 The Residue Theorem of Cauchy and its applications 84
11.1 The residue . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
11.2 The winding number . . . . . . . . . . . . . . . . . . . . . . . 87
11.3 The Residue Theorem of Cauchy . . . . . . . . . . . . . . . . 88
11.3.1 Applications of the Residue Theorem . . . . . . . . . . 89
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
ii
12 Some additional theorems on zeros and poles. . . 94
12.1 The number of zeros and poles . . . . . . . . . . . . . . . . . . 94
12.2 Entire functions - the theorem of Liouville . . . . . . . . . . . 96
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
13 The theorems of Weierstrass and Mittag-Leer 98
13.1 Innite products . . . . . . . . . . . . . . . . . . . . . . . . . 98
13.2 The Weierstrass Product Theorem . . . . . . . . . . . . . . . . 101
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
13.3 The Mittag-Leer expansion . . . . . . . . . . . . . . . . . . . 104
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
IV Special functions 108
14 Analytic extension 110
14.1 The notion of analytic extension . . . . . . . . . . . . . . . . . 110
14.2 The Riemann zeta function . . . . . . . . . . . . . . . . . . . 110
15 The Riemann zeta function 113
15.1 The values of at positive even integers . . . . . . . . . . . . 113
15.2 The connection between primes and the Riemann zeta function115
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
16 The Euler Gamma function 117
16.1 The connection between the and functions . . . . . . . . . 120
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
17 Conformal maps 122
17.1 Some examples of transformations on the complex plane . . . 122
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
17.2 Conformal mappings . . . . . . . . . . . . . . . . . . . . . . . 124
17.3 The Mbius transformations . . . . . . . . . . . . . . . . . . . 124
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
iii
Part I
Complex numbers, sequences,
series, topology
1
2
Complex function theory is a widely applicable branch of mathematics.
Its powerful results can be applied to solve problems in other parts of math-
ematics and even physics.
Complex analyis is based on complex numbers, which are extensions of
the real ones. Many notions of real analysis can be carried to this more gen-
eral situation. Therefore in the foregoing chapter we introduce the complex
numbers and (re)introduce many notions from the real analysis, such that
sequences, series, convergence, open and closed sets, compactness, etc.
1
Complex numbers
1.1 The denition of the complex numbers and
the basic operations
If we try to solve the quadratic equation
x
2
+ x + 1 = 0,
we nd that
x
1,2
=
1

1 4
2
.
In the square root there is 3, which is not a square of any real number.
Instead of saying that this equation cannot be solvable, we can try to calculate
with these type of numbers.
x
1,2
=
1

1 4
2
=
1
2

3
2

1.
As it is easy to see, any solution of a quadratic equation can be rewritten as
a + bi,
where a, b R and i is the symbol

1. Can we calculate with these
numbers as we did with the reals? Can we add, multiply divide them? Let
us take an example:
(5 + 3i) + (4 + 2i) = 5 + 3

1 + 4 + 2

1 = 9 + 5

1 = 9 + 5i.
(5 + 3i) (4 + 2i) = 5 4 + 5 2i + 3i 4 + 3i 2i = 20 + 10i + 12i + 6i
2
.
Since i =

1, i
2
must be 1. Therefore
(5 + 3i) (4 + 2i) = 20 + 22i + 6(1) = 14 + 22i.
3
4 1. chapter. Complex numbers
Now let us try the division. This is a bit more tricky:
5 + 3i
4 + 2i
=
5 + 3i
4 + 2i

4 2i
4 2i
=
(5 + 3i)(4 2i)
(4 + 2i)(4 2i)
=
20 10i + 12i 6i
2
16 8i + 8i 2i
2
=
26 + 2i
18
=
26
18
+
2
18
i =
13
9
+
1
9
i.
We can write the result of the division in the form a + bi.
Therefore we can see, that we can calculate with these quantities as with
the usual real numbers, so we have the right to call these quantities as "num-
bers".
1.1.1. Denition. The algebraic expressions of the form
a + bi,
where a, b R and i =

1 are called complex number. The symbol i is


called the imaginary unit.
The set of complex numbers is denoted by C. That is,
C = a + bi [ a, b R.
The next task after the denition is to give the rules, how we calculate
with these new numbers. As we have already seen, the four basic opeara-
tions (addition, substraction, multiplication, division) are easy to describe.
Formally, we can have the following proposition.
1.1.2. Proposition. Let a+bi C and c+di C are two complex numbers.
Then
(a + bi) + (c + di) = (a + c) + (b + d)i,
(a + bi) (c + di) = (a c) + (b d)i,
(a + bi) (c + di) = (ac bd) + (ad + bc)i,
a + bi
c + di
=
_
ac + bd
c
2
+ d
2
_
+
_
bc ad
c
2
+ d
2
_
i.
In the division at least one of c and d is not zero.
For one who is familiar with abstract algebra, we remark that the set
C is a eld with respect to the above dened addition and multiplication,
therefore C is often called as the eld of complex numbers.
It will be useful to dene three functions on the set of complex numbers:
1 : C R; 1(a + bi) = a,
: C R; (a + bi) = b,
: C C; a + bi = a bi.
1.1. The denition of the complex numbers and the basic operations 5
The rst function is called real part, the second one is the imaginary part.
The third function is the conjugate function.
It is worth to note the following simple identities.
1.1.3. Proposition. For any complex number z, w C
z = z,
zz = [z[
2
,
or, in algebraic form,
(a + bi)a + bi = (a + bi)(a bi) = a
2
+ b
2
,
[z[ = [z[,
z + w = z + w,
zw = zw,
1(z) =
1
2
(z + z),
(z) =
1
2i
(z z),
1(z + w) = 1(z) +1(w),
(z + w) = (z) +(w),
1
z
=
1
[z[
2
z.
Some useful inequalities can also be easily proven.
1.1.4. Proposition. For any complex number z, w C
[1(z)[ [z[,
[(z)[ [z[,
[[z[ [w[[ [z + w[ [z[ +[w[,
[z[ [1(z)[ +[(z)[.
Exercises
1.1.5. Exercise. Calculate (3 + 2i) + (5 7i), (2 + 3i)(4 6i), and
132i
78i
.
1.1.6. Exercise. Calculate z w
2
, where z =
1
2
4i and w = 5 + 5i.
1.1.7. Exercise. Calculate (2 + 3i)
4
.
1.1.8. Exercise. What is the square root of i?
1.1.9. Exercise. Determine i
2013
.
1.1.10. Exercise. Calculate the sum i + i
2
+ i
3
+ + i
100
.
6 1. chapter. Complex numbers
1.2 Polar form of complex numbers
Since the complex number z = a + bi have two coordinates, we can draw
them in coordinate system.
The angle is called the angle or argument. The quantity r is the absolute
value of z. The absolute value is also often called as the modulus of z, and
it is denoted by [z[.
Since
cos =
a
r
and
sin =
b
r
,
we can write that
z = a + bi = r cos + ir sin = r(cos + i sin ).
Thus we have got the following proposition.
1.2.1. Theorem. Every complex number z = a + bi can be written in the
form
z = r(cos + i sin ),
where r 0 is a real number (the length of z), and [0, 2[ is the angle
of z.
1.2.2. Denition. The above representation of z is called polar form. The
representation z = a + bi is the algebraic representation.
The polar form is extremely useful if we want to calculate the powers and
roots of a complex number. Let us x two complex numbers in polar form:
z
1
= r
1
(cos
1
+ i sin
1
),
z
2
= r
2
(cos
2
+ i sin
2
).
1.2. Polar form of complex numbers 7
Then their product is
z
1
z
2
= r
1
r
2
(cos
1
+ i sin
1
)(cos
2
+ i sin
2
) =
r
1
r
2
[(cos
1
cos
2
sin
1
sin
2
) + i(sin
1
cos
2
+ cos
1
sin
2
)].
Then we can use the well known addition theorems for the trigonometric
functions:
cos(
1
+
2
) = cos
1
cos
2
sin
1
sin
2
,
sin(
1
+
2
) = sin
1
cos
2
+ cos
1
sin
2
.
Then we get that
z
1
z
2
= r
1
r
2
[cos(
1
+
2
) + i sin(
1
+
2
)].
By an obvious reason, the division is as follows:
z
1
z
2
=
r
1
r
2
[cos(
1

2
) + i sin(
1

2
)].
We arrived at the next proposition.
1.2.3. Proposition. In polar form we multiply and divide the complex num-
bers
z
1
= r
1
(cos
1
+ i sin
1
),
z
2
= r
2
(cos
2
+ i sin
2
).
as
z
1
z
2
= r
1
r
2
[cos(
1
+
2
) + i sin(
1
+
2
)], (1.1)
and
z
1
z
2
=
r
1
r
2
[cos(
1

2
) + i sin(
1

2
)] (r
2
,= 0). (1.2)
In special, the nth (n 1 is an integer) powers and roots of a complex number
z = r(cos + i sin ) become
z
n
= r
n
[cos(n) + i sin(n)], (1.3)
and
n

z =
n

r
_
cos
_
+ 2k
n
_
+ i sin
_
+ 2k
n
__
(k = 0, 1, . . . , n 1).
(1.4)
The penultimate formula in the theorem
z
n
= r
n
[cos(n) + i sin(n)]
is the so-called de Moivres formula
1
.
1
Abraham de Moivre (1667-1754) french mathematician.
8 1. chapter. Complex numbers
1.2.1 The argument function
The angle of a complex number is important enough to introduce a new
function. We have seen, that for every z C there is an angle , which is
also called as argument. This is what the argument function arg maps to a
complex number. To be more formal, if we have
z = r(cos + i sin ),
then
arg : C R; arg(z) = .
Note that this angle is not unique, since +2k represents the same angle,
if k Z. Therefore the arg function maps innite values to a same complex
number z. For example, if z = 1 + i, then its argument is = 45

=

4
.
Hence
arg(1 + i) =
_

4
,

4
2,

4
4, . . .
_
.
If we restrict us to the interval [0, 2[, the argument becomes unique. To
be more general, we can say that if we restrict us to any interval of the form
[, + 2[, the argument becomes unique. This phenomenon is connected
to an important notion, which is called branch. Any interval on which the
argument function is unique, is a branch of this function. Hence the intervals
[0, 2[, [2, 4[, [4, 6[, . . . , [2, 0[, 4, 2[, . . .
are the branches of arg. From these branches it is very natural to choose
[0, 2[ which we shall call as the principal branch. On the principal branch
arg(1 + i) =

4
.
Remark that we can index the branches with the integers. That is, by an
obvious reason on the 1th branch
arg(1 + i) =

4
2 =
7
4
,
while on the 1th branch
arg(1 + i) =

4
+ 2 =
9
4
.
The zeroth branch is the principal branch.
Now we list the two most basic functional identities of the argument
function which help to calculate its value.
1.2. Polar form of complex numbers 9
1.2.4. Proposition. For any two complex numbers z
1
, z
2
we have
arg(z
1
z
2
) = arg(z
1
) + arg(z
2
),
arg
_
z
1
z
2
_
= arg(z
1
) arg(z
2
).
The proof of these identities comes from Proposition 1.2.3.
Finally we remark that we have already met with branches in the deni-
tion of the root function, see (1.4). The nth root function have n branches,
since a complex number have n pieces of nth roots.
Exercises
1.2.5. Exercise. Determine the absolute value, the argument and nally
the polar form of the complex numbers 1 + i, i, 1 + 2i.
1.2.6. Exercise. Determine [3 2i[.
1.2.7. Exercise. Find the algebraic and polar form of 1/i.
1.2.8. Exercise. Calculate

1 + i and (1 + i)
100
.
1.2.9. Exercise. Draw in the coordinate system the following sets:
A = z C [ 1(z) < 3,
B = z C [ (z) 2,
C = z C [ [z[ < 1,
D = z C [ [z[ 1,
E = z C [ [z 1[ < 3,
F = z C [ [z + 3 2i[ 4.
1.2.10. Exercise. What is the value of arg(i) on the principal branch, and
on the second branch?
1.2.11. Exercise. Determine the function value
arg
_
1 i
i
_
on the principal branch.
1.2.12. Exercise. Draw in the coordinate system the following sets:
A =
_
z C

[z[ > 1, 0 < arg(z) <



4
_
,
B =
_
z C

[z[ arg(z) =

2
_
.
2
Complex sequences and series
2.1 Complex sequences
Since we have the absolute value of complex numbers, which serve as a base
for measure distance, it is easy to dene the convergence of sequences in the
set C.
2.1.1. Denition. The function z : N C is called complex sequence. By
convenience, we write z
n
in place of z(n).
This sequence converges if there is a z C such that for every > 0
there is a natural number n
0
such that for any n > n
0
[z
n
z[ < . In this
case we say that z
n
converges to z, and we denote this fact by
lim
n
z
n
= z.
If the sequence converges, this existing z is called the limit of the sequence
z
n
.
Fortunately, we can trace back nding the complex limit of the sequence
to nding the limit of its real and imaginary part. This is the statement of
the following theorem.
2.1.2. Theorem. Let z
n
be a complex sequence with the limit z. Then
lim
n
z
n
= z
holds if and only if
lim
n
1(z
n
) = 1(z) and lim
n
(z
n
) = (z).
Proof. First let us suppose that
lim
n
z
n
= z.
10
2.1. Complex sequences 11
Then we will prove that the real and imaginary part of z
n
converges to 1(z)
and (z), respectively. The limit means that for a xed > 0 there is an
index n
0
for which
[z
n
z[ < for all n > n
0
.
Then
[1(z
n
) 1(z)[ = [1(z
n
z)[ [z
n
z[ < .
The rst equality comes from Proposition 1.1.3., the inequality comes from
Porposition 1.1.4. We have proven that for any xed > 0 [1(z
n
)1(z)[ <
for any n > n
0
. Thus
lim
n
1(z
n
) = 1(z).
The same can be proven for the imaginary parts.
To complete the proof, we have to show that the second statement implies
the rst one, that is, the convergence of the real and imaginary parts to 1(z)
and (z), respectively, imply the convergence of z
n
to z. Let > 0 such that
[1(z
n
) 1(z)[ <

2
,
and
[(z
n
) (z)[ <

2
.
and let n
0
and m
0
be the indices for this . Then
[z
n
z[ = [1(z
n
) + i(z
n
) (1(z) + i(z))[
= [1(z
n
) 1(z) + i((z
n
) (z))[
[1(z
n
) 1(z)[ +[i((z
n
) (z))[
= [1(z
n
) 1(z)[ +[i[[(z
n
) (z)[
= [1(z
n
) 1(z)[ +[(z
n
) (z)[
<

2
+

2
= .
holds for any n > max(n
0
, m
0
). We have proven the proposition.
As a demonstration, we show a plot of a convergent and a divergent
sequence. The rst plot belongs to the sequence
z =
1 + i
n
n
2
:
12 2. chapter. Complex sequences and series
The second plot below belongs to
z = (1 + i)
n
:
2.2. Complex series 13
Exercises
2.1.3. Exercise. Find the limit of the complex sequence
z
n
=

n + i(n + 1)
n
.
2.1.4. Exercise. Find the limit if exists of the following complex se-
quences
z
n
=
n + i
n
n
,
z
n
=
(n + i)(1 ni)
n
2
,
z
n
= i
1
n
.
(In the last example take the principal branch of the root function.)
2.1.5. Exercise. Show that the complex sequence z
n
= (1 + i)
n
diverges.
What is the limit of z
n
=
(1+i)
n
2
n
?
2.1.6. Exercise. Let z
n
be a sequence such that lim
n
z
n
= z. What is
the limit of z
n
?
2.2 Complex series
The denition of the convergence for complex series is totally the same as
for the real ones.
2.2.1. Denition. Let z
n
be a complex sequence, and let us dene the fol-
lowing (complex) sequence:
s
n
:= z
1
+ z
2
+ + z
n
.
If s
n
converges as a sequence, we say that the innite sum
z
1
+ z
2
+
exists and the limit of s
n
is the value of this sum. For short, this sum is also
denoted by

n=1
z
n
.
If s
n
does not converge, we say that the sum z
1
+z
2
+ diverges. This fact
is often symbolized by

n=1
z
n
= .
14 2. chapter. Complex sequences and series
Theorem 2.1.2. makes it clear that the sum or divergence of a complex
sequence can be determined by the behaviour of the real and imaginary part
of s
n
. That is, let

n=1
z
n
be a convergent series with sum z. Then

n=1
1(z
n
) = 1(z) and

n=1
(z
n
) = (z).
Closing this section we plot the partial sums of the innite series

n=0
_
1
k + 1 + i

1
k + i
_
.
It can be seen, that the values of the partial sums has real part close to
0, and imaginary part close to 1. (We have plotten up to n = 50.) So the
series seemingly converges ti i. One of the exercises is to prove this simple
fact.
Exercises
2.2.2. Exercise. Find the sum of the following complex series:

n=1
_
1
n
+
i
n
2
_
,

n=1
_
1
5
n
+
4i
3
n
_
,

n=1
1 + i
n
6
n
.
2.2. Complex series 15
2.2.3. Exercise. Show that

n=0
_
1
k + 1 + i

1
k + i
_
= i.
2.2.4. Exercise. Let

n=1
z
n
= z.
What is the sum of

n=1
z
n
?
2.2.5. Exercise. For which z C will be the series

n=1
1
n
2
+ z
2
and

n=1
arg(z + n)
1 + in
convergent?
2.2.1 Complex power series
An innite power series connected to a real sequence
n
(n = 0, 1, 2, . . . ) is
dened for a domain of the variable x for which the series around
f(x) =

n=0

n
(x )
n
is convergent.
From real analysis we know the next theorem.
2.2.6. Theorem. Let f be dened by a power series as above. Then the
domain of f is always is one of the following types.
1. ,
2. x R [ [x [ < for some > 0 union a (possibly empty) subset
of the boundary.
3. The whole set R.
2.2.7. Denition. The number in the theorem is called radius of conver-
gence of the power series.
Now we introduce a notation which we will frequently use as an abbrevi-
ation.
16 2. chapter. Complex sequences and series
2.2.8. Denition. The open circle on the complex plane with radius and
with center will be denoted by B(, ), that is,
B(, ) = z C [ [z [ < .
A real power series obviously generalizes to complex ones.
2.2.9. Denition. Let
n
be a complex series, and C. Then the series

n=0

n
(z )
n
is a function f of z. The series is called complex power series and f(z) is
called the sum of the series in a given point z for which the above series
convergent.
The theorem above obviously generalizes to the complex case, mutatis
mutandis: we have to change the absolute value to the complex absolute
value, and the set R to C.
The radius of convergence can be calculated as
=
1
limsup
n
[
n
[
1
n
.
If the rst case is valid in the theorem, then = 0, while if the third, = .
2.2.10. Example. The following power series have radius of convergence
= 0, = 1 and = , respectively:

n=0
n!z
n
,

n=1
z
n
n
,

n=0
z
n
n!
.
In the rst case
=
1
limsup n!
1
n
= 0,
in the second case
=
1
limsup(1/n)
1
n
=
1
limsup 1/
n

n
= 1,
while in the third case
=
1
limsup(1/n!)
1
n
=
1
limsup 1/
n

n!
= .
Now we present a very useful theorem which helps us to derive new func-
tions from old ones.
2.2. Complex series 17
2.2.11. Theorem. Let > 0 be the radius of convergence of the power series
f(z) =

n=0

n
(z )
n
.
Then the function f is dierentiable innitely many times on the set B(, ),
and
f

