Complex Analysis Christer Bennewitz
Complex Analysis Christer Bennewitz
Complex Analysis Christer Bennewitz
Fall 2006
Christer Bennewitz
Preface
These notes are basically a printed version of my lectures in complex
analysis at the University of Lund. As such they present a limited view
of any of the subject matters brought up, caused by the time constraints
one is faced by in a series of lectures. The core of the subject, presented
in Chapter 3, is very strongly influenced by the treatment in Ahlfors
Complex Analysis, one of the genuine masterpieces of the subject. Any
reader who wants to find out more is advised to read this book.
Mathematical prerequisites are in principle the mathematics courses
given in the first two semesters in Lund. Most importantly, this includes a reasonably complete discussion of analysis in one and several
variables and basic facts about series of functions including absolute
and uniform convergence. A course in topology is also useful, but not
essential. Primarily, a familiarity with the concept of a connected set
is of use.
Egevang, August 2006
Christer Bennewitz
Contents
Preface
1
1
4
5
7
10
16
23
23
27
33
Chapter 3. Integration
3.1. Complex integration
3.2. Goursats theorem
3.3. Local properties of analytic functions
3.4. A general form of Cauchys integral theorem
3.5. Analyticity on the Riemann sphere
37
37
40
45
48
51
Chapter 4. Singularities
4.1. Singular points
4.2. Laurent expansions and the residue theorem
4.3. Residue calculus
4.4. The argument principle
53
53
56
59
65
73
73
79
85
85
87
90
94
96
101
iv
CONTENTS
105
CHAPTER 1
Complex functions
1.1. The complex number system
Recall that a group (G, ) is a set G provided with a binary operation 1 satisfying the following properties:
(1) For all elements x, y and z G holds (x y) z = x (y z).
(associative law)
(2) There exists a neutral element e G with the properties xe =
e x = x for every x G.
(3) Every element x G has an inverse x1 with the properties
x x1 = x1 x = e.
Exercise 1.1. Show that a set provided with an associative binary
operation can have at most one neutral element.
Hint: Show that if the set has a left neutral element and a right
neutral element, they must coincide.
Exercise 1.2. Show that if a set has an associative binary operation with neutral element, then any element of the set has at most one
inverse.
Hint: Show that if an element has a left inverse and a right inverse,
then these must coincide.
A group may also have the property
(4) For all elements x and y G holds x y = y x. (commutative
law)
in which case the group is called commutative or Abelian (after Niels
Henrik Abel (18021829)). Familiar examples of Abelian groups are
(Z, +), the integers under ordinary addition; (R, +), the real numbers under addition; (Rn , +), the set of n-tuples of real numbers under
(vector) addition; and (R \ {0}, ), the non-zero real numbers under
multiplication. As an example of a non-Abelian group, consider the
set of all rotations around lines through the origin in 3-dimensional
space; the binary operation is the ordinary composition of maps. The
reader should check these examples carefully; in particular, find the
neutral elements and inverses in these groups.
1That
1. COMPLEX FUNCTIONS
1. COMPLEX FUNCTIONS
1+2i
3+4i
on standard form.
q
x = 1 ( u2 + v 2 + u),
q2
(1.2)
y = 1 ( u2 + v 2 u).
2
Note that all the expressions within square roots are non-negative no
matter what u and v are, so these are ordinary real square roots. (1.2)
therefore give all possible solutions of (1.1), and it is easily verified that
1. COMPLEX FUNCTIONS
that no choice of branch will be suitable for all curves. Since there is
no choice of branch
which will work best in all situations one must not
use the notation without specifying which branch of the square root
one is talking about.
The need to deal with several different branches occurs for all kinds
of other complex functions and is a major complicating factor in the
theory. There is a sophisticated and completely satisfactory solution
to the problem, namely the introduction of the concept of a Riemann
surface. Unfortunately we can not go into that here.
1. COMPLEX FUNCTIONS
2 Re z
x1 = 2
|z| + 1
2 Im z
x2 = 2
.
|z| + 1
|z|2 1
x3 = 2
|z| + 1
(|z|2
|z w|
.
+ 1)1/2 (|w|2 + 1)1/2
2
.
(|z|2 +1)1/2
Theorem 1.13. The image of a straight line in C under stereographic projection is a circle through N , with N excluded. The image
of a circle in C under stereographic projection is a circle not containing
N . The inverse image of any circle on S2 is a straight line together
with if the circle passes through N , otherwise a circle.
Proof. Since a straight line in the x1 x2 -plane together with N
determines a unique plane, the intersection of which with S2 is the
image of the straight line we only need to consider the case of a circle
in C. If it has center a and radius r its equation is |z a|2 = r2 or
1 +ix2
|z|2 2 Re(az) + |a|2 = r2 . Substituting z = x1x
into this, using that
3
2
2
2
x1 + x2 + x3 = 1 and x3 6= 1, we get 1 + x3 2x1 Re a 2x2 Im a +
(1 x3 )(|a|2 r2 ) = 0 which is the equation of a plane. Conversely,
a circle on the Riemann sphere is determined by three distinct points.
The inverse images of these three points determine a circle in C. The
image of this circle is clearly the original circle.
10
1. COMPLEX FUNCTIONS
The circles intersect also at the image of z on the sphere, and are
tangent to the images of the curves there. The angles at the two points
where the circles intersect are equal but of opposite orientation by
symmetry (the two angles are images of each other under reflection in
the plane through the origin and parallel to the normals of the planes
of the circles). The theorem now follows.
Although the proof above is very geometric in nature it is actually not difficult to make it analytic, using the fact that stereographic
projection is a differentiable map, but we will not do that here.
1.5. Mbius transforms
A Mbius transform (also called a linear fractional transformation)
az + b
is a non-constant mapping of the form z 7 f (z) =
for complex
cz + d
numbers a, b, c and d. To begin with we consider this defined in C
except, if c 6= 0, for z = d/c. The fact that the mapping is nonconstant means that (a, b) is not proportional to (c, d). This can be
expressed by requiring ad bc 6= 0 which is always assumed from now
on. Clearly we get the same mapping if we multiply all the coefficients
a, b, c, d by the same non-zero number so that although the mapping
is determined by the matrix ( ac db ) any non-zero multiple of this matrix
gives the same mapping. The requirement ad bc 6= 0 means that
the determinant is 6= 0 so multiplying by an appropriate number we
may always assume that the determinant is 1. This determines the
coefficients up to a change in sign of all of them.
It is clear that if c = 0, then f (z) as z . On the other
hand, if c 6= 0, then f (z) as z d/c and f (z) a/c as z .
We may therefore extend the definition of f to all of the extended plane
C in such a way that the extended function is a continuous function
of C into C . We will always consider Mbius transforms as defined
in the extended plane, or equivalently on the Riemann sphere, in this
way. We have the following interesting proposition.
Proposition 1.15. If f and g are Mbius transforms corresponding to the matrices A and B, then the composed map f g is a Mbius
transform corresponding to the matrix AB.
Exercise 1.16. Prove Proposition 1.15.
Since the set of all non-singular 2 2 matrices is a group under
matrix multiplication, it follows that so are the Mbius transforms.
This means that any Mbius transform has an inverse which is also a
Mbius transform.
Exercise 1.17. Find all Mbius transforms T for which T 2 = T .
11
Among other things this means that a Mbius transform is a homeomorphism of the extended plane onto itself, i.e., a continuous oneto-one and onto map whose inverse is also continuous. But Mbius
transforms have more surprising properties. Recall that we by a circle
in the extended plane mean either an actual circle in the plane or a
straight line together with .
Theorem 1.18. Mbius transforms are conformal and circle-preserving, i.e., any circle in the extended plane is mapped onto a circle
in the extended plane.
Proof. The theorem is obvious for certain simple special cases,
namely a translation z 7 z + b, a rotation z 7 az where |a| = 1
and a dilation z 7 az where a > 0. It is therefore also true for a
multiplication z 7 az where 0 6= a C, since this is the composite
a
of the rotation z 7 |a|
z and the dilation z 7 |a|z. Composing a
multiplication with a translation the theorem follows for a linear map
z 7 az + b where a 6= 0, and hence for any Mbius transform for which
1
c = 0. If c 6= 0 we have az+b
= bcad
+ ac so that this map is the
cz+d
c cz+d
composite of three maps, the first and last being linear and the middle
map is the inversion z 7 1/z. The theorem therefore follows if we can
prove it for an inversion.
If the image of z under stereographic projection is (x1 , x2 , x3 ), then
3
2
1 +ix2
so that 1/z = x1x
= (1 x3 ) xx12ix
we have z = x1x
Since
2 .
3
1 +ix2
1 +x2
(x1 , x2 , x3 ) is on the unit sphere we have x21 + x22 = 1 x23 so that
1 ix2
. Therefore 1/z is the inverse stereographic projection of
1/z = x1+x
3
(x1 , x2 , x3 ). The map that takes (x1 , x2 , x3 ) into this is a rotation
around the x1 -axis by an angle . This is obviously a circle-preserving
and conformal map, and since we know that also stereographic projection is circle-preserving and conformal it follows that the inversion has
the desired properties. The proof is now complete.
Exercise 1.19. Prove the theorem by calculation, not using stereographic projection.
Note that since removing a circle from the extended plane leaves
a set with exactly two components, and since Mbius transforms are
continuous in the extended plane, the interior of any circle in the plane
is mapped either onto the interior or onto the exterior, including ,
of another circle. This follows from the fact that continuous maps
preserve connectedness.
Exercise 1.20. Prove the statement above in detail.
Sets that are left invariant under a mapping are obviously important
characteristics of the map. For a Mbius transform one may for example ask which circles it leaves invariant, or conversely, which Mbius
transforms leave a given circle invariant. We will consider some such
12
1. COMPLEX FUNCTIONS
problems later. Right now we will instead ask for fixpoints of a given
transform, i.e., points left invariant by the map. By our definition of
the image of , this is a fixpoint if and only if the map is linear. A
linear map z 7 az + b also has the finite fixpoint z = b/(1 a), except
if a = 1. Thus, a translation which is not the identity has only the
fixpoint , but any other linear map which is not the identity has
exactly one finite fixpoint as well. For a Mbius transform z 7 az+b
cz+d
with c 6= 0 the equation for a fixpoint becomes z(cz + d) = az + b which
is a quadratic equation. It therefore has either two distinct roots or a
double root. We have therefore proved the following proposition.
Proposition 1.21. A Mbius transform different from the identity
has either one or two fixpoints, as a map defined on the extended plane.
Exercise 1.22. Find the fixed points of the linear transformations
z
2z
3z 4
z
w=
, w=
, w=
, w=
.
2z 1
3z 1
z1
2z
In particular, a Mbius transform that leaves three distinct points
invariant is the identity. It also follows that there can be at most
one Mbius transform that takes three given, distinct points into three
specified, distinct points. Because, if there were two, say f and g, then
f 1 g would be a transform different from the identity and leaving the
given points invariant. Conversely, we will prove that there actually
always exists a Mbius transform that takes the given points into the
specified ones. To see this, define the cross ratio of four distinct points
z0 , z1 , z2 , z3 in C by
z0 z2 z1 z2
(z0 , z1 , z2 , z3 ) =
/
z0 z3 z1 z3
when all the points are finite. If one of them is , the cross ratio is
defined as the appropriate limit of the expression above. The following
proposition follows by inspection.
Proposition 1.23. Suppose z1 , z2 , z3 are distinct points in C . The
unique Mbius transform taking these points to 1, 0, in order is z 7
(z, z1 , z2 , z3 ).
