You are on page 1of 114

An introduction to complex analysis

Michael Singer
University College London
Contents

Chapter I. Plane topology 5


1. Open sets, closed sets, neighbourhoods 5
2. Bounded sets and compactness 9
3. The Riemann sphere 12
4. Möbius transformations with prescribed properties 18
Chapter II. Analytic functions and holomorphic functions 25
1. Analytic functions 25
2. Holomorphic functions—complex differentiability 31
3. The Cauchy–Riemann equations 39
4. Connectedness 42
5. Harmonic functions 45
6. ‘Holomorphic at ∞’ 52
Chapter III. Complex Integration 55
1. Paths and line-integrals 55
2. Green’s Theorem, Cauchy’s Theorem and Cauchy’s Integral Formula 65
3. Holomorphic functions are analytic; higher derivatives 72
Chapter IV. Zeros, poles, and the residue theorem 77
1. Zeros of analytic functions 77
2. Poles and Laurent series 80
3. The Residue Theorem 84
4. Definite integrals 86
Chapter V. The argument principle 99
1. Counting zeros and poles 99
2. Winding number and argument principle 100
3. Rouché’s Theorem 104
Chapter VI. Green’s Theorem 107
1. Green’s Theorem for a rectangle 107
2. Green’s Theorem for a region bounded by a graph 109
3. Green’s Theorem for bounded domains with smooth boundary 112
4. Piecewise smooth boundaries 114

3
CHAPTER I

Plane topology

1. Open sets, closed sets, neighbourhoods


1.1. Open sets and neighbourhoods.
Notation I.1.1. If a ∈ C and r > 0, then
D(a, r) = {z ∈ C : |z − a| < r} (I.1.1)
is the open disc with centre a and of radius r. It is convenient to allow r = ∞, in which
case D(a, r) is taken to be the whole of C.
The closed disc centred at a and of radius r is the set
D(a, r) = {z ∈ C : |z − a| 6 r}. (I.1.2)
The open punctured disc is the set
D∗ (a, r) = {z ∈ C : 0 < |z − a| < r}. (I.1.3)
We sometimes write C∗ for the punctured plane C \ {0}.
Remark I.1.2. Notation for open and closed discs is standard, though you sometimes
see Dr (a) for D(a, r) or D replaced by B (for ‘ball’).
It is very convenient to have the following basic notions from topology.
Definition I.1.3. A subset U of C is called open if: for every point z ∈ U , there is
an open disc D(z, r) which is itself contained in U .
Definition I.1.4. If z ∈ C, a set N ⊂ C is called a neighbourhood of z if there is an
open disc D(z, r) which is contained in N .
Examples I.1.5. The whole complex plane, C as well as the empty set ∅, are open.
The set D(a, r) is open for any positive r. The punctured set D∗ (a, r) is open. The
half-plane {z : Re(z) > 0} is also open.
Non-Examples I.1.6. The closed disc D(a, r) is not open if r < ∞. The closed
half-plane {z : Re(z) > 0} is not open. The set of point {z = n : n ∈ Z} is not open.
The set of points {z = 1/n : n ∈ N} is not open.
Examples I.1.7. Although D(0, r) (with 0 < r < ∞) is not open, it is a neighbour-
hood of every point z with |z| < r. Indeed, for such z, D(z, r − |z|) is an open disc
contained in D(0, r) and containing z.
Similarly, although H = {z : Im(z) > 0 is not open, it is a neighbourhood of every
point z = x + iy with x > 0.
5
6 I. PLANE TOPOLOGY

Exercise I.1.1. Fill in the details of the above remark: if z = x + iy with x > 0,
find an open disc centred at z, which is contained in H. Convince yourself also that
H is not a neighbourhood of any point on the imaginary axis.

Non-Examples I.1.8. The set I = {z = n : n ∈ Z} is not a neigbourhood of any of


its elements. Indeed, D(n, r) contains points that are not integers (hence outside of I),
for any r > 0.
The set R = {z = 1/n : n ∈ N} is also not a neigbourhood of any of its elements.
Proposition I.1.9. A set U is open if and only if it is a neighbourhood of each of
its points.
Proof. Suppose that U is open and let z be a point of U . By definition, there is
some open disc D(z, r) ⊂ U . But this also means that U is a neighbourhood of z.
Conversely, suppose that U is a neighbourhood of each of its points. This means, by
definition, that if z ∈ U , there is some open disc D(z, r), such that D(z, r) ⊂ U . Thus
for each point z of U , we have z ∈ D(z, r) ⊂ U , for some r, which means that U is
open. 

Think of it like this: if N is a neigbourhood of z, then we can approach z in any


direction through points of N . If U is open, then it is a neighbourhood of each of its
points, and so any point z of U can be approached in any direction through points
of U .

This approaching in all directions idea makes open sets the natural domains of defi-
nition of continuous functions (and also differentiable, analytic, holomorphic functions).
However, if S is any subset of C, we define continuity as follows: a function f : S → C
is continuous at z0 ∈ S if given ε > 0, there exists δ > 0 such that
z ∈ S ∩ D(z0 , δ) =⇒ |f (z) − f (z0 )| < ε.
Further, f : S → C is continuous if it is continuous at each of its points.
Definition I.1.10. If X is any subset of C, the interior of X is the set
X ◦ = {z ∈ C : X is a neighbourhood of z}. (I.1.4)
Examples I.1.11. The interior of any open set U is, by definition, U itself. The
interior of D(a, r) is just D(a, r). The interior of Z ⊂ C is the empty set.

Exercise I.1.2. What is the interior of D∗ (a, r)? For which z is the set D(0, 1)
a neighbourhood of z? Same question for D(0, 1) and D∗ (0, 1).
Exercise I.1.3. *For enthusiasts: Show that if X is any subset, then X ◦ is
the union of all open sets U with U ⊂ X.
1. OPEN SETS, CLOSED SETS, NEIGHBOURHOODS 7

Exercise I.1.4. *For enthusiasts: Suppose that R = {z = 1/n : n ∈ N}


and f : R → C is a function. What conditions must f satisfy if f : R → C is
continuous? [You should unpack the definition of continuity in this case rather than
merely restating it.] Same question with R replaced by the closure R of R (Def. I.1.16
below.)

1.2. Closed sets and limit points.


Definition I.1.12. A subset F of C is called closed if the complement C \ F is open.
We shall see in a moment that this is equivalent to the following, which feels more
like ‘closure’: F is closed if and only if for any sequence (zn ) with zn → w in C, then if
zn ∈ F for all n, the point w is also in F .

Exercise I.1.5. Which of D(a, r), D(a, r) and D∗ (a, r) are closed?
Exercise I.1.6. Is Z ⊂ C closed? How about {1/n : n ∈ N}?
Exercise I.1.7. Is it the case that any given subset of C is either open or closed?
Reasons?

Definition I.1.13. If S is any set, then z0 is a limit point of S if for every r > 0,
the punctured disc D∗ (z0 , r) contains points of S (in other words, D∗ (z0 , r) ∩ S 6= ∅).
Limit points are also often called accumulation points.
Proposition I.1.14. Let (z1 , z2 , . . .) be a convergent sequence of complex numbers
zn → w. Then if the set S = {zn : n ∈ N} is infinite, the set of limit points of S is {w}.
Proof. The assumption that S is infinite means that infinitely many of the zn are
different from w. The definition of convergence implies that given r > 0, there exists N
such that |zn − w| < r for all n > N . Since also infinitely many of the zn are different
from w, this means that
S ∩ D∗ (w, r) 6= ∅.
Hence w is a limit point of S.
Conversely, suppose that v is a limit point. It means that given any r > 0,
S ∩ D∗ (v, r) 6= ∅.
Thus at least a subsequence of (zn ) converges to v. Since zn converges to w, every
subsequence of zn also converges to w, so v = w. 

We now show that the definition given above is equivalent to F containing all its
limit points.
Proposition I.1.15. A set F is closed if and only if it contains all its limit points.
8 I. PLANE TOPOLOGY

Proof. We shall use the definition that F is closed if and only if C \ F is open.
Suppose that F is closed and that z0 6∈ F . We shall show that z0 cannot be a limit
point; turning this around F contains all its limit points.
Another way of writing z0 6∈ F is z0 ∈ C \ F . But by definition, this is open, so for
some r > 0, D(z0 , r) ⊂ C \ F . Hence D(z0 , r) ∩ F = ∅ and so too D∗ (z0 , r) ∩ F = ∅.
Hence z0 cannot be a limit point of F . This proves ‘F closed’ ⇒ ‘F contains all its limit
points’.
For the converse, suppose that F contains all its limit points and let w ∈ C \ F .
Then w is not a limit point of F and this means that for some r > 0, D∗ (w, r) ∩ F = ∅,
or in other words D∗ (w, r) ⊂ C \ F . Since by hypothesis w ∈ C \ F , it follows that
D(w, r) ⊂ C \ F . Thus C \ F is a neighbourhood of w and since w was an arbitrary
point of C \ F , we conclude that C \ F is open. 
Definition I.1.16. If X ⊂ C is any set, then the closure of X is the union of X and
all limit points of X. This is denoted by X. The boundary of X, denoted ∂X, is the set
X \ X ◦.

Exercise I.1.8.
(a) What are the closures and boundaries of D(0, 1), D(0, 1) and D∗ (0, 1)?
(b) What are the closure and boundary of the unit circle {z : |z| = 1}?
(c) If A = {a1 , a2 , . . .} is a countable sequence of points, is A sometimes, always,
or never, open? Is it sometimes, always, or never closed? Give reasons.
(d) *For enthusiasts Show that if X is any subset of C then X is the inter-
section of the family of all closed sets F which contain X.

NB If X is a subset of C it does not need to be open or closed.

1.3. Behaviour under union and intersection. The following is useful back-
ground knowledge and very significant for the development of continuity in the context
of abstract topological spaces.
Proposition I.1.17. Let (Uα )α∈A be some family of open sets. (The ‘index set’ A
need not be finite or even countable.) Then
[
U= Uα
α∈A
is again open. If A is finite, then
\
V = Uα
α∈A
is open.
Similarly if (Fα )α∈A is a family of closed sets, then the intersection
\
F = Fα
α∈A
2. BOUNDED SETS AND COMPACTNESS 9

is closed. If A is finite, then


[
G= Fα
α∈A
is closed.
Proof. Suppose that z is a point of the union of the Uα . Then in particular, there
is some element α1 , say, of A, such that z ∈ Uα1 . Since Uα1 is open, there is a small disc
D(z, r) such that D(z, r) ⊂ Uα1 . But then D(z, r) ⊂ U as well (by definition of union),
so U is a neighbourhood of z. Since z was arbitrary, U is open.
If A is finite, let us write the Uα as U1 , . . . , UN . If z ∈ V , then there exists D(z, rj ) ⊂
Uj for each j = 1, . . . , N . If r is the smallest of the rj , then D(z, r) ⊂ Uj for all j, that
is, D(z, r) ⊂ V . Hence V is a neighbourhood of each of its points.
The corresponding results for closed sets follow from de Morgan’s laws: F is closed
if and only if C \ F is open. But
[
C\F = (C \ Fα )
α∈A

which is a union of open sets, hence open by the first part. So F is closed. Similarly for
the finite union G of closed sets. 

Exercise I.1.9.
(a) Look at the above proof and make sure you understand where we used that
A is finite in the proof that V is open.
(b) Find an example of a countable family of open sets whose intersection is
not open.
(c) Find a countable family of closed sets whose union is not closed.

2. Bounded sets and compactness


Recall from real-variable analysis courses that continuous functions behave better on
closed bounded intervals than on arbitrary intervals. For example, a continuous function
on [a, b], where −∞ < a < b < ∞ is bounded and attains its bounds.
Subsets of C which are closed and bounded are similarly important.
Definition I.2.1. A subset S ⊂ C is bounded if S ⊂ D(0, R) for some R < ∞.
Otherwise, we say that S is unbounded.
Remark I.2.2. It would be equivalent to define boundedness in terms of closed discs
or discs of the form D(a, R), where a ∈ C. It is also equivalent to define it in terms of
putting S inside a large square or rectangle.
Examples I.2.3. The complex plane itelf is unbounded, as is the whole of the x axis
and the set Z ⊂ C. Half-planes are also unbounded.
If r < ∞, then D(a, r), D(a, r) and D∗ (a, r) are all bounded.
10 I. PLANE TOPOLOGY

Proposition I.2.4. A set S is bounded if and only if there exists M > 0 such that
for any z ∈ S,
− M 6 Re(z) 6 M and − M 6 Im(z) 6 M. (I.2.1)
In other words, S is bounded if and only if it can be enclosed in some large square.
Proof. Suppose that S is bounded. By definition, it is contained in some disc
D(0, R). So it is enough to show that there is some M > 0 such that (I.2.1) holds for
all z ∈ D(0, R).
See Figure I.2.5. Geometrically, the set of points satisfying (I.2.1) is a closed square
with horizontal and vertical sides, centre at 0 and of side-length 2M . It is clear that any
given disc will be contained in such a square if we are allowed to take M as large as we
like.
How about the converse? We are given that S is contained in a large square of
the above type, of side-length 2M . To show it is bounded, we need to show that it is
contained in some large disc. Again, geometrically, we just have to choose the disc so
large that it contains√the square. This can be done by taking the centre of the disc a = 0
an the radius to be 2M . 
Remark I.2.5. The main idea of the proof is in Figure I.2.5

Figure 1. A square inside a circle and a circle inside a square

Note that in this proof, we used the definition of boundedness in terms of containment
within a disc and checked that the new condition (containment within a square) was
equivalent to this definition.
Definition I.2.6. A subset K of C is called compact if and only if it is closed and
bounded.
Remark I.2.7. This is a very important and powerful definition. We hope to per-
suade you of this with two propositions. You have seen precursors of both in 1101.
Theorem I.2.8. Let K be compact and let zn ∈ K be any sequence. Then there is a
convergent subsequence znj with znj → z0 ∈ K.
2. BOUNDED SETS AND COMPACTNESS 11

Exercise I.2.10. The set D(0, 1) is bounded but not closed, so not compact.
Find a sequence of points in D(0, 1) with no limit point in D(0, 1). The set C is
closed but unbounded. Find a sequence of points in C with no limit point in C.