(z) =

n=0
n
n
(z )
n1
,
f

(z) =

n=0
n(n 1)
n
(z )
n2
,
.
.
.
f
(k)
(z) =

n=0
n(n 1) (n k + 1)
n
(z )
nk
.
Here z B(, ) and k = 1, 2, 3 . . . .
2.2.12. Example. Let
f(z) =

n=0
z
n
.
It is easy to nd a closed form for this function:
f(z) zf(z) = 1 + z + z
2
+ z
3
+ z z
2
z
3
,
so
f(z) =
1
1 z
(z B(0, 1)).
Therefore, for example,
f
_
1
2
_
= 1 +
1
2
+
1
4
+ =
1
1
1
2
= 2.
Or, considering a complex number,
f
_
1 + i
2
_
= 1 +
1 + i
2
+
_
1 + i
2
_
2
+ =
1
1
1+i
2
= 1 + i.
(Note that
1+i
2
B(0, 1).)
2.2.13. Example. To see how we use the above theorem, we deduce the
formula

n=0
nz
n
=
z
(1 z)
2
.
18 2. chapter. Complex sequences and series
First, we start from the above proved result
f(z) =

n=0
z
n
=
1
1 z
(z B(0, 1)).
Then we derive f:
f

(z) =

n=0
nz
n1
=
1
(1 z)
2
(z B(0, 1)).
Finally we multiply by z:
zf

(z) =

n=0
nz
n
=
z
(1 z)
2
(z B(0, 1)).
2.2.2 The Polylogarithm functions
The Polylogarithm functions are important, often appearing special func-
tions. Hence we give a short introduction of them.
2.2.14. Denition. Let k be a xed integer. The kth Polylogarithm func-
tion is dened by
Li(z) =

n=1
z
n
n
k
.
By the above considerations we have the following special polylogarithms:
Li
1
(z) =

n=1
nz
n
=
z
(1 z)
2
,
Li
0
(z) =

n=1
z
n
=
z
1 z
,
Li
1
(z) =

n=1
z
n
n
= log(1 z).
The Polylogarithm functions of higher positive order does not have a simple
closed form. The most what we can do is to nd some special values. For
example, we shall see that
Li
2
(1) =

2
6
.
Note that for any parameter k
Li

k
(z) = z Li
k1
(z)
for all z within the domain of the convergence.
2.2. Complex series 19
Exercises
2.2.15. Exercise. Calculate the value of the sums
Li
1
_
i
2
_
=

n=1
n
_
i
2
_
n
, Li
2
_
1
2
_
=

n=1
n
2
_
1
2
_
n
.
2.2.16. Exercise. Theorem 2.2.11. works reversely. Find the general ex-
pression and radius of convergence for the power series
Li
1
(z) =

n=1
z
n
n
.
Then nd the sum
Li
1
(1) =

n=1
(1)
n
n
.
2.2.17. Exercise. Determine the radius of convergence for the power series
of the polylogarithm function for all integer k.
3
Topological basics
In this short chapter we reconsider the basic topological notions which are
needed in what follows. We suppose that the reader is already more or less
familiar with the foregoing topological facts, therefore we shall not prove the
presented statements. We will not go into the details and we do not try to
be general; the topological notions will be dened via the metric.
3.1 Metric spaces and their properties
3.1.1. Denition. A metric space is a pair (X, d), where X ,= and d :
X X [0, +[ is a function with the following properties
1. d(x, y) = d(y, x)
2. d(x, y) = 0 if and only if x = y,
3. d(x, z) d(x, y) + d(y, z)
for any x, y, z X. The function d is called metric.
Often if we are not specically interested in the form of the metric d, in
place of (X, d) we simply write X.
We know that if X = R and d(x, y) = [x y[ then (X, d) is a metric
space. It is also true that X = C and d(x, y) = [x y[ then (X, d) is a
metric space. This metric is the standard one for the purpose of complex
analyisis, therefore in the following we shall deal with this metric space in
detail. However, before we introduce some important notions on general
metric spaces.
3.1.2. Denition. Let r > 0 is a xed real number. We dene the sets
B(x, r) = y X [ d(x, y) < r,
B(x, r) = y X [ d(x, y) r.
These sets are called the open and closed balls with center x and radius r,
respectively.
20
3.1. Metric spaces and their properties 21
3.1.3. Denition. A point x A X in a subset of a metric space is an
inner point of A if it is contained in an open ball which is entirely in A. In
other words, the point of A is an inner point of A if there is an r > 0 such
that
B(x, r) A.
3.1.4. Denition. For a metric space (X, d) a subset A X is open if all
of its points are inner points. If the complement of A is open, we say that A
is closed.
3.1.5. Proposition. Let (X, d) be a metric space. Then the following propo-
sitions are true:
1. The sets X and are open and close.
2. Any nite intersection of open sets are open.
3. Any union (nite or innite) of open sets are open.
4. Any nite union of closed sets are closed.
5. Any intersection (nite or innite) of closed sets are closed.
3.1.6. Exercise. Which of the following sets are open in C?
The disk
A = z C [ [z[ < 1
with center 0 and radius 1,
the real line,
B = z C [ z
4
= 1.
C = z C [ z
n
= 1 for some n N.
3.1.7. Denition. The smallest closed set which contains A is called the
closure of A, and it is denoted by A. Moreover, the set of inner points of A
is called the interior of A and denoted by Int A. The boundary of a set A in
a metric space is denoted by A and is dened by
A = A Int A.
3.1.8. Proposition. Let A be a subset of a metric space X. Then:
1. A is open if and only if A = Int A.
2. A is closed if and only if A = A.
22 3. chapter. Topological basics
3. Int A = X X A.
4. A = X Int(X A).
5. A B = A B.
3.1.9. Exercise. Find the closure, the interior and the boundary of the sets
in the above exercise.
3.1.10. Denition. A subset A X of a metric space X is dense if its
closure equals to X, that is,
A = X.
It is obvious that X = X.
3.1.11. Exercise. Give examples for dense sets in C.
3.1.12. Denition. An open ball B(a, r) is a neighbourhood of a point x in
a metric space if x B(a, r).
3.1.13. Denition. A point x of A X is a limit point if for any r > 0
there is a neighbourhood of x which contains an element of A.
For example, all the points of C are limit points of C.
3.1.14. Proposition. A point x is a limit point of A X if and only if
there is a non-constant sequence x
n
in A which tends to x.
Note that the limit point does not necessarrily belong to the set A.
3.1.15. Exercise. Let (C, [ [) be the already considered metric space. Find
the limit point(s) of the sets
Q,
QQ,
QR,
[0, 1[ i,

__
1
n
+ i
1
n
2
_
C

n = 1, 2, 3, . . .
_
.
Now we deal with connectedness of sets in metric spaces.
3.1.16. Denition. Let A X be a subset of the metric space X. Then A
is connected if it is not possible decompose A in the form A = B C where
B and C are disjoint, nonempty and open. If A is not connected, we say that
disconnected.
3.2. Continuous functions on metric spaces 23
3.1.17. Example. The whole set C is connected in the metric space (C, [[).
The open and closed balls with any center and any radius in C are connected.
In special, any singleton set is connected.
3.1.18. Denition. If a set A is disconnected, it is possible represent it as
a union of maximal connected subsets. In this case these sets are called
components of A.
3.1.19. Example. The set A = [5, 2] [3, 4] i, i + 2 is disconnected.
Its components are
[5, 2], [3, 4], i, i + 2.
Finally, we introduce a notion which turned to be very important in
analysis.
3.1.20. Denition. Let A be a set in a metric space X, and let be an
index set. A collection of sets B

is a covering system for A if


A
_

.
If all of the B

sets are open, we say that the covering system is an open one.
A is compact if it has a nite covering system (i.e., is nite).
3.1.21. Denition. A subset A of the metric space X is bounded if there is
a radius r > 0 such that
A B(x, r)
for some x X.
Here is a very useful characterization of compactness in the complex plane.
3.1.22. Theorem. A subset A of C is compact if and only if it is bounded
and closed.
3.2 Continuous functions on metric spaces
3.2.1. Denition. Let X and Y be two metric spaces. The function f :
X Y is continuous in a point x X if for any sequence x
n
such that in
the case lim
n
x
n
x in X we have that lim
n
f(x
n
) f(x). If f is
continuous in any x X we say that f is continuous.
Now we list some basic and important equivalents of continuity.
3.2.2. Theorem. Let X and Y be two metric spaces, and let f : X Y be
a function. Then the followings are equivalent.
24 3. chapter. Topological basics
1. f is continuous,
2. if A Y is open then f
1
(A) is open,
3. if A Y is closed then f
1
(A) is closed.
It is also easy to see if f and g are two continuous functions from X into
C then the functions
f + g (, C)
are also continuous. Moreover, f/g is continuous provided g(x) ,= 0 for every
x X.
The next theorem connects compactness and connectedness to continuity.
3.2.3. Theorem. Let f : X Y be a continuous function between the
metric spaces X and Y . Then
1. If X is compact then f(X) is a compact subset of Y .
2. If X is connected then f(X) is a connected subset of Y .
Exercises
3.2.4. Exercise. Let us consider the metric space (Z, [ [). Characterize the
null sequences in this space. (A null sequence is a sequence with limit 0.)
3.2.5. Exercise. Let us consider the metric space (Z, [ [). Decide whether
the subset
0, 2, 4, . . .
is closed or open.
3.2.6. Exercise. Let us consider the metric space (Z, [ [). Look for the
dense sets in this space.
3.2.7. Exercise. We dene the set K R as
K =
_
2n
2m + 1

n, m Z
_
.
What is the closure of this set (with respect to the usual topology of R)?
3.2.8. Exercise. We dene the set K C as
K =
_
2n
2m + 1
+ i
2k + 1
2l

n, m, k, l Z
_
.
What is the closure of this set (with respect to the usual topology of C)?
3.2.9. Exercise. Give example for a subset in a metric space X which is
closed and bounded, but not compact.
Part II
Complex Dierentiability and the
Integral Formula of Cauchy
25
26
From now on we deal with the real topics of complex analysis. First we
turn to the question of dierentiability of complex functions. We shall see
that the complex dierential behaves a bit dierently than the real dieren-
tiability, because the complex dierentiability gives a rather strict restriction
for the function. Then we turn to the complex integral. We shall present
very powerful results which help to calculate integrals.
The well known real functions, like the exponential function, the sin, cos
and their derivatives will be extended to complex variable in this part.
4
Dierentiability
The denition of dierentiability does not dier in the complex case.
4.0.10. Denition. Let E be a nonempty set with a nonepty interior. Let
f : E C and z Int A. We say that f is dierentiable in z if the limit
lim
h0
f(z + h) f(z)
h
exists. If this limit exists, we denote it by f

(z) and we say that this is the


derivative of f in the point z.
We shall see the surprising fact that dierentiability in a class of functions
(the analytic functions) implies the innite dierentiability.
It is convenient to suppose that the set E is itself open (instead of suppos-
ing that has a nonempty interior) and it is connected (in contrary, we should
always handle the dierent components of E). Therefore we introduce the
next denition.
4.0.11. Denition. If a set E C is nonempty, open and connected, we
call it as domain.
4.0.12. Denition. Let D be a domain in C. The function f : D C is
analytic on D if it is dierentiable in all the points of D. The set of analytic
functions on this domain D will be denoted by /(D).
The denition of dierentiability immediately yields the following propo-
sition.
4.0.13. Proposition. Let f, g /(D) be two complex functions and C.
Then f, f + g, fg /(D). Moreover, (f)

= f

, (f + g)

= f

+ g

and
(fg)

= f

g + g

f.
27
28 4. chapter. Dierentiability
4.1 The Cauchy-Riemann equations
It is very interesting to check that what causes on the real and imaginary
part of a function if we suppose the dierentiability. Let f /(D) and let
us write f as
f(z) = u(z) + iv(z), where u = 1(f), v = (f).
We have the following very important theorem. This theorem in many cases
helps us to check the dierentiability.
4.1.1. Theorem. Let f /(D) with the above decomposition, and let z
D. The function f is dierentiable in z if and only if u and v are dieren-
tiables and they satisfy the partial dierential equations
u
x
(z) =
v
y
(z),
u
y
(z) =
v
x
(z).
These equations are called Cauchy-Riemann equations.
In other words, f is analytic on D if u and v are dierentiables and on
this domain they satisfy the Cauchy-Riemann equations.
Proof. We prove the only if part. Let z = x + iy D and
u(x, y) = 1f(x + iy), and v(x, y) = f(x + iy).
Then
f

(z) = lim
h0
f(z + h) f(z)
h
.
The fraction becomes
f(z + h) f(z)
h
=
f(x + iy + h) f(x + iy)
h
=
u(x + h, y) u(x, y)
h
+ i
v(x + h, y) v(x, y)
h
.
If h 0, the fractions tend to the respective partial derivatives:
f

(z) =
u
x
(x, y) + i
v
x
(x, y). (4.1)
Calculating f

(z) in an other way, we take the fraction


f(z + ih) f(z)
ih
= i
u(x, y + h) u(x, y)
h
+
v(x, y + h) v(x, y)
h
.
If h tends to 0, the left hand side tends to f

(z), on the right hand side the


following partial derivatives appear:
f

(z) = i
u
y
(x, y) +
v
y
(x, y).
Comparing this with (4.1), we are ready.
4.1. The Cauchy-Riemann equations 29
4.1.2. Example. The absolute value function is nowhere dierentiable. To
show thsi, let f(z) = [z[ =
_
x
2
+ y
2
. Then
u(x, y) =
_
x
2
+ y
2
, and v(x, y) = 0
Then it is straightforward to see that the partial derivatives of u are not zero
(where it is exists) but the partial derivatives of v are identically zero, so the
Cauchy-Riemann equations does not satisfy.
4.1.3. Example. The arg function is nowhere dierentiable. The reason is
the same as above.
4.1.4. Example. The function f(z) = z
2
is dierentiable on the whole set
C and
f

(z) = 2z.
If f(z) = z
2
then
u(x, y) = 1f(x + iy) = 1(x + iy)
2
= x
2
y
2
,
and
v(x, y) = f(x + iy) = (x + iy)
2
= 2xy.
Then (4.1) gives that
f

(z) =
u
x
(x, y) + i
v
x
(x, y)
= 2x + i2y = 2(x + iy) = 2z.
4.1.5. Remark. It can be derived in the same manner as the above example
that if f(z) = z
n
then
f

(z) = nz
n1
,
or, more generally,
(a
n
z
n
+a
n1
z
n1
+ +a
1
z +a
0
)

= na
n
z
n1
+ (n 1)a
n1
z
n2
+ +a
1
.
To close this section, we also show that the function f(z) = z is nowhere
dierentiable. Let
u(x, y) = 1x + iy = 1(x iy) = x,
and
v(x, y) = x + iy = (x iy) = y.
30 4. chapter. Dierentiability
Then the Cauchy-Riemann equations says that
u
x
(z) =
v
y
(z),
u
y
(z) =
v
x
(z).
Substituting the actual u and v, we get that
1 = 1, and 0 = 0.
The rst equation is a contradiction, so the function f(z) = z does not
dierentiable, as we stated.
5
The exponential and logarithm
function and their derivatives
We shall extend the well known exponential function to complex variable.
The real exponential function is dened via a power series:
e
x
= exp(x) =

n=0
x
n
n!
.
This series has convergence radius = . Without any problem, we can
extend the denition to complex variable, because in place of a real x, we
can put a complex number z: the nth power is unique.
5.0.6. Denition. The complex exponential function exp : C C is dened
by the power series
e
z
= exp(z) =

n=0
z
n
n!
.
This denition is valid for all z C. What gives the special importance
of this function? That its derivative is itself:
exp(z)

=
_

n=0
z
n
n!
_

n=1
nz
n1
n!
=

n=1
z
n1
(n 1)!
=

n=0
z
n
n!
.
Now we deal with the basic properties of this function. The real expo-
nential function has the familiar graph
31
32 5. chapter. The exponential and logarithm function and their derivatives
To get a picture on the complex exponential, we plot its real and imagi-
nary part when z = x + iy and x, y [10, 10]:
33
On the same intervals we plot the absolute value [ exp(z)[ to see how this
function is growing:
34 5. chapter. The exponential and logarithm function and their derivatives
The constant
exp(1) = e =

n=0
1
n!
= 2.718281828 . . .
has fundamental importance in mathematics. This number is called Euler
1
constant hence the letter e - or Napier
2
constant.
That the well known identity exp(x + y) = exp(x) exp(y) holds for com-
plex numbers as well, we can check easily with the Cauchy product and the
binomial theorem.
5.0.7. Theorem. The complex exponential function exp : C C satisfy the
identity
exp(z + w) = exp(z) exp(w).
Proof.
exp(z) exp(w) =
_

n=0
z
n!
__

n=0
w
n!
_
=

n=0
n

k=0
z
k
k!
w
nk
(n k)!
=

n=0
1
n!
n

k=0
n!
z
k
k!
w
nk
(n k)!
=

n=0
1
n!
n

k=0
_
n
k
_
z
k
w
nk
=

n=0
(z + w)
n
n!
= exp(z + w).
5.0.8. Corollary. The exp function does not have zero, that is, there is no
z C for which exp(z) = 0.
Proof. Let us suppose that exp(z) = 0 for a z. Then
exp(z) exp(z) = exp(0) = 1.
This is impossible, if exp(z) = 0.
Now let us check the real and imaginary parts of the exponential function.
5.0.9. Theorem. Let z = x + iy be a complex number. The real and imag-
inary parts of exp are
1(exp(z)) = exp(x)

n=0
(1)
n
(2n)!
y
2n
,
(exp(z)) = exp(x)

n=0
(1)
n
(2n + 1)!
y
2n+1
.
1
Leonhard Euler (1707-1783) swiss mathematician.
2
John Napier () scottish mathematician, physicist, astronomer. He introduced the
logarithm function to simplify calculations.
5.1. The sine and cosine functions 35
Proof. Using the above proved identity for the exp function,
exp(z) = exp(x) exp(iy) = exp(x)

n=0
(iy)
n
n!
.
The number exp(x) is real, moreover, (iy)
n
= i
n
y
n
is real if n is even, and
purely imaginary if n is odd. Hence
exp(z) = exp(x)
__

n=0
(1)
n
(2n)!
y
2n
_
+ i
_

n=0
(1)
n
(2n + 1)!
y
2n+1
__
.
Now the result follows.
5.1 The sine and cosine functions
Now we dene the complex sine and cosine functions. They naturally come
from the real and imaginary parts of the complex exponential function.
5.1.1. Denition. The functions
1(exp(z))/ exp(x) and (exp(z))/ exp(x)
are called cosine and sine function, respectively. They sign are cos and sin.
By the above theorem we have that cos, sin : C C, and
cos(z) =

n=0
(1)
n
(2n)!
y
2n
,
sin(z) =

n=0
(1)
n
(2n + 1)!
y
2n+1
.
These functions are the complex extensions of the real cosine and sine
functions. From the denition and by Theorem 2.2.11. it is obvious that
sin, cos /(C).
Here comes the graph of the real and imaginary part of the complex sine
function on the square [2, 2] [2i, 2i]:
36 5. chapter. The exponential and logarithm function and their derivatives
With these functions we have the following fundamental identity.
5.1. The sine and cosine functions 37
5.1.2. Corollary. Let z = x + iy. Then
exp(z) = e
x
(cos(y) + i sin(y)).
In special, if z = i, we have that
exp(i) = 1.
The other two familiar functions, the tangent and cotangent now will be
dened.
5.1.3. Denition. The function sin(z)/ cos(z) is called complex tangent func-
tion and it is denoted by tan(z). Its reciprocal is the complex cotangent
function, ctg(z).
However the real exponential function is bijective, the complex one is very
far from satisfying this property. To get a deeper picture on exp, we deal
with its mapping properties.
Let us x a set of complex numbers:
A := x + iy C [ y R.
(The real part x is xed.) The image of A by exp is
exp(A) = e
x
(cos(y) + i sin(y)) [ y R.
This set is nothing else but a circle with center 0 and radius r = e
x
. The
original set A is a vertical line.
If x is running and y is xed, that is, we have a horizontal line, a similar
argument gives that the image of this line is a half line which starts from the
origin (this point is not the part of the half line). The angle of this line is y.
Therefore we have got the next proposition.
5.1.4. Proposition. The complex exponential function maps vertical lines
of the complex plane with abscissa x to circles with center 0 and radius e
x
.
The images of horizontal lines with ordinate y are half line starting from the
origo with angle y. The origo is not a part of the image.
We have the next corollary.
5.1.5. Corollary. Th ecomplex function exp is dened for all complex z and
its codomain is the whole set C except 0:
exp : C C 0.
38 5. chapter. The exponential and logarithm function and their derivatives
Exercises
5.1.6. Exercise. Show that for the complex sine and cosine functions
sin(z) =
1
2i
(exp(iz) exp(iz)),
cos(z) =
1
2
(exp(iz) + exp(iz)).
For all z C.
Let us dene the hyperbolic sine and hyperbolic cosine functions as
sinh(z) =
1
2
(exp(z) exp(z)),
cosh(z) =
1
2
(exp(z) + exp(z)).
5.1.7. Exercise. Deduce the closed form value
cos(i) =
1
2
(e