It is now clear that to find the unique Mbius transform taking the
distinct points z1 , z2 , z3 into the distinct points w1 , w2 , w3 in order, one
simply has to solve for w in (w, w1 , w2 , w3 ) = (z, z1 , z2 , z3 ).
Exercise 1.24. Find the Mbius transformation that carries 0, i,
i in order into 1, 1, 0.
Exercise 1.25. Show that any Mbius transformation which leaves
R {} invariant may be written with real coefficients.
Exercise 1.26. Show that the map z 7 z1
maps the right halfz+1
plane (i.e., the set Re z > 0) onto the interior of the unit circle.
13
(z a)(z a) = R2 .
14
1. COMPLEX FUNCTIONS
In particular the fact that the center of a circle and are symmetric
with respect to the circle are often very helpful in trying to find maps
that take a given circle into another.
Exercise 1.30. Find the Mbius transform which carries the circle
|z| = 2 into |z + 1| = 1, the point 2 into the origin, and the origin
into i.
Exercise 1.31. Find all Mbius transforms that leave the circle
|z| = R invariant. Which of these leave the interior of the circle
invariant?
Exercise 1.32. Suppose a Mbius transform maps a pair of concentric circles onto a pair of concentric circles. Is the ratio of the radii
invariant under the map?
Exercise 1.33. Find all circles that are orthogonal to |z| = 1 and
|z 1| = 4.
We will end this section by discussing conjugacy classes of Mbius
transforms.
Definition 1.34. Two Mbius transforms S and T are called conjugate if there is a Mbius transform U such that S = U 1 T U .
Conjugacy is obviously an equivalence relation, i.e., if we write
S T when S is conjugate to T , then we have:
(1) S S for any Mbius transform S.
(reflexive)
(2) If S T , then T S
(symmetric)
(3) If S T and T W , then S W .
(transitive)
It follows that the set of all Mbius transforms is split into equivalence
classes such that every transform belongs to exactly one equivalence
class and is equivalent to all the transforms in the same class, but to
no others.
Exercise 1.35. Prove the three properties above and the statement
about equivalence classes. What are the elements of the equivalence
class that contains the identity transform?
The concept of conjugacy has importance in the theory of (discrete)
dynamical systems. This is the study of sequences generated by the
iterates of some map, i.e., if S is a map of some set M into itself,
one studies sequences of the form z, Sz, S 2 z, . . . where z M . This
sequence is called the (forward) orbit of z under the map S. One
is particularly interested in what happens in the long run, e.g., for
which zs the sequence has a limit (and what the limit then is), for
which zs the sequence is periodic and for which zs there seems to be
no discernible pattern at all (chaos). Note that if S = U 1 T U , then
S n = U 1 T n U so that all maps in the same conjugacy class behave
qualitatively in the same way, at least with respect to the properties
15
z + 1 for = 1 ,
T z =
z for 0 6= 6= 1 .
We have then proved most of the following theorem.
Theorem 1.36. For every Mbius transform S different from the
identity there exists 6= 0 such that S T . If T T , then either
= or = 1/.
Proof. It only remains to prove the last statement. But this is
clear if = 1, since this is the only value for which T has just one
fixpoint. We may therefore assume that and are both 6= 1 (and of
course non-zero). But if U T = T U and U z = az+b
we obtain
cz+d
az + b
az + b
=
cz + d
cz + d
for all z. Since ad bc 6= 0 we can not have d = c = 0. If d 6= 0, setting
z = 0 gives b/d = b/d so that b = 0 and therefore a 6= 0. If now c 6= 0,
setting z = d/c we get on the right but not the left. It follows
that c = 0 and (5.2) becomes = . On the other hand, if d = 0 we
must have c 6= 0 and so z = gives a/c = a/c. It follows that a = 0.
In this case (5.2) becomes = 1/ and the proof is complete.
(1.3)
16
1. COMPLEX FUNCTIONS
17
18
1. COMPLEX FUNCTIONS
X
(1.4)
an (z a)n
n=0
where a, a0 , a1 , a2 , . . . are given complex numbers and z a complex variable. In many respects such series behave like polynomials of infinite
order and that is actually how they were viewed until the end of the
19:th century. The very first question to ask is of course: For which
values of z does the series converge? In order to answer this question
we make the following definition.
Definition 1.43. Let the radius of convergence for (1.4) be
R = sup{r 0 | a0 , a1 r, a2 r2 , . . . is a bounded sequence } .
Then R is either a number 0 or R = .
The explanation for the definition is in the following theorem.
Theorem 1.44. For |z a| > R the series (1.4) diverges and for
|z a| < R it converges absolutely. The convergence is uniform on
every compact subset of |z a| < R.
In order to prove the theorem we need a few results which should
be well known in the context of functions of a real variable.
Theorem 1.45. An absolutely convergent complex series is convergent.
Proof. For any complex number z we
P have | Re z| |z| and
| Im z| |z| | Re z| + | Im P
z|. Hence, if P |an | is convergent, then
by comparison the real series
Re an and
Im an are absolutely convergent, to x and y say. The theorem now follows from
|
N
X
n=0
an x iy| |
N
X
n=0
Re an x| + |
N
X
Im an y| 0 as N .
n=0
19
N
X
n=0
fn (z)| = |
X
n=N +1
fn (z)|
X
n=N +1
|fn (z)|
an .
n=N +1
20
1. COMPLEX FUNCTIONS
Proof. Let L = lim |an |1/n . If r < 1/L, then |an |1/n < 1/r for all
n
21
X
X
2 2
(a)
nz
(b)
nn z n
(c)
n=0
X
n=0
z
(n!)2
X
zn
(e)
ln n
n=2
(g)
(i)
(k)
X
n=1
X
n=0
(d)
z
arctan n
(h)
cos(n/4)z n
(j)
(l)
n=1
(s)
(u)
(w)
(y)
()
n=0
X
n=1
n=0
X
n=1
(n)
(n2 + 2n )z n
2n + 2n n
z
n
(2n + (2)n + 1)z n
2n
zn
zn
1 + 2 + + n
zn
(r)
(t)
X
n=1
n=0
n=1
n=1
(v)
n=0
(sin n)z n
n=0
(arctan n1 )z n
n=1
X
2n n
z
n
n n
32n+1
zn
X
2n
z 2n
(p)
n
n=1
X
zn
n=0
zn
X
(n + 2)3
n=0
X
21/n
n=1
()
n2
n=0
n=0
(q)
X
2n
n=1
n1/n z n
2n z n
n=0
(f)
(m)
( n2 + 1 n2 1)z n
(o)
n=0
(1 + 21 + + n1 )z n
2
zn
n2
z 4n
2n + n2
X
nn
zn
n!
n=1
X
a n
(z)
z
n
n=0
(x)
()
X
(3n + 1)!
n=0
(h)
X
n=1
(n!)4
2
zn
(1 + n1 )n z n
CHAPTER 2
Analytic functions
2.1. Conformal mappings and analyticity
Definition 2.1. A map f : C, where is an open subset of
C, is called conformal if it satisfies the following conditions:
(1) As a map from a subset of R2 into R2 , f is differentiable.
(2) f preserves angles of intersection between smooth curves.
(3) f preserves orientation in the sense that the determinant of
the total derivative of the map is > 0.
To explain the definition in more detail, note that if z = x + iy,
where x and y are real, then f (z) = u(x, y) + iv(x, y) where u and v are
real-valued functions of two real variables,
24
2. ANALYTIC FUNCTIONS
25
z is differentiable for
a b
b a satisfying the Cauchy-Riemann equations means that
u(x + h, y + k) u(x, y)
a
b
h
= k(h, k)kk(h, k)k
v(x + h, y + k) v(x, y)
b a
k
where (h, k) 0 as (h, k) 0. But if f = u + iv and w = h + ik
the left hand sides of these two relations are equal so the theorem
follows.
26
2. ANALYTIC FUNCTIONS
27
f+
f=
a
Zb
Zc
Zb
Zb
(f + g) =
Zb
f +
g
a
for a < c < b and arbitrary constants and if f and g are both
integrable on [a, b]. We also have the triangle inequality
(2.2)
Zb
Zb
28
2. ANALYTIC FUNCTIONS
Rb
by choosing = arg( a f ).
As already mentioned we also have the chain rule dtd f (g(t)) =
0
f (g(t))g 0 (t) if f is analytic and g is a differentiable complex-valued
function of a real variable. Thus, if f is analytic in a region containing
the line segment connecting z and z + h, then dtd f (z + th) = hf 0 (z + th)
R1
so that h1 (f (z + h) f (z)) = 0 f 0 (z + th) dt if the derivative is continuous. An immediately consequence is the following lemma.
Lemma 2.12. Suppose f is analytic with continuous derivative in
a compact set K containing the line segment connecting z and z + h
where h 6= 0. Then we have | h1 (f (z + h) f (z)| supK |f 0 |.
Proof. By the triangle inequality we obtain
f (z + h) f (z) Z1
Z1
X
g(z) =
kak (z a)k1
k=1
P
k
has
the
same
radius
of
convergence
as
k=1 kak (z a) and since
k
k 1 as k , it follows from Theorem 1.50 that g has the same
radius of convergence as
Pf .
k1
converges, and
If r < R the series
k=1 k|ak |r
X
1
ak
(f (z + h) f (z)) =
((z + h a)k (z a)k ).
h
h
k=1
29
Now fix z, |z a| < r. Then the terms of this series are continuous
functions of h, with value kak (z a)k1 for h = 0. By Lemma 2.12
the terms have absolute value less than k|ak |rk1 if |z a| < r and
|z + h a| r, so according to Theorems 1.46 and 1.47 the sum is
a continuous function of h in |z + h a| r. For h 6= 0 its value
is h1 (f (z + h) f (z)) and for h = 0 the value is g(z). Thus f is
differentiable and f 0 (z) = g(z) for any z satisfying |z a| < R.
30
2. ANALYTIC FUNCTIONS
cos2 z + sin2 z = 1
31
One may easily continue to define strictly all the usual (real) functions of elementary calculus and prove all the usual properties of them.
We will assume this done; in particular x is the arclength of the arc of
the unit circle beginning at 1 and ending at eix = cos x + i sin x, so x is
the angle between the rays through these points starting at 0. We will
also use the common properties of the inverse tangent function.
If we want to extend the definition of the logarithm to the complex domain, we should find the inverse of the exponential function.
However, since the exponential function is periodic it has no inverse
unless we restrict its domain appropriately (cf. the definition of the
inverse trigonometric functions). To see how to do this, let us attempt
to calculate the inverse of the exponential function, i.e., to solve the
equation z = ew for a given z.
We first note that we must assume z 6= 0, since the exponential
function never vanishes. Taking absolute values we find that |z| = eRe w
so that Re w = ln |z|, where ln is the usual natural logarithm of a
z
positive real number. We also obtain |z|
= ei Im w . Now cos 0 = 1 and
cos = 1 and since cos is continuous, it takes all values in [1, 1] in
z
the interval [0, ]. Since Re |z|
[1, 1] we can find x [0, ] such that
z
z
cos x = Re |z| . It follows that sin x = Im |z|
. Changing the sign of x
changes the sign on sin x but leaves cos x unchanged. Therefore either
z
eix or eix equals |z|
.
We may therefore solve the equation for w given any z 6= 0. If w1
and w2 are two solutions, it follows that ew1 w2 = 1, so that w1 and
w2 differ by an integer multiple of 2i. We call any permissible value
of Im w an argument for z, and denote any such number by arg z.