In fact, the theorem is if and only if.

Proof. Let zn = xn + iyn . Because K is bounded, we have −M 6 xn 6 M for


some M . By a theorem from 1101, the sequence of real numbers xn has a convergent
subsequence xnj . Similarly, the sequence ynj is bounded, hence has a convergent subse-
quence. So there is a convergent subsequence zn0j of zn with limit w, say, in C. But K
is closed, so contains all its limit points. Hence w ∈ K and zn has a subsequence which
converges to a point of K. 
Remark I.2.9. This is known as the Bolzano–Weierstrass or sequential compactness
theorem. Note that we used the 1-dimensional version of this theorem to construct the
proof. The proof extends readily to prove that if K is any closed and bounded subset of
Rn , then every sequence in K has a convergent subsequence.
Theorem I.2.10. If K is a compact subset of C and f : K → C is continuous, then
f is bounded, that is, there exists M such that |f (z)| 6 M for all z ∈ K.
Proof. If f is unbounded, for each n there is a point zn ∈ K such that |f (zn )| > n.
By the previous theorem, zn has a limit point w ∈ K.
Now f is continuous at w so we have a r > 0 such that
z ∈ K ∩ D(w, r) =⇒ |f (z) − f (w)| < 1. (I.2.2)
(We took ε = 1 in the definition of continuity at w.) So for all such z,
|f (z)| 6 |f (w)| + 1. (I.2.3)
Since zn → w, there exists N such that zn ∈ D(w, r) for all n > N . There is a
contradiction for the points zn with n > |f (w)| + 1. 

*Challenge Questions
(a) A more abstract and general definition of compactness is the following. K
is compact if any cover of K by open sets has a finite subcover. In other
words, whenever we have a family {Uα }α∈A of open subsets such that
[
K⊂ Uα ,
α∈A
one can find α1 , . . . , αN such that
K ⊂ Uα1 ∪ · · · ∪ UαN .
Show that it is equivalent to K being closed and bounded.
12 I. PLANE TOPOLOGY

(b) Let A and B be subsets of C. Let


dist(A, B) = inf{|a − b| : a ∈ A, b ∈ B}. (I.2.4)
Show that if A is compact, B is closed and A ∩ B = ∅, then dist(A, B) > 0.
Show that if A ∩ B = ∅ and A and B are merely closed, then it is possible
that dist(A, B) = 0.

3. The Riemann sphere


...or extended complex plane.
For many purposes of complex geometry and analysis, it is useful to introduce an
‘ideal point’ or ‘point at infinity’. This is not a number!
Algebraically, we adjoin a symbol ∞ and insist on the rules:
• a + ∞ = ∞ + a = ∞ if a ∈ C;
• a · ∞ = ∞ · a = ∞, a/0 = ∞ if a ∈ C, a 6= 0;
• a/∞ = 0 if a ∈ C
We cannot consistently define 0/0, ∞/∞ or ∞ + ∞.
The extended complex plane C∞ := C∪{∞} is just C with this ‘ideal point’ adjoined.
Again, C∞ is not a field or even a ring.

3.1. Stereographic projection. Nonetheless, C∞ can be pictured very nicely in


terms of the Riemann sphere S .
This is the unit sphere in 3D, x21 + x22 + x23 = 1. Denote by N the ‘north pole’, the
point with coordinates (0, 0, 1). The complex plane is identified with the plane x3 = 0,
so we make the point z = x + iy correspond with the point (x, y, 0).
Definition I.3.1. Stereographic projection σ : S \{N } → C is defined geometrically
by decreeing that for P ∈ S , N , P and σ(P ) are collinear.

You can play with a geogebra visualization of stereographic projection here


https://www.geogebra.org/m/AdgpuuT4.

Theorem I.3.2. The map σ defines a bijection between S \ {N } and C. We have


the formulae
p1 + ip2
σ(p1 , p2 , p3 ) = (I.3.1)
1 − p3
and
2 Re(z) 2 Im(z) |z|2 − 1
 
−1
σ (z) = , , (I.3.2)
1 + |z|2 1 + |z|2 |z|2 + 1

Proof. The line in 3-space which contains (0, 0, 1) and (p1 , p2 , p3 ) is given by
(x1 , x2 , x3 ) = (1 − t)(0, 0, 1) + t(p1 , p2 , p3 ), (t ∈ R). (I.3.3)
3. THE RIEMANN SPHERE 13

This is uniquely defined if (p1 , p2 , p3 ) 6= (0, 0, 1). It meets x3 = 0 when (1 − t) + tp3 = 0


i.e. t = 1/(1 − p3 ). For this value of t, we have
p1 + ip2
x1 + ix2 = , (I.3.4)
1 − p3
proving the first formula.
The key idea to proving the inverse formula is to take the mod-squared of (I.3.4),
giving
p2 + p22 1 − p23 1 + p3
|z|2 = 1 2
= 2
= (I.3.5)
(1 − p3 ) (1 − p3 ) 1 − p3
where we used p21 + p22 + p23 = 1 for the second equality. It is now easy to obtain the
formula for σ −1 .
The fact that we have constructed σ −1 means that we have the claimed bijection. 

Note that as (p1 , p2 , p3 ) → (0, 0, 1), |σ(p1 , p2 , p3 )| → ∞. This makes the following
reasonable:
Definition I.3.3. We extend σ to define a bijection between S and C∞ by setting
σ(N ) = ∞.

Exercise I.3.11.
(a) Show that the equation a + ∞ = ∞ is consistent with the definition σ(N ) =
∞ by calculating σ −1 (a + z) and letting |z| → ∞.
(b) What transformation of S corresponds to the map z 7→ −1/z?
Exercise I.3.12. The bijection between C∪∞ and a sphere is sometimes defined
in terms of stereographic projection of the unit sphere S 0 which is tangent to the
(x1 , x2 ) plane at the origin, i.e. with equation
x21 + x22 + (x3 − 1)2 = 1.
If the map is geometrically defined as before by taking the straight line joining the
north pole (0, 0, 2) to a variable point P on S 0 and defining σ 0 (P ) as the intersection
of this line with the plane x3 = 0, find the formula for this mapping analogous to
that of Theorem I.3.2.

One of the interesting properties of stereographic projection is that it maps circles


in S to circles or straight lines in C.
Theorem I.3.4. Any circle C on S maps under stereographic projection to a circle
in C if (0, 0, 1) 6∈ C or to a straight line in if (0, 0, 1) ∈ C.

Click on this link https://www.geogebra.org/m/rB4MEevn for a geogebra visu-


alization of stereographic projection of a family of circles.
14 I. PLANE TOPOLOGY

Proof. The circles on S are the (non-trivial) intersections with S of planes in R3 .


Such a plane Π has an equation of the form
a1 x1 + a2 x2 + a3 x3 = b (I.3.6)
where we may suppose
a21 + a22 + a23 = 1. (I.3.7)
With this condition imposed, Π meets in S if −1 6 b 6 1, this intersection being a
point if b = ±1 and a circle if inequality is strict.
Substituting the formula for inverse stereographic projection (I.3.2) into (I.3.6), we
obtain the equation
2a1 x + 2a2 y + a3 (x2 + y 2 − 1) = b(x2 + y 2 + 1) (I.3.8)
after clearing the denominator and writing z = x + iy. Rearranging, we obtain
(b − a3 )(x2 + y 2 ) − 2a1 x − 2a2 y + b + a3 = 0. (I.3.9)
This is a circle if b 6= a3 , with

a1 + ia2 1 − b2
centre = , radius = . (I.3.10)
b − a3 |b − a3 |
Note that the radius is positive because |b| < 1. The condition a3 = b is precisely the
condition that (0, 0, 1) ∈ P and this is the case that (I.3.9) reduces to the straight line
2a1 x + 2a2 y − b − a3 = 0. (I.3.11)


An important heuristic here is that straight lines in C are ‘ideal’ circles. From the
point of view of S , most circles map by σ to circles in C; it’s just we get a straight line
in C as the image if the original circle happens to pass through the north pole of S .
This idea is further illustrated by the following example.

Exercise I.3.13. For fixed real λ, consider the locus


|z + 1| = λ|z − 1|
Describe this locus for all positive λ. What happens when λ = 1, λ → 0, λ → ∞?
Click on this link https://www.geogebra.org/m/a9aAjZRz for a geogebra visualiza-
tion of this family of circles.

3.2. Distance on S . The sphere S inherits from R3 a notion of distance: if P


and Q are points of S , we have
|P − Q|2 = (p1 − q1 )2 + (p2 − q2 )2 + (p3 − q3 )2 . (I.3.12)
where (p1 , p2 , p3 ) and (q1 , q2 , q3 ) are the cartesian coordinates of P and Q.
Then we have:
3. THE RIEMANN SPHERE 15

Proposition I.3.5. Suppose P, Q ∈ S \ {N } and


σ(P ) = z, σ(Q) = w. (I.3.13)
Then
2|z − w|
|P − Q| = p p . (I.3.14)
1 + |z|2 1 + |w|2

Proof. For the moment denote σ(P ) = P 0 and σ(Q) = Q0 . From the definition
of σ it is a pleasant exercise to show that the triangles N P Q and N Q0 P 0 are similar.
Moreover,
|N P | p p
0
= 1 − p 3 1 − q3 . (I.3.15)
|N Q |
It follows that
|P Q| p p
= 1 − p 3 1 − q3 . (I.3.16)
|z − w|
Since
1 − p3 = 2/(1 + |z|2 ) (I.3.17)
and similarly for 1 − q3 (see (I.3.2)), the result follows. 

Exercise I.3.14. Carry out the ‘pleasant exercise’ mentioned in the proof.

3.3. Continuity on S . If F : S → C is a function, and P ∈ S is a point, we say


that F is continuous at P if
lim F (Q) = F (P ) (I.3.18)
Q→P

(Here it is understood that Q → P means that Q ∈ S and |P − Q| → 0, the distance


being defined by (I.3.12).)
Proposition I.3.6. Stereographic projection σ is continuous at all points of S \{N }
and its inverse σ −1 is continuous at all points z ∈ C.
Proof. Let z = σ(P ), w = σ(Q). For the continuity of σ, we have to show that if
|Q − P | is small then so is |z − w|. For the continuity of σ −1 we need to show that if
|z − w| is small, then so is |P − Q|. Both follow from Prop I.3.5. 

Continuity of a function on S at N also has a nice interpretation in C:


Proposition I.3.7. Suppose that F : S → C is continuous and let F (N ) = a. Let
f : C∞ → C be defined as follows:
f (z) = F (σ −1 (z)), f (∞) = F (N ). (I.3.19)
Then f is continuous in C and lim|z|→∞ f (z) = F (N ).
16 I. PLANE TOPOLOGY

Proof. The continuity of f at ‘finite points’ z ∈ C follows from Proposition I.3.6,


for F = f ◦ σ −1 is a composite of continuous functions. To prove the other part, note
that if z = σ(P ) then
q √ p 2
|N − P | = p21 + p22 + (p3 − 1)2 = 2 1 − p3 = p . (I.3.20)
1 + |z|2
(See (I.3.17) for the last equality.)
Hence |N − P | → 0 if and only if |z| → ∞. Thus F (P ) → F (N ) as |N − P | → 0
(continuity of F at N ) if and only if f (z) → F (N ) as |z| → ∞. 
3.4. Möbius transformations.
Definition I.3.8. Let a, b, c, d ∈ C, ad − bc 6= 0. Then the Möbius transformation
with coefficients a, b, c, d is
az + b
T (z) = . (I.3.21)
cz + d
Initially T is defined for all z 6= −d/c, but we always extend T to a transformation
C∞ → C∞ by defining
T (∞) = a/c, T (−d/c) = ∞. (I.3.22)
0 0 0 0
Two sets of coefficients (a, b, c, d) and (a , b , c , d ) define the same Möbius transformation
T if and only if there exists λ ∈ C \ {0} such that
a0 = λa, b0 = λb, c0 = λc, d0 = λd. (I.3.23)
Möbius transformations are sometimes also called fractional linear transformations.
Note that we can always assume ad − bc = 1 by multiplying all coefficients by a well
chosen constant. With this normalization, (a, b, c, d) will be determined up to overall
sign, (a, b, c, d) → (−a, −b, −c, −d).
Remark I.3.9. We have extended the definition of σ to define a bijection S → C∞ .
Using this bijection, we may equally picture a Möbius transformation as a transformation
S → S . Thus we treat S and C∞ as more-or-less interchangeable via stereographic
projection σ.
The set of all Möbius transformations forms a group under composition which we
denote by Aut(S ). In particular, the inverse of (I.3.22) is
dz − b
T −1 (z) = . (I.3.24)
−cz + a
We shall use the group property in the proofs of the next couple of results.

Exercise I.3.15. Describe, in terms of simple geometric transformations of the


plane, the transformation T (z) = az + b, where a 6= 0 and b ∈ C.

Proposition I.3.10. If (z1 , z2 , z3 ) is an ordered set of (pairwise distinct) points of


C ∪ ∞, then there is a unique Möbius transformation T with
T (z1 ) = 1, T (z2 ) = 0, T (z3 ) = ∞. (I.3.25)
3. THE RIEMANN SPHERE 17

Proof. Suppose first that none of the zj is equal to ∞. Then if T (z2 ) = 0 and
T (z3 ) = ∞, then T must have the form
z − z2
T (z) = λ . (I.3.26)
z − z3
Then
z1 − z2
T (z1 ) = λ , (I.3.27)
z1 − z3
Setting this to 1, we obtain
z 1 − z3 z − z2
T (z) = . (I.3.28)
z1 − z2 z − z3
It is clear from the construction that T is unique.
If any one of the zj is equal to ∞, we obtain the corresponding T by taking suitable
limits of (I.3.28). For example, if z1 = ∞, we get
z − z2
T (z) = ,
z − z3
if z2 = ∞ we get
z1 − z3
T (z) =
z − z3
and if z3 = ∞, we get
z − z2
T (z) = .
z1 − z2

Corollary I.3.11. If (z1 , z2 , z3 ) and (w1 , w2 , w3 ) are two ordered triples of points
in C∞ , with no two of the zj equal to each other and no two of the wj equal to each
other, then there is a unique Möbius transformation T such that
T (zj ) = wj for j = 1, 2, 3. (I.3.29)

Proof. From the previous Theorem, let S1 map (z1 , z2 , z3 ) to (1, 0, ∞) and let S2
map (w1 , w2 , w3 ) to (1, 0, ∞). Then T = S2−1 ◦ S1 does the job. 