+ e

) = cosh() = 11.59195328 . . . .
5.1.8. Exercise. Show that sin

= cos, cos

= sin, sinh

= cosh and
cosh

= sinh.
5.1.9. Exercise. Derive the addition theorems for the complex sine and
cosine:
sin(z + w) = sin(z) cos(w) + cos(z) sin(w),
cos(z + w) = cos(z) cos(w) sin(z) sin(w).
for all z, w C.
5.2 The logarithm function
In real analysis the logarithm function is the inverse of the exponential. This
inverse exists, since the real exponential function is strictly increasing. But,
the complex exponential is not bijective as we had seen, hence the complex
logarithm as a function cannot be dened. What we can do? Let us try to
solve the equation
exp(z) = w. (5.1)
Let us suppose that z = x + iy, Then via Corollary 5.1.2. [w[ = exp(x)
and y = arg(w) + 2k for som k Z. Hence we have that the solution of
(5.1) is not a denite number but a set:
log [w[ + i(arg(w) + 2k) [ k Z.
Now we can dene the logarithm functions:
5.2. The logarithm function 39
5.2.1. Denition. Let
log
k
(w) = log [w[ + i(arg(w) + 2k), (5.2)
where k Z.
Note that since exp(z) is never zero, we cannot dene log function as
an inverse of exp at zero.
In addition, we have innitely many logarithm functions or if we want
to use the notion of branch the log function have innitely many branches.
From now on we consider the branch belonging to k = 0 as the principal
branch and we denote this function with log
5.2.2. Example. Let us calculate log(1 + i) and log(i) and log(1).
Since [1 + i[ =

2, and arg(1 + i) = /4, we have that


log(1 + i) = log(

2) + i

4
=
1
2
log(2) + i

4
.
Next, log(i) = i

2
, because [i[ = 1 and arg(i) = /2.
Finally, log(1) = i, because [ 1[ = 1 and arg(1) = .
All of these values are on the principal branch.
We can see that the derivative of log has the same form as in the real
analysis:
5.2.3. Proposition. We have that
log

(z) =
1
z
for any branch.
Proof. Since exp(log(z)) = z, the well known rule for the derivative of the
inverse function
log

(z) =
1
exp

(log(z))
=
1
exp(log(z))
=
1
z
.
This has an important corollary:
5.2.4. Theorem. The log function (with the principal branch) has the power
series expansion
log(z) =

n=1
(1)
n+1
n
(z 1)
n
.
40 5. chapter. The exponential and logarithm function and their derivatives
Proof. Let us dene the function g(z) as
log(z)

n=1
(1)
n+1
n
(z 1)
n
.
The series on the right has convergence radius 1, so this function is well
dened on the open ball B(1, 1). Its derivative is
g(z) =
1
z

n=1
(1)
n+1
(z 1)
n1
.
Here we used the above theorem and Theorem 2.2.11. The series on the right
is a geometric series with sum
1
1+(z1)
, whence g(z) = 0. This gives the
statement.
5.2.5. Corollary. We have the famous expansion for the constant log(2):
log(2) =

n=1
(1)
n+1
n
= 0.6931471806 . . . .
This series is called alternating harmonic series.
In general, the series
log(1 + z) =

n=1
(1)
n+1
n
z
n
is the Mercator
3
series.
We close this section with pointing out that the well known identity
log(ab) = log(a) + log(b)
is not valid in general for the complex logarithm (on the principal branch).
5.2.6. Theorem. If z and w are two complex numbers such that 1(z) > 0,
1(w) > 0 and 1(zw) > 0, then
log(zw) = log(z) log(w).
This relation is not valid if the above assumptions do not hold.
3
Nicholas Mercator (1620-1687) netherlandic mathematician and inventor. He pub-
lished rstly this series for the logarithm in 1668, but Isaac Newton and Gregory Saint-
Vincent are also revealed this formula independently.
5.3. General powers 41
Proof. We begin with the denition of the log function:
log(z) = log [z[ + i arg(z).
Then
log(zw) = log [zw[ + i arg(zw) = log([z[[w[) + i(arg(z) + arg(w)).
This latter equality comes from Proposition 1.2.4. We continue:
log([z[[w[) + i(arg(z) + arg(w)) = log [z[ + log [w[ + i(arg(z) + arg(w)) =
log(z) + log(w).
This relation holds, but if the assumptions are not valid the argument steps
over 2, and we arrive at an other branch.
5.3 General powers
It is known, how we dene the power z
n
, when z is an arbitrary complex
number and n is a nonnegative integer. This denition can be extended to
negative integers: z
n
= 1/z
n
. Fractions also can be involved via formula
(1.4): z
1/n
=
n

z. But how can we dene the power for arbitrary exponent?


What is the value, for example, i
i
? In this section we deal with this question.
In real analysis we dened powers as e
b log(a)
, since
e
b log(a)
=
_
e
log(a)
_
b
= a
b
.
We can use this denition in the complex case as well.
5.3.1. Denition. Let z and w be two complex number. The power of z
with exponent w is dened as
(z
w
)
k
= exp(z log
k
(w)).
Here log
k
(z) is the logarithm function dened in (5.2). If k = 0, i.e. we use
the principal branch, then we omit the zero and write
z
w
= exp(z log(w)).
5.3.2. Example. We calculate i
i
. Already we know that log(i) = i

2
, so we
have that
i
i
= exp
_
i
2

2
_
= exp
_

2
_
= 0.2078795764 . . . .
42 5. chapter. The exponential and logarithm function and their derivatives
Exercises
5.3.3. Exercise. Determine the value log(2) and log(2 + i).
5.3.4. Exercise. Calculate i
i
on the branch k = 1.
5.3.5. Exercise. What is the value of the power (1)
i
?
5.3.6. Exercise. Describe the set z C [ e
z
= i.
5.3.7. Exercise. We know that the real sin function have codomain [1, 1].
However, the complex sin takes the value, for example, 2. Therefore solve
the equation
sin(z) = 2.
6
Path integrals
If we would like to calculate the area of a shape under a function with more
than one variable, we use path integrals. Introducing the complex path
integral which is a central notion in complex function theory we shall
extensively use the notions introduced in real analysis. This also means that
we will not prove results which are straight generalizations of the real analitic
case (like additivity of the path integral, parametrization invariance, etc.)
6.1 Curves in the complex plane
6.1.1. Denition. A function of the form : [a, b] C is called curve on
the complex plane or simply complex curve or complex path. (The domain
interval can be opened, we can exclude the point a or b as well.)
6.1.2. Example. The curve
1
: [0, 1] C,

1
(t) = (2 + i)t + 3(1 t)
is a straight line connecting the points 3 and 2 + i on the plane C.
The curve
2
: [0, 2[C,

2
(t) = 1 + i + 3 exp(it)
is a circle with radius 1 + i and radius 3.
6.1.3. Denition. The complex curve : [a, b] C is called smooth if as
a function can be dierentiated continuously innitely many times.
If is not smooth, but it is smooth except a nitely many point, we say
that it is piecewise smooth. The restrictions of the curve to the intervals on
which it is smooth called the smooth components of
6.1.4. Example. The curves in the above example are smooth. But the
curve : [0, 2] C,
(t) =
_
t(1 + i) t [0, 1],
(2 t)(1 + i) + (t 1)2 t ]1, 2]
43
44 6. chapter. Path integrals
is not a smooth one. It is piecewise smooth, and it has two smooth compo-
nents.
Now we dene a very important class of curves. The real variant should
be familiar to the reader, therefore we just declare our statements, we will
not prove them.
Let : [a, b] C is a curve on the complex plane. Let P = t
0
, t
1
, . . . , t
n

be a partition of the interval [a, b], which means that


a = t
0
< t
1
< < t
n1
< t
n
= b.
Let T[a, b] be the set of all the partitions of [a, b]. We introduce the notation
T(, P) =
n

j=1
[(t
j
) (t
j1
)[.
6.1.5. Denition. The total variation of is the value
V () = supT(, P) [ P T[a, b]
(if it exists).
If the total variation exists for , it is said to be of bounded variation or
rectiable.
The set of the curves with bounded variation is denoted by BV [a, b]:
BV [a, b] = : [a, b] C [ V (f) < .
The total variation is nothing else but the arc length of the curve. From
the denition directly is very hard to calculate the total variation. Fortu-
nately there is a very useful theorem, which helps.
6.1.6. Theorem. Let : [a, b] C be a complex curve such that
BV [a, b]. Then
V () =
_
b
a
[

(t)[dt.
6.1.7. Example. Let : [0, 2[C be the curve which represents the unit
circle, that is, (t) = exp(it). Then
V () = 2,
as we expect. To prove this, we just calculate the integral
_
2
0
[ exp

(it)[dt =
_
2
0
[i exp(it)[dt =
_
1
0
1dt = 2i.
6.2. The path integral 45
Now we calculate the arc length of the curve in Example 6.1.4. : [0, 2]
C,
(t) =
_
t(1 + i) t [0, 1],
(2 t)(1 + i) + (t 1)2 t ]1, 2]
We have that
V () =
_
2
0
[

(t)[dt =
_
1
0
[1 + i[dt +
_
2
1
[(1)(1 + i) + 2[dt =
_
1
0

2dt +
_
2
1
[1 i[dt =

2 +
_
2
1

2 = 2

2.
This result is also obvious if we sketch the graph of our function.
The orientation of a curve is a very picturesque notation.
6.1.8. Denition. If : [a, b] C is a curve, then its reverse orientation
is the curve
(t) = (a + b t).
It is obvious that and have the same arc length.
Two more notations will come.
6.1.9. Denition. A curve : [a, b] C is closed if
(a) = (b).
It is simple if it is injective.
6.1.10. Denition. The curve
(t) = exp(it) (t [0, 4])
is closed but not simple.
6.2 The path integral
In this section we shall extensively use the notion of Riemann-Stieltjes inte-
gral, which is considered to be familiar to the reader.
6.2.1. Denition. Let f : A C C be a continuous function and :
[a, b] C be a complex rectiable path such that ([a, b]) A. Then the
path integral of f with respect to the curve is denoted by
_

f(z)dz and
dened by the Riemann-Stieltjes integral
_

f(z)dz =
_
b
a
(f )(t)d(t).
46 6. chapter. Path integrals
By a well known theorem of the theory of Riemann-Stieltjes integrals we
have that the path integral equals to
_

f(z)dz =
_
b
a
(f )(t)

(t)dt.
By utilizing the basic properties of Riemann-Stieltjes integrals, we can
state the next theorem, which collects the basic properties of path integrals.
(We often omit the part (z)dz if we do not use other variables.)
6.2.2. Theorem. Let f : A C C and g : B C C be two contin-
uous functions and : [a, b] C BV [a, b] a complex rectiable path such
that ([a, b]) A, B. Then
1.
_

f =
_

f ( C),
2.
_

(f + g) =
_

f +
_

g,
3.
_

f =
_

f.
Proof. The rst two statements follow from the properties from the Riemann-
Stieltjes integral, so we focus just on the third one.
_

f =
_
b
a
(f ())(t)()

(t)dt =
_
b
a
(f )(a + b t)

(a + b t)dt =

_
b
a
(f )(t)

(t)dt =
_

f.
A useful approximation can be proven easily.
6.2.3. Theorem. We have the next estimation (estimacin) for path inte-
grals

f(z)dz

V () sup
z([a,b])
[f(z)[.
6.2. The path integral 47
Proof. The well known property of the Riemann-Stieltjes integrals give that

f(z)dz

_
b
a
(f )(t)

(t)dt

_
b
a
[(f )(t)[[

(t)[dt sup
z([a,b])
[f(z)[
_
b
a
[

(t)[dt = sup
z([a,b])
[f(z)[V ().
6.2.4. Denition. An analytic function F is a primitive function of f if
F

= f.
Any function which has a primitive function has path integral zero, if the
path is closed. This is a special consecuence of the next theorem.
6.2.5. Theorem. If f : A C has a primitive function F, and : [a, b]
C is a curve such that ([a, b]) A, then
_

f(z)dz = F((b)) F((a)).


Proof. We have
_

f(z)dz =
_
b
a
(f )(t)

(t)dt =
_
b
a
(F

)(t)

(t)dt =
_
b
a
(F )

(t)dt = F((b)) F((a)).


6.2.6. Example. Let f(z) = z, and (t) = exp(2it) is the closed unit
circle, where t [0, 1]. Then
_

zdz =
1
2
exp
2
(2i 1)
1
2
exp
2
(2i 0) =
1
2

1
2
= 0.
Note that this is true not just for f(z) = z but for any polynomial.
Exercises
6.2.7. Exercise. Calculate the next path integrals:
_

1(z)dz, where (t) = t (t [0, 2]),


_

(z)dz, where (t) = it (t [0, 3]),


_

zdz, where (t) = exp(2it) (t [0, 1]).


48 6. chapter. Path integrals
6.2.8. Exercise. Taking care on the branches of the log(z) function, let us
calculate the integral
_
i
0
log(z)dz.
(This is a path integral, where the path is the line segment [0, i].)
7
The theorem of Cauchy
In this chapter we present the most important, central theorem of Cauchy.
This theorem says, roughly speaking, that if a function f is analytic on a
domain and the closed path is contained in this domain, then
_

f(z)dz = 0.
To prove the theorem, we rst prove it for an easier closed path, a triangle.
Then we shall extend this and specify the domain of f.
We follow the idea of the proof what can be found in the book of J.
Duncan (The Elements of Complex Analysis, John Wiley & Sons, 1968).
7.1 The theorem of Cauchy for triangle paths
On the complex plane we dene a triangle T by three non collinear point

1
,
2
,
3
C:
T =
1

1
+
2

2
+
3

3
[
1
,
2
,
3
0;
1
+
2
+
3
= 1.
The path running around this triangle (with the orientation
1
,
2
,
3
) is
denoted by
T
.
7.1.1. Theorem. If T is a triangle in the domain D and f : D C is
analytic, then
_

T
f(z)dz = 0.
The assumption that T is in D implies that D is convex. Later we shall
generalize this theorem to starlike sets and in place of triangles we can have
any simple closed paths.
Proof. Let
A =

T
f(z)dz

.
49
50 7. chapter. The theorem of Cauchy
We decompose the triangle to four smaller ones:

1
=
1
2
(
2
+
3
),
2
=
1
2
(
3
+
1
),
3
=
1
2
(
1
+
2
).
Let the triangles (
1
,
3
,
2
), (
2
,
1
,
3
), (
3
,
2
,
1
) and (
1
,
2
,
3
) be de-
noted by T
1
1
, T
2
1
, T
3
1
, T
4
1
, respectively.
Then we have that
_

T
f(z)dz =
_
[
1
,
3
]
f(z)dz +
_
[
3
,
2
]
f(z)dz +
_
[
2
,
1
]
f(z)dz+
_
[
1
,
3
]
f(z)dz +
_
[
3
,
2
]
f(z)dz +
_
[
2
,
1
]
f(z)dz.
Since
_
[
j
,
k
]
f(z)dz =
_
[
k
,
j
]
f(z)dz
for all indices, we have that
_

T
f(z)dz =
_

T
1
1
f(z)dz +
_

T
2
1
f(z)dz +
_

T
3
1
f(z)dz +
_

T
4
1
f(z)dz.
Now we choose a triangle T
r
1
from T
i
1
(i = 1, 2, 3, 4) such that
A 4

T
r
1
f(z)dz

.
This can be done by the previous integral decomposition. This triangle will
be denoted by T
1
(so we leave the upper index, because in what follows we
does not deal with them).
This triangle T
1
has sidelengths a/2, b/2, c/2. Now consider this triangle
as a whole triangle. We repeat the process, and win a sequence of triangles
T
1
, T
2
, . . . such that
1. T
m+1
T
m
,
2. the sidelengths of T
m
are 2
m
a, 2
m
b, 2
m
c,
3. A 4
m

T
m
f(z)dz

.
If we take the intersection of all of these triangles, we have that

n=1
T
m
= .
That is, it contains only one point. (This is not a trivial statement. There
is a theorem, which says that if we have a decreasing sequence of compact
7.2. Cauchys theorem for starlike domains 51
sets such that the limit of the diameters tends to zero, then the intersection
contains one and only one point.)
This point is in T. Since D is open, there is a > 0 such that B(, )
D. Since f is dierentiable on D, it is dierentiable in the point , so there
is a function : B(, ) C such that
f(z) f()
z
= f

() + (z).
This is continuous and () = 0.
Hence for all > 0 there is a such that
[(z)[ < for all z B(, ).
It is possible choose an m such that T
m
B(, ). Hence, utilizing the
denition of ,
_

T
m
f(z)dz =
_

T
m
(f() + (z )f

(z))dz +
_

T
m
(z )(z)dz.
The rst integrand has a primitive function:
f()z +
1
2
f

()(z )
2
,
whence it follows that the rst integral vanishes by Theorem 6.2.5. Using the
inequality for A, and the estimation of Theorem 6.2.3. we have that
0 A 4
m