We should therefore define log z = ln |z| + i arg z. To get an actual
(single-valued) function, we must make particular choices of arg z for
each z. We shall see later that in order to be able to this and obtain
a continuous function, we can not allow all of C \ {0} in the domain.
Intuitively it is clear that we must choose the domain such that there
are no closed curves in it that go around the origin, since following
such a curve we would have changed the argument continuously by an
32
2. ANALYTIC FUNCTIONS
33
d log z
e
dz
34
2. ANALYTIC FUNCTIONS
|z + 3| < 10,
|z 2| < 5,
onto the interior of the first quadrant.
Exercise 2.23. Map the region
0 < arg z < / ,
|z 1| > 1,
|z| < 2,
onto the upper half plane.
Exercise 2.26. Construct a conformal map of the region
(R > 1)
onto the interior of the unit circle, so that z = 1 is mapped onto the
origin. Calculate the length of the arc of the unit circle which is the
image of the arc
/4 arg z /4 ,
|z| = R .
Exercise 2.27. What is the image of the unit disk under the map
w = F (z) = (z + 1/z)/2, F (0) = .
Hint: Introduce W = w1
and Z = z1
.
w+1
z+1
35
Exercise 2.28. Map the region a < arg z < b, where 0 < a <
b < 2, conformally onto the complex plane with the positive real axis
removed.
Exercise 2.29. Find a conformal map that takes
|z 1 i| < 2 ,
0
=
i(z z) > 2 .
Exercise 2.30. Consider the conformal map given by cos z. What
are the images of lines parallel to the real and imaginary axes? What
is the image of the strip < Re z < ?
CHAPTER 3
Integration
3.1. Complex integration
Complex integration is at the core of the deeper facts about analytic
functions. Here we will discuss the basic definitions.
Let be a piecewise differentiable curve in C. This means a complexvalued, continuous function defined on a compact real interval which is
continuously differentiable except at a finite number of points, where
at least the left and right hand limits of the derivative exist. Thus it
is described by an equation z = z(t) where a t b for some real
numbers a and b and z 0 is continuous except for a finite number of jump
discontinuities. For convenience we will in the sequel call such a curve
an arc.
If f is a continuous, complex-valued function of a complex variable
defined on an arc , then the composite function f (z(t)) is continuous
and we make the following definition.
Z
Definition 3.1.
Zb
f (z(t))z 0 (t) dt.
f (z) dz =
If you know
R about line integrals and f = u + iv, z = x + iy you will
realize that f (z) dz is the line integral
Z
Z
u dx v dy + i v dx + u dy,
but we will not use this. It is, however, very important that the
complex integral is independent of the parametrization of the arc .
This means the following. A change of parameter is given by a piecewise differentiable, increasing function t(s) mapping an interval [c, d]
onto [a, b]. The usual change of variables formula then shows that
R
Rd
f (z) dz = c f (z(t(s)))z 0 (t(s))t0 (s) ds. Here z 0 (t(s))t0 (s) is, by the
chain rule, the derivative of z(t(s)), so that the definition of the complex integral gives the same value whether we parametrize by z(t) or
z(t(s)).
Note that the arc has an orientation, in that it begins at z(a) and
ends at z(b). If t(s) is a decreasing piecewise differentiable function,
mapping [c, d] onto [a, b], then the equation z = z(t(s)) will give a
parametrization of the opposite arc to , which we denote by , in
37
38
3. INTEGRATION
that the initial pointRis now z(t(c)) =Rz(b) and the final point z(t(d)) =
z(a). Thus we have f (z) dz = f (z) dz.
It is clear by the definition that the integral is linear in f , and also
that if we divide an arc into two sub-arcs 1 and 2 by splitting
the parameter interval into two subintervals with
no common
interior
R
R
points (keeping the correct orientation), then f (z) dz = 1 f (z) dz +
R
f (z) dz. It is now an obvious step to consider the sum of two (or
2
more) arcs 1 and 2 even if they are not sub-arcs of another arc, and
define the integral over such a sum as the sum of the integrals over the
individual terms. Such a formal sum of arcs is called a chain. Given arcs
1 , . . . , n we may integrate over chains of the form = a1 1 + +an n ,
where the coefficients
a1 , . . .R, an are arbitrary integers, indicating that
R
the integral j f enters in f with the coefficient aj . If aj = 0 the
arc j can of course be left out of .
Note that our notation for the opposite of an arc makes sense, in
that integrating over amounts to integrating over (1). Very often
we will integrate over closed arcs. This means an arc where the initial
and final points coincide. A simple arc is one without self-intersections;
for a closed arc this means no self-intersections apart from the common
initial and final point.
There is also a triangle inequality for complex integrals. From the
definition of integral and the triangle inequality (2.2) it immediately
follows that
Z
Z
Rb
where the last integral is defined by |f (z)| |dz| := a |f (z(t)||z 0 (t)| dt
and is called an integral
R with respect
R b 0 to arc-length. The reason for
this is, of course, that |dz| = a |z (t)| dt gives the length of the arc
. If you dont know this already, you may take it as a definition of
length. Note that a very similar calculation to the one we did earlier
shows that an integral with respect to arc length is independent of the
parametrization, and in this case also of the orientation of the arc.
Example 3.2. Suppose is the circle |z a| = r, oriented by
running through it counter-clockwise. A parametrization is z(t) =
a + reit , 0 t 2. We obtain
z 0 (t) = ireit so that |z 0 (t)| = r. The
R 2
length of the circle is therefore 0 rdt = 2r, as expected.
It is possible to integrate along more general curves than those
that are piece-wise differentiable, so called rectifiable curves. There
is seldom any reason to do this in complex analysis, however. In fact,
when integrating analytic functions the integral is, as we shall see later,
independent of small changes in the path we integrate over, so it is
practically always enough to consider piece-wise differentiable arcs.
39
Zb
f (z) dz =
Zb
0
(F (z(t)))0 dt
f (z(t))z (t) dt =
a
b
= F (z(t)) a = F (z1 ) F (z0 ),
40
3. INTEGRATION
f (z) dz = 0.
41
Z
=
Z
f (w) dz +
Z
0
f (w)(z w) dz +
n
(z)(z w) dz.
n
Now, the constant f (w) has primitive f (w)z and the first order polynomial f 0 (w)(z w) has primitive 21 f 0 (w)(z w)2 , so that the first two
integrals in the second line vanish. The third integral is estimated as
follows:
Z
Z
(z)(z w) dz |z w||dz| dn Ln ,
n
f (z) dz = 0.
42
3. INTEGRATION
boundary of R0 we obtain
Z
Z
f (z) dz
0
|dz|
8.
|z p|
The last inequality is due to the facts that |z p| `/2 if ` is the side
length of R0 , and that the length of 0 is 4`.
Now extend the sides of R0 until they cut R into 9 rectangles, one
of which is R0 . RThe other
R 8 satisfy the assumptions of Theorem 3.4.
It follows that | f | = | 0 f | 8, and since > 0 is arbitrary the
integral over must be 0.
f (z) dz = 0 .
43
The conclusion of Corollary 3.6 can not be drawn with weaker assumptions on f at the point p. To illustrate this, let f (z) = 1/(z p)
which is analytic in any disk centered at p, except at z = p, and let be
the positively oriented boundary of such a circle. We may parametrize
by z(t) = p + reit , 0 t 2. Then z 0 (t) = ireit so that
Z2 it
Z2
Z
ire dt
dz
=
= i dt = 2i 6= 0.
(3.2)
zp
reit
Note that the range of is a compact set, being finite union of continuous images of the compact parameter intervals, so its complement
is open. An open set may be split into open, connected components 2.
Clearly there is precisely one unbounded component in the complement
of .
Lemma 3.8. The index has the following properties.
(1) n(, p) is always an integer.
(2) n(, p) = n(, p).
(3) n(1 + 2 , p) = n(1 , p) + n(2 , p) if 1 and 2 are both cycles
not containing p.
(4) n(, p) is constant as a function of p in any connected component of the complement of the range of .
2This
44
3. INTEGRATION
(5) n(, p) = 0 for all p in the unbounded component of the complement of the range of .
Proof. Let z = z(t), a t b, be a parametrization of a closed
R t 0 (s) ds
for t [a, b]. Then g(b) = 2i n(, p) and
arc and set g(t) = a zz(s)p
0
0
g (t) = z (t)/(z(t) p) so that the derivative of h(t) = eg(t) (z(t) p)
is identically 0. We have h(a) = z(a) p, so h is constant equal to
z(a) p. For t = b we obtain eg(b) (z(b) p) = z(a) p. Since is a
closed arc we have z(b) = z(a) 6= p so that eg(b) = 1. Thus g(b) is an
integer multiple of 2i. Since a finite sum of integers is an integer this
proves (1).
(2) and (3) are obvious from the definition, and it is also obvious that n(, p) depends continuously on p
/ (give detailed reasons
yourself!). But a continuous, real-valued function in a region assumes
intermediate values, so since the index is integer-valued, (4) follows.
Finally, it is clear that n(, p) 0 as p , and since n(, p) is
independent of p for p in the unbounded component of the complement
of , (5) follows.
45
k
expand the function in a power series g(z) =
k=0 ak (z p) with
radius of convergence Rat least equal to r. The coefficients in the series
f (z)
1
are given by ak = 2i
dz.
(zp)k+1
zp
Proof. The denominator in the integral is z = ( p)(1 p
).
The reciprocal of this is the sum of a convergent geometric series since
| zp
| < 1, z being closer to p than . A partial sum of this series has
p
the sum
n1
X
k=0
( p)
z p k
p
= ( p)
1 ( zp
)n
p
1
zp
p
1
(z p)n
.
z ( p)n ( z)
|d|
r
2
| z|
46
3. INTEGRATION
The power series is called the Taylor series for f at p and the
formula
Z
n1
X
(z p)n
f () d
k
,
f (z) =
ak (z p) +
2i
( p)n ( z)
k=0
3We
47
p)
.
Since
Theorem 3.11 we may expand f in power series
k=0 k
f (p) = 0 the first term in the series vanishes, and if n is the first index
for whichP
an 6= 0 we obtain f (z) = (z p)n g(z), where g is the analytic
k
function
k=0 an+k (z p) , so that g(p) = an 6= 0. The positive integer
n is called to order or multiplicity of the zero p.
48
3. INTEGRATION
The fact that the only bounded, entire functions are constants is
often very useful. We can for example now give a very simple proof of
the fundamental theorem of algebra.
Theorem 3.17 (Fundamental theorem of algebra). Any non-constant polynomial has at least one zero.
Proof. Suppose P is a polynomial without zeros. Then 1/P (z) is
an entire function, and we shall see that it is bounded, so that Liouvilles theorem will show it to be constant.
If P (z) = an z n + an1 z n1 + + a0 with an 6= 0 we may write
P (z) = z n (an + an1 /z + + a0 /z n ). Here the expression in brackets
tends to an as z . Since z n if n > 0 we have 1/P bounded
for large |z|. Thus P is constant. The theorem follows.
49
is a topological fact
50
3. INTEGRATION
We end this section with a very important consequence of the previous theorem.
Corollary 3.20. Suppose f is analytic and has no zeros in a
simply connected region . Then one may define a branch of log(f (z))
in .
Proof. Since f has no zeros in the function f 0 (z)/f (z) is analytic in so that Cauchys integral theorem applies to it. According
to Theorem 3.3 there is therefore a primitive g of this function defined
0 (z)
d
in , and dz
(f (z)eg(z) ) = f 0 (z)eg(z) f (z) ff (z)
eg(z) = 0, so that
f (z)eg(z) = C, where C 6= 0 since neither f nor the exponential function vanishes. Thus we may find A C so that eA = C. It follows that
f (z) = eg(z)+A , so that g(z) + A is a branch of log(f (z)).