Exercise I.3.16.
(a) What is the most general Möbius transformation which maps 0 to ∞?
(b) What is the most general Möbius transformation which maps the upper
half-plane H = {z ∈ C : Im(z) > 0} to itself?
Exercise I.3.17. *For enthusiasts: Show that any Möbius transformation
viewed as a mapping T : S → S (cf. Remark I.3.9) is continuous.

Theorem I.3.12. Let C be any circle on S and let T be any Möbius transformation.
Then T (C) is again a circle on S .
18 I. PLANE TOPOLOGY

Proof. We know that C corresponds under stereographic projection to a circle or


straight line.
Now if T (z) = az + b, T is geometrically a composition of rotation, enlargement
and translation. These transformations all map circles and straight lines to circles and
straight lines.
If w = S(z) = 1/z and C is the circle |z − α| = r, then the image circle has the
equation
|1/w − α| = r. (I.3.30)
This can be rearranged in the form
1 − αz − αz + (|α|2 − r2 )|z|2 = 0.
This is a circle unless |α| = r in which case it is a straight line. Similarly, if C is a
straight line of the form λz + λz = t (real), then a similar calculation shows that S(C)
is either a circle or a straight line.
Now, we build a general Möbius as a composition of two of the form az + b and
S(z) = 1/z. If
az + b
T (a) =
cz + d
The result is already proved if c = 0, so assume c 6= 0. then
az + b (a/c)(cz + d) + b − ad/c a 1 ad − bc
= = − .
cz + d cz + d c c cz + d
Thus if T1 (z) = cz + d and T2 (z) = − ad−bc 1
c z + c , we see that

T (z) = T2 ◦ S ◦ T1 (z).
Since each of the three transformations maps circles/straight lines to circles/straight
lines, the result is true also of their composite T . 

Exercise I.3.18. Make sure you understand why, geometrically, the image of
the circle C under S(z) = 1/z is a straight line if |α|2 = r2 .
Exercise I.3.19. *For enthusiasts:
Show that if T has the form
az + b
T (z) = , |a|2 + |b|2 = 1, (I.3.31)
−bz + a
then T acts as a rotation of S .

4. Möbius transformations with prescribed properties


We know that a Möbius transformation T maps circles and straight lines in C to
circles and straight lines in C. Equivalently (since σ maps circles in S to circles and
straight lines in C) T , regarded as a map S → S , maps circles to circles.
4. MÖBIUS TRANSFORMATIONS WITH PRESCRIBED PROPERTIES 19

Can we build a picture of what T does? To discuss this, it is convenient to imagine


having two ‘copies’ of C which we shall refer to as the z-plane and the w-plane. Then
we think of the equation
w = T (z) (I.4.1)
as setting up a bijection or transformation of the z-plane to the w-plane. Since T is
invertible, we can also think of fixing w, so that (I.4.1) determines z implicitly.
4.1. Families of circles. If T (z) does not have the simple form az + b, we can
write (I.4.1) as
z−α
w=λ (I.4.2)
z−β
where λ 6= 0 and α 6= β. Consider the family of straight lines through the origin of the
w-plane. If such a line is inclined at angle θ with the real axis, it will have the form
arg(w) = θ. (I.4.3)
Strictly speaking, this defines a ray rather than a line. The full straight line is given by
the union of the sets
{w 6= 0 : arg(w) = θ} ∪ {w = 0} ∪ {w 6= 0 : arg(w) = θ + π}. (I.4.4)
Since T −1 maps this straight line to a circle or a straight line, we have
Proposition I.4.1. For any real number θ, the set
   
z−α
Aθ = z ∈ C : arg λ =θ (I.4.5)
z−β
is a circular arc, or possibly a straight line, containing α and β in its closure. The arcs
Aθ and Aθ+π are the two arcs of a circle containing α and β (degenerating to parts of
the straight line joining α and β if θ = 0 or θ = π).
Proof. If θ = 0, the equation is
arg(z − α) = arg(z − β).
This means that z is on the straight line joining α to β and not on the segment between
them. If θ = π, similarly, z is on the line-segment joining α to β.
In every other case Aθ is part of a circle, as has already been argued. Strictly
speaking the condition does not make sense if z = α or z = β. But letting w approach
0 along a fixed ray corresponds to z → α and letting w approach ∞ along the same ray
corresponds to z → β. This verifies the first statement of the Proposition. We’ve seen
that the straight line through 0 inclined at angle θ to the real axis is the union of sets
(I.4.5). This straight line maps to a circle which is the union of Aθ , Aθ+π (and the point
α). 

Exercise I.4.20. What is the geometric interpretation of the condition


arg ((z − α)/(z − β)) = θ in terms of angles? Does this help you to understand
why Aθ is a circular arc?
20 I. PLANE TOPOLOGY

There is another family of circles in the w-plane whose images are easy to understand:
the concentric circles |w| = r. This locus corresponds in the z-plane to a circle, and so
we have
Proposition I.4.2. For any r > 0, the set
 
z−α r
Br = z ∈ C : = (I.4.6)
z−β |λ|
is a circle in the z-plane (unless r = |λ|, in which case it is a straight line).

Click on this link for a geogebra picture of a family of these orthogonal circles
https://www.geogebra.org/m/z8H4mPWh

The circles Br are called circles of Apollonius. The A- and B- circles form a ‘net’ in
the z-plane. Every point z (6= α, β) lies on the intersection of precisely one A- and one
B-circle. Moreover, every A-circle and every B-circle meet orthogonally.
The first of these claims is clear because the corresponding statement is clear in the
w-plane, namely that every point w 6= 0 lies on a unique ray and a unique circle with
centre the origin. (Polar coordinates!) The second claim about orthogonality will follow
once we have proved that Möbius transformations are conformal or angle-preserving, for
in the w-plane rays through 0 cut the circles |w| = r orthogonally. For now we take it
on trust or take the time to prove it directly.

Exercise I.4.21. Prove that for any r > 0 and θ, the circles Aθ and Br meet
each other orthogonally.

Key point: If α and β are two points of C then the Möbius transformation
w = (z − α)/(z − β) maps any circle containing α and β to a straight line through
the origin in the w-plane.

C1 C2
−1 z2 R z1 1
0

Figure 2. The region R defined by the intersection of two discs


4. MÖBIUS TRANSFORMATIONS WITH PRESCRIBED PROPERTIES 21

4.2. Applications.
Example I.4.3. Suppose it is desired to map the ‘lens-shaped region’
√ √
R = {|z − 1| < 2} ∩ {|z + 1| < 2} (I.4.7)
(Figure 2) by a Möbius transformation to a sector of the w-plane. There are several
ways to do this: the following is perhaps the simplest.
If R is to map to a sector, then we need C1 and C2 to map to straight lines through
0. From the previous section, a Möbius of the form w = (z − α)/(z − β) will map a
circle to a straight line if α and β are any two points on the circle. If we want the same
Möbius to map both C1 and C2 to straight lines, then we had better choose α and β to
be on both of C1 and C2 . From the picture, the intersection points of C1 and C2 are i
and −i.
So our Möbius transformation will be
z−i
w = T (z) = . (I.4.8)
z+i
The images of C1 and C2 under T are straight lines through the origin. Which ones? To
find out, we compute the images of two other points, one on C1 and one on C2 : since the
images of C1 and C2 are definitely going to be straight lines through 0, knowing these
images will determine these lines. We
√ shall follow this through.

Convenient points are z1 = 1 − 2 on C1 and z2 = 2 − 1 on C2 . We calculate

√ 1− 2−i √ −1 + i
T (z1 ) = T (1 − 2) = √ = 2( 2 − 1) √ (I.4.9)
1− 2+i 1 + ( 2 − 1)2
and √
√ 2−1−i √ −1 − i
T (z2 ) = T ( 2 − 1) = √ = 2( 2 − 1) √ (I.4.10)
2−1+i 1 + ( 2 − 1)2
So w1 lies on the ray arg(w) = 3π/4 and w2 lies on the ray arg(w) = 5π/4. Thus the
images of C1 and C2 are the lines y = x and y = −x with the arcs bounding R mapping
to the parts of these lines in the left half-plane: see Figure 3.
The original circles C1 and C2 cut the plane up into 4 pieces: the region R, the
complement of R in each of the two discs, and the exterior region. These four regions
are mapped by T to the four sectors cut out by the two lines y = x and y = −x. The
image of R must be one of these sectors, and by the above discussion, the evidence is
that R will map to the region shaded in Figure 3,
T (R) = {w : 3π/4 < arg(w) < 5π/4}. (I.4.11)
This is confirmed by computing T (0) (or T of any other point of R—but 0 is always
easiest if available!). We have T (0) = (−i)/(i) = −1 which lies in (I.4.11) as expected.
Note that the angle between T (C1 ) and T (C2 ) is π/2. This is consistent with the
angle between C1 and C2 being π/2 at their two intersection points. (This can be seen
easily from the picture.)
We can map R to the standard positive quadrant 0 < arg(w) < π/2 by applying
a clockwise rotation which maps arg(w) = 3π/4 to the real axis. This is achieved by
22 I. PLANE TOPOLOGY

arg(w) = 3π/4

T (C1 )

T (0) = −1
0

T (C2 )

arg(w) = 5π/4

Figure 3. The image T (R) in the w-plane



multiplying by e−3πi/4 = (−1 + √ i)/ 2. Since we don’t mind if we apply a real scaling as
well, we can drop the factor of 2, just multiply by −1 + i and define
z−i
w = T1 (z) = −(1 + i)
z+i
This maps C1 to the real axis in the w-plane, C2 to the imaginary axis in the w-plane
and 0 to the point 1 + i in the first quadrant.

Exercise I.4.22. Work through this example with S(z) = z+i z−i instead of T (z).
Just to convince yourself that the order of the intersection points of the arcs doesn’t
matter too much.

Conclusion: To map a region bounded by two circular arcs C1 and C2 to a sector,


use the Möbius transformation T (z) = (z − α)/(z − β), where α and β are the
intersection points of C1 and C2 .

Example I.4.4. Suppose it is required to map the half-space H above the line L =
{y = 2x} in the z-plane to the unit disc in the w-plane.
One possibility is to pick three points on the line y = 2x and three points on the
unit circle, and use Corollary I.3.11 to write down a Möbius transformation S which
maps the line to the circle. S will either map H to the inside or the outside of the unit
circle. If it maps to the inside, our work is done. If note, replace S(z) by 1/S(z), since
w 7→ 1/w interchanges the inside and outside of the unit circle in the w-plane.
A more elegant solution is to realise the line as the set of points equidistant from two
auxiliary points in the z-plane. For example, the line M , 2y + x = 0 cuts L orthogonally
at 0. If we parameterize M in the form
z(t) = (2 − i)t, (t ∈ R) (I.4.12)
4. MÖBIUS TRANSFORMATIONS WITH PRESCRIBED PROPERTIES 23

L : y = 2x

z2 = −2 + i

0
M : 2y + x = 0
z1 = 2 − i

Figure 4. The geometry for the construction in Example I.4.4. The


half-space H is shaded.

then for any pair of points z(t), z(−t) (t 6= 0), L is the set of points equidistant from
z(t) and z(−t). For example if t = ±1, we obtain the points
z1 = z(1) = 2 − i, z2 = z(−1) = −2 + i (I.4.13)
and
L = {z ∈ C : |z − 2 + i| = |z + 2 − i|}. (I.4.14)
(See Figure 4.)
The two open half-spaces determined by L arise by replacing = with < or > in
(I.4.14). To see
√ which is which, note that z = i is in the half-space above the line and
|i − 2 + i| = 2 2 whereas |i + 2 − i| = 2. Thus
H = {z ∈ C : |z − 2 + i| > |z + 2 − i|}. (I.4.15)
It follows at once from this that if
z+2−i
w= (I.4.16)
z−2+i
then |w| < 1 corresponds precisely to z ∈ H. So (I.4.16) is a Möbius transformation
that does the job.
Remark I.4.5. Note that this Möbius transformation is not unique: we get one for
every choice of points z1 and z2 for which it is true that
H = {z ∈ C : |z − z1 | < |z − z2 |} (I.4.17)

Exercise I.4.23. Find the most general Möbius transformation which maps the
half-space H of Example I.4.4 onto the open unit disc in the w-plane.
24 I. PLANE TOPOLOGY

Exercise I.4.24.
(a) Find a Möbius transformation which maps the right half-plane {z : Re(z) >
0} onto the open unit disc {w : |w| < 1}.
(b) Find a Möbius transformation which maps the set
{z : Im(z) > 0, |z| < 1}
onto an open sector of the w-plane.
(c) Find a Möbius transformation which maps the half-space G above the line
x+y = 1 in the z-plane onto the open unit disc {w : |w| < 1} in the w-plane.