T
m
f(z)dz

= 4
m

T
m
(z )(z)dz

4
m
V (
T
m
) sup
z
T
m
[z [[(z)[
4
m
2
m
(a + b + c)2
m
(a + b + c) = (a + b + c)
2
.
The last line comes, since the two farest points in the triangle T
m
is not
greater than 2
m
(a+b +c). Since was arbitrary, we have that A = 0, what
gives the statement.
7.2 Cauchys theorem for starlike domains
Now we go to present and prove a more general version of the theorem of
Cauchy. We have said that the theorem remains valid for any simple closed
path in a starlike domain. Now we specify what does it mean.
52 7. chapter. The theorem of Cauchy
7.2.1. Denition. A domain D C is starlike if there is a point x
0
in D
such that any point x in D can be connected to x
0
by a line segment.
7.2.2. Example. The ball B(0, 1) is starlike in C but the annulus B(0, 2)
B(0, 1) is not.
Now we prove a theorem, from which the more general version of Cauchys
theorem will immediately follow.
7.2.3. Theorem. If D is a starlike domain on the plane C, then any f
/(D) has a primitive function.
Proof. Let be a point from which every point can be seen (this exists, since
D is starlike). Then we dene a function g as
g(z) =
_
[,z]
f (z D).
Let D . Since D is open, there is a h such that + h D. We
have that
g( + h) =
_
[,+h]
f (z D).
Let us choose r such that
[, + h] B(, r) D.
Since D is starlike, for any point in w [, +h] the line segment [, w] D.
Hence the triangle given by , , +h is entirely in D. Hence Theorem 7.1.1.
gives that
_
[,]
f +
_
[,+h]
f +
_
[+h,]
f = 0.
Hence
g( + h) g() =
_
[,+h]
f
_
[,]
f(t)dt =
_
[,+h]
f.
After an algebraic rearrangement we take absolute value and win that

g( + h) g()
h
f()

1
h
_
[,+h]
f
f()
h
_
[,+h]
1

=
=

1
h
_
[,+h]
(f(z) f())dz

1
h
_
[,+h]
[(f(z) f())[ dz

1
h
h sup
z[,+h]
[f(z) f()[.
If h tends to 0, f(z) tends to f(), because f is analytic, so dierentiable,
so continuous. Hence we get that g is dierentiable and its derivative in
is f(). This means that g is a primitive function of f, so the theorem is
proven.
7.3. Cauchys theorem on simply connected domains 53
Now the generalized version of Cauchys theorem follows.
7.2.4. Theorem (Theorem of Cauchy). Let D be a starlike domain in C
and let f /(D) be an analytic function on D. If the path : [a, b] is simple
and closed such that ([a, b]) D, then
_

f(z)dz = 0.
Proof. The proof now comes from Theorem 6.2.5., because f has a primitive
function.
7.3 Cauchys theorem on simply connected do-
mains
We present a more general version of Cauchys theorem. This says that the
path integral of an analytic function vanishes if the curve is closed; not just
when the domain of the function is starlike, but it can have a more general
shape: simply connected.
Two curves are called homotopic, if , informally, it is possible to deform
one to the other with a continuous transformation. Here comes the mathe-
matically precise denition.
7.3.1. Denition. Let D C be a domain. Two paths
0
,
1
: [a, b] D
are called homotopic if there is a continuous function H : [a, b] [0, 1] D
such that
H(t, 0) =
0
(t), and H(t, 1) =
1
(t).
In this case we write that
0

1
.
A sample of homotopic curves can be seen on the following graphic:
(This picture is come from the WikiMedia Commons; homotopy curves.svg)
Note that the homotopy is an equivalence relation.
54 7. chapter. The theorem of Cauchy
7.3.2. Denition. If the path
0
: [a, b] D is homotopic to a point, it is
called null-homotopic.
If in the domain D C every closed path is null-homotopic, this domain
is called simply connected.
Intuitively, a simple connectivity means that in the domain there are no
holes.
7.3.3. Example. A circle, or more general, an ellipse is always null-homoto-
pic on convex domains. But, for example, the circle with origin 0 and radius 1
is not nullhomotopic on the domain C0. For this reason the domain C0
is not simply connected. The same is true for the annulus B(0, 2) B(0, 1).
The above example leads to the next proposition.
7.3.4. Proposition. Any two closed paths
0
,
1
: [a, b] D are homotopic
if D is convex.
Proof. Let H : [a, b] [0, 1] D is dened by
H(t, s) = s
0
(t) + (1 s)
1
(t).
Since D is convex, all the points in the codomain of H (which are the connect-
ing line segments between two points of the curves) is in D. H is continuous
and it is also holds that
H(t, 0) =
0
(t), and H(t, 1) =
1
(t).
So the two paths are homotopic.
7.3.5. Proposition. Every starshaped domain is simply connected. (Thus
convex domains are simply connected.)
The above two proposition shows that the homotopy is at least as strongly
connected to the domain D as to the curves itself.
The above theorem is very important. It says that for any two homotopic
curves the path integrals of an analytic function are equal.
7.3.6. Theorem (Deformation Theorem). Let D C be a domain and
f /(D). Moreover, let
0
,
1
: [a, b] D are homotopic curves. Then
_

0
f(z)dz =
_

1
f(z)dz.
The corollary is the generalized version of Cauchys theorem:
7.3. Cauchys theorem on simply connected domains 55
7.3.7. Theorem (Cauchys theorem on simply connected domains).
Let D C is a simply connected domain. Moreover, let f /(D) be an
analytic function on this domain, and let : [a, b] D be a closed path.
Then
_

f(z)dz = 0.
Proposition 7.3.5. tells us why this is a generalization of Theorem 7.2.4.
Proof. D is a simply connecte domain, so our curve is homotopic to a point
p D. If we consider this point p as a curve
p
: [a, a] p, it is obvious
that the theorem holds by the Deformation Theorem.
If in the domain where our path runs and our function is dened, there
are holes, the theorem loses its validity. For example, the function f(z) =
1
z
is dened on C0. If we take the circle (t) = exp(2it), the path integral
is
_

1
z
dz =
_
1
0
1
exp(2it)
exp

(2it)dt =
log(exp(2it))[
t=1
t=0
= 2it[
t=1
t=0
= 2i.
And this is not zero. The reason as we said is that there is a hole in the
domain C 0, that is, it is not simply connected.
8
Cauchy Integral formula
In what follows we shall present some important consequences of Cauchys
theorem. The rst application, with which this chapter deals, is the Cauchy
Integral Formula, which helps us to calculate complicated circle integrals
(circle integral is a path integral where the path is a circle). This theorem
says that the values of a function in a ball is determined by the values of the
function on the boundary of this ball.
8.1 A special version of the Cauchy Integral
Formula
To have shorter phrasing of our theorems, we introduce a technical denition.
8.1.1. Denition. The closed curve : [0, 1] C which is dened by
(t) = a + r exp(2it) (t [0, 1]),
describes a circle with centre a and radius r. This curve is shortly will be
denoted by
(a, r).
8.1.2. Theorem (Cauchys Integral Formula). Let f : D C is an
analytic function and let r > 0 such that B(r, a) D for some a D. Then
for any z B(r, a)
1
2i
_
(a,r)
f(w)
w z
dw = f(z).
Proof. Let us x a point z B(a, r). The function
f(w)
w z
56
8.1. A special version of the Cauchy Integral Formula 57
is dierentiable on D z, since f and
1
wz
are. Now we choose an 0 < <
r [z a[ (we have to substract [z a[ to keep us in D). Then, because
(a, r) (z, ), we have by the Deformation Theorem that
_
(a,r)
f(w)
w z
dw =
_
(z,)
f(w)
w z
dw. (8.1)
An algebraic trick gives that
_
(z,)
f(w)
w z
dw =
_
(z,)
f(z)
w z
dw +
_
(z,)
_
f(w) f(z)
w z
f

(z)
_
dw +
_
(z,)
f

(z)dw.
Let us deal with the rst integral.
_
(z,)
f(z)
w z
dw = f(z)
_
(z,)
1
w z
dw =
f(z)
_
1
0
1
z + exp(2it) z

t
(z + exp(2it))dw =
f(z)
_
1
0
1
exp(2it)
2i exp(2it)dw = 2if(z)
_
1
0
1dw = 2if(z).
The function under the second integral is bounded, because f is dierentiable.
Let this bound be M, that is,

f(w) f(z)
w z
f

(z)

< M.
Then the absolute value of the second integral is estimated by

_
(z,)
_
f(w) f(z)
w z
f

(z)
_
dw

M
_
(z,)
1dw = M2. (8.2)
The third integral is zero, because it equals to
_
(z,)
f

(z)dw = f((1)) f((0)) = 0.


If tends to zero, the second integral also vanishes by (8.2).
Altogether, we have that (8.1) becomes
_
(a,r)
f(w)
w z
dw = 2if(z).
58 8. chapter. Cauchy Integral formula
8.2 Example circle integrals
As we have mentioned, this theorem can be very useful in the calculations of
circle integrals. We now take some examples to demonstrate this statement.
8.2.1. Example. Let us evaluate the integral
_
(0,1)
e
w
w
dw.
(We note that the non-complex integral
_
e
x
x
dx
does not have an expression with classical functions.) We can use the function
f(z) = exp(z) (the domain D = C). By the integral formula of Cauchy:
_
(0,1)
e
w
w
dw =
_
(0,1)
f(w)
w 0
dw = 2if(0) = 2i.
8.2.2. Example. Let us determine the value of the integral
_
(0,1)
e
w
+ w
w + 2
dw.
Since the function
e
w
+ w
w + 2
is analytic on the ball B(0, 1) circumscribed by (0, 1), and this ball is a
simply connected domain, Cauchys theorem on simply connected domains
gives that this integral is zero:
_
(0,1)
e
w
+ w
w + 2
dw = 0.
Now let us calculate the integral
_
(0,3)
e
w
+ w
w + 2
dw.
The previously used theorem cannot be applied, since the integrand is not
analytic in the point w = 2, which point is included in the ball B(0, 3).
However, the nominator f(z) = e
z
+ z is analytic on B(0, 3) (in fact, on the
whole C), so Cauchys integral formula can be applied. It gives that
_
(0,3)
e
w
+ w
w + 2
dw = 2i(e
2
2).
8.2. Example circle integrals 59
8.2.3. Example. Let us calculate the integral
_
(i,1)
1
w
2
+ 1
dw.
If we want to apply Cauchys integral formula, we have to have an analytic
function on B(i, 1) in the nominator. To reach this aim, we observe that
1
w
2
+ 1
=
1
(w + i)(w i)
,
so
f(w) =
1
w + i
and z = i. (The other choice
f(w) =
1
w i
and z = i is not applicable, since this f is not analytic on B(i, 1)!) Now
_
(i,1)
1
w
2
+ 1
dw =
_
(i,1)
1/(w + i)
w i
dw = 2i
1
2i
= .
Now we rephrase the Integral Formula of Cauchy in an other way.
8.2.4. Theorem. With the same conditions as in Theorem 8.1.2., we have
that
f(a) =
1
2
_
2
0
f(a + re
it
)dt.
Proof. Let us apply the denition of the path integral in the Cauchy integral
formula in the point z = a:
f(a) =
1
2i
_
(a,r)
f(w)
w a
dw =
1
2i
_
2
0
f(a + re
it
)
a + re
it
a

t
(a + re
it
) =
1
2i
_
2
0
f(a + re
it
)
re
it
ire
it
=
1
2
_
2
0
f(a + re
it
).
As an application, we calculate an interesting integral in the next exam-
ple.
8.2.5. Example. The so-called log-sine integral equals to /2 log(2):
_
/2
0
log(sin(x))dx =

2
log(2).
60 8. chapter. Cauchy Integral formula
To prove this nice identity, we begin with the integral
_
2
0
log(1 + re
it
)dt = 0.
This is the straight consequence of the above theorem (0 < r < 1). Now let
us take the real and imaginary part in the integrand:
0 =
_
2
0
log(1 + re
it
)dt =
_
2
0
_
log [1 + re
it
[ + i arg(1 + re
it
)
_
dt
on the principal branch. Note that we does not pass through the branches,
because t [0, 2]. Let us calculate the real part of the integral. The
argument times i is surely real, so we can omit it:
1(0) = 0 = 1
__
2
0
log [1 + re
it
[dt
_
.
Let us calculate the absolute value:
[1 + re
it
[ = [1 + r cos(t) + ir sin(t)[ =
_
(1 + r cos(t))
2
+ (r sin(t))
2
=
_
1 + 2r cos(t) + r
2
cos
2
(t) + r
2
sin
2
(t) =
_
1 + r
2
+ 2r cos(t).
Now specify r = 1:
_
1 + r
2
+ 2r cos(t) =
_
2 + 2 cos(t) =

2
_
1 + cos(t). (8.3)
We use the trigonometric identity
2 cos
2
(t) = 1 + cos(2t)
to proceed in (8.3):

2
_
1 + cos(t) =

2
_
2 cos
2
(t/2) = 2
_
cos
2
(t/2).
The variable t runs from 0 to 2, so t/2 runs from 0 to . On the interval
[0, /2] cos is nonnegative, on ]/2, ] is nonpositive. Thus we can restrict
us to the interval [0, /2], and use the variable t in place of t/2. On the
other part of the interval cos and thus its square root and its logarithm is
imaginary. Hence
2
_
cos
2
(t) = 2 cos(t) (t [0, /2]).
Now we can proceed with our integral.
1(0) = 0 = 1
__
2
0
log [1 + re
it
[dt
_
=
_
/2
0
log(2 cos(t))dt =
8.3. The general version of Cauchys integral formula 61
_
/2
0
log(2)dt +
_
/2
0
log(cos(t))dt.
A rearrangement gives that
_
/2
0
log(cos(t))dt =

2
log(2).
This is what we wanted to prove.
8.3 The general version of Cauchys integral
formula
The Cauchy Integral Formula says that
1
2i
_
(a,r)
f(w)
w z
dw = f(z).
for an analytic function f on a circle contained in the domain of f. If we
derive both of the sides of the equation, we have that
1
2i
_
(a,r)
f(w)
(w z)
2
dw = f

(z).
Taking the second derivative:
2
2i
_
(a,r)
f(w)
(w z)
3
dw = f

(z).
And, in general,
n!
2i
_
(a,r)
f(w)
(w z)
n+1
dw = f
(n)
(z).
Of course, we have to justify that the derivative can be carried into the
integral sign. This always can be done if on the domain of the integral the
integrand is continuous. Since f is analytic, it is continuous on its domain,
and this domain encloses the circle on which we integrate, so our steps are
correct. We have the next theorem.
8.3.1. Theorem (Cauchy). If f is analytic on a domain D and this do-
main contains the open ball B(a, R), then f is dierentiable innitely many
times on B(a, R) and for any 0 < r < R
f
(n)
=
n!
2i
_
(a,r)
f(w)
(w z)
n+1
dw (z B(a, r)).
62 8. chapter. Cauchy Integral formula
We would like to emphasize the deep meaning of this theorem: this says
that if a function is analytic, that is, dierentiable one time in its domain,
then it is dierentiable innitely many times. This is very far from to be
true in real analysis.
In what follows, we would not like to say always that the domain of D
contains a ball on which our theorems hold, so in place of a general do-
main D we always will specify D to be an open ball. This is just a virtual
modication, since we will immediately see that analytic functions on a do-
main always have power series expansions, and these power series are always
absolutely convergent on a circular domain (see Theorem 2.2.6.).
The theorem of Cauchy implies that an analytic function always have a
power series expansion.
8.3.2. Theorem (Taylor). Let f /(B(a, R)). Then
f(z) = f(a) +

n=1
f
(n)
(a)
n!
(z a)
n
(z B(a, R)).
Proof. We shall use the following modication of the geometric series:
1
w z
=

n=0
(z a)
n
(w a)
n+1
(w B(a, r)), (8.4)
where r is xed such that [z a[ < r < R, moreover, z B(a, R).
By Theorem 8.1.2.,
f(z) =
1
2i
_
(a,r)
f(w)
w z
dw =
1
2i
_
(a,r)
f(w)

n=0
(z a)
n
(w a)
n+1
=

n=0
(z a)
n
1
2i
_
(a,r)
f(w)
(w a)
n+1
=
f(a) +

n=1
f
(n)
(a)
n!
(z a)
n
.
8.3.3. Denition. The coecients
f
(n)
(a)
n!
are called Taylor coecients of f around a. The series is called Taylor series
of f around a.
8.3. The general version of Cauchys integral formula 63
The Taylor coecients are unique.
8.3.4. Example. Example 2.2.13. says that the Taylor coecients of the
function 1/(1 z) are all 1 around a = 0 and R = 1.
The Taylor coecients of the function z/(1z
2
) are the natural numbers
around a = 0 and R = 1.
The Taylor coecients of exp(z) around a = 0 are the reciprocals of
factorials, and R = .
8.3.5. Example. Let us nd the Taylor coecients of the function
1
1 z z
2
around a = 0. Let us denote this function by f(z) and denote the Taylor
coecients by F
n
. Then
f(z) =
1
1 z z
2
=

n=0
F
n
z
n
.
This means that
(1 z z
2
)f(z) = f(z) zf(z) z
2
f(z) = 1.
Using the Taylor coecients,
1 = f(z) zf(z) z
2
f(z) =

n=0
F
n
z
n

n=0
F
n
z
n+1

n=0
F
n
z
n+2
=

n=0
F
n
z
n

n=1
F
n1
z
n

n=2
F
n2
z
n
=
F
0
+ (F
1
F
0
)z +

n=2
(F
n
F
n1
F
n2
)z
n
.
Since this sum is 1, the coecients of z, z
2
,. . . must cancel. This means that
for the coecients F
n
the following recursion holds:
F
n
= F
n1
+ F
n2
.
We need the initial values F
0
and F
1
. It is obvious that F
0
= 1, hence F
1
F
0
gives that F
1
= 1. For F
2
we have that
F
2
= F
1
+ F
0
= 2,
F
3
= F
2
+ F
1
= 3,
64 8. chapter. Cauchy Integral formula
F
4
= F
3
+ F
3
= 5,
and so on. This sequence is called the Fibonacci sequence.
We left to the reader showing that the radius of convergence of
1
1 z z
2
=

n=0
F
n
z
n
is
R =

5 1
2
.
Exercises
8.3.6. Exercise. Find the Taylor expansion of log(1 + z) around the point
a = 0. What is the radius of convergence?
8.3.7. Exercise. Let us consider the function f, which satises the dier-
ential equation
f

(z) = 1 + 2zf(z).
Find the Taylor series of f around 0 and determine the radius of convergence.
8.3.8. Exercise. Prove that for the Fibonacci sequence
F
n
=
1