Since one may define a branch of the logarithm one may also define
branches of any power function in . We shall use this in proving the
Riemann mapping theorem in Chapter 7.
Remark 3.21. To obtain a version of Cauchys theorem valid in
arbitrary regions we would have to discuss homology of cycles, and we
will abstain from this. We sometimes have to deal with regions which
are not simply connected, but the cycles we integrate over are then
51
always very simple and explicitly given and therefore never cause any
problem.
For example, suppose f is analytic in a circular ring defined by
0 r0 < |z a| < R0 and suppose r0 < r < R < R0 and let be
the cycle consisting of the two circles |za| = r and |za|
= R, the first
R
negatively and the second positively oriented. Then f (z) dz = 0, and
R f (z) dz
1
for any w satisfying
we also have Cauchys formula f (w) = 2i
zw
r < |w a| < R.
To see this, let 1 be the vertical ray going upwards from a and
2 the opposite ray. Then \ 1 is simply connected. Let 1 be a
positively oriented cycle obtained by taking the part of in the set
Im z Im a and connecting
R the pieces by radial line segments. Then
Theorem 3.19 shows that 1 f = 0.
Similarly, the region \ 2 is simply connected, and if 2 is the
part of in Im z Im a, made into a positively
oriented cycle by
R
adding radial line segments, we also have 2 f = 0. But = 1 + 2 ,
since the radial line segments will be run through twice and in opposite
directions.
It is also easy to see that if Im w > Im a, then n(1 , w) = 0 and
n(2 , w) = 1 so that n(, w) = 1. The reader should modify the construction for other locations of w to see that n(, w) = 1 as soon as
r < |w a| < R.
3.5. Analyticity on the Riemann sphere
Viewing analytic functions as defined on the Riemann sphere, where
all points including look the same, one should be able to define
analyticity at infinity. This leads to the following definition.
Definition 3.22. Suppose f is analytic in a neighborhood |z| >
r > 0 of Then we say that f is analytic at if z 7 f (1/z), which
is analytic in 0 < |z| < 1/r, extends to a function analytic also at 0.
Similarly, if f is analytic in a neighborhood of a and f (z) as
z a it would be tempting to say that f is analytic at a if 1/f (z)
extends to a function analytic at a. We will not use this terminology
since it may lead to confusion, but it is a perfectly reasonable point of
view. In fact, in the next section we will show that if f (z) as
z a, then 1/f (z) always has an analytic extension to a.
Any Mbius transform is in this sense analytic everywhere on the
Riemann sphere, and the reader should should carry out the simple
verification, and also show that the same is true for any rational function.
CHAPTER 4
Singularities
4.1. Singular points
An isolated singularity of a complex function f is a point a such
that it has a neighborhood O with f analytic in O \ {a} (a so called
punctured neighborhood of a). In some cases a is an isolated singularity
simply because we do not know that f is analytic there, or that f is
not analytic at a but will become so provided we assign the correct
value to f (a). In that case a is said to be a removable singularity for
f . A typical example would be z 7 sinz z which is not defined at 0,
but where it is clear from the power series expansion of sin z that the
function becomes entire once we assign it the value 1 at the origin. The
main fact about removable singularities is contained in the following
theorem.
Theorem 4.1. Suppose that f is analytic in a punctured neighborhood of a. Then a is a removable singularity for f if and only if
(z a)f (z) 0 as z a.
Thus the singularities we allowed in Corollaries 3.4, 3.6 are actually
removable, and may be ignored.
Proof. The only if part of the theorem is trivial, since in that
case f must have a finite limit at a. To prove the other direction, let
and be the positively oriented boundaries of disks centered at a
and such that f is analytic in the punctured disks. If is the smaller
disk f is analytic in the ring-shaped region between and , so by
Remark 3.21
Z
Z
1
1
f ()
f ()
f (z) =
d
d
2i
z
2i
z
54
4. SINGULARITIES
that |( a)f ()| < if | a| < . Then, if the radius of is r < and
r < |z a|/2, we obtain | z| |z a|| a| = |z a|r |z a|/2
so that
Z
Z
f ()
4
|d|
z d | a|| z| |z a| .
n
power series expansion around a of (z a) f (z) is k=0 ak (z a)k it
follows that
1
X
X
k
bk (z a)k ,
bk (z a) +
f (z) =
k=n
k=0
where bk = an+k . The first sum above is called the singular part of f
at a. Note that the singular part is analytic everywhere (even at )
except at a. Therefore, if we subtract the singular part from f we get a
function which is analytic wherever f is, and also at a. Subtracting the
singular part at a therefore removes the singularity at a. The fact that
the singular part, in this case, consists of a finite sum of very simple
functions makes this type of singularity rather harmless. It is called a
pole of order n.
A pole of order n > 0 is characterized by the fact that (z a)n f (z)
has a non-zero limit as z a, just as a zero of order n is characterized
by the fact that (z a)n f (z) has a non-zero limit as z a. Note
that f (z) as z approaches a pole so that 1/f has a removable
singularity there. We may therefore view a pole as a point where f
is analytic with the value ; this agrees completely with our point
of view when we discussed functions analytic on the Riemann sphere.
Also note that poles, like zeros, are isolated points. We finally note
that if f has a pole or zero of order |n| at a, then case (1) holds exactly
if > n and case (2) holds exactly if < n.
Now let us consider case (2). If n is the smallest integer , then
(z a)n f (z) as z a so that (z a)n /f (z) has a removable
55
56
4. SINGULARITIES
|a|=r
f ()
d,
z
|a|=r
then f (z) = f1 (z)+f2 (z) for such values of z. However, by Lemma 3.10
f1 is analytic in |z a| < R. It follows that f2 is analytic in r <
|z a| < R. Actually, f2 is analytic in |z a| > r, even at ,
which is seen similarly to the proof of Lemma 3.10. In fact, setting
57
X
wk+1
f2 (a + 1/w) =
f ()( a)k d,
2i
k=0
|a|=r
P1
k=
ak (z a)k .
X
ak (z a)k ,
f (z) =
k=
58
singularities,
4. SINGULARITIES
Z
f (z) dz = 2i
Similar, even more complicated, formulas hold for higher order poles.
P
k
Proof. According to assumption f (z) =
k=n ak (z a) for z
n
close to a, so g(z) = (z a) f (z) has a removable singularity at a and
a1 is the coefficient of (z a)n1 in the corresponding power series
expansion. But this coefficient is g (n1) (a)/(n 1)! and since g (n1) is
continuous at a the first claim follows. If now q has a simple zero at
za
a, then (z a)p(z)/q(z) = p(z) q(z)q(a)
p(a)/q 0 (a) since q(a) = 0,
q 0 (a) 6= 0. Finally, if q has a double zero at a, then q(z) = (z a)2 q2 (z)
where q2 (a) = q 00 (a)/2 and q20 (a) = q 000 (a)/6, as is easily verified. Hence
59
((z a)2 f (z))0 = (p(z)/q2 (z))0 = (p0 (z)q2 (z) p(z)q20 (z))(q2 (z))2 . Letting z a (4.2) follows, and the final claim is left to the reader to
verify.
The conclusion of all this is that simple poles cause little problem
in determining the residue, whereas higher order poles are considerably
more messy to deal with. In the next section we shall see how one may
use the residue theorem to calculate certain real integrals.
4.3. Residue calculus
In this section we shall see how one may use the residue theorem
to calculate certain real integrals. We will only discuss a few types of
integrals that can be handled; many others exist. R
2
1. Let us first consider an integral of the form 0 (cos , sin ) d.
Here (x, y) is a rational function of two variables with no poles for
(x, y) on the unit circle. If we think of this integral as the result of calculating an integral around the unit circle by the parametrization z = ei ,
0 2, we note that by Eulers formulas we have cos = 21 (z+1/z)
and sin = 2i1 (z1/z). Furthermore, dz = iei d so that d = i dz/z.
R
The integral therefore equals i |z|=1 ( z+1/z
, z1/z
) dz/z. The inte2
2i
grand is rational and it only remains to find those poles that are inside
the unit circle and evaluate their residues.
R 2 d
Example 4.9. Consider the integral 0 a+cos
where a > 1. Set
R
i
ting z = e as above, the integral equals i |z|=1 (a+ 12 (z+1/z))1 dz/z
which after simplification becomes
Z
dz
2i
.
z 2 + 2az + 1
|z|=1
60
4. SINGULARITIES
On the other hand, along the part of which is a half circle we can
estimate the integral by
Z
Z
|dz|
2
= 2 sup |z 2 (z)|/R 0
(z) dz sup |z (z)|
2
|z|
|z|=R
|z|=R
|z|=R
Im z>0
|z|=R
Im z>0
Im z>0
R
2
Example 4.10. Consider the integral x4x+1 dx which satisfies
all the requirements above. The poles of the integrand are given by the
zeros of its denominator, so they are the roots of z 4 + 1 = 0. Setting
z = Rei we easily see that the roots are zk = ei(/4+k/2) , k = 0, 1, 2, 3.
The roots in the upper half plane are the two first ones. Since the zeros
z2
1
are simple ones, the residues are obtained by evaluating 4z
3 = 4z at
these points. It follows that
Z
x2
i/4
3i/4
dx
=
2i(e
+
e
)/4
=
2/2 .
x4 + 1
R
3. We next consider an integral (x)eiax dx where a is real and
a rational function without real poles. This is the Fourier transform of
the function at a. We assume that the degree of the denominator of
is higher than the degree of the numerator. This does not guarantee
absolute convergence of the integral, but as we shall see it does imply
conditional convergence if a 6= 0. In the calculations below we shall
assume that a > 0, but the case a < 0 can be treated very similarly.
This is done either by replacing the upper half plane by the lower half
plane in the considerations below, or else by first making the change of
variable t = x in the integral, which has the effect of replacing a by
a.
Let A, B and C be positive real numbers. We consider a contour
starting with a segment [A, B] of the real axis, continuing with a
vertical line segment from B to B + iC, then a horizontal line segment
from B+iC to A+iC and finally a vertical line segment from A+iC
to A. If A, B and C are sufficiently large, this rectangle will contain
all poles of which are in the upper half plane so that
Z
X
(z)eiaz dz = 2i
Res((z)eiaz ) .
Im z>0
61
provided A > R. Similarly, the integral over the other vertical side
M
may be estimated by aB
provided B > R. Note that these estimates
are independent of C. Assuming that C > R and parametrizing the
upper side of the rectangle by z = t + iC, B t A, we can
similarly estimate the corresponding part of the integral by
ZA
M
M (A + B)
.
eaC dt
| t + iC|
CeaC
M 1
1
iax
iaz
(x)e dx 2i
Res((z)e
( + ).
a A B
Im z>0
A
This shows that the original integral indeed converges, at least conditionally, and that its value is given by the residues in the upper half
plane.
R
Example 4.11. Consider the integral cos(x)
dx, where is real.
x2 +1
First note that the cos function is even. We may therefore replace
by || without affecting the value of the integral. Next, note that the
R ix||
integral is the real part of xe 2 +1 dx which we may evaluate using
the method above, and then take the real part of. Actually, since it is
easily seen that the integrand of the imaginary part is an odd function,
the imaginary part is zero anyway. Note, however, that we can not
evaluate the present integral, or integrals similar to it, by considering
the residues of cos(z)
, since the function cos(z) is large for large | Im z|,
z 2 +1
in both upper and lower half planes.