Remark I.4.6. When it comes to these kinds of mapping problems, the transforma-
tion is generally not unique. If you compare your answer with a friend’s, they may look
different yet still both be right.
There are also several different ways of solving any given problem.
A general principle is that if you have a region bounded by two circular arcs, find the
intersection points α and β, say. Then the image in the w-plane by w = (z − α)/(z − β)
will be a sector, with the intersection at α being mapped to w = 0. Figuring out exactly
which sector it is can be done in different ways, see the above example.
When it comes to mapping a half-plane G to the unit disc, you can proceed as in
the worked example. Another way is to map G by a combination of a rotation and a
translation, to the standard upper half-plane {z 0 : Im(z 0 ) > 0}. This can always be
0 −i
accomplished by a transformation of the form z 0 = az + b. Now w = zz 0 +i is a standard
0
Möbius transformation from {Im(z ) > 0} onto {|w| < 1}. This is an application of the
principle of solving a problem by breaking it up into manageable pieces, which is often
a good ploy.
Remark I.4.7. Using a Möbius transformation to transform figures in the z-plane to
different figures in the w-plane is an example of conformal mapping. (In mathematics,
‘conformal’ means ‘angle-preserving’.) We shall return to the topic later in the course.
It has many applications in mathematics and ‘physical applied mathematics’.
CHAPTER II

Analytic functions and holomorphic functions

1. Analytic functions
Let Ω be an open set of C. We shall define A (Ω), the set of analytic functions in
Ω to be the set of functions f which are ‘locally expandable in power series’ about any
point of Ω.
The space A (Ω) is a vector space, and in fact a ring. Moreover, if f 6= 0 in Ω then
1/f ∈ A (Ω) also.

1.1. Power series. The power series (centred at 0) with coefficients (a0 , a1 , . . .) (a
sequence of complex numbers) is by definition

X
f (z) = an z n (II.1.1)
n=0

Recall that any such power series has a radius of convergence r ∈ [0, ∞] with the
property that (II.1.1) is convergent for |z| < r and divergent for |z| > r. In fact,
Proposition II.1.1. For each s let M (s) = sup{|an |sn : n = 0, 1, . . .}, so that
M (s) ∈ [0, ∞]. Then the radius of convergence of (II.1.1) is equal to
r := sup{s : M (s) < +∞}. (II.1.2)
Examples II.1.2. The radius of convergence of each of
∞ ∞ ∞
X
n
X zn X zn
nz , , (II.1.3)
n n2
n=1 n=1 n=1

is r = 1. The radius of convergence of the exponential series



X zn
(II.1.4)
n!
n=0
is r = +∞.

Exercise II.1.1.
(a) Calculate M (s) for each of the three series in (II.1.3). Note that M (1) is
not always finite.
(b) Discuss the convergence of these three series at points z with |z| = 1.

25
26 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

In general nothing can be said about convergence of a power series at points on


the boundary of its disc of convergence.

For the proof of the following important fact, we refer the reader to a previous course.
Theorem II.1.3. Let (II.1.1) have radius of convergence r > 0. Then f (z) is con-
tinuous for all z ∈ D(0, r).
Actually, this was probably only proved for real power series in a real variable x.
The proof, however, is the same, with a little care putting in | · | signs appropriately.
Remark II.1.4. If z0 is any point of C we can also consider power series centred at
z0 and the e
an are complex coefficients,
X
fe(z) = an (z − z0 )n
e (II.1.5)
centred at z0 . The radius of convergence re is defined using M (s) and the quantities
an |sn exactly as before. (II.1.5) will converge in the disc D(z0 , re), where re is the radius
|e
of convergence.

P∞ n
Exercise II.1.2. If n=0 an z has radius of convergence r, write down the disc
of convergence of
∞  n
X z−1
an .
10
n=0

Definition II.1.5. Let Ω be an open set of C. A function f : Ω → C is called


analytic if for each point z0 ∈ Ω, we have

X
f (z) = an (z − z0 )n , (II.1.6)
n=0
in some disc D(z0 , δ) ⊂ Ω.
The set of all analytic functions defined in Ω is denoted A (Ω).
Remark II.1.6.
(a) It is part of the definition that the power series in (II.1.6) has a positive radius
of convergence.
(b) In (II.1.6) the coefficients an will generally (almost always) depend upon the
point z0 , as will the number δ. Even f (z) = z has this property.
Remark II.1.7. A soundbite way of describing this definition is to say that f is
analytic if it is ‘locally expandable in a power series’ around every point (of its domain
of definition).

Exercise II.1.3.
1. ANALYTIC FUNCTIONS 27

(a) Show explicitly that 1 + 6z − 10z 2 is analytic in C. [Look at the definition.


What do you need to check?]
(b) Generalize your proof to show that any polynomial
p(z) = a0 + a1 z + · · · + an z n
(where the coefficients aj are given complex numbers) lies in A (C).
(c) Show that f (z) = 1/z is analytic in C \ {0}.

The next result shows that a convergent power series is analytic in its disc of con-
vergence. This may seem obvious, but is not quite trivial.
Proposition II.1.8. Let f (z) = ∞ n
P
n=0 an z be convergent in D(0, r), r > 0. Then
f ∈ A (D(0, r)).
Proof. From the definition, we need to check that for any given point z0 , f can
be expanded in a convergent power series centred at z0 . This follows from the absolute
convergence of power series in their disc of convergence. Write z = z0 + w. Then by the
binomial theorem,

X
f (z) = an (z0 + w)n (II.1.7)
n=0
∞ Xn  
X n n−j j
= an z w . (II.1.8)
j 0
n=0 j=0

Since the sum is absolutely convergent for |z0 + w| < r, we may rearrange at will, to
obtain
1
f (z0 + w) = f (z0 ) + f 0 (z0 )w + f 00 (z0 )w2 + · · · (II.1.9)
2!
where

n(n − 1) · · · (n − j + 1)an z0n−j .
X
(j)
f (z0 ) = (II.1.10)
n=j

Here we just collected all terms multiplying wj in (II.1.9). (The notation in (II.1.10) will
be justified as the j-th derivative of f at z = z0 in due course.) Then (II.1.22) converges
for |z0 + w| < r, in particular for w ∈ D(z0 , r − |z0 |). 
Of course, there is a precisely similar statement for a power series centred at some
point a rather than 0.

Examples II.1.9.
(a) Any polynomial
p(z) = a0 + a1 z + · · · + an z n
is analytic in C.
28 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

(b) Any rational function


p(z)
R(z) = ,
q(z)
where p and q are polynomials without common factors (and with q not
identically zero) is analytic in the set Ω = C \ {z : q(z) = 0}. In particular
the Möbius transformation T (z) = az+bcz+d is analytic in the complement of
the point −d/c.
(c) Elementary transcendental functions like exp z, sin z, cos z, are analytic in
C.
(d) Functions like tan z, sec z are analytic at all points where they are defined
(so in the case of these two, at all points apart from z = (n + 1/2)π.
(e) As per Proposition II.1.8, any convergent power series is analytic in its disc
of convergence.

Statements (b) and (d) will follow from Theorem II.1.10 below.
For (c), recall the formulae

z
X zn
exp(z) = e = (II.1.11)
n!
n=0

and
eiz − e−iz eiz + e−iz
sin z = , cos z = . (II.1.12)
2i 2
Combining (II.1.11) with (II.1.12), we have
∞ ∞
X (−1)n z 2n+1 X (−1)n z 2n
sin z = , cos z = . (II.1.13)
(2n + 1)! (2n)!
n=0 n=0

These power series, like (II.1.11), are convergent in the whole of C, showing that sin z
and cos z are analytic in C by Proposition II.1.8.
The power series expansion of tan z at z = 0 is much more subtle: the coefficients
involve the so-called Bernoulli numbers.
The complex logarithm log z, fractional powers such as z 1/2 and even more generally
complex powers z a where a ∈ C are tricky to deal with. In some sense they want to be
analytic where defined, but they are ‘multi-valued’ and this causes serious problems (or
leads to new and fascinating mathematics, depending upon your point of view).
The next theorem shows that analytic functions are plentiful and that they behave
well under the usual algebraic operations—as long as you don’t divide by zero.

Theorem II.1.10. For any open set Ω, A (Ω) is an infinite-dimensional vector space.
It is also a ring in the sense that f, g ∈ A (Ω) implies f g ∈ A (Ω) (pointwise product).
Furthermore, if f ∈ A (Ω) and f (z) 6= 0 for all z ∈ Ω, then 1/f ∈ A (Ω).

Proof.
1. ANALYTIC FUNCTIONS 29

(a) Suppose that f, g ∈ A (Ω) and let z0 ∈ Ω. We know that f and g are given by
convergent power series

X ∞
X
f (z) = an (z − z0 )n , g(z) = bn (z − z0 )n (II.1.14)
n=0 n=0

in D(z0 , δ), for some δ > 0. (If the expansion for f is valid in D(z0 , δ1 ) and the
expansion for g is valid in D(z0 , δ2 ), then we take δ to be the smaller of δ1 and
δ2 .)
Then it is clear that
X∞
λf (z) + µg(z) = (λan + µbn )(z − z0 )n (II.1.15)
n=0
for any complex numbers λ and µ, and the RHS is convergent in the same disc
D(z0 , δ). Since z0 was an arbitrary point of Ω, this implies that λf +µg ∈ A (Ω).
Thus A (Ω) is a vector space. (Clearly the zero-function 0 lies in A (Ω).
(b) We have seen that every polynomial
p(z) = a0 + · · · + an z n
is analytic in the whole of C and hence in A (Ω). The monomials 1, z, z 2 , . . . are
linearly independent, and so the vector space A (Ω) is infinite-dimensional.
(c) In a previous course, we saw the Cauchy Product formula for power series. With
f and g as in (II.1.14) this says that

X
f (z)g(z) = cn (z − z0 )n (II.1.16)
n=0
where
n
X
cn = aj bn−j (II.1.17)
j=0
and (II.1.15) is convergent in D(z0 , δ). It follows that f g (the pointwise product)
is in A (Ω).
(d) Finally, suppose that f ∈ A (Ω) is non-zero everywhere in Ω. As before suppose
X
f (z) = an (z − z0 )n , (II.1.18)
convergent in D(z0 , δ). We know f (z0 ) = a0 6= 0. The argument is simpler if
we replace f (z) by f (z)/f (z0 ) and z − z0 by w. This reduces us to considering

X
g(w) = 1 + an w n (II.1.19)
n=1

which is supposed convergent in D(0, δ). If we can find a power series expansion
of 1/g(w) valid in some D(0, δ 0 ), then we can also find a power series expansion
of 1/f valid in D(z0 , δ 0 ). P
Suppose that h(w) = ∞ n
n=0 bn w satisfies
h(w)g(w) = 1. (II.1.20)
30 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

Then we know what the bn have to be in terms of the an by the Cauchy product
formula. Indeed we have
n
X
a0 b0 = 1, a0 b1 + a1 b0 = 0, . . . , a0 bn + aj bn−j = 0. (II.1.21)
j=1

since the coefficients of the expansion of the function 1 are c0 = 1, cn = 0 for


n > 1. We have assumed a0 = 1 so it follows that b0 = 1, b1 = −a1 and
n
X
bn = − aj bn−j for n > 1. (II.1.22)
j=1

Notice that (II.1.22) determines bn in terms of a1 , . . . , an and b0 , . . . , bn−1 . Thus


if b0 , . . . , bn−1 are known then (II.1.22) determines bn uniquely, so that the bn
are uniquely determined by the an .
*For enthusiasts: It remains to show that 1 + ∞ n
P
n=1 bn w has a positive
radius of convergence. How can we prove such a thing? We need to show that
for some positive s, |bn |sn is uniformly bounded in n. Since b0 = 1, let us
suppose, aiming for an inductive proof, that there exists s > 0 such that
|bn |sn 6 1 for n = 0, . . . , m − 1. (II.1.23)
This is true for every s if m = 1, so let us see whether we can choose s so that
if (II.1.23) is true, then it is also true with m replaced by m + 1.
Now we know that (II.1.19) is convergent so that there exists M such that
|an |rn 6 M (II.1.24)
for some positive r and all n. (In fact any r < δ will do. r = δ will not do!)
Now we can use the inductive formula for bm in terms of the an and the previous
bj :
Xm m
X
m m
|bm s | 6 |aj bm−j |s = (|aj |sj ))(|bm−j |sm−j ). (II.1.25)
j=1 j=1

Using the inductive hypothesis to bound |bm−j |sm−j by 1 we obtain


m
X m
X
m j
|bm |s 6 |aj |s 6 M (s/r)j . (II.1.26)
j=1 j=1

Now if s < r, this geometric progression is bounded by its sum to ∞, so


s
|bm |sm 6 M . (II.1.27)
r−s
s
If we pick s so small that M r−s 6 1 (keeping 0 < s < r), then we obtain
(II.1.23) with m replaced by m + 1. Notice that this choice is independent of
m: it is determined entirely by the bound M and r. Thus there is s > 0,
indepdendent of n, such that (II.1.23)Pholds for all m. This s gives a lower
bound for the radius of convergence of bn wn .

2. HOLOMORPHIC FUNCTIONS—COMPLEX DIFFERENTIABILITY 31

Remark II.1.11. The proof of (d) is a little tricky but there are some important
ideas in it that I want to draw your attention to. Note first of all the use of simplifying
assumptions, passing from f (z) to g(w). These were not essential, but they save on
notational complication later in the proof.
The idea of the rest of the proof can be summarized as follows: the Cauchy product
tells us what the coefficients of a power series expansion of 1/g(w) must be, which is a
good start, but then we need to see that these coefficients give a series which actually
converges for some non-zero value of w. For this we use the convergence of g(w), which
gives the estimates (II.1.24) and feed these in to the inductive definition of the coefficients
of the expansion of 1/g.
In manipulations with power series, bounding a series by a geometric series, which
we did in (II.1.26) is a very common technical device (which goes back at least as far as
Cauchy!) Note also that we could not have reached our conclusion with s = r; taking
s < r gave us extra room to manoeuvre.
In a similar vein, note the remark that we could not take r = δ, the radius of
convergence of g. The numbers |an |rn may well be unbounded if r is the radius of
convergence (see Exercise II.1.1). One always needs to come in just a little bit from the
boundary of the disc of convergence to be safe!
At first, it may be surprising that the radius of convergence of h = 1/g may be smaller
than that of g. But note the following example: If g(w) = 1 + w, which converges
everywhere, then the radius of convergence is ∞, but the power series for 1/(1 + w)
(centred at w = 0) converges only in D(0, 1).

Exercise II.1.4. Can you construct a power series centred at 0, with f (0) = 1
and an infinite radius of convergence, but such that the power series for 1/f (z) has
radius of convergence 1/2? 1/10? Any number δ > 0? (Try making small changes
to the example just given.)
Exercise II.1.5. *For enthusiasts:
Prove Hadamard’s characterization of the radius of convergence:
Theorem II.1.12. For the power series ∞ n
P
n=0 an z , we have the formula
1
= lim sup |an |1/n
r n→∞
for the radius of convergence r.