5
_
_
_
1 +

5
2
_
n+1

_
1

5
2
_
n+1
_
_
(n = 0, 1, 2, . . . ).
Part III
Local and global analysis
65
66
Having the Cauchy theorems of analytic functions we are able to look the
dierent analytic properties of complex functions. We are able to analyze
the zeroes and poles of complex functions. In this part we shall present the
maybe most beautiful results of complex analysis, the Weierstrass Product
Representation Theorem and the Mittag-Leer Expansion Theorem. These
says that if we prescribe the zeros of an analytic function, this function can
be represented as a nite or innite product of simple factors. If, in spite of
the zeros, we deal with the poles, the given function can be expressed as a
nite or innite sum of simple terms.
9
The Laurent series
We have seen that an analytic function f /(B(a, r)) have a Taylor series
expansion. If our function is not analytic in a point (like 1/z in 0), then
around this point there is no Taylor series representation for it. (To see
this let us try to expand 1/z around a = 0.) In place of a Taylor series
representation the Laurent expansion comes.
9.1 The Laurent series. Main part of a function
9.1.1. Theorem (Laurent). Let f be an analytic function on the punctured
disk B(a, R) a. Then
f(z) =

n=
a
n
(z a)
n
for all z in the punctured disk B(a, r) a. The coecients a
n
are
a
n
=
1
2i
_
(a,r)
f(w)
(w a)
n+1
dw (0 < r < R, n Z).
Moreover, there are functions f
1
/(B(a, R)) and f
2
/(C a) such
that
f(z) = f
1
(z) + f
2
(z) (z B(a, R) a).
Proof. Let us choose R
1
and R
2
such that 0 < R
1
< [z a[ < R
2
< R. Let
> 0 be a radius such that B(z, ) B(a, R
2
) B(a, R
1
). Then by Cauchys
theorem
f(z) =
1
2i
_
(z,)
f(w)
w z
dw =
1
2i
_
(a,R
2
)
f(w)
w z
dw
1
2i
_
(a,R
1
)
f(w)
w z
dw.
With the same argument as in Theorem 8.3.2., we can see that
1
2i
_
(a,R
2
)
f(w)
w z
dw =

n=0
(z a)
n
2i
_
(a,R
2
)
f(w)
(w a)
n+1
dw.
67
68 9. chapter. The Laurent series
To calculate the other integral, we use the modication of (8.4):

1
w z
=

n=0
(w a)
n
(z a)
n+1
([w a[ < [z a[).
Now we continue with the second integral:

1
2i
_
(a,R
1
)
f(w)
w z
dw =
1
2i
_
(a,R
1
)
f(w)

n=0
(w a)
n
(z a)
n+1
=

n=0
(z a)
n1
1
2i
_
(a,R
1
)
f(w)
(w a)
n
=

n=1
(z a)
n
1
2i
_
(a,R
1
)
f(w)
(w a)
n+1
.
Altogether,
f(z) =

n=0
(z a)
n
2i
_
(a,R
2
)
f(w)
(w a)
n+1
dw+

n=1
(z a)
n
1
2i
_
(a,R
1
)
f(w)
(w a)
n+1
.
It can be seen that the value of the path integrals is independent from the
choice of R
1
and R
2
iw we stay in the domain B(a, R). Therefore we can
choose a common r ]0, R[. For this reason let us dene the coecients a
n
as
a
n
=
1
2i
_
(a,r)
f(w)
(w a)
n+1
(n Z).
Then we can rewrite the expression for f(z) as
f(z) =

n=0
(z a)
n
a
n
+

n=1
(z a)
n
a
n
=

n=
a
n
(z a)
n
.
The proof of the rst point is done.
To prove the second part, let us separate the sum
f(z) =

n=
a
n
(z a)
n
to two parts:

n=
a
n
(z a)
n
=

n=0
a
n
(z a)
n
+

n=1
a
n
(z a)
n
.
9.1. The Laurent series. Main part of a function 69
The rst part as a function is analytic on B(a, R), because this is the Taylor
part of the function. The second sum is convergent on C a. Thus, if we
dene the functions f
1
and f
2
as
f
1
(z) : B(a, R) C; f
1
(z) =

n=0
a
n
(z a)
n
,
f
1
(z) : C a C; f
2
(z) =

n=1
a
n
(z a)
n
,
then f(z) = f
1
(z) + f
2
(z) on B(a, R), as we stated.
9.1.2. Denition. The Laurent coecients are the coecients
a
n
=
1
2i
_
(a,r)
f(w)
(w a)
n+1
dw (0 < r < R, n Z) (9.1)
in the Laurent expansion of the function f.
The function f
2
in the theorem can be considered as the complement part
of the Taylor series for a function, which is not analytic in a point just in a
punctured disk around this point. This function f
2
will play an important
role later, so we introduce the next denition.
9.1.3. Denition. The function f
2
in the Laurent theorem is called the
main part of the function f.
9.1.4. Example. We said at the begin of this chapter that the function
1/z does not have a Taylor expansion, since it is not analytic on a disc
which contain the origin. But it is analytic on any punctured disk of the
form B(0, R) 0, 0 < R < . Hence it has a Laurent series expansion.
Of course, this Laurent expansion coincide with 1/z, since it is already the
Laurent form of this function. The Laurent coecients are all zero, except
a
1
. Indeed,
a
n
=
1
2i
_
(0,R)
1/w
(w 0)
n+1
dw =
1
2i
_
(0,R)
1
w
n+2
dw =
1
2i
_
2
0
1
(e
Rit
)
n+2
Rie
Rit
dt =
R
2
_
2
0
1
(e
Rit
)
n+1
dt =
R
2
_
2
0
e
Ri(n+1)t
dt =
R
2
_

e
Rit
Ri(n + 1)
_
2
0
=
1
2i(n + 1)
_
e
Rit

2
0
=
1
2i(n + 1)
(1 1) = 0.
70 9. chapter. The Laurent series
Of course, if n = 1 we cannot use this primitive function. Instead, we go
back to the integral:
a
1
=
1
2i
_
(0,R)
1
w
dw.
This integral, by Cauchys theorem equals to 1, since f(z) 1 is analytic on
our domain.
So
1
z
=

n=
a
n
z
n
= 1 z
1
=
1
z
.
9.1.5. Example. To take a more advanced example, let us look for the
Laurent series for the function
f(z) =
1
1 z
2
.
This function is analytic on C 1. Thus, it has a Taylor series in any
disk not containing 1. For example, around a = 0 it has the Taylor series
expansion
f(z) =

n=0
z
2n
.
Now we look the Laurent expansion around a = 1, for example. We can
separate the expression for f(z) to two parts:
f(z) =
1
1 z
2
=
1
2
1
1 + z
+
1
2
1
1 z
.
The rst part has a Taylor expansion (since it is analytic around a = 1):
1
2
1
1 + z
=
1
2
1
2 + (z 1)
=
1
4
1
1
_

z1
2
_ =
=
1
4

n=0
_
1
2
_
n
(z 1)
n
.
Hence we have found a
n
where n 0. Now we nd out that the negatively
indexed a
n
s are all zero, except a
1
. This is so, because the term
1
2
1
1 z
=
1
2
(z 1)
1
is yet a Laurent series around a = 1.
Altogether, we have that
f(z) =
1
1 z
2
=

n=0
1
4
_
1
2
_
n
(z 1)
n
+
_

1
2
_
(z 1)
1
.
9.1. The Laurent series. Main part of a function 71
Hence the Laurent coecients are
a
n
=
1
4
_
1
2
_
n
if n = 0, 1, 2, . . . and
a
n
=
1
2
if n = 1 and
a
n
= 0
for n = 1, 2, . . . .
To see how the Laurent theorem works, let us calculate the Laurent co-
ecients by the formula (9.1) as well. This formula says that
a
n
=
1
2i
_
(1,R)
1
2
1
1w
(w 1)
n+1
dw
for all n Z and for an radius R > 0. This radius cannot be greater than
2, since otherwise we reach the point 1, where the function has an other
non-analyticity point. For the sake of simplicity, let R = 1.
1
2i
_
(1,R)
1
2
1
1w
(w 1)
n+1
dw =
1
4i
_
(1,1)
1
(w 1)
n+2
dw =
1
4i
_
2
0
1
(1 + e
it
1)
n+2
ie
it
dt =
1
4
_
2
0
e
(n+1)it
dt =
1
4
_

e
(n+1)it
i(n + 1)
_
2
0
=
1
4
1
n + 1
(1 1) = 0
if n ,= 1. If n = 1, the integral is 2, so a
1
=
1
2
, as we have seen before.
Using the denition of the principal part, we say that it is
f
2
(z) =
1
2
1
1 z
(z C 1).
Exercises
9.1.6. Exercise. Find the main part of the functions with respect to the
given points.
f(z) =
1
1+z
2
, a = i,
f(z) = exp
_
z +
1
z
_
, a = 0.
72 9. chapter. The Laurent series
9.1.7. Exercise. Look for all the points in which the following functions are
not analytic:
[z[,
1(z) (z),

1
1+z
4
,

1
sin(z)
.
9.1.8. Exercise. Find the Laurent expansion of the function
f(z) =
1
z
2
z 2
in the point a = 2. What is the principal part of this function at this point?
9.1.9. Exercise. Find the Laurent expansion of the function
f(z) = exp
_
1
z
_
in the point a = 0. What is the principal part of this function at this point?
9.1.10. Exercise. Find the Laurent expansion of the function
f(z) = exp
_
z +
1
z
_
in the point a = 0. What is the principal part of this function at this point?
10
Singularities and zeroes of a-
nalytic functions
10.1 Classication of singularities
The zeroes and singularities are important points of a function. In what
follows, we deal with these specic points of functions. This theory has
important applications what we shall present as well.
On singular point we mean that the function is not analytic in this point.
For example, the function f(z) = 1/z is singular in the point zero. If we
go close to zero from the left (z = , where > 0 a small quantity),
the function tends to . If we go close to zero from the right (z = ),
the function diverges to +. If we consider complex values, like z = i, the
function tends to
1
i
. We can see that the dierent directions give dierent
innities.
Now we dene precisely what we mean on singularity.
10.1.1. Denition. If a function f : D C C is analytic on its domain
but not in the point D, we say that f is singular in this point. The
point is called singular point of f.
We note that it could be happen that a function is not analytic on a
domain like B(a, R) B(a, r), where 0 < r < R. In this case we also could
dene singularity. But we will not go into these details. However, considering
these possibilities, we could distinguish our above dened singularity and we
could say that is an isolated singularity because is an isolated point in
which our function is not analytic.
For example, the function 1/z is singular in the point = 0. To take an
other example,
f(z) =
1
z + 2
+
1
z 3
is singular in the points = 2 and = 3.
73
74 10. chapter. Singularities and zeroes of analytic functions
We know that if we have an analytic function f on a domain D, then
f has Taylor series expansion, or, in other words, the principal part of f is
identically zero (on the domain of analyticity D). This shows that singularity
is connected to the principal part of the function. Using this observation, we
dene the order of singularity.
10.1.2. Denition. Let f : D C C be an analytic function with
singularity in . Then by the theorem of Laurent f has the Laurent series
expansion
f(z) =
1

n=

n
(z )
n
+

n=0

n
(z )
n
=
+
2
1
(z )
2
+
1
1
(z )
1
+
0
+
1
(z ) +
2
(z )
2
+
The singular point can be of three dierent types:
1. is a removable singularity if
1

2
=
3
= = 0.
2. is pole of order n if
n
,= 0 but
m
= 0 if m > n.
3. is an essential singularity if innitely many of
n
,= 0.
Hence we have an important observation: a singularity can appear if the
principal part is zero, but if and only if the singularity is removable.
10.1.3. Example. Let f(z) be dened as
f(z) =
sin(z)
z
.
The function is not dened in z, but the limit exists:
lim
z0
f(z) = 1,
as it can be easily seen by the Taylor expansion of sin (see Section 5.1):
sin(z) =

n=0
(1)
n
(2n + 1)!
z
2n+1
= z
z
3
6
+ .
If we divide by z, we have that
f(z) =
sin(z)
z
= 1
z
2
6
+ .
And the limit really converges to 1.
10.1. Classication of singularities 75
The function g(z) =
1
z+1
has a singularity of order 1 in the point = 1,
since this is yet a Laurent expansion
1
z + 1
= 1
1
z (1)
,
so
1
= 1 ,= 0 but
n
= 0 if n > 1.
The function
h(z) =
1
(z 3)
3
+
3
(z + 1)
2
+
5
z
has three singular points: = 3 with order 3, = 1 with multiplicity 2
and = 0, which is a rst order singularity.
To see an example of essential singularity, we can consider the function
exp
_
1
z
_
=

n=0
1
z
n
n!
= 1 +
1

n=
1
n!
z
n
= 1 +
1
1

1
z
+
1
2

1
z
2
+
1
6

1
z
3
+
The removable singularity can be caracterized as follows: the function f
has a removable singularity in the point if and only if f is bounded in this
point. This is so, because if is removable, the principal part is zero of f
around , and there is no fractions which make the series giving f divergent.
The essential singularity is the most curious one, as we shall see later. The
theorem of Casorati and Weierstrass will discribe how a function behaves near
to its essential singularity. Now we just plot the absolute value of exp(1/z)
and the real part of this function, respectively.
76 10. chapter. Singularities and zeroes of analytic functions
10.2. The connection between zeros and poles 77
10.1.1 The denition of zeros
In this point we dene the zero of a function, because it is strongly connected
to singularities. In the following sections after dealing with singularities
we turn back to zeros and we shall give a deeper description of them.
10.1.4. Denition. Let f : D C C be a function. We say that the
point is a zero or root of f if f() = 0.
We can dene the order of a zero. is a zero of f of order n if
f() = 0, f

() = 0, f

() = 0, but f
(n)
,= 0.
If f is analytic on the domain D, then f has a Taylor series in this domain
with the center (since is contained in the domain D, we can have a Taylor
series expansion around this point):
f(z) =

n=0
f
(n)
()
n!
(z )
n
(z D).
If D is a zero of order n, the rst n term of the Taylor series cancel, so
we have that
f(z) =
f
(n)
()
n!
(z )
n
+
f
(n+1)
(n + 1)!
(z )
n+1
+
We can see that it is possible carry the common factor (z )
n
. Hence we
have the following proposition.
10.1.5. Proposition. The function f : D C C is analytic on D and
has a zero of order n if and only if then there is a function g : D C C
which is analytic on D, g() ,= 0 and
f(z) = (z )
n
g(z).
10.2 The connection between zeros and poles
The very strong connection between the zeros and poles now will be discuted.
10.2.1. Theorem. Let f : D C C be analytic on D . Then the
following two propositions are true
1. If f has a pole of order n, then
1
f
has a zero of order n.
2. If f has a zero of order n, then
1
f
has a pole of order n.
78 10. chapter. Singularities and zeroes of analytic functions
Proof. Let us suppose that f has a pole of order n in . Then in its Laurent
expansion around the terms with coecients a
n1
, a
n2
are all zero, so
f(z) = a
n
1
(z )
n
+ a
n+1
1
(z )
n1
+ + a
1
1
(z )
+
We can carry out (z )
n
to get
f(z) =
1
(z )
n
_
a
n
+ a
n+1
(z ) + + a
1
(z )
n1
+
_
.
Let us denote the function in the brackets by g(z). Then
f(z) =
g(z)
(z )
n
.
The function g is analytic on D, since it is dened by a Taylor series. Now
1
f
(z) =
(z )
n
g(z)
.
The function g is not zero in the point (it equals to g() = a
n
,= 0),
and by analiticity it is continuous, so in a neighbourhood of it is not zero.
Hence, by Proposition 10.1.5. f has a zero of order n in .
The proof of the second point is also follows from Proposition 10.1.5. If
f has a zero of order n in D, then f has the form
f(z) = (z )
n
g(z),
where g(z) ,= 0 in a neighbourhood of . Hence
1
f
(z) =
1
g(z)
1
(z )
n
.
The analiticity of g implies the analiticity of 1/g, so this latter function has
a Taylor expansion around
1
g
(z) = a
0
+ a
1
(z ) +
Multiplying this with
1
(z)
n
, we get that
f(z) =
a
0
(z )
n
+
a
1
(z )
n1
+
This shows that f has a pole of order n in the point .
We said before that removable singularity means that the function has
a nite limit in the given point. If the singularity is a pole, however, the
function has innity as a limit. This is the content of the next theorem.
10.2. The connection between zeros and poles 79
10.2.2. Theorem. The point D is a pole of f : D C C if and only
if
lim
z
f(z) = . (10.1)
Proof. Let be a pole of f of order n, say. Then, using our previous consid-
erations
f(z) = g(z)
1
(z )
n
,
for some g which is analytic in a neighbourhood of . It is obvious that the
fraction tends to innity if z . Moreover, lim
z
g(z) = g() ,= 0, so
(10.1) holds.
Now we show that if (10.1) holds then is a singularity of f. Indeed, if
lim
z
f(z) = ,
then [f(z)[ > 1 in a neighbourhood of . So 1/f is bounded and analytic on
this range. We dene
h(z) =
1
f
(z),
if z ,= and h(z) = 0 if z = . Hence h(z) will be continuous in and on
a neighbourhood around it. By the second point of the previous theorem we
know that f = 1/h has a pole in , since h has a zero in this point.
Exercises
10.2.3. Exercise. Determine the poles and the order of the poles of the
next functions (if they have):
1
z
2
1
,
z
2
+ 1
z
2
(z
4
+ 1)
,
1
exp z
,
exp(z)
(exp(z) 1)
2
.
10.2.4. Exercise. We know that sin(0) = 0, so the function
f(z) =
1
sin(z)
has a pole in = 0. Determine the order of this pole.
10.2.5. Exercise. Find the order of the zero of the function
f(z) = cos(z
4
) 1.
10.2.6. Exercise. What kind of singularity does the function
f(z) = exp
_
z + 1
z 1
_
have in the point z = 1?
80 10. chapter. Singularities and zeroes of analytic functions
10.3 The Theorem of Casorati and Weierstrass
Now we describe, how a function behaves near by its essential singularity. Let
f be analytic on D , and we suppose that it has an essential singularity
in . Then if we take an arbitrary small disk B(, ) around the point ,
then the set f(B(, ) ) is dense in C.
10.3.1. Theorem (Casorati
1
-Weierstrass
2
). Let f /(D ) and let
be an essential singularity of f. Then for any > 0
f(B(, ) ) = C.
Proof. Let us suppose that the statement is not true. Then there is an
such that [f(z) [ > for all z C and for some positive . We dene the
function g as
g(z) =
1
f(z)
.
This function os bounded, since
[g(z) =
1
[f(z) [
<
1