According to our deliberations above, we have
Z ix||
X
eiz||
e
dx = 2i
Res 2
.
x2 + 1
z +1
Im z>0
In the upper half plane there is only one singularity, a simple pole at
1 iz||
z = i. Thus, the residue is obtained by evaluating 2z
e
at z = i.
e||
This gives 2i so that we obtain
Z
cos(x)
dx = e|| .
x2 + 1
62
4. SINGULARITIES
Since there are no poles in the upper half plane we only need to consider
the integral overR. If we parametrize by z = rei , 0 , the
integral
i 0 exp(irei ) d which tends to i as r 0. It follows
R equals
that sinx x dx = .
R
4. Next consider the integral 0 x (x) dx where 0 < < 1 and is
rational with no positive real poles. For convergence we must assume
that the degree of the denominator in is at least 2 more than the
degree of the numerator. Similarly, we may allow at most a simple
pole at the origin. If we want to relate the value of this integral to the
residues of the function z (z), note that we now have a branch point
at the origin. However, instead of causing difficulties this is actually
what will allow us to evaluate the integral.
Suppose we choose the branch of z where the plane is cut along
the positive real axis and for which 0 < arg z < 2. This means that
z = e log z where log is the corresponding branch of the logarithm. As
we approach a point x > 0 on the real axis from above we then obtain
the usual real power x . However, as we approach the same point from
below we instead obtain e(ln x+i2) = ei2 x .
Intuitively, we would therefore like to choose a contour which
starts at r > 0 on the upper edge of the real axis, continues to R >
r, then follows the circle with radius R and centered at the origin,
counterclockwise, until we reach R on the lower edge of the real axis,
then back to r, still on the lower edge, and finally along the circle
with radius r and centered at the origin, clockwise, until we reach the
initial point again (you will need to draw a picture of this). The two
contributions from integrating along the real axis will not cancel since
the power has different values along the upper and lower edges. The
catch is, there is no such thing as an upper or lower edge of the positive
real axis; in fact, since we have cut the plane along the positive real
axis, we cant integrate along it at all.
63
The problem can be avoided in the following way. First cut the
plane along some ray from the origin other than the positive real axis.
In the remaining part of the plane define a branch of z by requiring its
values to be real on the positive real axis. Now pick a contour, starting
at r > 0, continuing to R > r, then along the circle of radius R as
before, but stop before you reach the branch cut and go back along a
ray , which does not contain any pole of , until you reach the circle
with radius r again. Finally, continue clockwise along this circle until
you reach the point r again.
Next, pick another branch of z by cutting the plane along a ray
which comes before , counted counterclockwise from the positive real
axis. Fix the branch by requiring its values on to coincide with the
values of the earlier branch. This will give the branch the value ei2 x
for a real, positive x. We integrate the new branch along a contour
which starts at R, continuous along the positive real axis to r, then
follows the circle with radius r clockwise until it reaches the ray ,
then follows this ray outwards until it reaches the circle with radius R.
Finally, it follows this circle counterclockwise until it reaches the point
R on the positive real axis again.
If we add the two integrals constructed above, the contributions to
them along the ray will cancel, and the total effect will be exactly
as if we could integrate as we originally did, letting z have different
values along the upper and lower edges of the positive real axis. In
view of this, we will not commit any errors of we think of this as being
possible.
Now let us estimate the integrals along the circles. Note that |z | =
|z| whatever branch we use (as long as is real). Our assumption on
the degree of means that z 2 (z) is bounded, say by M , for large |z|.
If R is sufficiently large we therefore get
Z
Z
z (z) dz M R
|dz| = 2M R1 0
|z|=R
|z|=R
|dz| = 2mr 0
z (z) dz mr
|z|=r
|z|=R
as r 0. The
along the positive real axis together contribute
R R integrals
i2
(1 e
) r x (x) dx. It follows that
Z
x (x) dx = (1 ei2 )1 2i
0
X
z6=0
Res(z (z)) .
64
4. SINGULARITIES
=
.
2
x +1
2 cos(/2)
0
This is actually true for 1 < < 0 too, since in that case we may write
x+1
the integrand as x(x
2 +1) where now 0 < +1 < 1 so all the assumptions
are satisfied. But the residues remain the same since z +1 /z = z , using
the same branches of the powers. The formula is of course also true for
= 0, for elementary reasons.
R
5. We finally consider an integral 0 (x) ln x dx, where again is
rational without positive real poles. We still need to assume that the
degree of the denominator in is at least 2 more than that of the
numerator. In contrast to integrals of type 4, however, we can no longer
allow a pole at the origin. On the other hand, we use the same contour,
justifying the use of different values of the logarithm along the upper
and lower edges of the positive real axis as before. If we consider
(z) log z, using the branch of the logarithm where 0 < arg z < 2, its
values at x > 0 on the upper edge of the positive real axis is (x) ln x,
where ln is the usual real logarithm. For x > 0 on the lower edge
we instead get (x)(ln x + 2i). The difference is therefore 2i(x),
so we will not get the integral we are looking for. So, instead we
consider the function (log z)2 (z) which is (ln x)2 (x) on the upper
and (ln2 x + 4i ln x 4 2 )(x) on the lower edge of the positive real
axis. The difference is therefore 4i(x) ln x + 4 2 (x). It follows
that
Z
Z
1X
(x) ln x dx + i (x) dx =
Res((log z)2 (z)) .
2 z6=0
0
If, as we normally assume, has real coefficients we can therefore calculate the desired integral by taking the real part of the right hand
side. Otherwise, we would first have to calculate the second integral
by integrating (z) log z as in our first attempt.
65
R x
Example 4.14. Consider the integral 0 xln2 +1
dx which satisfies
the assumptions above. Using the appropriate branch in the plane cut
z)2
along the positive real axis, the function (log
has simple poles at i
z 2 +1
2
so the residues are the values of (log2zz) at these points. The sum of
the residues is therefore 2i1 ((i/2)2 (i3/2)2 ) = i 2 which is purely
imaginary. It follows that
Z
ln x
dx = 0 .
x2 + 1
0
R
0
dx
1+x2
= /2,
x
lower half plane, centered at 1. As an example, consider 0 xln2 1
dx.
The integral along the half circle is then
Z2
i
as r 0 .
66
4. SINGULARITIES
We normally choose so that the index of each zero and pole with
respect to equals one, and then the right hand side becomes the
difference between the number of zeros and the number of poles in ,
each counted by multiplicity.
Proof. If f (z) = (z a)n g(z) where n is a non-zero integer, g is
analytic near a and g(a) 6= 0, then f 0 (z) = n(za)n1 g(z)+(za)n g 0 (z)
0 (z)
0 (z)
n
so that ff (z)
= za
+ gg(z)
. The last term is analytic near a, so the residue
of the left hand side at a is n. Since |n| is the multiplicity of a as a zero
respectively pole, the theorem follows from the residue theorem.
67
68
4. SINGULARITIES
69
1
2i
f 0 (z)
dz
f (z) a
|zz0 |=
We restate the most important part of the conclusion of Theorem 4.21 as the open mapping theorem.
Corollary 4.22. Suppose f is analytic in some region and not
constant. Then f is an open mapping, i.e., the images of open sets are
open.
Proof. If z0 is in the domain of f , then by Theorem 4.21 the image
of any sufficiently small neighborhood of z0 contains a neighborhood of
f (z0 ). Hence f is an open mapping.
70
4. SINGULARITIES
71
CHAPTER 5
Harmonic functions
5.1. Fundamental properties
Suppose f is analytic in some region and u, v are its real and
imaginary parts, so that f (x + iy) = u(x, y) + iv(x, y). Then u and v
are harmonic in , according to the following definition.
Definition 5.1. A function u defined in an region C is called
harmonic if it is twice continuously differentiable in and satisfies
2
2
u = 0, where = x
2 + y 2 .
This follows since u, v satisfy the Cauchy-Riemann equations
ux = v y ,
uy = vx .
Since f is infinitely differentiable, we can differentiate the first equation
with respect to x, the second with respect to y, and add the results to
obtain u = 0, using that vxy = vyx . Similarly one shows that v is
harmonic.
If a function u, harmonic in , is given, then another harmonic
function v is called a conjugate function to u in if u + iv is analytic
in . Note that if u has a conjugate function in some region, then it
is determined up to an additive real constant. For suppose u + iv and
u + i
v are both analytic. Then so is the difference i(v v) which has
real part 0. It follows that the imaginary part v v is constant (this
follows from the Cauchy-Riemann equations, but also directly from the
open mapping theorem).
Note that if v is the harmonic conjugate of u, then u is the harmonic conjugate of v, since v iu = i(u + iv) is analytic if u + iv
is. A harmonic function does not necessarily have a conjugate
function
p
2
2
defined in all of its domain; consider for example ln x + y which is
the real part of any branch of the logarithm and therefore harmonic in
R2 \ {(0, 0)}. It can not have a conjugate function in this set, because
that would imply that we could define a single-valued branch of the
logarithm in the plane with just the origin removed. But we cant. On
the other hand, locally there is always a conjugate function. In fact,
the following theorem holds.
Theorem 5.2. If u is harmonic in a disk, then it has a conjugate
function there.
73
74
5. HARMONIC FUNCTIONS
Zy
=
75
Z2
f (z + rei ) d .
0
Taking the real part of this gives the desired formula with R replaced
by r. By the continuity of the integrand, however, we may now let
r R and so obtain the desired result.
Clearly one can calculate mean values in the above sense for any
continuous function. Interestingly enough, any continuous function
having the mean value property has to be harmonic (and is therefore
also infinitely differentiable). We will show this in Theorem 5.11.
Theorem 5.7 (Maximum principle). Suppose u is continuous on
the closure of a bounded region and satisfies the mean value property
in . Then u takes its largest and smallest value in on , and if
either is assumed in an interior point, then u is constant.
Proof. Suppose a and sup = u(a). There is a disk |z a| <
R contained in , and u(a + rei ) u(a) for all and 0 < r < R.
If there is strict inequality for some choice of r, , Rthen there is strict
2
1
inequality in a neighborhood by continuity, and 2
u(a + rei d <
0
u(a), violating the mean value property.
Thus the set M = {z | u(z) = u(a)} is open, as is the complement {z | u(z) 6= u(a)} by continuity. Since is connected and
M 6= it follows that M = , so that u is constant.
Since u satisfies the mean value property if u does, the statement
about smallest value follows as well.
76
5. HARMONIC FUNCTIONS
Z2
1
u(T (e )) d =
2i
u(T ())
||=1
d
.
77
1
u(a) =
2
Z2
u(Rei )
R2 |a|2
d .
|Rei a|2
What we have just seen is this: If Dirichlets problem for the disk
|z| < R has a solution, it must be given by Poissons integral formula.
2 |a|2
z+a
Note that R|za|
2 = Re za and that for |a| < R the integral
(5.2)
1
2
Z2
u(Rei )
Rei + a
d
Rei a
Z2
|u(Rei ) u(Rei )|
R2 |a|2
d .
|Rei a|2
Given > 0 we may find > 0 so that |u(Rei ) u(Rei )| < for
< < + . The integral over [0, ] [ + , 2] (if = 0,
over [, 2 ]) clearly tends to 0 as a Rei , and the integral over
[ , + ] (respectively [0, ] [2 , 2]) is < . It follows that
|Pu (a) u(Rei )| 0 as a Rei so that actually Pu tends to the
correct boundary values. We have proved the following theorem.