2. Holomorphic functions—complex differentiability


We have introduced the class of analytic functions of z. Now we consider an appar-
ently completely different class of functions.
Definition II.2.1. Let Ω be an open set of C and let f : Ω → C be a complex-valued
function. We say that f is complex-differentiable (or C-differentiable) at the point z0 of
Ω, if there exists a complex number A such that
ε(h) := f (z0 + h) − f (z0 ) − Ah (II.2.1)
32 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

has the property


|ε(h)|
→ 0 as |h| → 0. (II.2.2)
|h|
(Generally ε(h) will also depend upon z0 , but we do not make this dependence explicit
unless it is absolutely necessary.) If this is the case, A is called the derivative of f at
z0 and denoted f 0 (z0 ). f is said to be holomorphic in Ω if it is complex-differentiable at
every point of Ω. Then we have
f (z0 + h) = f (z0 ) + f 0 (z0 )h + ε(z0 , h) whenever z0 , z0 + h ∈ Ω (II.2.3)
and
|ε(z0 , h)/h| → 0 as h → 0.
The set of holomorphic functions in Ω is denoted H(Ω).
An equivalent definition is in terms of difference quotients: f is complex-differentiable
at z0 if and only if
f (z0 + h) − f (z0 )
lim
h→0 h
exists. Whichever way we think about it, a crucial point is that the limit exists no
matter how h goes to 0. In a vertical or horizontal direction, or on some curve that is
spiralling in: we are always supposed to get the same answer.

z0 + h

z0 z0 + h z0
z0 + h
z0 + h
0 0

Figure 1. Direction-independence: for f to be complex-differentiable,


(f (z0 + h) − f (z0 ))/h has the same limit however z0 + h approaches z0 .

Notation II.2.2 (Big O, little o.). If f and g are functions, g real-valued and non-
negative, we write f = O(g) to mean that |f | is bounded by a multiple of g. More
precisely, if S is a set we write
f (z) = O(g(z)) for z ∈ S (II.2.4)
if and only if there is some constant C, independent of z, such that
|f (z)| 6 Cg(z) for all z ∈ S. (II.2.5)
The notation is often used for limits:
f (z) = O(g(z)) as z → z0 (II.2.6)
2. HOLOMORPHIC FUNCTIONS—COMPLEX DIFFERENTIABILITY 33

if and only if, for any sufficiently small δ > 0, there is a constant C, independent of z,
such that
|f (z)| 6 Cg(z) for all |z − z0 | < δ. (II.2.7)
We write f = o(g) as z → z0 if
f (z)
lim =0 (II.2.8)
z→z0 g(z)

Note carefully the difference between this and (II.2.5).


• f = O(g) as z → z0 means
f (z)
6 C for z → z0 ;
g(z)
• f = o(g) as z → z0 means
f (z)
→ 0 for z → z0 .
g(z)

Thus another way of writing the condition for C-differentiability of f at z0 is


f (z0 + h) = f (z0 ) + f 0 (z0 )h + o(|h|) for |h| → 0. (II.2.9)
This is the natural complex analogue of the real picture of a graph and its tangent
line. (II.2.9) says that for small h, the function h 7→ f (z0 + h) is well approximated by
the (affine) linear map
h 7→ f (z0 ) + f 0 (z0 )h
by composition of a translation, rotation and enlargement.
The usual rules for working with derivatives apply:
(a) Any constant is complex-differentiable with derivative equal to 0;
(b) f (z) = z is complex-differentiable everywhere, with derivative equal to 1;
(c) If f and g are complex-differentiable at z0 and λ and µ are constants, then
λf + µg is complex-differentiable at z0 with derivative λf 0 (z0 ) + µg 0 (z0 ).
(d) With the same hypotheses, f g (the product, not the composition) of f and g is
complex-differentiable at z0 with derivative f 0 (z0 )g(z0 ) + f (z0 )g 0 (z0 ).
(e) Chain rule: if f is holomorphic near w0 and g is holomorphic near z0 , with
g(z0 ) = w0 , then f ◦ g is holomorphic near z0 with derivative
(f ◦ g)0 (z0 ) = f 0 (g(z0 ))g 0 (z0 ). (II.2.10)
Here ‘near’ means ‘in a suitably small disc centred at’.
We shall not give the proofs of these. They imply the first part of the following
important
Theorem II.2.3. If Ω is an open set of C, then H(Ω) is an infinite-dimensional
vector space which is in fact a ring. Moreover, if f, g ∈ H(Ω), then f /g is holomorphic
at all points of Ω for which g(z) 6= 0.
The space H(Ω) contains A (Ω) and is contained in the space C(Ω) of complex-valued
continuous functions.
34 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

Proof.
(a) The properties listed just before the statement of the theorem imply that H(Ω)
is a ring, for they show that if f, g ∈ H(Ω) then λf + µg and f g are again
in H(Ω). Similarly the derivative of f /g is (f 0 g − g 0 f )/g 2 at any point where
g(z) 6= 0.
(b) Applying the product rule repeatedly (or by direct calculation), f (z) = z n is
in H(Ω) for every positive integer n and f 0 (z) = nz n−1 . Thus we have an
infinite set of linearly independent elements of H(Ω), showing it is an infinite-
dimensional vector space.
(c) If f ∈ A (Ω), then for each point z0 ∈ Ω, there exists δ > 0 such that
X∞
f (z) = an (z − z0 )n for z ∈ D(z0 , δ).
n=0
The theorem about term-by-term differentiation of power series seen in a pre-
vious course implies here that f 0 (z) exists and

X
f 0 (z) = nan (z − z0 )n−1 for z ∈ D(z0 , δ).
n=1
(The proof you’ve seen was probably only for real variables, but it goes through
in the complex case almost unchanged.) Thus every function in A (Ω) is also in
H(Ω).
(d) The usual real-variable proof that ‘differentiable ⇒ continuous’ goes through
here to show that H(Ω) ⊂ C(Ω).
(e) This is also a simple adaptation of the usual real-variable proof.


We shall see, after a lot of further work, that in fact H(Ω) = A (Ω).
This is remarkable, compared with the corresponding situation for functions of a
real variable! If a function is complex-differentiable, then with no further hypotheses,
it can be expanded in a convergent power series about every point!

Remark II.2.4. Note that it is much easier to check that a given function is in
H(Ω): one simply needs to check that the derivative exists at every point of Ω. This is
much better than trying to check that a given function can be locally expanded in power
series. Take, for example, the Γ function or the ζ function,
Z ∞ ∞
−t z−1
X 1
Γ(z) = e t dt, ζ(z) = . (II.2.11)
0 nz
n=1
It is relatively easy to check that these are holomorphic (complex-differentiable) (in
Re(z) > 0 and Re(z) > 1 respectively, since you asked). It would be much more arduous
to check that they are locally expandable as power series. In fact, though we did do
it in Theorem II.1.10, it is much harder to check that product and quotient of two
analytic functions are again analytic than it is to check the corresponding properties of
holomorphic functions.
2. HOLOMORPHIC FUNCTIONS—COMPLEX DIFFERENTIABILITY 35

Exercise II.2.6.
Show directly that z n is holomorphic with derivative nz n−1 .
Show that f (z) = z is not complex-differentiable.
Decide whether or not f (z) = |z|2 is complex-differentiable.

2.1. Higher derivatives, Taylor’s Theorem. If f is holomorphic in an open set


Ω, then f 0 (z) is a function in Ω. We can ask if f 0 is again holomorphic, and if it is we will
have its derivative, the complex second derivative f 00 (z). In general if we can differentiate
our holomorphic function n times, we denote its n-th derivative in the standard way by
f (n) (z).

If f ∈ A (Ω), that is f is analytic, then all derivatives f (n) exist and f (n) ∈ A (Ω).
This is because when we differentiate the power series expansion term by term as in
the proof of Theorem II.2.3, the radius of convergence of the expansion of f 0 is the
same as for f . Hence f 0 can be differentiated term by term without changing the
radius of convergence and so on.
So those holomorphic functions that are in A (Ω) all are infinitely differentiable.
We have mentioned before that H(Ω) = A (Ω) (to be proved later). It follows that
if f ∈ H(Ω) then f is infinitely differentiable.

Remark II.2.5. While the coefficients of the expansion of a holomorphic function


at different points may not have anything very obvious to do with each other, a given
holomorphic function has one and only one expansion at any given point z0 .
Proposition II.2.6. If f ∈ A (Ω) and z0 ∈ Ω, then the power-series expansion of f
at z0 is the Maclaurin–Taylor expansion of f at that point,

X f (n) (z0 )
f (z) = (z − z0 )n . (II.2.12)
n!
n=0
In particular the coefficients in the expansion at z0 are uniquely determined by f near
z0 .
Proof. This follows directly from term-by-term differentiation. If the expansion is

X
f (z) = an (z − z0 )n
n=0

then f (z0 ) = a0 . Differentiating once and substituting z = z0 , we see that f 0 (z0 ) = a1 .


Differentiating m times we obtain

X
f (m) (z) = n(n − 1) · · · (n − m + 1)an (z − z)n−m (II.2.13)
n=m

and evaluation at z = z0 gives am = f (m) (z0 )/m!. 


36 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

Exercise II.2.7. What is the 100th derivative of 1/(1 − z) at z = 0?

Another result that is proved in the same way as in the real-variable case is the chain
rule:

Exercise II.2.8. **For enthusiasts: Can you prove directly the statement
corresponding to (II.2.10) for analytic functions? In other words, if you have two
convergent power series and you substitute one into the other, can you prove that
the resulting power series is convergent? This may be very challenging!

2.2. Holomorphic functions as mappings: conformality.


Definition II.2.7. A mapping F : Ω → Ω,e where Ω and Ω e are open subsets of C is
said to be conformal if it maps regular C curves in Ω to regular C 1 curves in Ω0 and
1

preserves angles between curves.


We shall make this definition more precise in the course of proving the next result.
Theorem II.2.8. If f : Ω → C is holomorphic, then it is conformal at z0 ∈ Ω if
f 0 (z0 ) 6= 0.
Proof. The main point of the proof is the definition of regular C 1 curve and angle!
A C 1 curve through z0 ∈ Ω is a map γ : (−δ, δ) → Ω with γ(0) = z0 and such that
the real and imaginary part of γ are continuously differentiable functions:
γ(t) = X(t) + iY (t), −δ < t < δ
where X(t) and Y (t) are real and continuously differentiable. γ is regular at z0 = γ(0)
if the derivative γ 0 (0) is non-zero. The curve is regular if γ 0 (t) 6= 0 for all t ∈ (−δ, δ).
The interpretation of γ 0 (0) is as the tangent vector to γ at the point z0 (i.e. the
vector with cartesian components (X 0 (0), Y 0 (0))).
Now we have a tangent vector, we can define the angle this vector makes with the
positive real axis as arg γ 0 (0).
Now suppose that γ1 and γ2 are C 1 regular curves both through z0 ,
γ1 (0) = γ2 (0), γ1 (t) = X1 (t) + iY1 (t), γ2 (t) = X2 (t) + iY2 (t)
The angle between γ1 and γ2 at z0 is by definition the difference of arguments
φ = arg γ20 (0) − arg γ10 (0). (II.2.14)
Now suppose that f is holomorphic near (i.e. in an open ball containing) z0 , that f (z0 ) =
w0 , and that f 0 (z0 ) 6= 0. The image curves are
Γ1 (t) = f (γ1 (t)) and Γ2 (t) = f (γ2 (t)).
By the chain rule,
Γ01 (t) = f 0 (γ1 (t))γ10 (t), Γ02 (t) = f 0 (γ2 (t))γ20 (t).
2. HOLOMORPHIC FUNCTIONS—COMPLEX DIFFERENTIABILITY 37

Setting t = 0, we obtain
Γ01 (0) = f 0 (z0 )γ10 (0), Γ02 (0) = f 0 (z0 )γ20 (0).
Thus the tangent vectors Γ01 (0) and Γ02 (0) to the curves Γ1 and Γ2 at w0 are obtained
from the corresponding tangent vectors to γ1 and γ2 at z0 by multiplication by the
complex number f 0 (z0 ).
If f 0 (z0 ) = 0 the tangent vectors are 0, so the image curves are not regular. But if
0
f (z0 ) 6= 0 we may take args and use arg(zw) = arg(z) + arg(w) (mod 2π, for non-zero
complex numbers z and w) to obtain
arg Γ02 (0) − arg Γ01 (0) = (arg f 0 (z0 ) + arg γ20 (0)) − (arg f 0 (z0 ) + arg γ10 (0)) = φ.
So the angle between the image curves is equal to the angle between the original curves
as long as f 0 (z0 ) 6= 0. 

Think of it this way. If f 0 (z0 ) 6= 0, then a good approximation to f (z) should be


f (z0 + h) ' f (z0 ) + f 0 (z0 )h (II.2.15)
if |h| is small. Thus the effect on the small vector h which gives a displacement
from z0 to the nearby point z0 + h is approximately the standard transformation
h 7→ f 0 (z0 )h + f (z0 ). If f 0 (z0 ) 6= 0, this is a combination of dilation, rotation
and translation all of which preserve angles. This gives some intuition for why
holomorphic functions are conformal at points where f 0 (z0 ) 6= 0.