.
Since f is not dened in , the same is true for g, so is a singularity of g.
But g is bounded, so this singularity is removable. This means that the limit
lim
z
g(z)
exists. Dene the function g(z) as
g(z) =
_
g(z) z ,= ,
lim
z
g(z) z =
There are two possibilities: g() = 0 or g() ,= 0. If the rst possibility is
valid, then g(z) has the form
g(z) = (z )
n
h(z)
for some positive integer n. Hence, by denition
f(z) =
1
g(z)
=
1
(z )
n
h(z)

in the points z ,= . This shows that f has a singularity of order n in .
If g() ,= 0, then f has a removable singularity (since the limit lim
z
f(z)
exists).
Anyway, we have that the singularity of f is not essential, which is a
contradiction.
10.4. Further properties of zeros and poles 81
We remark that Picard
3
proved a stronger theorem. Namely, the theorem
of Picard says that the set
f(B(, ) )
not just dense, but it is always the whole complex set C, excluding maximum
one point. This theorem says, that using the above notations , the ecuation
f(z) =
can always be solvable in z for all C, with maximum one exception.
To see a particular example for the Picard theorem, we try to solve the
equation
exp
_
1
z
_
= .
The Picard theorem says that this can be solvable for any lambda with at
most one exception (since the function on the left has an essential singularity
in 0). The one exception is = 0, since we have seen earlier that the
exponential function never attend the value 0.
Taking logarithm,
z =
1
log()
=
1
log [[ + i(arg() + 2k)
.
We see that z exists, and taking great values for k, z can be arbitrarily small.
10.4 Further properties of zeros and poles
Now we go further in the investigation of the properties of zeros and poles.
The topological properties of the sets of zeros and poles will be revealed.
Then we will discuss several important consecuences, like the identity theo-
rem.
10.4.1. Denition. Let be a zero of f : D C C. Then the multi-
plicity of is denoted by mult
f
(). Moreover, the set of zeros is denoted by
Z
f
:
Z
f
= z D [ f(z) = 0.
10.4.2. Theorem. Let f : D C C is analytic on D. If we suppose that
f is not identically zero, then the next statements are valid.
1. mult
f
() < for all zeroes of f,
2. The set Z
f
does not contain limit points in D,
3
Charles mile Picard (1856-1941) French mathematician.
82 10. chapter. Singularities and zeroes of analytic functions
3. If K is a compact subset of D, then K Z
f
is nite,
4. [Z
f
[ [N[.
Proof. 1) Let us suppose that mult
f
() = for a zero . Since f is analytic,
it has a Taylor expansion
f(z) =

n=0
f
(n)
()
n!
(z )
n
(z D).
By our assumption all the Taylor coecients are zero, so the function is iden-
tically zero which contradicts with our assumption that f is not identically
zero.
2) Let Z
f
be a zero with multiplicity n. Then
f(z) = (z )
n
g(z),
where g is not zero in and a neighbourhood of . Since (z )
n
neither
zero in a neighbourhood of , we get that is an isolated zero. Since Z
f
was arbitrary, we get that every point of Z
f
is isolated in D. This also
implies that Z
f
is a closed set. Hence it cannot contain limit point, since
this limit point would be a member of Z
f
and it would be isolated, which is
a contradiction.
3) Let K D and Q = K Z
f
. If Q would be innite, then it would
have limit point in K. But this also would mean that this limit point is a
limit point of Z
f
as well. And we have just proved that this is impossible.
4) Let us choose a sequence of compact subset such that K K
n+1
and
K
n
= D. Then
Z
f
= Z
f
K
n
[ n = 1, 2, . . . .
We know from the previous point that Z
f
K
n
is nite. Countable union of
nite subsets is countable, so the last point is also proved.
10.4.3. Corollary. [Identity Theorem]Let f, g /(D) such that f(z) =
g(z) for all z H D, where H has a limit point in D. Then f(z) = g(z)
for all z D.
Proof. The function h(z) = f(z) g(z) is zero on a subset (H), which has a
limit point in D. By the previous theorem it can happen just if h is identically
zero on D.
As an illustration how strong this corollary is, let us take two analytic
function f and g on D = C. If the two functions are equal on the points 1/n,
f
_
1
n
_
= g
_
1
n
_
,
then the two functions are equal on the whole plane C. (This is so, because
the set Z
fg
has a limit point (0) in C).
Two other spectacular consequences are the following.
10.4. Further properties of zeros and poles 83
10.4.4. Corollary. If f, g /(D) such that f(z) = g(z) on a line segment
or a nonempty open set, then f(z) = g(z) on the whole set D.
10.4.5. Corollary. If f, g /(D), then fg 0 holds if and only if f 0
or g 0.
Proof. Let us suppose that f , 0. Then H = f
1
(C 0) is a nonempty,
and open set by Theorem 3.2.2. (since f is continuous). Our assumption is
that fg 0, so g(z) = 0 must hold on the set H. The previous corollary
gives the statement.
Finally we deal with the topological properties of the set of poles of a
function, which does not have essential singularities.
10.4.6. Denition. Let us denote the set of poles of a function f by P
f
.
10.4.7. Denition. Let f be a function on the domain D, which does not
have essential singularity. Then f is called meromorphic. The set of mero-
morphic functions on the domain D is denoted by /(D).
10.4.8. Theorem. Let f be a meromorphic function on D. Then
1. The set P
f
does not have limit points in D,
2. If K D and K is compact, then K P
f
is nite,
3. [P
f
[ [N[.
The proof is the same as the proof of the respective theorem for zeros.
Exercises
10.4.9. Exercise. Let f is analyic on the domain D, where B(0, 1) D.
Let us suppose that
f
_
1
n
_
= 0.
Show that f 0 on D.
10.4.10. Exercise. Let
f(z) = sin
_
1
z
_
sin
_
1
1 z
_
.
Determine the set Z
f
and the limit point(s) of Z
f
.
11
The Residue Theorem of Cauchy
and its applications
We have arrived to the probably most important theorem of complex analy-
sis. This theorem is the residue theorem of Cauchy. With this theorem we
will be able to prove some important additional theorem on zeros and poles
(Argument Principle, Theorem of Rouch).
11.1 The residue
11.1.1. Denition. Let f /(D a) and let us write the Laurent series
of f around a. Then the coecient a
1
is called the residue of f (with respect
to a). It is denoted by Res(f, a).
Formula (9.1) gives that
Res(f, a) =
1
2i
_
(a,r)
f(w)dw,
where r is an arbitrary radius such that B(a, r) does not contain other sin-
gularities other than a.
11.1.2. Example. The function
f(z) =
1
(1 + z)
2
has a pole at a = 1 of order 2. Since f is already written as a Laurent
series, we can see that a
1
= 0, so
Res(f, 1) = 0.
84
11.1. The residue 85
11.1.3. Example. The function f(z) = exp(z)/z
3
has a third order pole at
a = 0. Considering the Taylor expansion of exp(z), we have that
exp(z)
z
4
=
1
z
4
_
1
0!
+
1
1!
z +
1
2!
z
2
+
1
3!
z
3

_
=
1
z
4
+
1
z
3
+
1
2
1
z
2
+
1
6
1
z
+
1
24
+ .
Hence a
1
=
1
6
. So
Res
_
exp(z)
z
3
, 0
_
=
1
6
.
The calculation of residues is not always a trivial task. The next theorem
helps us, if we know the order of the pole.
11.1.4. Theorem. 1. Let f /(D a) and let us suppose that f has
a rst order pole at a. Then
Res(f, a) = lim
za
(z a)f(z).
2. If, specially, the function f has the form h/k, where h(a) ,= 0 but
k(a) = 0 (which is necessary if f has a pole in a), then
Res(f, a) =
h(a)
k

(a)
.
3. If f has a pole in a of order m, then let g(z) = (z a)
m
f(z). The
residue of f at a is
Res(f, a) =
g
m1
(a)
(m1)!
.
Proof. 1) By our assumption f(z) has a Laurent series around a of the form
f(z) =
a
1
z a
+ a
0
+ a
1
(z a) +
So
(z a)f(z) = a
1
+ a
0
(z a) + a
1
(z a)
2
+
Taking the limit z a, we get that
lim
za
(z a)f(z) = a
1
+a
0
lim
za
(z a) +a
1
lim
za
(z a)
2
+ = a
1
= Res(f, a).
2) Let f is of the form h/k with a rst order pole. Then
Res(f, a) = lim
za
(z a)f(z) = lim
za
(z a)
h(z)
k(z)
= lim
za
(z a)
h(z)
k(z) k(a)
=
86 11. chapter. The Residue Theorem of Cauchy and its applications
lim
za
h(z)
k(z)k(a)
za
=
h(a)
k

(a)
.
3) Now let the order of the pole a is m. Then
f(z) =
a
m
(z a)
m
+
a
m+1
(z a)
m1
+ +
a
1
(z a)
+ a
0
+ a
1
(z a) +
By the denition of g(z) we have that
g(z) = (z a)
m
f(z) =
a
m
+a
m+1
(z a) + +a
1
(z a)
m1
+a
0
(z a)
m
+a
1
(z a)
m+1
+
Then the m1th derivative of g becomes
g
(m1)
(z) = (m1)!a
1
+ m!a
0
(z a) +
In the point z = a
g
(m1)
(a) = (m1)!a
1
.
The result hence comes.
As an application, let us see the next examples.
11.1.5. Example. Let us calculate the residue of the function
z
3
+ z
2
2
z
2
+ 2z 1
in all the singularities.
The denominator equals to
z
2
+ 2z 1 = (z (1 +

2))(z (1

2)),
so this function has two rst order poles:
a
1
= 1 +

2, and a
2
= 1

2.
We will calculate
Res
_
z
3
+ z
2
2
z
2
+ 2z 1
, 1 +

2
_
.
Since the poles has order one, we can use the rst part of the theorem.
Res
_
z
3
+ z
2
2
z
2
+ 2z 1
, 1 +

2
_
= lim
z1+

2
(z (1 +

2))
z
3
+ z
2
2
z
2
+ 2z 1
=
lim
z1+

2
(z (1 +

2))
z
3
+ z
2
2
(z (1 +

2))(z (1

2))
=
lim
z1+

2
z
3
+ z
2
2
(z (1

2))
=
(1 +

2)
3
+ (1 +

2)
2
2
(1 +

2 (1

2))
=
(1 +

2)
3
+ (1 +

2)
2
2
2

2
=
6 + 3

2
2

2
=
3
2

3

2
.
11.2. The winding number 87
Exercises
11.1.6. Exercise. Show that if f has a removable singularity in a, then
Res(f, a) = 0.
11.1.7. Exercise. Calculate the residues
Res
_
1 + z
1 z
, 1
_
, Res
_
sin(z)
z
, 0
_
, Res
_
exp(z)
(exp(z) 1)
2
, 0
_
.
11.1.8. Exercise. Look for the poles of the function
z
2
+ 2
z
3
+ 3z
2
4z
and calculate the residues in these points.
11.2 The winding number
Later we shall need the winding number of a closed curve hence now we
deal with this notion, which is interesting itself.
Roughly speaking, the winding number of a curve counts how many
times runs around a curve a point.
Let z(r, ) = z = re
i
. Then the total derivative of z is
dz = e
i
dr + ire
i
d.
Dividing by z = re
i
,
dz
z
=
dr
r
+ id.
Integrating this along the curve we get that
_

dz
z
=
_

dr
r
+
_

id = log((b) (a)) + i
_

d = i
_

d.
The latter integral (if we divide by i) is the change in the angle . If we
divide by 2, we get the winding number of around z = 0. If we would
like to calculate the winding number in a point dierent from the origin, we
have to shift the denominator. Here comes the general denition.
11.2.1. Denition. The winding number of a closed curve around the
point p is dened as
wn(, p) =
1
2i
_

1
z p
dz.
88 11. chapter. The Residue Theorem of Cauchy and its applications
11.2.2. Example. Let us calculate the winding number of the circle (0, 1)
around p = 0.
wn((0, 1), 0) =
1
2i
_
(0,1)
1
z p
dz =
=
1
2i
_
2
0
1
exp(it) 0
exp

(it)dt =
i
2i
_
2
0
1
exp(it)
exp(it)dt =
1
2
_
2
0
1dt =
2
2
= 1,
as we waited.
The winding number is always integer, as we wait: if is a simple closed
curve, and p C Im(), then
wn(, p) Z.
11.3 The Residue Theorem of Cauchy
Now we turn to the statement of the residue theorem.
11.3.1. Theorem (Cauchy). Let f be meromorphic on the domain D. Let
us suppose that the path runs in D, closed and starlike. Moreover, let
us suppose that f does not contain poles on Im(), and inside Im() it has
nitely many poles:
1
, . . . ,
n
. Then
_

f(z)dz = 2i
n

i=1
wn(,
i
) Res(f,
i
).
In special, if all th poles are encircled just one time,
_

f(z)dz = 2i
n

i=1
Res(f,
i
).
Proof. Let us denote the principal part of f around
i
by p

i
(z). Then
g(z) := f(z) p

1
(z) p

2
(z) p

n
(z)
is an analytic function on D, without any poles. Therefore
_

(f(z) g(z))dz = 0,
by Cauchys theorem (Theorem 7.3.7.). Hence
_

f(z)dz =
_

g(z)dz =
n

i=1
_

i
(z).
11.3. The Residue Theorem of Cauchy 89
Every p

i
(z) has the form
p

i
(z) = a
n,i
1
(z
i
)
n
+ a
n+1,i
1
(z
i
)
n1
+ + a
1,i
1
(z
i
)
.
If we integrate these functions, all of the terms disappear:
_

a
k,i
1
(z
i
)
k
dz = 0
for all k = n, n + 1, . . . , 2. This is so, because these functions have
primitive functions (see also Theorem 6.2.5.). For k = 1 we have that
_

a
1,i
1
(z
i
)
dz = a
1,i
_

1
(z
i
)
dz.
This last integral is nothing else but the winding number of the pole
i
with
respect to the path , and a
1,i
= Res(f,
i
). Altogether, we have that
_

f(z)dz =
n

i=1
wn(,
i
) Res(f,
i
).
The theorem is proven.
11.3.1 Applications of the Residue Theorem
11.3.2. Example. Let us calculate the integral
_
(0,1)
1
z
2
4z + 1
dz.
The integrand has two singularities:
2 +

3 and 2

3.
The rst one is outside the unit circle (0, 1), so its residue will not appear
in the residue formula. Hence
_
(0,1)
1
z
2
4z + 1
dz = 2i Res
_
1
z
2
4z + 1
, 2

3
_
. (11.1)
This residue can be calculated by the rst point of Theorem 11.1.4.:
Res
_
1
z
2
4z + 1
, 2

3
_
= lim
z2

3
(z (2

3))
1
z
2
4z + 1
=
lim
z2

3
(z (2

3))
(z (2 +

3))(z (2

3))
= lim
z2

3
1
z (2 +

3)
=
1
2

3
.
Substituting this into (11.1), we get the answer:
_
(0,1)
1
z
2
4z + 1
dz = 2i
1
2

3
=
i

3
.
90 11. chapter. The Residue Theorem of Cauchy and its applications
11.3.3. Example. The second integral is
_
(0,2)
z + 7
z
4
+ 4z
3
+ 3z
2
dz.
The denominator has the form
z
4
+ 4z
3
+ 3z
2
= z
2
(z + 1)(z + 3).
This shows that the integrand has three singular points: 0, 1 and 3. The
third is outside of the path. Hence
_
(0,2)
z + 7
z
4
+ 4z
3
+ 3z
2
dz =
2i
_
Res
_
z + 7
z
4
+ 4z
3
+ 3z
2
, 0
_
+ Res
_
z + 7
z
4
+ 4z
3
+ 3z
2
, 1
__
.
The residues can be calculated as follows: we take the partial fraction de-
composition:
z + 7
z
4
+ 4z
3
+ 3z
2
=
7
3
1
z
2

25
9
1
z
+ 3
1
1 + z

2
9
1
3 + z
.
So the rst residue above is
Res
_
z + 7
z
4
+ 4z
3
+ 3z
2
, 0
_
=
25
9
,
while the second one is
Res
_
z + 7
z
4
+ 4z
3
+ 3z
2
, 1
_
= 3.
Collecting these and substituting, we get the result:
_
(0,2)
z + 7
z
4
+ 4z
3
+ 3z
2
dz = 2i
_

25
9
+ 3
_
= 2i
2
9
=
4i
9
.
Now we turn to a very useful and common trick to calculate real integrals
with contour ones. To this end, let us take the next example.
11.3.4. Example. Let us determine the value of the real integral
_
2
0
1
2 + cos(t)
dt.
We transform this integral to a path integral. First, let us realize that
_
2
0
1
2 + cos(t)
dt =
_
2
0
1
2 +
e
it
+e
it
2
dt
11.3. The Residue Theorem of Cauchy 91
holds, because
1
2
(e
it
+ e
it
) =
1
2
(cos(t) + i sin(t) + cos(t) + i sin(t)) =
1
2
2 cos(t).
Continuing the transformation, we can see that
_
2
0
1
2 +
e
it
+e
it
2
dt =
_
(0,1)
1
2 +
z+1/z
2

1
iz
dz.
The second one is an integral, for which the residue theorem already can be
applied:
_
(0,1)
1
2 +
z+1/z
2

dz
iz
=
1
i
_
(0,1)
1
2z +
z
2
2
+
1
2
dz.
The denominator has the form
2z +
z
2
2
+
1
2
=
1
2
(z (2

3))(z (2 +

3)).
The rst root, 2

3 is outside (0, 1), so we get that


_
2
0
1
2 + cos(t)
dt =
1
i
_
(0,1)
1
2z +
z
2
2
+
1
2
dz =
1
i
2i Res
_
1
2z +
z
2
2
+
1
2
, 2 +

3
_
=
2 lim
z2+

3
(z (2 +

3))
1
1
2
(z (2

3))(z (2 +

3))
=
2 lim
z2+

3
1
1
2
(z (2

3))
=
2

3
.
Another very useful trick comes what often helps to nd improper inte-
grals. To show how this trick works, let us take the next improper integral.
11.3.5. Example. We will show that
_

1
(x
2
+ 4x + 13)
2
dx =

54
.
The trick works as follows: we take the semicircle contour
= [R, R] z C [ [z[ = R, 0 < arg(z) < .
Here R > 0 is a xed radius. We determine
_

1
(x
2
+ 4x + 13)
2
dx
92 11. chapter. The Residue Theorem of Cauchy and its applications
rst. The integrand has two singularities, +2 3i and 2 + 3i. The rst
does not belong to the domain of . Hence
_

1
(x
2
+ 4x + 13)
2
dx = 2i Res
_
1
(x
2
+ 4x + 13)
2
, 2 + 3i
_
.
The pole is of second order, hence, following the third point of Theorem
11.1.4., we take the function
g(z) = (z (2 + 3i))
2
1
(x
2
+ 4x + 13)
2
,
and calculate
g
21
(2 + 3i)
(2 1)!
.
The result is
1
108i
. Hence
_

1
(x
2
+ 4x + 13)
2
dx = 2i
1
108i
=

54
.
On the other hand,
_

1
(x
2
+ 4x + 13)
2
dx =
_
R
R
1
(x
2
+ 4x + 13)
2
dx +
_

R
1
(x
2
+ 4x + 13)
2
dx,
where
R
is the semicircle part of . The rst integral tends to
_

1
(x
2
+ 4x + 13)
2
dx,
the integral we are looking for. The second integral tends to zero, if R :

R
1
(x
2
+ 4x + 13)
2
dx

sup
zIm(
R
)
1
(x
2
+ 4x + 13)
2
V (
R
)
1
(R
2
4R 13)
2R
2
0.
This gives the result.
Exercises
11.3.6. Exercise. Evaluate the integral
_
(0,3)
3z + 8
z
4
+ 4z
3
5z
2
dz.
11.3. The Residue Theorem of Cauchy 93
11.3.7. Exercise. Show that
_
2
0
1
3 cos(t) + sin(t)
dt =
2

7
by converting the integral to a path integral.
11.3.8. Exercise. Prove that the integral
_
(0,)
z tan(z)dz = 0.
11.3.9. Exercise. Taking the above used semicircle contour, show that
_

x sin(x)
x
2
+ 1
dx =

2e
.
Hint: use the function
g(z) =
z exp(z)
z
2
+ 1
with the semicircle contour, and nally take the imaginary part.
11.3.10. Exercise. Prove that
_

1
1 + x
6
dx =
2
3
.
12
Some additional theorems on
zeros and poles of meromor-
phic functions
12.1 The number of zeros and poles
The residue theorem is very useful in determination of integrals. In this
section we give new applications.
12.1.1. Theorem. Let f : D C C be a meromorphic function. Let
be a closed simple, starlike curve in D. Let us suppose that f does not have
zeros and poles in Im(). Then
_

(z)
f(z)
dz = 2i(N P),
where N is the number of zeros and P is the number of poles in Im(),
including the multiplicities.
Proof. Since Im() is a compact subset of D, N and P are nite (see Theo-
rems 10.4.2. and 10.4.8.). We also know that
f(z) = (z a)
k
g(z) (z D),
where a is a zero and k is its multiplicity. Then
f

(z)
f(z)
=
k
z a
+
g

(z)
g(z)
,
whence we see that f

/f has a rst order pole at z = a, and its residue is k.