Theorem 5.10. Suppose u is a continuous function defined on |z| =
R. Then the function which equals Pu (z) for |z| < R and u(z) for
|z| = R is harmonic in |z| < R and continuous in |z| R.
In the process of solving Dirichlets problem we also obtained (5.2)
which expresses the values of a function analytic in the disk |z| < R in
78
5. HARMONIC FUNCTIONS
terms of the boundary values of its real part, in the case when these are
assumed continuous. This is a well known theorem by H. A. Schwarz.
Theorem 5.11. Suppose u is continuous in a region C and
has the mean value property there. Then u is harmonic.
Proof. Let |z z0 | < R be an open disk with closure contained in
and Pu the Poisson integral applied to u( + z0 ). Then Pu ( + z0 ) is
harmonic in the disk so that Pu u satisfies the mean value property
in the disk and is continuous in its closure. Therefore Pu u satisfies
the maximum principle Theorem 5.7 in the closed disk. But Pu u
vanishes on the boundary of the disk and is therefore identically 0.
Thus Pu = u in the disk, so that u is harmonic.
79
that this function is identically zero so that g has the appropriate symmetry. Since f g has zero imaginary part in the upper half circle it
is a real constant there. It follows that f can be extended analytically
as claimed.
.
means that v u
That v u satisfies the maximum principle in
80
5. HARMONIC FUNCTIONS
z,
vF
81
Theorem 5.16 (Harnacks principle). Suppose u1 , u2 , . . . is an increasing sequence of functions harmonic in a region . Then either
un + locally uniformly in , or else un converges locally uniformly to a function u which is harmonic in .
Proof. Suppose u is harmonic in a closed disk |z z0 | . The
Poisson integral formula then states that for z in the open disk
1
u(z) =
2
Z2
0
2 r 2
u(z0 + ei ) d ,
i
2
|e (z z0 )|
i
2
+r
|e (z z0 )|
r
If now u is non-negative in the disk we obtain Harnacks inequality
r
+r
u(z0 ) u(z)
u(z0 ) ,
+r
r
by the Poisson integral formula and the mean value property. Now
suppose r /2. Then Harnacks inequality shows that
1
u(z0 ) u(z) 3u(z0 ) .
3
Now consider the sequence u1 , u2 , . . . . Since the sequence is increasing it has a pointwise limit everywhere in , which is either finite or
+. If n > m the function un um is positive and harmonic in
so we can apply Harnacks inequality to it. It follows from (5.3) that
if un (z0 ) +, then un + uniformly in a neighborhood of z0 .
It also follows that the set where un tends to + is an open subset
of . Similarly, if un (z0 ) has a finite limit, then the limit is finite in
a neighborhood of z0 so the set where the limit is finite is also open.
Since is connected it follows that either un tends locally uniformly
to + in , or else the limit function u is finite everywhere. Applying
(5.3) to un um and letting n we get
(5.3)
82
5. HARMONIC FUNCTIONS
83
3z0
3z0
follows.
Let > 0 and choose a neighborhood O of 0 such that |f ()
f (0 )| < for O . Furthermore, let w0 be the minimum
of w over the (compact) set \ O. By the properties of w we have
w0 > 0. Now put V (z) = f (0 ) + + w(z)
(M f (0 )). Then V is
w0
harmonic in and continuous in the closure. For O we have
V () f (0 ) + > f (). For \ O we have w() w0 so we
get V () M + > f (). If v F and we therefore have
lim (v(z) V (z)) < 0 so by Lemma 5.17 v V in . It follows that
3z
also u V in so that
lim u(z) V (0 ) = f (0 ) + .
3z0
3z0
84
5. HARMONIC FUNCTIONS
CHAPTER 6
Entire functions
6.1. Sequences of analytic functions
In this section we shall consider sequences of analytic functions
which are uniformly convergent. We will use the notation H() for
the functions holomorphic (analytic) in the region C. By a region
we will always mean an open, connected set. Recall that we say that
a sequence f1 , f2 , . . . of real or complex-valued functions defined on a
set E is uniformly convergent on E to another function f defined on E
provided that for each > 0 we can find a number N such that if n N
then |fn (z)f (z)| < for every z E. If one introduces the maximumnorm k kE by setting kf kE = supzE |f (z)| the uniform convergence
of fn to f on E is equivalent to kfn f kE 0 as n . When
dealing with functions defined in an open set C (or Rn ) one
often talks about locally uniform convergence. A sequence of functions
f1 , f2 , . . . defined in is said to converge locally uniformly to f in
if every x has a neighborhood in which the sequence converges
uniformly to f . Equivalently, this means that fn f uniformly on
every compact subset of . This is an immediate consequence of the
Heine-Borel lemma.
Exercise 6.1. Show this equivalence!
Recall that in Flervariabelanalys it is proved that the uniform limit
of continuous functions is continuous. It immediately follows that the
same is true of locally uniform limits of continuous functions (explain
why this is obvious!). When dealing with analytic functions one can
say a lot more.
The main result of the section is the following.
Theorem 6.2. Suppose fn H(), n = 1, 2, 3, . . . and that fn f
(j)
locally uniformly in . Then f H(). Furthermore, fn f (j)
locally uniformly in for j = 1, 2, 3, . . . .
Proof. Let be a positively oriented circle such that the corresponding closed disk is contained in . For z in the open disk we then
have (Corollary 3.13)
Z
j!
fn (w) dw
(j)
.
(6.1)
fn (z) =
2i
(w z)j+1
85
86
6. ENTIRE FUNCTIONS
fn (w) dw
f (w) dw
j+1
(w z)
(w z)j+1
Z
(r/2)j1 |fn (w) f (w)| |dw| 2r(r/2)j1 kfn f k .
87
1
1
|f (z) fn (z)|
2
|=
2 kfn f k .
fn (z) f (z)
|fn (z)f (z)|
m
As an almost immediate consequence we have the following interesting corollary about so called univalent functions. A univalent function
is an injective (one-to-one) analytic function.
Corollary 6.5. Suppose fn H(), n = 1, 2, . . . , and fn f
locally uniformly in . If all fn are univalent, then so is f , unless it is
constant.
Proof. Assume f is not constant. Then, if f (z0 ) = w, we must
show that f (z) w 6= 0 for z \ {z0 }. Setting gn (z) = fn (z) fn (z0 )
we have gn f w locally uniformly in . Since by assumption gn
does not vanish in \ {z0 }, neither does f w by Theorem 6.4.
Exercise 6.6. Show that for any > 0 there exists N such that all
P (1)k 2k+1
Taylor polynomials of sin x (partial sums of
x
) of degree
(2k+1)!
at least N has exactly one zero in ( , + ).
Exercise 6.7. A famous theorem by Weierstrass states that any
function continuous on a real interval [a, b] is the uniform limit of a
sequence of polynomials. Why does this not contradict Theorem 6.2?
6.2. Infinite products
Any analytic function may be expanded in a power series centered
at any point of the domain of analyticity; the radius of convergence is
such that on the boundary of the disk of convergence there is at least
one singularity of the function. If the function is analytic everywhere
in C, the radius of convergence is therefore infinite. Such a function
is called entire (in Britain often also integral ). A power series used to
be viewed as a polynomial of infinite order, especially if the radius of
convergence is infinite. The reason is of course that many properties
of polynomials have their counterpart for entire functions.
One of the more fundamental properties of a polynomial is that,
according to the fundamental theorem of algebra and the factor theorem, it may be factored into a product of first degree polynomials,
each of which vanishes at one of the zeros of the polynomial.
If p is a
Qn
polynomial of degree n one usually writes p(z) = A k=1 (z zk ), where
z1 , z2 , . . . are the zeros of p, repeated according to multiplicity, and A
88
6. ENTIRE FUNCTIONS
Y
(6.3)
(1 + ak ) ,
k=1
so that the necessary condition for convergence just derived takes the
following form.
Proposition 6.9. A necessary (but not sufficient) condition for
convergence of the infinite product (6.3) is that ak 0 as k .
Since a sequence has a non-zero limit precisely if the sequence of
logarithms has a finite limit it is natural to compare the infinite product
(6.3) with the series with terms log(1 + ak ). Recall that the principal
branch of the logarithm is Log z = ln |z|+i arg z, where < arg z .
Theorem 6.10. If ak 6= 1, k = 1, 2, . . . , then the infinite product
(6.3) converges if and only if the series
X
(6.4)
Log(1 + ak )
k=1
89
(6.5)
log(1 + |ak |) .
k=1
Log(1+z)
z
Noting that
1 as z 0 (using the principal branch of the
logarithm) it follows by a standard comparison theorem that the series
(6.4) (omitting
Pterms for which ak = 1) is absolutely convergent if
and only if
k=1 ak is absolutely convergent (note that if either of
the two series are convergent, then we must have ak 0Pas k
). In particular it follows that (6.5) converges if and only
k=1 |ak |
converges. So we have proved the following proposition.
Proposition
6.11. The product (6.3) converges absolutely if and
P
only if
k=1 ak converges absolutely. This is also equivalent to the
series (6.4) converging absolutely, after omitting (the finite number of )
terms for which ak = 1.
We now turn to the case when the factors of (6.3) are functions of
z C. By inspection of the proofs it is clear that all the results obtained so far remain true if we replace convergence by locally uniform
convergence. So by Theorem 1.2, if ak H() for every
P k, then (6.3)
converges locally uniformly to a function in H() if
|ak | converges
locally uniformly in . In particular, by Weierstrass majorization theorem (Weierstrass M -test
P in most English language books) it follows
that this is the case if
kak kK converges for every compact K .
We can now return to the problem of generalizing the polynomial
factorization (6.2) to an arbitrary entire function. Suppose that we
have an entire function for which 0 is a zero of multiplicity j which
also has other zeros a1 , a2 , . . . , repeated according to multiplicity. By
analogy with (6.2) our candidate for this function would then be a
90
6. ENTIRE FUNCTIONS
Q
z
constant multiple of z j
k=1 (1 ak ). This may not be so, however.
First of all, there are entire functions with no zeros at all. One example
is ez ; more generally, eg(z) is such a function for any entire function g.
We would certainly have to allow such a factor in front of the product
to obtain a generally valid factorization. Furthermore, for the product
P 1
to converge absolutely for some z 6= 0 we must require that
|ak |
converges; this may not always hold, although it is true that we always
have ak as k (Exercise P
6.12). For example, the function
sin(z) has zeros 0, 1, 2, . . . and
1/k is divergent. A little more
effort is therefore required to obtain a general factorization formula for
entire functions. We will carry this out in the next section.
Exercise 6.12. Prove that if a1 , a2 , . . . are the zeros of an entire
function, repeated according to multiplicity, then ak as k .
6.3. Canonical products
Consider first the case of an entire function f with only finitely
many non-vanishing zeros a1 , . . . , an , as always counted with multiplicities. If the multiplicity of 0 as a zero is j 0 it is clear that
j
h(z) = f (z)z /
n
Y
(1
k=1
z
)
ak
f (z) = z j eg(z)
(1
k=1
z
),
ak
X
1
2
(6.7)
=
n2
6
n=1
91
sin z Y
z
(6.8)
=
(1 2 2 ) .
k
z
k=1
It is at least clear
the product converges absolutely and locally
P that
2
uniformly since
1/k converges.
P The partial product Pn is a polynomial and its z-coefficient is nk=1 (k)2 . This is the value of Pn0 (0),
and according to Theorem 6.2 this converges to the derivative at 0 of
the infinite product. Assuming (6.8) the power series for f shows that
this is 1/6 and so Euler would be vindicated.