What happens at points where f 0 (z0 ) = 0? We are not going to study this system-
atically, but let’s consider f (z) = z 2 as an example. f 0 (z) 6= 0 if z 6= 0, so it is conformal
everywhere except at z = 0. If z = reiθ then w = z 2 = r2 e2iθ . This means that the ray
{arg(z) = α} is mapped into the ray {arg(w) = 2α}. The angles between rays through
the origin in the z-plane are doubled by the mapping w = z 2 .
More generally, if w = z n , n a positive integer, then angles between rays through
the origin in the z-plane are multiplied by n.
Note that there is a subtlety here which is that in these cases, regular C 1 curves
through 0 in the z-plane are not mapped to regular C 1 curves in the w-plane. For
example, the curve
γ(t) = t, (−δ < t < δ)
in the z-plane is mapped to γ e(t) = t2 in the w-plane. As t increases, γ(t) comes in to
the origin of the z-plane from the left along the real axis, passes through and emerges
along the positive real axis.
On the other hand, γ e(t) comes in to the origin along the positive real axis of the
w-plane as t approaches 0 from below, then doubles-back at t = 0 and goes out again
along the positive real axis. The tangent vector at t = 0 is 0 so γ e is not regular at t = 0.
2.3. Real functions and complex functions. In applications (for example con-
tour integration) we often have a function φ(x), say, on the real line and we want to
find a holomorphic function f (z) defined at least near (i.e. in an open neighbourhood
of) the real line which extends φ into the complex, in other words f (x) = φ(x). This
is not always possible (I don’t want to get into the technical details here) but for the
38 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

kinds of simple functions that you learned about at school this is rather easy, simply by
substituting z for x in the formula.

Examples II.2.9.
(a) Suppose we have a real rational function of x such as

x5 − 7x3 + 1
φ(x) =
x8 + 4x4 + 1
Then simply replacing x by z = x + iy yields the holomorphic function

z 5 − 7z 3 + 1
f (z) = .
z 8 + 4z 4 + 1
Note that this is not holomorphic in the whole of C—it fails to be holomorphic
at the zeros of the denominator, but it is holomorphic in an open strip which
contains the real axis.
(b) Similarly if φ(x) involves the elementary transcendental functions exp, sin, cos
etc., we can simply replace x by z since we already have meanings for exp z,
sin z, cos z etc. For example
2
e−x
φ(x) =
5 + sin x
yields the holomorphic function
2
e−z
f (z) = .
5 + sin z
Note that even though φ(x) is continuous for all x (because | sin x| 6 1 for all
real x, so the denominator in φ is never 0), f (z) is not holomorphic in the whole
of C. There are infinitely many (non-real) values of z with sin z = −5 and f (z)
is singular at all these points.

Exercise II.2.9. Find a holomorphic function f (z) with the property that
1
f (x) = √ .
5 − cos x
What is the largest subset of C on which f (z) is defined?
Exercise II.2.10. Suppose that φ(x) is a smooth function on the real line. For
real (x, y), define
F (x + iy) = φ(x).
Make sure you understand the difference between this and the kind of extensions we
discussed in the examples above. What do you need to assume about φ for F to be
holomorphic? This problem will be easier after the next topic!
3. THE CAUCHY–RIEMANN EQUATIONS 39

3. The Cauchy–Riemann equations

If Ω is an open set of C and u is a real- or complex-valued function in Ω, the partial


derivatives of u at z0 = x0 + iy0 ∈ Ω are defined as the limits
∂u u(x0 + t, y0 ) − u(x0 , y0 ) ∂u u(x0 , y0 + t) − u(x0 , y0 )
(x0 , y0 ) = lim , (x0 , y0 ) = lim .
∂x t→0 t ∂y t→0 t
(II.3.1)
(Here t is real.) We also use the subscript notation ux = ∂u/∂x, uy = ∂u/∂y for partial
derivatives.
If u is a function on Ω such that ux and uy exist at every point of Ω and these
partial derivatives are continuous, then we write u ∈ C 1 (Ω). Similarly, C 2 (Ω) is the
set of functions with the property that the second partial derivatives uxx , uxy , uyx , uyy
exist and are continuous in Ω. It is a theorem that if u ∈ C 2 (Ω), then the mixed partial
derivatives are equal,
   
∂ ∂u ∂ ∂u
uxy = = = uyx . (II.3.2)
∂y ∂x ∂x ∂y
Definition II.3.1. Suppose that f : Ω → C is a function and we write the real and
imaginary parts of f as u and v,
f (x, y) = u(x, y) + iv(x, y). (II.3.3)
The Cauchy–Riemann equations (in Ω) are the conditions
∂u ∂v ∂u ∂v
= , =− (II.3.4)
∂x ∂y ∂y ∂x
or
∂f ∂f
+i = 0. (II.3.5)
∂x ∂y
(II.3.4) is called the real form of the Cauchy–Riemann equations, (II.3.5) is called the
complex form.
The real and complex forms of the CR equations are equivalent. Indeed, we have
fx + ify = (ux + ivx ) + i(uy + ivy )
= (ux − vy ) + i(uy + vx ).
So the real and imaginary parts of the complex form of the CR equations comprise the
real form of the CR equations.
Theorem II.3.2. Let f (z) be holomorphic in an open set Ω. Then it satisfies the
Cauchy–Riemann equations in Ω.
Conversely, if (II.3.4) hold at all points of Ω and the first derivatives ux , uy , vx , vy
are continuous in Ω, then f = u + iv is holomorphic in Ω.
Proof. Deriving (II.3.4) is relatively easy. In the definition II.2.1 of ‘holomorphic’
take h first to be t and then it, where t is real. We obtain two equations:
f (z0 + t) − f (z0 ) − f 0 (z0 )t = o(|t|) (II.3.6)
f (z0 + it) − f (z0 ) − if 0 (z0 )t = o(|t|). (II.3.7)
40 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

Now by definition of partial derivative,


f (z0 + t) − f (z0 )
= fx (z0 ) + o(1), (II.3.8)
t
f (z0 + it) − f (z0 )
= fy (z0 ) + o(1). (II.3.9)
t
for t → 0, with t real. Combining (II.3.6) with (II.3.8) we obtain
f 0 (z0 ) = fx (z0 ). (II.3.10)
Similarly, by combining (II.3.7) with(II.3.9), we find
f 0 (z0 ) = −ify (z0 ). (II.3.11)
Eliminating f 0 (z0 ) between (II.3.10) and (II.3.11) yields the complex form of the Cauchy–
Riemann equations.
Let us turn to the slightly fiddly converse, which requires the continuity of the first
derivatives.
*This proof is not examinable. Suppose that f satisfies the CR equations in Ω. Pick
z0 = x0 + iy0 ∈ Ω and let h = s + it be small (s, t real). By the complex CR equations,
fx (z) = −ify (z) =: A(z), say. (II.3.12)
Then by continuity of the first partial derivatives, A(z) is continuous in Ω.
We have
f (z0 + h) − f (z0 ) = f (x0 + s, y0 + t) − f (x0 , y0 )
= f (x0 + s, y0 + t) − f (x0 + s, y0 )
+ f (x0 + s, y0 ) − f (x0 , y0 ).
By the definition of partial derivatives, we may replace the difference on the second line
by the partial derivative with respect to x and that on the third line by the partial
derivative with respect to y, up to small errors:
f (z0 + h) − f (z0 ) = fy (x0 + s, y0 )t + fx (x0 , y0 )s + o(|s|) + o(|t|). (II.3.13)
By (II.3.12), this gives
f (z0 + h) − f (z0 ) = A(z0 )s + iA(z0 + s)t + o(|s|) + o(|t|)
= A(z0 )(s + it) + i(A(z0 + s) − A(z0 ))t + o(|s|) + o(|t|)
= A(z0 )h + i(A(z0 + s) − A(z0 ))t + o(|s|) + o(|t|) (II.3.14)
Looking back at the definition of complex-differentiability, (II.3.14) is very close to what
we need. All that remains is to show that the ‘error term’
ε(h) := i(A(z0 + s) − A(z0 ))t + o(|s|) + o(|t|) (II.3.15)
is o(h). Now |s| 6 |h| and |t| 6 |h| so o(|s|) + o(|t|) = o(|h|). So that term is fine.
Because A is assumed continuous at z0 ,
(A(z0 + s) − A(z0 )) = o(1) for s → 0.
Again using |t| 6 |h|, we see that
|(A(z0 + s) − A(z0 ))t| = o(1)O(|h|) = o(|h|).
3. THE CAUCHY–RIEMANN EQUATIONS 41

Hence we can rewrite (II.3.14) in the form


f (z0 + h) − f (z0 ) = A(z0 )h + o(|h|) (II.3.16)
for small h, showing that f is complex-differentiable at z0 , with derivative A(z0 ) given
by (II.3.12). 
Remark II.3.3. In this proof we have exploited that the limit for defining the de-
rivative exists as h → 0 in any direction (and the limit is independent of the direction).
3.1. The ∂ and ∂-operators. It is usual to define
   
∂ 1 ∂ ∂ ∂ 1 ∂ ∂
∂ = ∂z = = −i , ∂ = ∂z = = +i . (II.3.17)
∂z 2 ∂x ∂y ∂z 2 ∂x ∂y
Then the complex form of the CR equations is ∂z f = 0. If the partial derivatives of
f are continuous and ∂z f = 0, it is complex-differentiable and f 0 = ∂z f .
Inverting (II.3.17),
∂ ∂
= ∂ + ∂, = −i(∂ − ∂). (II.3.18)
∂x ∂y

Exercise II.3.11. Show that if f is holomorphic, then f 0 (z) = ∂f (z).


Exercise II.3.12. Show that the real and imaginary parts of a holomorphic
function are harmonic (uxx + uyy = 0, vxx + vyy = 0).

3.2. Some consequences of the CR equations. The Cauchy–Riemann equa-


tions have some surprising consequences, for example:
Proposition II.3.4. If f is holomorphic in C and real-valued, then f 0 (z) = 0 and
f must be a constant.
Proof. Writing f = u + iv, we have that v = 0. The CR equations immediately
give that ux = uy = 0. Hence f 0 (z) = 0 for all z.
The fact that f 0 (z) = 0 for all z in C implies that f is constant will be shown in
Proposition II.3.5 below. 

There are many variants of this. For example:


Exercise II.3.13. Suppose that f is holomorphic in Ω. Show that any of the
following imply that f 0 (z) = 0 in Ω.
(a) f (z) imaginary for all z ∈ Ω;
(b) |f (z)| constant for all z ∈ Ω;
(c) arg(f (z)) constant for all z ∈ Ω.

Proposition II.3.5. If f : C → C and f 0 (z) = 0 for all z ∈ C, then f (z) is a


constant.
42 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

Proof. Let z0 = x0 + iy0 ∈ C. Since f 0 (z) = 0, both partial derivatives of f vanish


everywhere. In particular,
Z x0
f (x0 , 0) − f (0, 0) = fx (t, 0) dt = 0 (II.3.19)
0
(where we have written f as a function of two real variables for clarity). This equation
says that f is constant along the real axis.
Similarly, Z y0
f (x0 , y0 ) − f (x0 , 0) = fy (x0 , t) dt = 0. (II.3.20)
0
So f is constant along the vertical line through (x0 , 0). Combining the two results,
f (z0 ) = f (0), and so f is constant in C. 
It is not true that f 0 (z) = 0 for all z ∈ Ω implies that f is constant in Ω for arbitrary
open subsets of C. For example, if Ω = D(−1, 1) ∪ D(1, 1), let

3 for z ∈ D(−1, 1);
f (z) = (II.3.21)
−5 for z ∈ D(1, 1).
(See Figure 2.) Then f is holomorphic and f 0 (z) = 0 for all z ∈ Ω, but f is not constant.
In some sense this is the only thing that can go wrong. The problem here is that
Ω consists of two ‘disconnected pieces’ that don’t talk to each other. We therefore now
embark on a digression on (dis)connectedness.

0
−1 1

Figure 2. A disconnected set. The function f equal to 3 in the left-hand


disc and −5 in the right-hand disc satisfies f 0 = 0 everywhere but is not
constant. (The discs are open, so the origin is in neither of them.)

4. Connectedness
We have seen that if f 0 (z) = 0 for all z in some open set Ω, it may or may not be true
that f is constant. (For Ω = C it is true, for Ω the disjoint union of discs in Figure 2,
it is not.) So a natural question is: is there a useful characterization of the (say) open
sets Ω for which it is true that f 0 = 0 implies f is constant?
In the proof of Proposition II.3.5 what we used is that any two points in C can be
connected by a path. In the counterexample in Figure 2, on the other hand, there is no
path which connects z = −1 to z = 1 while remaining entirely in Ω. We therefore define
connectedness in terms of paths.
Definition II.4.1.
4. CONNECTEDNESS 43

(a) A path connecting z0 to z1 is a continuous map


γ : [0, 1] → C such that γ(0) = z0 , γ(1) = z1 . (II.4.1)
(b) If S ⊂ C is a subset, we say that γ is a path in S if γ(t) ∈ S for all t ∈ [0, 1] (or
in brief, γ : [0, 1] → S.
(c) The subset S is said to be path-connected if for any pair of points z0 , z1 ∈ S,
there is a path γ in S connecting z0 to z1 .
Example II.4.2. The three kinds of discs D(a, r), D(a, r) and D∗ (a, r) are all path-
connected. In the case of the first two, this is clear, for if S = D(a, r) or D(a, r) and z0
and z1 are two points in S then the straight-line segment between z0 and z1 connects
these points in S.
The case of D∗ (0, r) is not so easy. The line segment joining z0 to z1 lies in D∗ (0, r)
unless z0 and z1 are on ‘opposite sides of 0. In this case, we can connect them by a path
consisting of a segment from z0 to w and then w to z1 where w is a suitably chosen
intermediate point.
Example II.4.3. A set S is convex if for any two points z0 , z1 ∈ S, the straight-line
segment joining z0 to z1 is also contained in S. It is clear from this definition that any
convex set is path-connected.
Convex ⇒ path-connected, but the converse is certainly false: there are non-convex
sets that are path-connected (such as D∗ (a, r)) and other that are not (such as the
disjoint union of discs in Figure 2).
With the definition of path-connectedness in hand, we can prove a result which will
eventually answer the question at the beginning of this section.
Definition II.4.4. A function f : S → C is said to be locally constant if for every
z0 ∈ S, there exists δ > 0 such that
f (z) = f (z0 ) for all z ∈ S ∩ D(z0 , δ). (II.4.2)
Note that the function defined in (II.3.21) is locally constant. The next theorem
says, however, that if S is path-connected, then locally constant in S implies constant
in S.
Theorem II.4.5. Let S ⊂ C be any subset. If S is path-connected then any locally
constant function f : S → C must be constant.
Proof. Let f be locally constant and let S be path-connected. Let z0 and z1 be
any two points of S and let γ : [0, 1] → S be a path connecting z0 to z1 . Let
F (t) = f (γ(t)) (II.4.3)
From the definition, F is locally constant on [0, 1]. Let
T = sup{t ∈ [0, 1] : F (t) = F (0)}. (II.4.4)
The set is non-empty (it contains 0) and is bounded above by 1. So the supremum T
exists and is T 6 1. Note that since F is locally constant at t = T ,
F (T ) = F (T − δ)
44 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

for all sufficiently small positive δ. Hence F (T ) = F (0) because F (T − δ) = F (0). If