Similarly, if a is a pole of order k, then this point is a pole of f

/f of
order one and with residue m.
From this point our theorem is the consequence of the Residue Theorem.
94
12.1. The number of zeros and poles 95
12.1.2. Corollary (Argument Principle). If f : D C C is an ana-
lyticfunction, then with the same assumptions
_

(z)
f(z)
dz = 2iN.
As an application, we take the next example.
12.1.3. Example. Let us show that f(z) = cos(z) 1 +
z
2
2
has four roots
in the domain B(0, 1). This function does not have poles, so the Argument
Principle will help us.
To this end, we need to show that
_
(0,1)
z sin(z)
cos(z) 1 +
z
2
2
dz = 2i 4,
We calculate this integrals by the residue theorem.
First, let us analyze the integrand. The denominator has the Taylor series
expansion
cos(z) 1 +
z
2
2
= 1
z
2
2
+
z
4
24

z
6
720
+ 1 +
z
2
2
=
z
4
24

z
6
720
+ ,
while for the nominator we have
z sin(z) = z
_
z
z
3
6
+
z
5
120
+
_
=
z
3
6
+
z
5
120
+
Hence
The integrand has a pole in the point 0. The Laurent series about this
point is
z sin(z)
cos(z) 1 +
z
2
2
=
z
3
6
+
z
5
120
+
z
4
24

z
6
720
+
,
so, if we divide by z
3
, we see that z = 0 is a rst order pole. Its residue can
be calculated by the limit
Res
_
z sin(z)
cos(z) 1 +
z
2
2
, 0
_
= lim
z0
z
z sin(z)
cos(z) 1 +
z
2
2
.
If we use tha above-calculated series expansions, we have that
lim
z0
z
z sin(z)
cos(z) 1 +
z
2
2
= lim
z0
z
z
3
6
+
z
5
120
+
z
4
24

z
6
720
+
= lim
z0
z
4
6
+
z
5
120
+
z
4
24

z
6
720
+
=
24
6
= 4.
Hence our integral equals to
_
(0,1)
z sin(z)
cos(z) 1 +
z
2
2
dz = 2i 4,
as we said. Hence the result follows from the Argument Principle.
96 12. chapter. Some additional theorems on zeros and poles. . .
12.2 Entire functions - the theorem of Liouville
There are no analytic functions on the whole set C, which are constant. This
is the theorem of Liouville. As we shall see, the Fundamental Theorem of
Algebra immediately follows from this fact.
12.2.1. Denition. A function f /(C) is called entire.
12.2.2. Theorem (Liouville
1
). If f is entire and bounded, then it is con-
stant.
Proof. Let us suppose that f is entire, and [f[ M. Then for any R > 0
f

(a) =
1
2i
_
(a,R)
f(w)
(w a)
2
dw,
so
[f

(a)[
1
2i
_
(a,R)

f(w)
(w a)
2

dw
M
2
V ((a, R))
1
(R a)
2

M
2
2R
1
R
2
=
M
R
.
Since R can be arbitrarily large, f

(a) = 0. Since a C is arbitrary, we get


the statement.
Now we prove the next fundamental theorem.
12.2.3. Theorem (The Fundamental Theorem of Algebra). Any po-
lynomial of degree at least one has a complex root.
Proof. The polynomial p(z) is entire, and not constant if its degree is greater
than zero. By the Liouville theorem p(z) cannot be bounded. Hence 1/p(z)
is not constant. Let us suppose that p(z) does not have zero at all. By this
reason, its reciprocal is bounded, and entire. Liouvilles theorem gives that
it is constant, which is impossible, since p(z) is not constant. The reason is
that we supposed that p(z) does not have zero.
12.2.4. Corollary. Every polynomial can be written in the form
p(z) = c(z
1
)
n
1
(z
2
)
n
2
(z
k
)
n
k
.
Here n
i
is the multiplicity of
i
, and
i
C are the roots of p, moreover,
c C.
Proof. We have seen that p(z) has a root, say
1
. Hence, by Proposition
10.1.5. we know that
p(z) = (z
1
)
n
1
g(z),
where g(z) is a polynomial and does not have zero in
1
. We can continue
by induction.
12.2. Entire functions - the theorem of Liouville 97
We remark that this theorem can be extended from polynomials to func-
tions with nitely many zeros:
12.2.5. Theorem. Let us suppose that f is entire, and it has nitely many
roots, say
1
, . . . ,
n
. Then f can be written in the form
f(z) = (z
1
)
n
1
(z
2
)
n
2
(z
k
)
n
k
exp(g(z)) (z C),
where g is entire.
Exercises
12.2.6. Exercise. Let us suppose that f is entire such that 1(f(z)) 0 for
any z C. Prove that f is constant.
12.2.7. Exercise. Let us suppose that f is entire such that (f(z)) 0 for
any z C. Prove that f is constant.
13
The Weierstrass Product The-
orem and the Mittag-Leer ex-
pansion
The two main theorems of this chapter describes when and how it is possible
construct a meromorphic function which have the prescribed zeros and poles,
with prescribed multiplicities.
First we deal with the construction of meromorphic functions having pre-
scribed set of zeros with given multiplicities.
It is straightforward to see, that if we would like to construct a function
f(z) having nite zeros,
1
, . . . ,
n
, we can take the form
f(z) = (z
1
) (z
n
).
One can guess that it is working with innitely many zeros. This is true, if
we take care about the convergence of innite sums.
To see how we can deal with innite products, we go into details now.
13.1 Innite products
13.1.1. Denition. Let a
n
be a sequence of real, nonzero numbers. We
construct the next sequence:
b
n
:= a
1
a
2
a
n
.
If the limit of b
n
as a sequence exists and it is A, say, then we say that the
product of the sequence a
n
exists, or that the product converges. This limit
of b
n
is denoted by
A =

k=1
a
k
.
98
13.1. Innite products 99
13.1.2. Example. The product

k=1
_
1
1
k
_
converges to zero, since
n

k=1
_
1
1
k
_
=
_
1
1
1
__
1
1
2
_

_
1
1
n
_
=
1
2
2
3
3
4

n 1
n
=
1
n
0.
13.1.3. Lemma. If a product converges to a nonzero number, then its fac-
tors must tend to 1.
Proof. Let
b
n
= a
1
a
2
a
n
.
Let us suppose that

k=1
a
k
= A.
Then
a
n
=
a
1
a
2
a
n1
a
n
a
1
a
2
a
n1
=
b
n
b
n1

A
A
= 1.
The next two theorems are very useful deciding the convergence of prod-
ucts of the form

k=1
(1 + a
k
).
13.1.4. Theorem. Let a
n
be a nonnegative sequence. The product

k=1
(1 + a
k
).
converges if and only if the sum

k=1
a
k
converges.
100 13. chapter. The theorems of Weierstrass and Mittag-Leer
Proof. Let us dene the partial products and partial sums of a
n
as
b
n
:=
n

k=1
(1 + a
k
),
and
s
n
:=

k = 1
n
a
k
.
Since a
n
is nonnegative, b
n
and s
n
are nondecreasing. This implies that they
converge if and only if they are bounded.
Since e
x
1 + x for any x 0, we have that
b
n
=
n

k=1
(1 + a
k
)
n

k=1
e
a
k
= e
a
1
+a
2
++a
n
= e
s
n
.
This gives that if s
n
is bounded, then b
n
is bounded, too. On the other hand,
b
n
= (1 + a
1
)(1 + a
2
) (1 + a
n
) 1 + a1 + a
2
+ + a
n
= 1 + s
n
.
This gives that if b
n
is bounded, then s
n
is bounded, too. We are done.
13.1.5. Example. The product

k=1
_
1 +
1
k
a
_
is convergent for any a > 1, because the sum

n=1
1
n
a
is convergent. But the product

k=1
_
1 +
1

k
_
is divergent by the same reasons.
Technically somewhat harder to prove that the nonnegativity of a
n
can
be omitted in the above theorem:
13.1.6. Theorem. Let a
n
be a real sequence. The product

k=1
(1 + a
k
).
converges if and only if the sum

k=1
a
k
converges.
13.2. The Weierstrass Product Theorem 101
13.2 The Weierstrass Product Theorem
To see how the idea of the construction of functions with innite prescribed
zeros works, rst we take an example. The sin(z) function has zeros on the
set
Z
sin
= 0, , 2, 3, . . . ,
as we have seen before. Hence, one could think that sin can be written as
sin(z) = z
_
1
z

__
1
z
2
_

_
1 +
z

__
1 +
z
2
_
.
We have two innite products which are divergent. We can resolve this
problem if we pair the factors:
sin(z) = z
_
1
z

__
1 +
z

__
1
z
2
__
1 +
z
2
_
=
sin(z) = z
_
1
z
2

2
__
1
z
2
4
2
_
.
Now we have one product, which is convergent, since

n=1
z
2
k
2

2
=
z
2

n=1
1
k
2
=
z
2
6
<
for all z C. Thus we have (at least in principle) that
sin(z) = z

n=1
_
1
z
2

2
k
2
_
.
This result was known by Euler 250 years ago.
It can happen that a function f has zeros on the negative integers:
Z
f
= 0, 1, 2, 3, . . . .
Then our approach would say that
f(z) = z
_
1 +
z
1
__
1 +
z
2
_
.
Sadly the above product does not converge, however, there is no theorem
which shows that the function f does not exists. Weierstrass was who could
resolve the problem. He introduced the so-called prime factors:
E
0
(z) = 1,
E
1
(z) = e
z
,
E
2
(z) = e

z+
z
2
2

,
E
3
(z) = e

z+
z
2
2
+
z
3
3

,
.
.
.
102 13. chapter. The theorems of Weierstrass and Mittag-Leer
Weierstrass said that the product with which we would like to represent a
function is not convergent, we can use the appropriate prime factor to make
the product nite. Since the exponential factors in E
n
(z) are never zero, we
cannot involve unwanted zeros into our product.
For example, as we have seen, the product
_
1 +
z
1
__
1 +
z
2
_

is not convergent. Hence, if we would like to represent the above function f,
then we can involve some Weierstrass prime factors:
f(z) = z
_
1 +
z
1
_
E
n
1
_
z
1
__
1 +
z
2
_
E
n
2
_
z
2
_

It is easy to see that it is enough to choose n
1
= n
2
= = 1. Indeed,
_
1 +
z
1
_
E
1
_
z
1
__
1 +
z
2
_
E
1
_
z
1
_
=
_
1 +
z
1
_
e
(z/1)
_
1 +
z
2
_
e
(z/2)

We shall prove later that the above product is already convergent. This
function
z
_
1 +
z
1
_
e
(z/1)
_
1 +
z
2
_
e
(z/2)

apart from a factor is a very important function, which is called the Euler
Gamma function. We shall deal with it in more details in the followings.
Here is the time to phrase the Weierstrass theorem.
13.2.1. Theorem (Weierstrass). If f : C C is an entire function with
zeros
1
,
2
, . . . (none of them is zero) with respective multiplicity
1
,
2
, . . . ,
and a zero in z = 0 with multiplicity 0 then
f(z) = z
k
exp(h(z))

k=1
__
1
z

k
_

k
E
m
k
_
z

k
__
.
Here h : C C is an entire function. The prime factors E
m
k
are chosen
such that the product has to be convergent.
If f is not entire but meromorphic on C, then it can be written as a
fraction g/h, where g and h can be written as in the above theorem. (Now
the zeros of h will be the poles of f.)
13.2.2. Example. The function e
z
does not have zeros, so e
z
is already a
product representation with an empty product. But the function e
z
1 have
zeros, innitely many. The seemingly straightforward way (looking for the
zeros directly and then forming th product) will not be suitable. To see why
13.2. The Weierstrass Product Theorem 103
not, let us solve the equation, that is, we look for the zeros of the function
e
z
1:
e
z
1 = 0 =e
z
= 1.
Taking the logarithm,
z = log(1) = log [1[ + i(arg(1) + 2k) = 2ki (k Z).
Then we see that the zeros of e
z
1 are all have multiplicity one. Separating
the zero z = 0,
e
z
1 = z exp(h(z))
_
1
z
2i
__
1 +
z
2i
__
1
z
4i
__
1 +
z
4i
_
=
z exp(h(z))
_
1 +
z
2
4
2
__
1 +
z
2
16
2
_

= z exp(h(z))

n=1
_
1 +
z
2
4n
2

2
_
.
Here comes the problem: the theorem of Weierstrass do not say nothing on
the function h(z). Hence we need a trick.
Using the formula
sin(z) =
1
2i
(exp(iz) exp(iz)),
we have that
e
z
1 = 2ie
z/2
sin
_

iz
2
_
.
Using the innite product for sin
sin
_

iz
2
_
=
iz
2

n=1
_
1 +
z
2
4
2
k
2
_
,
we get that
e
z
1 = 2ie
z/2
_

iz
2
_

n=1
_
1 +
z
2
4
2
k
2
_
=
ze
z/2

n=1
_
1 +
z
2
4
2
k
2
_
.
Which is the nal answer.
104 13. chapter. The theorems of Weierstrass and Mittag-Leer
Exercises
13.2.3. Exercise. Show that

k=1
_
1
1
k
2
_
=
1
2
.
13.2.4. Exercise. Let a
n
= e
1/n
2
Find the value of the product

n=1
a
n
.
13.2.5. Exercise. Show that
cos(z) =

k=1
_
1
4z
2

2
(2k 1)
2
_
,
sinh(z) = z

k=1
_
1 +
z
2

2
k
2
_
,
cosh(z) =

k=1
_
1 +
4z
2

2
(2k 1)
2
_
.
13.3 The Mittag-Leer expansion
Now we go to present the pair of the Weierstrass Theorem. This said that it
is possible to construct a function with a given set of zeros. The theorem of
Mittag-Leer
1
gives us a tool to construct functions with prescribed set of
singularities.
If we would like to construct a function with a singularity of z = 1 with
order 2 and a singularity at z = 3 of order one, say, then the next function
is suitable:
1
(z + 1)
2
+
1
z 3
.
This is a composition of the Laurent expansions around the singularities,
too. The terms of the sum are the principal parts of our function around the
singularities.
Hence one could argue that if the function has innitely many singulari-
ties, we just simply add the principal parts. But this does not work always.
1
Magnus Gustaf Mittag-Leer (1846-1927) Swedish mathematician
13.3. The Mittag-Leer expansion 105
For example, if the function f has singularities in the points 1, 2, 3, . . . , then
this approach would give that
f(z) =
1
z 1
+
1
z 2
+
1
z 3
+ =

n=1
1
z n
.
But this series is divergent for any z C. What we can do is that we
substract a polynomial from all the terms what makes the series convergent.
In this case, for example, we can do the following:
1
z n
=
1
n
1
1
z
n
=
1
n

k=0
_
z
n
_
k
.
Hence if we subtract the partial sums

1
n
n

k=0
_
z
n
_
k
from
1
zn
, the dierence will tend to zero, and the sum will be convergent:

n=1
_
1
z n

_

1
n
_
n

k=0
_
z
n
_
k
_
=

n=1
_
1
z n
+
1
n
n

k=0
_
z
n
_
k
_
.
Using the fact that the inner sum has the form
n

k=0
_
z
n
_
k
=
n z(z/n)
n
n z
,
we get a simpler form:

n=1
_
1
z n
+
1 (z/n)
n1
n z
_
.
Further simplication is possible:

n=1
_
1
z n
+
(z/n)
n1
1
z n
_
=

n=1
_
1
z n
+
(z/n)
n1
1
z n
_
=

n=1
(z/n)
n1
z n
.
This function has the wanted properties and the series is convergent on
the whole set C minus the original singularities 1, 2, 3, . . . . Finally we can
realize that we can add any entire function h(z).
We can generalize this idea which is the theorem of Mittag-Leer.
106 13. chapter. The theorems of Weierstrass and Mittag-Leer
13.3.1. Theorem (Mittag-Leer). Let f /(C) be a meromorphic func-
tion on C with singularities 0 =
0
,
1
, . . . with given multiplicities. Then
f(z) = h(z) +

n=1
_
p
n
_
1
z
n
_
h
n
(z)
_
(z C).
Here the polynomials p
n
are the principal parts of f and (h
n
(z)) are a suitable
sequence of polynomials.
13.3.2. Example. We shall use the theorem to prove a very useful sum
representation for the cotangent function
cot(z) =
cos(z)
sin(z)
.
It is more easy to deal with cot(z) in place of cot(z). The function cot(z)
has singularities where sin(z) has zeros. Namely:
P
cot
= 0, 1, 2, 4, . . . .
These roots are all of rst order. The residues:
Res(cot(z), k) = lim
zk
(z (k))
cos(z)
sin(z)
= cos(k) lim
zk
z k
sin(z)
=
cos(k)
1
cos(k)
=
1

.
This holds for all k Z. Hence the principal parts around the zero k is
1

1
z k
(k Z).
This means that
cot(z) =
1
z
+

k=1
1

_
1
z k
+
1
z + k
_
=
1
z
+

k=1
1

2z
z
2
k
2
.
Multiplying by , we get the next nice formula:
cot(z) =
1
z
+

k=1
2z
z
2
k
2
.
Substituting z =
1

2
, we get the interesting sum formula:

k=1
1
k
2

1
2
= 1

2
2
cot
_

2
_
.
13.3. The Mittag-Leer expansion 107
13.3.3. Example. The hyperbolic cotangent can be dened as coth(z) =
i cot(iz). Therefore the zero set of coth is
P
coth
= 0, i, 2i, 4i, . . . .
Running through the steps in the last example now with the hyperbolic
cotangent function we get the next identity