Instead of proving (6.8) we will show the equivalent statement
sin z Y
z2
(6.9)
(1 2 ) ,
=
z
k
k=1
obtained by replacing z in (6.8) by 2 z 2 . Since the infinite product
converges absolutely locally uniformly we at least know from (6.6) that
Y
z2
sin z
g(z)
(1 2 ) ,
=e
z
k
k=1
where g is entire. Taking the logarithmic derivative of both sides we
obtain
X
1
1
0
+
,
(6.10)
cot z 1/z = g (z) +
zk z+k
k=1
and differentiating once more we obtain
X
2
1
00
=
g
(z)
+
.
2
2
(z
+
k)
sin z
k=
However, both the infinite sum and the left hand side are of period 1
here, so g 00 also has period 1. Writing z = x + iy it is easy to see that
both the infinite sum and the left hand side tend to 0 locally uniformly
in x as y . Thus the same is true of g 00 which is therefore bounded
and entire, hence by Liouvilles theorem constant; since g 00 (iy) 0 as
y + the constant is 0. Therefore g 0 is constant. But from (6.10)
follows that g 0 is odd, so that also g 0 vanishes. Thus g is constant, so
that (6.9) holds apart from a constant factor. Since both sides tend to
1 as z 0 this factor is 1, and we have finally established (6.9), and
thereby Eulers formula.
92
6. ENTIRE FUNCTIONS
Exercise 6.14. Justify all unproved claims at the end of Example 6.13.
What is one to do to obtain a factorization for an entire function
where the sum of the reciprocals of the zeros is not absolutely convergent? The idea is to replace the factor 1 azk in the product by
(1 azk )epk (z) , where pk is an entire function which promotes convergence without introducing new zeros. As we shall see, one can always
choose pk to be a polynomial. Convergence is obtained by choosing pk
so that (1 azk )epk (z) is sufficiently close to 1, so the ultimate choice
would be log(1 azk ). Unfortunately this is not an entire function.
It is therefore natural to attempt to choose pk as a Taylor polynomial of this function of sufficiently high degree. Now for the princiP z k
pal branch of the logarithm Log(1 z) =
k=1 k and the series
Pn z k
converges for |z| < 1. In fact, if we set Ln (z) =
k=1 k we have
Log(1 z) = Ln (z) + Rn (z), where an easy estimate gives
X
X
|z|k
|z|n+1
|Rn (z)|
|z|k =
k
1 |z|
k=n+1
k=n+1
(1
k=1
z pk (z)
)e
ak
X
z
{Log(1 ) + pk (z)}
ak
k=1
does (check this carefully!). We assume now |z| < R. There are only
finitely many factors in (6.11) for which |ak | < 2R (Exercise 2.5) so excluding these factors from the product will not affect convergence. We
may thus also assume that |ak | 2R. If we choose pk (z) = Lnk (z/ak )
the absolute value of the term in (6.12) is |Rnk (z/ak )| and may therefore be estimated by 21nk , using the facts that R/|ak | 1/2 and
1 R/|ak | 1/2. We conclude that (6.12) converges absolutely and
uniformly
P n in |z| < R if we can choose nk for every k such that the series
k converges. A obvious choice that works is n
k = k. Since
k=1 2
R is arbitrary we conclude that the choice pk (z) = Lk (z/ak ) makes
(6.11) absolutely and locally uniformly convergent. We have proved
the following theorem by Weierstrass.
Theorem 6.15. There exists an entire function with arbitrarily prescribed non-vanishing zeros a1 , a2 , . . . (repeated according to multiplicity), provided they are either finitely many or else ak as k .
93
Every entire function with these and no other zeros may be written
Y
z L (z)
(6.13)
f (z) = z j eg(z)
(1 )e nk ak ,
ak
a 6=0
k
P
k
where g is an entire function, Ln (z) = nk=1 zk , and nk , k = 1, 2, . . .
are certain (sufficiently large) positive integers. A possible choice is
nk = k.
The theorem has a very important corollary concerning meromorphic functions. Recall that a function f is called meromorphic in if
it is analytic in except for isolated singularities which are poles.
Corollary 6.16. Every function which is meromorphic in the
whole plane is the quotient of two entire functions.
Proof. If f is meromorphic in the whole plane we may, according
to Theorem 3.3, find an entire function g so that all the poles of f are
zeros of g, and with the same multiplicities. Thus h = f g is an entire
function and f = h/g.
R
for
any
R.
Since
ak
P
P
h+1
h+1
h+1
this happens if (R/|ak |)
=R
1/|ak |
converges. In other
words, if the zeros do not tend too slowly
to
infinity.
Suppose now
P
h+1
that h is the smallest integer for which
1/|ak |
converges (so that
h 0). Then
Y
z
(1 )eLh (z/ak )
ak
k=1
is called the canonical product associated with the sequence a1 , a2 , . . . ,
and the integer h is called the genus of the canonical product. If possible we use the canonical product in the expansion (6.13). In that case
the expansion becomes uniquely determined by f . If it then happens
that g is a polynomial, one says that the function f has finite genus,
and the genus of f is the degree of g or the genus of the canonical
product, whichever
is the largest. This means for example that the
function sin z/ z considered in Example 6.13 and with the product
expansion (6.8) is of genus 0.
Example
P 6.17. The function
P sin z has all the integers as its zeros,
and since 1/n diverges but 1/n2 converges we obtain an expansion
of the form
Y
z
sin z = zeg(z)
(1 )ez/k .
k
k6=0
If we group the factors for k together and compare the result to (6.9)
it follows that g is the constant log . Consequently, sin z is of genus
94
6. ENTIRE FUNCTIONS
Y
ak ak z
|ak | 1 ak z
k=1
n
X
k=1
Pk (
1
)
z ak
X
f (z) = g(z) +
(Pk (
k=1
1
) qk (z)) ,
z ak
95
} d .
h(){
k+1
k!
2i
k=0
k=0
1
Supposing |Pk ( za
)| Mk for |z| = |ak |/2 we obtain
k
2|z| nk +1
|ak |
for |z| |ak |/3. Consider a disk |z| R. There are only finitely
many of the ak with |ak | < 3R, and it is clear that after removing the
corresponding
terms (6.15) converges uniformly in |z| R if the series
P
2R nk
Mk ( |ak | ) converges. We may consider this a power series in R, and
it will then have infinite radius of convergence if the terms tend to 0 for
)nk
every R > 0. Choosing nk log Mk we have Mk ( |a2Rk | )nk ( 2eR
|ak |
0 as k since ak . Thus the sum in (6.15) represents a
meromorphic function with the same singular parts as f in all poles,
so the theorem follows.
X
1
2
.
=
2
sin z k= (z k)2
As a final example, consider the function sinz which has simple poles
at the integers, with residue 1 at even and 1 at odd integers. The
partial fractions expansion is thus of the form
(1)k
= g(z) +
sin z
zk
k=
96
6. ENTIRE FUNCTIONS
1
1 X
1
+
+
(1)k
z k=1
zk z+k
1 X 1
1
1
1
1
= +
+
+
z k=1 z 2k z + 2k
z 1 k=1 z 1 2k z 1 + 2k
z
(z 1)
g(z) = cot
cot
.
2
2
2
2
sin z
But the right hand side is easily seen to vanish identically, so we have
the partial fractions expansion
(1)k
=
sin z k= z k
Exercise 6.22. Evaluate
1
.
(z + k)2 + a2
k=
This means that is the smallest number such that |f (z)| < e|z|
for any given > 0 as soon as |z| is sufficiently large. Consequently,
polynomials have order 0, ez and sin z have order 1, ep(z) has order n
z
if p is a polynomial of degree n, and ee has infinite order. Note that
whereas the genus is always a natural number (or infinity), the order
may be anynon-negative number (or infinity); for example, the entire
function sinz z we discussed earlier has order 1/2 (show this!).
97
a
2
e
z
k)
k=1
0
1
log |f (0)| =
log
+
log |f (eit )| dt
|ak | 2
k=1
0
2 a
z zak
zak
has absolute
then F has no zeros in |z| < and |F (z)| = |f (z)| for |z| = . Thus
(6.17) follows on applying Poissons integral formula to log |F (z)|.
We can now turn to Hadamards theorem.
Proof of Theorem 6.24. AssumeP
first that the entire function
f has finite genus h. This means that
1/|ak |h+1 converges, where
a1 , a2 , . . . are the zeros of f . The exponential factor in (6.13) is clearly
of order h, and since the order of a product clearly does not exceed
98
6. ENTIRE FUNCTIONS
(6.18)
for all z. This is true for h = 0, since log |1 z| log(1 + |z|) |z|.
By the definition of Rh it is obvious that we have
| Re Rh (z)| | Re Rh1 (z)| + |z|h
(6.19)
for all z. If | Re Rh1 (z)| (2h 1)|z|h then clearly (6.18) follows if
|z| 1. But if |z| < 1 we have the estimate (1|z|)| Re Rh (z)| |z|h+1
from page 92. Multiplying (6.19) by |z| and adding we get | Re Rh (z)|
|z|| Re Rh1 (z)| + 2|z|h+1 from which again (6.18) follows by the induction assumption. We can now estimate
log |P (z)| =
X
X
( Re Rh (z/ak )) (2h + 1)|z|h+1
1
|ak |h+1
Z2
log |f (2eit )| dt log |f (0)| ,
0
n()
X
1
f 0 (z) X 1
ak
=
+
+
2
f (z)
z ak k=1 ak z 2
k=1
Z2
0
2eit
log |f (eit )| dt .
it
2
(e z)
99
n()
X
k=1
h! ak h+1
1
+
(2 ak z)h+1 2
Z2
2(h + 1)!eit
log |f (eit )| dt .
it
h+2
(e z)
log
M ()
dt .
|f (eit )|
R 2
1
By Jensens formula we have 2
log |f | log |f (0)| and since by
0
log(M ())
assumption h+1 0 as , it follows that the integral in
(6.20) vanishes as . Similarly, the penultimate term in (6.20)
may, for |z| /2 be estimated by n()/h+1 which, as we have already
seen, tends to 0 as . It follows that
X
h!
dh f 0 (z)
=
.
h
dz f (z)
(ak z)h+1
k=1
If we write f (z) = eg(z) P (z), where P is the canonical product, then
dh P 0 (z)
(h+1)
clearly the sum to the right is dz
(z) =
h P (z) , so that it follows that g
0. Thus g is a polynomial of degree at most h, and the proof is finally
complete.
100
6. ENTIRE FUNCTIONS
an integer, the genus, and thus the form of the factorization, is uniquely
determined, whereas there is an ambiguity if the order is an integer.
Exercise 6.27. Let f be entire and M (r) as before. Suppose
lim log rM (r) = A is finite and not 0. Show that f is of order , but
r
that the existence of the limit does not follow from assuming f to have
order . An entire function for which A is finite and > 0 is said to be
of order and normal type. Extend Corollary 6.26 to show that an
entire function of finite order has infinitely many zeros unless it is of
integer order and normal type.
CHAPTER 7
102
if n and m > N .
103
d
2i
z w
|z0 |=2r
|z w|
=
2
f ()
d
( z)( w)
|z0 |=2r
M |z w|
2r2
|d| =
2M
|z w|
r
|z0 |=2r
r
makes
since | z| > r, | w| > r. It follows that choosing = 2M
|f (z) f (w)| < if z and w B(r, z0 ) and |z w| < . The local
equicontinuity of the family F follows and the theorem is therefore a
corollary to Theorem 7.5.