T = 1 we are done, so suppose if possible that T < 1. Using the local constancy of F at
T again, we have that there exists δ > 0 such that
F (t) = F (T ) for all t ∈ (T − δ, T + δ). (II.4.5)
Hence F (T + δ/2) = F (T ) = F (0) by the previous argument. This contradicts the
definition of T as the sup in (II.4.4). Hence T = 1 and so f (z0 ) = f (z1 ).
and F (t) = F (0) for all t < T . So T = 1. Applying the same argument, that F is
locally constant at t = 1, we learn that F (1) = F (0). Thus f (z0 ) = f (z1 ) and since this
is true for every pair of points, the proof is complete. 
Remark II.4.6. The main point of the proof is that the interval [0, 1] has the property
that locally constant functions on [0, 1] are constant. That is the ‘deep’ part, which
requires use of sup (and hence the completeness axiom of the reals). The auxiliary
construction of F as the composite f ◦ γ allows us to transfer this property of [0, 1] to
our path-connected set S. We shall see other proofs which follow essentially the same
strategy later in the course.
The following gives a criterion for a set not to be path-connected.
Corollary II.4.7. If f : S → Z is continuous but non-constant, then S is not
path-connected.
Proof. This follows from the previous theorem, because a continuous map to Z
is automatically locally constant. (Apply the definition of continuity with ε = 1/2,
say.) 
Finally we answer the original question:
Theorem II.4.8. Suppose that Ω is a path-connected open set of C, that f : Ω → C
is holomorphic and f 0 (z) = 0 for all z ∈ Ω. Then f is constant in Ω.
Proof. By the previous result, it is enough to show that f is locally constant. Pick
z0 ∈ Ω and D(z0 , δ) ⊂ Ω. Every point z in D(z0 , δ) can be joined to z0 by a path which
runs first parallel to the x-axis, then parallel to the y-axis. The vanishing of partial
derivatives implies, as in the proof of Proposition II.3.5, that f (z) = f (z0 ) for all such
z. Thus f is indeed locally constant. 

IMPORTANT
Definition II.4.9. A subset Ω of C is called a domain if it is open and connected.

Remark II.4.10. Both Ahlfors and Priestley use the term ‘region’ instead of domain.
Domain is also standard terminology in analysis, however.
4.1. Path components. Let S ⊂ C be any set. On S we may define an equivalence
relation: z0 ∼ z1 if and only if z0 can be connected to z1 by a path in S. Let S1 , S2 , S3 , . . .
(in general infinitely many) be the equivalence classes. Any such equivalence class is
called a path-component of S. The path-components are pairwise disjoint (this is true
5. HARMONIC FUNCTIONS 45

of equivalence classes for any equivalence relation) and by definition each of the Si is
path-connected.
Example II.4.11. The two path-components of the set in Figure 2 are the two discs.
Hence:
Theorem II.4.12. Any subset S ⊂ C is a disjoint union of path-connected subsets.
Moreover, if S is open, each of the path-components of S is again open.
Proof. The statement about openness of the path-components is missing from the
above discussion. Suppose that S is open and that S1 is a path-component. Let z0 be a
point of S1 . Since S is open, some small disc D(z0 , δ) ⊂ S. But any point of z ∈ D(z0 , δ)
can clearly be connected to z0 (by a straight line segment, for example) so D(z0 , δ) ⊂ S1 .
This is what we need for S1 to be open. 

Let Ω be an open set of C. Then Ω = Ω1 ∪ Ω2 ∪ · · · where the path components


Ωi are open and connected, and Ωi ∩ Ωj = ∅ for all i 6= j.

A generalization of Theorem II.4.8 can be stated as: if Ω is an open set of C and


f : Ω → C is holomorphic and f 0 = 0 in Ω then f is a constant on each path-component
of Ω.

Exercise II.4.14. *For enthusiasts: Check that the relation ∼ defined at the
beginning of this section is an equivalence relation.

5. Harmonic functions
We have now discussed analytic functions (those locally expandable as convergent
power series in z) and holomorphic functions (those which are complex-differentiable).
Closely related also are harmonic functions.
Definition II.5.1. Let Ω be an open set of C. A real- or complex-valued function
u in C 2 (Ω) is called harmonic if
∆u = 0, (II.5.1)
where ∆, the Laplace operator or Laplacian is defined as
∂2 ∂2
∆= + . (II.5.2)
∂x2 ∂y 2
Example II.5.2. The functions x, y, x2 − y 2 , xy are all harmonic.

Exercise II.5.15. Let


u(x, y) = Ax3 + Bx2 y + Cxy 2 + Dy 3
where A, B, C, D are constants. What is the most general such function which is
harmonic?
46 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

Exercise II.5.16. Show that 12 log(x2 + y 2 ) and tan−1 (y/x) are harmonic where
they are defined (the latter, for example in the half space {x > 0} with the choice
of tan−1 lying in (−π/2, π/2)).

Theorem II.5.3. Let f = u + iv be holomorphic in an open set Ω ⊂ C. Suppose that


u and v are twice continuously differentiable. Then u and v are harmonic in Ω.

Proof. From the Cauchy–Riemann equations,

uxx = −vyx = −vxy = −uyy .

The first equality is the CR equation ux = vy , the second is the symmetry of mixed
partial derivatives, and the third is the CR equation −vx = uy . Thus the continuity of
the second partials is needed for vxy = vyx and hence for the proof to work. The proof
that v is harmonic is precisely similar. 

We shall see later that f ∈ H(Ω) implies automatically that the second derivatives
of u and v exist and are continuous. So in fact the real and imaginary parts of f are
harmonic without further assumption.
Conversely, if u is a real harmonic function, then there is a v such that u + iv is
holomorphic—at least under some assumptions on the domain Ω.

Theorem II.5.4. Suppose that u is harmonic in C or in an open disc D(z0 , r) and all
second order partial derivatives are continuous. Then there exists a function v, unique
up to the addition of a constant, such that

f = u + iv

is holomorphic.

Remark II.5.5. Being the imaginary part of a holomorphic function, v is also har-
monic. If u is harmonic and u + iv is holomorphic, then v is called a harmonic conjugate
of u.

Proof. The proof for the disc is the same, so we just give the proof for C.
We have seen that if u+iv is holomorphic then u and v must satisfy the CR equations.
Given u, we define v so that the first CR equation holds, by integration. Let
Z y
w(x, y) = ux (x, t) dt. (II.5.3)
0

For this to make sense, it is good to have the continuity of ux . By the fundamental
theorem of calculus, we have wy = ux . That is the pair (u, w) satisfies the first of the
real CR equations. This pair may not satisfy the other CR equation, though. If we
define v(x, y) = w(x, y) + φ(x), then vy = wy and so the pair (u, v) also satisfies the first
CR equation.
5. HARMONIC FUNCTIONS 47

How about the second CR equation? We calculate


Z y
∂v
(x, y) = uxx (x, t) dt + φ0 (x)
∂x 0
Z y
= − uyy (x, t) dt + φ0 (x)
0
= −uy (x, y) + uy (x, 0) + φ0 (x).
In going from the first to the second line we used the hypothesis that u is harmonic
and in going from the second to the third, we used the fundamental theorem of calculus
again. So the second CR equation is satisfied if we choose φ0 (x) = −uy (x, 0). Such a φ
is unique up to a constant. 
So there is a very close relationship between holomorphic functions and harmonic
functions. In fact in Riemann’s approach to complex function theory, he emphasised this
aspect. He used the existence of harmonic conjugates to pass from well chosen harmonic
functions to interesting holomorphic functions.
Extending Theorem II.5.4 to domains more general than C is possible, but requires
care. In particular the analogous result is true provided the domain is simply connected
which roughly speaking means ‘no holes’.

Exercise II.5.17. What is the harmonic conjugate of u = ex cos y? Show that


the two functions in Exercise II.5.16 are harmonic conjugates.
Exercise II.5.18. Relate your answer to Exercise II.5.15 to the holomorphic
function z 3 .

Harmonic functions arise frequently in ‘physical’ applied mathematics, for example


in electrostatics, Newtonian gravity, steady fluid flow, and steady temperature distribu-
tions. The next result has practical value in transforming a harmonic function in one
domain to one in another domain:
Theorem II.5.6. Suppose that u e → R is a harmonic function and that f : Ω → Ω
e:Ω e
is holomorphic. Then u : Ω → R,
u(z) = u
e(f (z)), (II.5.4)
is again harmonic.
Proof. There are two proofs. The first is direct calculation with the chain rule and
the CR equations.
For the second, suppose that ve is a harmonic conjugate of u
e, defined near some point
z1 = f (z0 ) of Ω.
e Let g = u
e +ie
v , which is then holomorphic. The function F (z) = g(f (z))
is then holomorphic near z0 and its real part is u(z). Hence u is harmonic near z0 . This
argument works near any given point of z0 , showing that u is harmonic in Ω. 
For practical applications of this theorem, suppose that Ω is some perhaps com-
plicated domain of physical interest. We want a harmonic function in u with certain
properties (typically conditions at the boundary ∂Ω). If we can solve the corresponding
48 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

problem for ue in some simple domain Ω e (for example Ωe = H, the open upper half-space,
or the disc D(0, 1), then ‘all’ we need to do is find a holomorphic function f : Ω → H or
D(0, 1) and define u(z) = u e(f (z)) as in the Theorem.
Thus many (idealized, two-dimensional) physical problems are reduced to conformal
mapping of the domain of physical interest to a simpler one (and solving the problem in
this simple domain).
Example II.5.7. The function u e(w) = log |w| in the disc D∗ (0, 1) is harmonic. A
physical interpretation is the (steady-state) temperature distribution in D∗ (0, 1) when
w = 0 is heated to temperature −2π and the boundary |w| = 1 is held at temperature
0.
We have seen that w = f (z) = (z − i)/(z + i) maps the half-space H = {Im z > 0}
bijectively (and holomorphically) to {|w| < 1}. By the previous theorem, it follows that
z−i
u(z) = log |f (z)| = log
z+i
is harmonic in H \ {i} and vanishes on the boundary Im z = 0. Thus u(z) has the
physical interpretation of the steady-state temperature distribution in H if the point
z = i is held at temperature −2π, along with the side conditions that the real axis is
held at temperature 0, and that the temperature goes to 0 as Im z → +∞.

Exercise II.5.19. Let a be a point in the upper half-space. Find the harmonic
function u which is defined in {Im z > 0} \ {a}, which vanishes along Im z = 0 and
as Im z → +∞, and with a logarithmic singularity near a:
u(z) = log |z − a| + O(1) for |z − a| → 0.

5.1. The mean-value property and the maximum principle. We prove in


this section two fundamental properties of harmonic functions.
Theorem II.5.8 (Mean-value property). Let Ω be an open set of C and let u : Ω → R
be harmonic. If z0 ∈ Ω then for every sufficiently small circle |z − z0 | = ρ, we have
Z 2π
1
u(z0 ) = u(z0 + ρeiθ ) dθ. (II.5.5)
2π 0

Proof. By a translation, we may suppose that z0 = 0. Let C be the curve |z| = ρ,


traversed once anticlockwise, let D be the closed disc {z : |z| 6 ρ}. Green’s theorem in
the plane (which you have seen in a previous course) states that for any C 1 functions P
and Q defined in D, we have
Z Z
P (x, y)dx + Q(x, y)dy = (Qx (x, y) − Py (x, y)) dxdy. (II.5.6)
C D
If we take f and g to be C2 and
P = −f gy , Q = f gx (II.5.7)
5. HARMONIC FUNCTIONS 49

then we obtain
Z Z
(−f gy dx + f gx dy) = (fx gx + fy gy + f ∆g) dxdy. (II.5.8)
C D
Interchanging the roles of f and g and subtracting, the fx gx + fy gy term drops out,
leaving
Z Z
(gfy − f gy ) dx + (f gx − gfx ) dy = (f ∆g − g∆f ) dxdy. (II.5.9)
C D

(This is one of so-called Green’s Identities). If we take f = 1, g = u, all first derivatives


of f on the LHS disappear and the RHS vanishes identically because 1 and u are both
harmonic. Hence,
Z
(ux dy − uy dx) = 0. (II.5.10)
C
Afterp f = 1, the next simplest harmonic function in some sense is log r, where as usual
r = x2 + y 2 . If we try to apply (II.5.9) with f = log r, g = u, then we have the
problem that f is not even continuous, far less C 2 , near the origin.
Let us therefore replace D by Dδ = D \ D(0, δ). Then we can apply (II.5.9), but the
LHS also has a boundary contribution from the small ‘inner boundary’ Cδ , the circle
|z| = δ (Figure 3).

Dδ = {δ 6 |z| 6 ρ}

Figure 3. The set Dδ with its inner and outer boundary circles C =
{|z| = ρ} and Cδ = {|z| = δ}.