2z
coth(z) =
1
2z
2
+

k=1
1
z
2
+ k
2
.
If we substitute z = 0, we get the famous formula of Euler:

n=1
1
n
2
=

2
6
.
The substitution z = 1 gives that

n=1
1
n
2
+ 1
=
1
2
+

2
coth().
Exercises
13.3.4. Exercise. Find the Mittag-Leer expansion of the functions
f(z) =
1
exp(z) 1
,
g(z) = tan(z).
Part IV
Special functions
108
109
The functions exp, sin, tan, etc appear often in many contexts in math-
ematics and every sciences in which they use mathemaitcs. There are some
additional functions which often appear in a bit more advanced level. These
functions are the Riemann zeta function, the Euler gamma function, to men-
tion just two of them. Having the necessary complex function theoretical
tools, we now turn to these functions. (Without the theory of analytic func-
tions, it is almost impossible study these functions.)
14
Analytic extension
In this chapter we shall frequently use our observation, which we called Iden-
tity Theorem (see Corollary 10.4.3.). This says that if we have two functions
on a domain D such that they are equal on a set H which has a limit point
in D, then the two functions are equal on the whole domain D.
In special, if the set H contains an (even arbitrarily small) disk or line
segment, then the two functions coincide on the whole set D.
14.1 The notion of analytic extension
14.1.1. Denition. Let g : /(D) be an analytic function on the domain D.
Let us suppose that E is a greater domain, containing D properly: D E.
If f /(E) such that f D = g, then we say that f is an analytic extension
of g.
14.1.2. Example. The most basic and non trivial example for analytic ex-
tension is of the geometric series. We dene the function g by the series
g(z) :=

n=0
z
n
.
Then g is analytic on the open disk D = B(0, 1). We also know that
g(z) =
1
1 z
(z B(0, 1)).
This latter function on the right, however, is dened on the whole complex
plane minus the point z = 1. This means that f(z) = 1/(1z) is an analytic
extension of g. Moreover, this is the largest possible extension.
14.2 The Riemann zeta function
We go forward, and take a famous example.
110
14.2. The Riemann zeta function 111
14.2.1. Denition. The function is dened by the series
(s) =

n=1
1
n
s
.
This function is called the Riemann
1
zeta function.
The variable s in the denition is a complex variable. Now we show that
the above series dening the function is convergent for any complex number
s, if 1(s) > 1. Indeed, let us suppose that s = + it, where and t are
reals. Then

n=1
1
n
s

n=1
[n
s
[ =

n=1
[n

[[n
it
[ =

n=1
[n

[ =

n=1
1
n

.
And this latter series, as we know from real analysis (by the integral test) is
convergent if > 1.
Hence we have the next proposition:
14.2.2. Proposition. The function is analytic on the half plane
H = + it [ 1() > 1, t R.
In the next we show that it is easy to nd an analytic extension of the
Riemann zeta function. To this end let us take the next denition.
14.2.3. Denition. The function
(s) =

n=1
(1)
n
n
s
is called Dirichlet eta function.
It is easy to see, by using the Leibniz test, that this function (as a series)
is convergent if 1(s) > 0.
To see how this can be related to the Riemann zeta function, let us x
an s from the above set H. Then

n=1
(1)
n
n
s
=

n=1
(1)
2n
(2n)
s
+

n=0
(1)
2n+1
(2n + 1)
s
=
1
2
s

n=1
1
n
s

n=0
1
(2n + 1)
s
.
Hence we have that
(s) =
1
2
s
(s)

n=0
1
(2n + 1)
s
. (14.1)
1
Bernhard Riemann (1826-1866), German mathematician.
112 14. chapter. Analytic extension
The last sum can be expressed by . To this end, let us do the following.
Every even natural number n can be written as 2
k
m, where m > 0 is odd.
Then
(s) =

k=1

m = 1

1
(2
k
m)
s
+

n=0
1
(2n + 1)
s
=

k=1
1
2
ks

m=1
1
m
s
+

n=0
1
(2n + 1)
s
= (s)
2
s
1 2
s
+

n=0
1
(2n + 1)
s
.
After a rearrangement we get the next formula.
14.2.4. Proposition.

n=0
1
(2n + 1)
s
= (s)
_
1
1
2
s
_
.
Substituting this into (14.1), and doing a simplication, we arrive at the
next theorem.
14.2.5. Theorem. We have
(s) =
1
2
1s
1
(s).
Hence we have the analytic extension of to the plane
H
1
= + it [ 1() > 0, t R.
14.2.6. Remark. It can be seen, that the singularities coming from the
factor
1
2
1s
1
are not real singularities of , since this factor is singular if
2
1s
= 1, that is, on the set
_
1 +
2ik
log(2)

k Z
_
.
But on this set we can use the original denition of .
15
The Riemann zeta function
We go a bit deeper in the discussion of the properties of the Riemann zeta
function.
15.1 The values of at positive even integers
We deduce the famous formula (2n) (n > 0). To do this, we need a deni-
tion.
15.1.1. Denition. The Taylor coecients times n! of the function
z
e
z
1
around z = 0 are called Bernoulli numbers:
z
e
z
1
=

n=0
B
n
n!
z
n
.
It can be easily seen that
B
0
= 1, B
1
=
1
2
, B
2
=
1
6
, B
3
= 0, B
4
=
1
30
, . . .
These numbers are necessary to express (2n) in a closed form.
15.1.2. Theorem.
(2n) =
(1)
n1
B
2n
2
2n1

2n
(2n)!
(n = 1, 2, . . .).
Proof. The next proof comes from Knuth, Graham, Patashik: Concrete
Mathematics. First, observe that by the Mittag-Leer expansion of the
cotangent function
x cot(x) = 1 + 2

n=1
x
2
x
2
n
2
= 1 2

n=1
x
2
n
2
x
2
. (15.1)
113
114 15. chapter. The Riemann zeta function
The members of the sum can be expanded into a Taylor series:
x
2
n
2
x
2
=
x
2
n
2
+
x
n
n
4
+ =

m=1
x
2m
n
2m
.
Substituting
x

in (15.1) and using the expansion above, we have


cot(x) = 1 2

n=1

m=1
x
2m

2m
n
2m
= 1 2

m=1

x
2m

2m

n=1
1
n
2m
=
1 2

m=1
x
2m
(2m)

2m
.
This shows the surprising result that x cot(x) is the generating function of
the even zeta values.
Expanding x cot(x) on a dierent way we shall get our proposition. Take
the function
x
e
x
1
. Since
B
1
1!
=
1
2
x, we subtract this term (there is no such
a term with x in our series involving (2m),
x
e
x
1
+
x
2
e
x
+ 1
e
x
1
=
x
2
e
x
2
+ e

x
2
e
x
2
e

x
2
.
Since
e
x
+ e
x
e
x
e
x
=
cosh(x)
sinh x
= coth(x), we have
x
2

e
x
2
+ e

x
2
e
x
2
e

x
2
=
x
2
coth
x
2
.
Note that
x
2
coth
x
2
=
x
2
cot
x
2
, so this is an even function.
Therefore, and because of our former rearrangements,
x
2
coth
x
2
=
x
e
x
1
+
x
2
=

n=0
B
2n
x
2n
(2n)!
, so
x coth(x) =

n=0
B
2n
4
n
(2n)!
x
2n
.
Finally, since cot x =
cos x
sin x
=
i cosh(ix)
sinh(ix)
= i coth(ix), we have that
x cot x = ix coth(ix) =

n=0
B
2n
4
n
(2n)!
(ix)
2n
=

n=0
B
2n
4
n
(1)
n
(2n)!
x
2n
.
Since the coecients must be equal, the two representations
x cot x = 1 2

m=1
(2m)

2m
x
2m
=

n=0
B
2n
(4)
n
(2n)!
x
2n
,
the theorem follows.
15.2. The connection between primes and the Riemann zeta function 115
15.1.3. Corollary. We get the next special values:
(2) =

2
6
, (4) =

4
90
, (6) =

6
945
, . . .
The rst value we deduced earlier.
15.2 The connection between primes and the
Riemann zeta function
The zeta function is strongly connected to number theory via the formula
rstly proved by Euler.
15.2.1. Theorem (Euler). If 1(s) > 1, then
1
(s)
=

n=1
_
1
1
p
s
n
_
,
where p
n
is the nth prime: p
1
= 2, p
2
= 3, . . .
Proof. From Proposition 14.2.4. we have that

n=0
1
(2n + 1)
s
= (s)
_
1
1
2
s
_
.
On the left the sum runs over odd numbers. With other words, over numbers
which have no prime factor p
1
= 2. With the same argument one can also
show that

2n,3n
1
n
s
= (s)
_
1
1
2
s
__
1
1
3
s
_
.
Continuing the process, we get that on the left all the positive numbers are
excluded except 1. Then
1 = (s)
_
1
1
2
s
__
1
1
3
s
__
1
1
5
s
_
.
Dividing by (s), we get the result.
This formula of Euler is one of the most important formulas in mathe-
matics. This shows that number theory and complex analysis are strongly
connected.
116 15. chapter. The Riemann zeta function
Exercises
15.2.2. Exercise. Using the relation between the cotangent function and
the zeta, show the next identity:

n=1
(2m)
2
m
=
1
2


2

2
cot
_

2
_
.
15.2.3. Exercise. With the aid of Theorem 14.2.5. show that in s = 1 there
is a rst order pole of . Moreover, prove that
Res(, 1) = 1.
15.2.4. Exercise. Show that

n=2
((n) 1) = 1.
16
The Euler Gamma function
The other function appearing very often in dierent parts of mathematics is
the Euler Gamma function. This function is the extension of the factorial
function ! : N N, as we shall see soon.
16.0.5. Denition. Let us dened the function (z) by the next improper
integral:
(z) =
_

0
t
z1
e
t
dt.
This function is called the Euler Gamma function.
By taking the absolute value of the integrand, we can easily see that the
integral is nite for any z from the set
D = x + iy [ x > 0, y R.
Hence the function is dened and analytic on the set D.
The next theorem has two main importance: shows that as we said
is the extension of the factorial function, and it will helps us to nd the
analytic extension of .
16.0.6. Theorem. For any z D, we have that
(z + 1) = z(z).
Proof. By partial integration,
(z) =
_

0
t
z1
e
t
dt =
1
z
[t
z
e
t
]

0
+
1
z
_

0
t
z
e
t
dt =
0 +
1
z
(z + 1).
117
118 16. chapter. The Euler Gamma function
16.0.7. Corollary. We easily get that
(1) =
_

0
e
t
= 1.
Moreover,
(2) = 1(1) = 1,
(3) = 2(2) = 2,
(4) = 3(3) = 6,
(5) = 4(3) = 24,
.
.
.
Hence, in general,
(n + 1) = n!.
Now we show how it is possible to extend the Gamma function the the
whole complex plane, excluding the points 0, 1, 2, 3, . . . .
16.0.8. Corollary. The above theorem also shows that
(z + n) = z(z + 1)(z + 2) (z + n 1)(z).
Rearranging, we get that
(z + n)
z(z + 1)(z + 2) (z + n 1)
= (z).
Hence we can calculate smaller values using greater values on the real line.
But in the denominator we can see that z cannot be 0, 1, . . . , n+1. Since
n can be arbitrary positive integer, the function has poles in these points (the
nominator does not cancel the singularity). For example, we can calculate
the value
_

1
2
_
using the original function denition. To this end, we can
substitute n = 1 in the above formula:

1
2
_
=

1
2
+ 1
_

1
2
_

1
2
+ 1
_ =

_
1
2
_

1
4
= 4
_
1
2
_
.
The above considerations show that the following theorem is true.
16.0.9. Theorem. The function is meromorphic on the whole set C with
rst order poles at the nonpositive integers, with residues
Res(, n) =
(1)
n
n!
.
119
Proof. All the statements had been proven earlier, except the residues. For
the residues we have that
Res(, n) = lim
zn
(z+n)(z) = lim
zn
(z+n)
(z + n)
z(z + 1)(z + 2) (z + n 1)
=
lim
zn
(z + n + 1)
z(z + 1)(z + 2) (z + n 1)
=
(1)
(n)(n + 1)(n + 2) (1)
=
(1)
n
n!
.
The next theorem is the consequence of the previous one, and of the
Mittag-Leer theorem.
16.0.10. Theorem. For any complex z, we have the next formula
(z)(1 z) =

sin(z)
.
Proof. The function
f(z) = (z)(1 z)
1
z
,
as we can easily see, has rst order poles at all the nonzero integers, and these
poles are of rst order. Having the residues, we can use the Mittag-Leer
theorem. The residues are
Res(f, n) = lim
zn
(z + n)
_
(z)(1 z)
1
z
_
=
Res(, n)(n + 1) = (1)
n
.
The same for n:
Res(f, n) = (1)
n
.
This shows that, by the Mittag-Leer theorem, the function f can be
rewritten as
f(z) =

n=1
_
(1)
n
1
z n
+ (1)
n
1
z n
_
= 2z

n=1
(1)
n
z
2
n
2
.
Using the same argument as in Example 13.3.3., we easily get the expansion

sin(z)
=
1
z
+ 2z

n=1
(1)
n
z
2
n
2
.
With this we nished the proof.
120 16. chapter. The Euler Gamma function
16.0.11. Corollary. The above formula immediately gives the special value

_
1
2
_
=

.
Using the factorial notion
(n + 1) = n!,
we formally get that
_

1
2
_
! =

.
Also, using the original denition of the Gamma function, we get the
value of the improper integral
_

0
t
1
2
e
t
dt =
_
3
2
_
=
1
2

_
1
2
_
=

2
.
16.1 The connection between the and func-
tions
Using the denition the function and making the substitution t = nu, we
get that
(z) =
_

0
(nu)
z1
e
nu
ndu = n
z
_

0
u
z1
e
nu
du.
Hence
1
n
z
=
1
(z)
_

0
u
z1
e
nu
du.
Summing over n = 1, 2, . . . , we get the next theorem.
16.1.1. Theorem.
(z) =
1
(z)
_

0
u
z1
e
u
1
du.
Exercises
16.1.2. Exercise. Show that
(z) = (z) (z C).
16.1. The connection between the and functions 121
It is often very useful consider several geometric transformation on the
complex plane. The maybe most important class of the transformations are
the angle preserving ones, they called conformal maps. In this part we intro-
duce the notion of conformal maps, and we study one class of these transfor-
mations, the Mbius transformations. Such transformations are extremely
useful not just in complex function theory but in physics as well.
17
Conformal maps
We begin with some well known transformations, like translation and rota-
tion, and then we dene conformal maps, which are the angle preserving
extension of the above mentioned ones.
17.1 Some examples of transformations on the
complex plane
Let us take the transformation
f : C C; f(z) = z + a,
where a C is a xed constant. It is easy to see that the function (transfor-
mation) f is one to one and onto, and it acts as a translation by the vector
a.
The next transformation
g : C C; f(z) = e
i
z
is a rotation with the xed R as the angle of the rotation. To see this,
it is better to write z in polar coordinates:
z = [z[e
it
.
Then the action of the transformation g on z is as follows:
g(z) = e
i
[z[e
it
= [z[e
i(t+)
,
which shows that we wanted to demonstrate.
Now let us torn to a less picturesque transformation:
h : C 0 C; f(z) =
1
z
.
122
17.1. Some examples of transformations on the complex plane 123
What we can immediately see, is that this transformation maps the unit disk
to its complement, and its complement to the unit circle. (This can be proven
by taking the absolute values.) More generally, h maps circles to circles wit
reciprocal radius.
To get more information on this map, let us determine the images of the
real and imaginary axes under the transformation h: the imaginary axis can
be paramtrized by
t + i 0 (t R),
so
h(t + i 0) =
1
t
+ 0 i.
The image is again the real line, except the origin. The origin is mapped to
innity. What is interesting is that the line segments ]0, 1] is sent to [1, +[,
and [1, 0[ is sent to ] , 1].
Similarly, for the imaginary axis we have the standard parametrization
0 + i t (t R),
and therefore
h(0 + i t) =
1
it
= i
1
t
.
This shows that we can say the same on the image of the imaginary line. It
is not hard to see what happens with an arbitrary line crossing the origin: if
z = te
i
,
where t R and is xed, then
h(z) =
1
t
e
i
,
what means that the original line is mirrored to the real axis, and the dierent
line segments are transformed as before. As before, the origin is sent to
innity.
Altogether, h maps circles to circles and lines crossing the origin to lines.
(That also can be seen that, in general, lines are mapped onto lines or circles.)
17.1.1. Denition. The above dened map h is called inversion.
Exercises
17.1.2. Exercise. Describe the tranformation
f : C C; f(z) = z.
124 17. chapter. Conformal maps
17.1.3. Exercise. Prove that the above described transformation
h : C 0 C; h(z) =
1
z
maps every line to a line or a circle. Especially, prove that the vertical line
z = c +it (c is xed and t R) has the circle with centre (c/2, 0) and radius
c/2 as its image.
17.2 Conformal mappings
17.2.1. Denition. The maps that preserves angles are called conformal .
In other words, if we have a conformal transformation f and two arbitrary
point on the plane (now it is better to consider them as vectors) z
1
and z
2
such that the angle between them is , then the angle between f(z
1
) and
f(z
2
) is also . The examples of the previous section are all conformal. the
proofs are obvious, except for the inversion.
17.2.2. Proposition. The inversion is a conformal map.
Proof. Since
1
z
=
1
[z[
2
z,
we can see that the image of point is nothing else just its mirror image with
respect to the real axis (this is conformal) and then a scaling by the factor
1
|z|
2
, which is again conformal.
17.3 The Mbius transformations
There is a very important class of conformal mappings: the Mbius trans-
formations.
17.3.1. Denition. Let a, b, c, d C be xed such that ad bc ,= 0. The
transformation
f(z) : C C; f(z) =
az + b
cz + d
is called Mbius transformation.
We could think that one could exclude z = d/c from the domain of f.
In what follows we shall consider the image of the point z = d/c as a point
at the innite.
The above properties are the most basic and most important ones of
Mbius transformations.
17.3. The Mbius transformations 125
17.3.2. Theorem. Let f be an arbitrary Mbius transformation. Then
1. f can be expressed as a composition of an ane transformation and an
inversion.
2. f is conformal.
3. f maps circles and lines to circles and lines.
4. f is one to and and onto (except the only one singular point mentioned
above)
5. f is continuous (except the only one singular point mentioned above),
Proof. Let us write f as
f(z) =
az + b
cz + d
=
a
c
+
b ad/c
cz + d
.
Then we see that f is a composition of a scaling cz then a translation cz +d,
then an inversion 1/(cz + d) which is followed by a scaling with the factor
b ad/c. Then nally a translation comes with the real vector a/c. This
gives the rst point.
The second and third points follow from the rst one.
The last two points are straightforward.
From th rst point of the proof we can see why we excluded the case
ad bc = 0 from the denition: in this case our transformation would be
constant.
17.3.3. Example. Let us consider the Mbius transformation
f(z) =
z
z 1
.
Determine the set which has 1/2 + it as an image.
We know that the solution will be a circle or a line. To get this set, we
have to solve the equation
z
z 1
=
1
2
+ it.
It easily follows that
z =
1 + 2it
2it 1
.
It can easily be seen that this can be converted to
z =
2t i
2t + i
as well. Anyway, the absolute value of z is 1. Hence, we get that the unit
circle is mapped to the line 1/2 + it by f.
126 17. chapter. Conformal maps
17.3.4. Example. Let us nd the image of the interior of the circle with
centre 2 and radius 2 under the map
f(z) =
z
2z 8
.
A circle is determined by three points on the plane, so it is enough to
determine the image of three points on the circle to get the result. These
three points can be 0, 4, 2 + 2i. We have that
f(0) = 0,
f(4) =
4
0
,
f(2 + 2i) =
2 + 2i
4 + 4i
=
1
2
1 + i
2 + 2i
=
1
2
i.
Hence the image cannot be else but the imaginary line.
But this is the boundary, we have to nd the image of the interior. The
image of the interior is a half plane. Since the boundary of the disk has the
imaginary axis as an image, the resulting half plane is the left or the right
one. Taking the center of the circle, we have that
f(2) = 2/(4 8) =
1
2
.
Hence we get that
z C [ [z 2[ = 2
f
z C [ 1(z) < 0..
Exercises
17.3.5. Exercise. Construct a Mbius map which sends the unit disk to
the right half plane.
17.3.6. Exercise. Prove that if an analytic function is bounded on the left
half plain such that it is zero in the negative integer points, then this function
is constant.

Potrebbero piacerti anche