Theorem 7.6 (Riemann mapping theorem). Given a simply connected region which is not the entire complex plane C and a point
z0 there is precisely one univalent conformal map f of onto the
unit disk such that f (z0 ) = 0 and f 0 (z0 ) > 0.
Note that Liouvilles theorem shows that it is not possible to map
the entire plane C conformally onto the unit disk; the only bounded
entire functions are the constants.
Proof. We have already proved the uniqueness in Chapter 4.4
after Schwarz lemma (p.7071). To see how to get existence, note that
if g solves the problem and f is a map of into the unit disk mapping
z0 onto 0 and with positive derivative at z0 , then f g 1 satisfies the
conditions of Schwarz lemma so |(f g 1 )0 (0)| 1. Calculating the
derivative we see that this means that f 0 (z0 ) g 0 (z0 ). If we have
equality it follows from Schwarz lemma that f g.
Now let F be the family of univalent functions f analytic in such
that f (z0 ) = 0, |f (z)| 1 for z and f 0 (z0 ) > 0. We just saw that
if our problem has a solution it is the element of F which maximizes
the derivative at z0 . To complete the proof along these lines we need
to: (1) Show that F is not empty, (2) See that F has an element
f maximizing the derivative at z0 and, finally, (3) Show that this f
actually solves the mapping problem.
(1) Since is not all of C there is a (finite) point a
/ .
Since is
simply connected we can define a single-valued branch h of z a in
. Clearly h can not take the value w if it somewhere takes the value
104
h(z) h(z0 )
2
= h(z0 )
h(z) + h(z0 )
h(z0 ) h(z) + h(z0 )
0
h (z0 )
maps z0 to 0 and is bounded by 4|h(z0 )|/. Its derivative at z0 is 2h(z
.
0)
If we now put
|h0 (z0 )| h(z0 ) h(z) h(z0 )
g(z) =
4 |h(z0 )|2 h0 (z0 ) h(z) + h(z0 )
it follows that g is univalent, g(z0 ) = 0, |g(z)| 1, and g 0 (z0 ) > 0 so
that g F. Hence F 6= .
(2) Since all elements of F have their values in the unit disk it
follows that F is an equibounded family, and therefore by Theorem 7.3
a normal family. Now let B = supf F f 0 (z0 ) so that 0 < B .
We can then find a sequence f1 , f2 , . . . in F so that fj0 (z0 ) B as
j . Since F is normal we can find a locally uniformly convergent
subsequence; call the limit function f . It is then clear that f 0 (z0 ) = B
so that actually B < and f is not constant. By Corollary 6.5 f is
univalent. It is clear that f (z0 ) = 0 and f has its values in the closed
unit disk; but by the open mapping theorem the values are then in the
open unit disk.
(3) We need to prove that f () is the unit disk. Suppose to the
contrary that w0 is in the unit disk but w0
/ f (). Since is simply
connected we may define a single-valued branch of
s
f (z) w0
G(z) =
.
1 w0 f (z)
ww0
Since the Mbius transform w 7 1w
preserves the unit disk, the
0w
function G maps univalently into the unit disk. To obtain a member
of F we now set
|G0 (z0 )| G(z) G(z0 )
.
F (z) = 0
G (z0 ) 1 G(z0 )G(z)
It is again clear that F has its values in the unit disk and maps z0 to 0.
2
0 )|
The derivative at z0 is easily calculated to be F 0 (z0 ) = B 1+|G(z
>B
2|G(z0 )|
so that F F. But this contradicts the definition of B.
Note that it is no accident that we get F 0 (z0 ) > f 0 (z0 ); this just
expresses the fact that the inverse of the map f 7 F takes the unit
disk into itself with 0 fixed so that Schwarz lemma shows that the
derivative at 0 is < 1 (clearly the map is no rotation).
CHAPTER 8
(z + 1) = z (z)
is valid for Re z > 0 (check this). Since clearly (1) = 1 it follows by induction that (n + 1) = n! for natural numbers n, so one may view the
gamma-function as an extension of the factorial to non-natural numbers. Another very important consequence of (8.1) is that it allows one
to extend analytically to the left of Re z = 0. If is already defined
in z + 1 we may define (z) = z1 (z + 1). Clearly this works as long
as z 6= 0. By induction we may therefore define everywhere except
at the non-positive integers. In these points the extended gammafunction has simple poles. In this way the gamma-function is extended
to a meromorphic function in the whole complex plane, with poles at
0, 1, 2, . . . and nowhere else.
Exercise 8.1. Calculate the residues of at the non-positive integers!
To obtain a product expansion of , let us first construct
an entire
P
function
with
simple
zeros
where
has
poles.
Since
1/n
diverges
P
2
but
1/n does not, we set
F (z) = z
z
(1 + )ez/k .
k
k=1
105
106
k6=0 (1
z sin z
.
(8.2)
(8.3)
F (z)F (z) =
F (z)/z = e(z) F (z + 1)
X
X
1
1
1
1
1
0
(
) = (z) +
+
(
).
z+k k
z + 1 k=1 z + 1 + k k
k=1
If we replace k by k + 1 in the first sum we obtain, after simplification
X
1
1
0 (z) =
(
)1 .
k
k
+
1
k=1
Since the series telescopes to the sum 1 we have 0 (z) = 0 so that
is constant. To determine the value of , we note that F (z)/z 1 as
z 0 so we obtain from (8.3) that 1 = e F (1). But the n:th partial
product of F (1) is
n
X
1
23
n + 1 1 1 1
2
n
...
e
= (n + 1) exp(
),
12
n
k
k=1
P
so that = limn ( nk=1 k1 log n). The constant is called Eulers
constant and equals approximately 0.5772. As far as I know it is not
known whether is rational (though it seems unlikely). If we set
G(z) = ez /F (z) we have the expansion
ez Y
z
(8.4)
G(z) =
(1 + )1 ez/k ,
z k=1
k
and from (8.3) follows
G(z + 1) = zG(z) ,
the same functional equation that satisfies. One might now guess
that G = . We will show this, which is surprisingly difficult. Note
that we obtain from (8.2) and the functional equation the so called
reflection formula
.
G(z)G(1 z) =
sin z
Since F has no poles the function G has no zeros, and since it has
the same poles as , the function (z)/G(z) is entire. If we can show
that it is bounded, then by Liouvilles theorem it is constant and since
(1) = G(1) we would be done. Note that by the functional equations
that G and satisfy we have (z + 1)/G(z + 1) = (z)/G(z) so that
107
(8.5)
G(z) = 2z z1/2 ez eJ(z) ,
where J(z) 0 as z in a half-plane Re z c > 0. We will
prove this formula later; for the moment let us show that it implies the
desired lower bound for G and hence the identity of G and . If (8.5)
is true we obtain, for z = x + iy,
log |G(z)| = 12 log 2 x + (x 12 ) log |z| y arg z + Re J(z) .
All terms are here bounded from below except y arg z, which is at
least bounded from below by |y|/2. It follows that /G is bounded
in the period strip by a constant multiple of e|y|/2 . For a function of
period 1 this is enough to show boundedness, since such a function may
1
be viewed as a function of = e2iz , the possible values of z = 2i
log
differing by integers. As a function of the function /G has isolated
singularities at 0 and , but our bound on /G is e| log |||/4 , i.e., for
small || a multiple of ||1/4 and for large || a multiple ||1/4 . Thus
both singularities are removable (see the following exercise), /G is
bounded, and we are done.
Exercise 8.2. Recall that if f is analytic with an isolated singularity at z = w, then the singularity is removable if (and only if)
(z w)f (z) 0 as z w. State a similar condition for singularities
at infinity.
Hint: Look at the discussion just before Theorem 4.2.
Let us now turn to Stirlings formula, so assume Re zP
> 0. Accord1
1
ing to (8.4) the logarithmic derivative of G is z1
1 ( z+k k )
and differentiating once more we get
d G0 (z) X
1
=
.
dz G(z)
(z + k)2
k=0
For fixed z in the right half-plane the terms of the sum are the residues
cot
. Note that
in the right half-plane of the function H() =
(z + )2
= z is not in the right half-plane, and sincos is analytic and equals
1 at 0. Thus the residue at 0 is z12 . By periodicity the residue at k is
1
; thus the residues of H are as stated.
(z+k)2
Now let be the contour consisting of a rectangle with corners iY
and n+ 21 iY , except for avoiding = 0 by a small semicircle of radius
R
1
H.
r centered at 0, such that 0 is inside the contour. Consider 2i
108
P
1
This is independent of r for small r and equals nk=0 (z+k)
2 . On the
horizontal sides the factor cot tends uniformly to i as Y , and
the other factor (z + )2 tends uniformly to 0, so the corresponding
integrals also tend to 0. Our contour now consists of two infinite vertical
lines, apart from the little semicircle. On the line Re = n+ 21 the factor
cot is bounded, independentlyR of the integer n, so the corresponding
integral is less than a multiple of Re =n+ 1 |z+|2 d Im which tends to
2
0 as n . The integrals over the straight line parts of the remaining
part of the contour may be written
1
Zr
1
cot(i)
d
2
(i + z)
2
cot(i)
d
(i + z)2
1
=
2
cot(i)
1
1
d ,
(i z)2 (i + z)2
d G0 (z)
1
1
= + 2+
dz G(z)
z 2z
Z
( 2
4z
d
.
2
2
2
+z ) e 1
G(z)
2z
Z
2
d
2
.
2
2
+z e 1
( + z )
0
109
Z
2
z
log(1 e2 ) d .
+ z2
Z
2
1
z
log
d ,
2
+z
1 e2
so it only remains to show that J(z) has the claimed behavior and to
determine the constants of integration C, D. But we have | 2 + z 2 | =
|z i||z + i| c|z| if Re z R c > 0 so the integral over [N, ) may be
1
estimated by the integral c
log 1e12 d which is convergent and
N
therefore < for sufficiently large N . But if |z| > N we can estimate
R
log 1e12 d, which tends to 0
the integral over (0, N ] by (|z||z|
2 N 2 ) 0
as z . Thus J(z) 0 if z in Re z c > 0. The functional
equation for G may be expressed log G(z + 1) = log z + log G(z), at
least if z > 0. Substituting (8.6) in this gives, after simplification,
1
1
C = (z + ) log(1 + ) + J(z) J(z + 1) .
2
z
Letting z + it follows that C = 1. To determine D we substitute
(8.6) in the reflection formula G(z)G(1 z) = / sin z for z = 12 + iy
to obtain, after simplification,
= (eD )2
cosh y
exp(1 + iy(log( 12 + iy) log( 12 iy)) + J( 12 + iy) + J( 12 iy)),
where the logarithms have their principal value. Further simplification
gives
(eD )2 = 2 exp(1 + 2y arctan(2y) J( 12 + iy) J( 12 iy))/(ey + ey )
1
= 2 exp(y 2y arctan 2y
+ 1 J( 12 + iy) J( 12 iy))/(ey + ey )
2 as y + .
.
(z) (1 z) =
sin z
and Stirlings formula
110
Exercise 8.4. Verify the calculations above. Then show that the
integrand in J(z) may be developed as a finite sum of odd powers of
1/z plus a remainder and that the result may be integrated to yield an
expansion
n
X
Ak
J(z) =
+ Jn (z) ,
2k1
z
k=1
where the remainder Jn (z) may be estimated by a constant multiple of
1/z 2n+1 for large z satisfying Re z c > 0. Also show that for fixed
z the remainder Jn (z) has no limit as n . An expansion of this
kind is called an asymptotic expansion (as z in Re z c > 0).
One may express the constants Ak explicitly in terms of the so called
Bernoulli numbers.