Since f and u are harmonic in Dδ , the RHS of (II.5.9) is zero, so we obtain


Z Z
u(fy dx−fx dy)−f (uy dx−ux dy) = u(fy dx−fx dy)−f (uy dx−ux dy) (II.5.11)
Cδ C

To simplify this, first observe that f = log r is constant on Cδ and on C. Thus


Z Z
f (uy dx − ux dy) = log δ (uy dx − ux dy) = 0
Cδ Cδ

by (II.5.10). Similarly
Z Z
f (uy dx − ux dy) = log ρ (uy dx − ux dy) = 0.
C C
50 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

Hence (II.5.11) simplifies to


Z Z
u(fy dx − fx dy) = u(fy dx − fx dy) (II.5.12)
Cδ C
where f is still log r.
R We iθclaim now that the LHS approaches 2πu(0) as δ → 0 and the RHS is equal to
C u(ρe ) dθ. This is a matter of calculation:
fx = x/r2 , fy = y/r2 , (r2 = x2 + y 2 ). (II.5.13)
If we parameterize the circle of radius s by (x, y) = (s cos θ, s sin θ), we have
cos θ sin θ
fx = , fy = , dx = −s sin θ dθ, dy = s cos θ dθ (II.5.14)
s s
and so
fy dx − fx dy = −dθ
on the circle of radius s (but the result is independent of the radius!). Hence (II.5.12)
reduces to Z 2π Z 2π
u(δeiθ ) dθ = u(ρeiθ ) dθ. (II.5.15)
0 0
The RHS is 2π times the RHS of the equation of the Theorem. As for the LHS, since u
is C 2 near z = 0, we have, for all sufficiently small δ,
|u(δeiθ ) − u(0)| 6 M δ
for some M and all θ ∈ [0, 2π]. Hence
Z 2π Z 2π
2πu(0) − u(δeiθ ) dθ = (u(0) − u(δeiθ )) dθ
0 0
Z 2π
6 |u(0) − u(δeiθ )| dθ
0
6 2πM δ. (II.5.16)
Hence the LHS of (II.5.15) approaches 2πu(0) as δ → 0. Since the RHS is independent
of δ, we can take this limit, completing the proof.

Remark II.5.9. The function log r is a Green’s function for the Laplacian with
singularity at 0. However, log r − C has the same singularity for any C, and for the disc
D(0, ρ), log(r/ρ) is the unique choice which vanishes when r = ρ. The proof could have
been carried out by choosing f = log(r/ρ) instead of f = 1 and f = log r.
Remark II.5.10. The previous proof is quite long, so let me summarize the main
steps. The key identity is (II.5.9). This is a standard result whose proof I put in for
completeness. Given this identity, the important idea is to let g = u be our unknown
harmonic function and to choose f in a clever way. The simplest choice of f , f =
constant, yields the identity (II.5.10) which is interesting but not what we need. So—
and this is the really important idea about this proof—we use f = log r as another
function in (II.5.9). f is harmonic away from its singularity so the RHS of (II.5.9)
should go away. But the singularity of f at r = 0 means that we have to modify D and
5. HARMONIC FUNCTIONS 51

then compute what happens as δ → 0. This combination of substituting a function with


a singularity and changing D is also important elsewhere. The calculations which lead
to (II.5.15) are fairly straightforward (no new conceptual idea). It is intuitively obvious
that as δ goes to 0 in this equation, the LHS should go to 2πu(0). The last part of the
proof is just a way of making that intuition precise.
Remark II.5.11. If we assume that u is the real part of an analytic function, and
also some material about integration of infinite series that has not yet been covered then
there is a different proof, as follows.
We suppose as before that z0 = 0. Then there is a power series

X
f (z) = an z n (II.5.17)
n=0

such that u(z) = Re(f (z)). Then



1X n
u(ρeiθ ) = r (an einθ + an e−inθ ). (II.5.18)
2
n=0
We simply integrate both sides with respect to θ. It is legitimate to interchange the
order of summation and integration here, though we have not proved this yet. This
gives
Z 2π
u(ρeiθ ) dθ = π(a0 + a0 ) (II.5.19)
0
R 2π
because for n 6= 0, 0 eint dt
= 0. (Remember this idea from Fourier series?)
From (II.5.17), f (0) = a0 and so u(0) = Re a0 . Thus (II.5.19) is equivalent to (II.5.5).
The next result is one of the most important theoretical properties of harmonic
functions:
Theorem II.5.12 (Maximum Principle). Let Ω be a domain and let u : Ω → R be a
harmonic function. If z0 ∈ Ω is such that u(z0 ) > u(z) for all z ∈ Ω, then u is constant.

Remark II.5.13. The point z0 is a global maximum of u. The Maximum Princi-


ple is often summarized as: a non-constant harmonic function cannot have an interior
maximum.
Proof. We use the mean-value property to show first that u must be equal to u(z0 )
in some disc D(z0 , r0 ), for some positive r0 . Then we shall use the connectedness of Ω
to show that u is constant in the whole of Ω.
To apply the mean-value property, write u(z0 ) in the stupid way
Z 2π
1
u(z0 ) = u(z0 ) dθ. (II.5.20)
2π 0
Subtracting the RHS from the LHS in (II.5.5)
Z 2π  
u(z0 ) − u(z0 + ρeiθ ) dθ = 0, (II.5.21)
0
52 II. ANALYTIC FUNCTIONS AND HOLOMORPHIC FUNCTIONS

for every ρ with D(z0 , ρ) ⊂ Ω. On the other hand, the integrand has a sign since z0 is
a maximum. More precisely, in (II.5.21) we are integrating a continuous non-negative
function over [0, 2π]: the only way to get zero is if this non-negative function is identically
zero. Hence if D(z0 , r0 ) ⊂ Ω,
u(z0 ) = u(z0 + ρeiθ ) for all 0 < ρ < r0 and θ ∈ [0, 2π]. (II.5.22)
Hence u(z) = u(z0 ) is identically zero in D(z0 , r0 ).
We now use the path-connectedness of Ω to conclude that u(z) = u(z0 ) for all z ∈ Ω.
If z1 ∈ Ω, choose a path γ(t) connecting z0 to z1 . Define
T = sup{t ∈ [0, 1] : u(γ(t)) = u(z0 )}. (II.5.23)
The set of t is non-empty and bounded above. So the supremum exists and is 6 1. Note
that
u(γ(T )) = lim u(γ(t)) = u(z0 ).
t%T
(The arrow in the limit means that T is approached from the left.)
Thus if T = 1 we are done, so suppose if possible that T < 1. In this case, let
z = γ(T ). We have u(z 0 ) = u(z0 ) and so z 0 is a maximum of u. Now we apply the first
0

part of the proof with z0 replaced by z 0 to conclude that u(z) = u(z 0 ) for all z in some
disc D(z 0 , r). This disc contains points of the form γ(T + δ) for sufficently small δ, and
so
u(γ(T + δ)) = u(z0 )
contradicting the definition of T . This contradiction shows that T = 1. So u(z0 ) = u(z1 )
for any given z1 in Ω, and u is constant. 
Remark II.5.14. The second part of the proof exploits the path-connectedness of Ω
to prove that u is constant in a way that is very similar to the proof of Theorem II.4.5.

Exercise II.5.20. Suppose that f is holomorphic near a point z0 . Show that


∆|f |2 = 2|fx |2 + 2|fy |2 (II.5.24)
and in particular ∆|f |2 > 0 at z0 with equality if and only if f 0 (z0 ) = 0. (A real
function u with ∆u > 0 is called subharmonic.
Exercise II.5.21. Suppose that f is holomorphic near a point z0 and that
f (z0 ) 6= 0. Show that u = log |f |2 is harmonic near z0 .

6. ‘Holomorphic at ∞’
In Chapter 1, we introduced the Riemann sphere S . We defined what it means for
f : S → C to be continuous. In this section we shall explain what it means for such a
function to be analytic or holomorphic. We shall give the main point with a minimum
of technical baggage.
The idea is simply to use a new variable ζ = 1/z. Then when |z| is very large, |ζ| is
close to 0, and we define f (z) to be holomorphic ‘near z = ∞’ if and only if fe(ζ) = f (1/ζ)
is holomorphic near ζ = 0.
6. ‘HOLOMORPHIC AT ∞’ 53

To be a little more precise, suppose that for some δ > 0, f is defined in {z : |z| > δ −1 }.
Suppose moreover that f (z) → L as |z| → ∞.
Define fe(ζ) = f (1/ζ) for ζ 6= 0, fe(0) = L. Then
Lemma II.6.1. So defined, fe(ζ) is continuous for ζ ∈ D(0, δ).
Definition II.6.2. f is holomorphic for z ∈ {z : |z| > δ −1 } ∪ {∞} if and only if fe(ζ)
is holomorphic for ζ ∈ D(0, δ).
Similarly, f is analytic for z ∈ {z : |z| > δ −1 }∪{∞} if and only if fe(ζ) is holomorphic
for ζ ∈ D(0, δ).
Examples II.6.3. With this definition z −1 is holomorphic in {|z| > 1} ∪ {∞} (with
value 0 at ∞). For in this case, fe(ζ) = ζ which is holomorphic in D(0, 1).
A rational function R(z) = p(z)/q(z) is holomorphic in a neighbourhood of z = ∞
if and only if deg p 6 deg q. For if
a0 + a1 z + · · · + am z m
R(z) =
b0 + b1 z + · · · + bn z n
with m 6 n, and am 6= 0, bn 6= 0, then
a0 + a1 ζ −1 + · · · + am ζ −m
R(ζ)
e = (II.6.1)
b0 + b1 ζ −1 + · · · + bn ζ −n
a0 ζ m + a1 ζ m−1 + · · · + am
= ζ n−m (II.6.2)
b0 ζ n + b1 ζ n−1 + · · · + bn
and this is holomorphic for |ζ| sufficiently small (so small as to avoid the zeros of the
denominator).
There is something to think about with this definition. Namely, we know that if f
is holomorphic in an open set Ω then it is also holomorphic in any open subset Ω0 ⊂ Ω.
Suppose that f is holomorphic in {z : |z| > δ −1 } ∪ {∞} in the sense of Definition II.6.2.
Combining with the above remark, this should imply that f is also holomorphic in the
subset {z : |z| > δ −1 }. But this set does not contain ∞ and we already have a definition
of holomorphic in this set, namely complex-differentiable at every point z with |z| > δ −1 .
The new definition says that f is holomorphic if fe(ζ) is complex differentiable for all ζ
with 0 < |ζ| < δ, where ζ = 1/z. Why are these two definitions consistent?
The answer is that if fe(ζ) = f (1/ζ) then by the chain rule, fe is holomorphic
(complex-differentiable) at ζ0 6= 0 if and only if f is holomorphic at z0 , where z0 = 1/ζ0 .
The key point is that ζ = 1/z is itself a holomorphic change of variables and this is
why Definition II.6.2 is consistent with the previous definition of holomorphic. There
are similar remarks with ‘holomorphic’ replaced by ‘analytic’ throughout.
CHAPTER III

Complex Integration

1. Paths and line-integrals


We have already met paths in our discussion of path-connectedness and in the dis-
cussion of conformality (where we needed to define the angle between two regular C 1
curves).
In order to avoid over-complicated notation later, we shall make the following defi-
nition
Definition III.1.1. Let S ⊂ C. A smooth curve or path in S is a map
γ : [a, b] → S (III.1.1)
where [a, b] is a closed bounded interval such that:
• if γ(t) = X(t) + iY (t) (where X and Y are real-valued functions) then X(t) and
Y (t) are smooth functions of t and
• γ 0 (t) = X 0 (t) + iY 0 (t) 6= 0 for all t ∈ [a, b].
Remark III.1.2. ‘Smooth functions’ here means that all derivatives exist. One could
get away with less (replacing ‘smooth’ by ‘continuously differentiable’) but the above will
be enough for our applications.
Definition III.1.3. The curve γ in the previous definition is closed if γ(a) = γ(b).
Whether or not it is closed, γ is simple if
γ(t1 ) 6= γ(t2 ) for all a < t1 < t2 < b.
Note in particular that the definition of ‘simple’ does allow for γ(a) = γ(b).

Examples III.1.4.
(i) γ(t) = Reit , t ∈ [0, 2π] is the circle centred at 0, radius R, traversed once
anticlockwise. This is an example of a simple, closed curve.
(ii) γ(t) = Reit , t ∈ [0, π] is the semi-circlular arc centred at 0, radius R, which
joins z = R to z = −R in the upper half-plane. This is an example of a
simple curve.
(iii) γ(t) = z0 + (z1 − z0 )t, t ∈ [0, 1] is the straight-line segment joining z0 at
t = 0 to z1 at t = 1. This is a simple curve.

Notation III.1.5. It is handy to denote by [z0 , z1 ] the straight-line segment from z0


to z1 , parameterized as in (iii) of the previous example.
55
56 III. COMPLEX INTEGRATION

It is also handy to denote the circular path γ(t) = z0 + Reit , 0 6 t 6 2π, simply by
|z − z0 | = R. In other words, when we write ‘the curve |z − z0 | = R’ we mean ‘the curve
|z − z0 | = R traversed once anticlockwise’.
We shall extend our definitions to allow for piecewise smooth curves in §1.4.
Note that the condition γ 0 (t) 6= 0 for all t means that the tangent vector is non-zero
at every point and that γ is locally well approximated by its tangent vector1, and there
are no sudden changes of direction along γ. If the assumption is dropped, this is not the
case. For example,  3
t , −1 6 t 6 0;
γ(t) = (III.1.2)
it3 , 0 6 t 6 1.
is C 2 , but γ 0 (0) = 0. This curve is runs along the negative real axis for −1 6 t < 0, hits
the origin z = 0 at t = 0 and then turns a sharp corner to go up the positive imaginary
axis for 0 < t 6 1, ending up at z = i.
Suppose that Ω is an open subset of C and that f : Ω → C is continuous. If
γ : [a, b] → C is a smooth curve in Ω, we define the integral of f along γ,
Z Z Z b
f = f (z) dz := f (γ(t))γ 0 (t) dt. (III.1.3)
γ γ a

γ0
Note that is continuous because of the assumption that γ is smooth. So the integrand
on the RHS is a continuous function on a closed bounded interval, so definitely exists
(as a Riemann integral).
Note that here we have the integral of a complex-valued function F (t) = f (γ(t))γ 0 (t)
of the real variable t. If F (t) = F1 (t) + iF2 (t) this is by definition
Z b Z b Z b Z b
F (t) dt = (F1 (t) + iF2 (t)) dt = F1 (t) dt + i F2 (t) dt. (III.1.4)
a a a a
Rb
Note that a is then a complex-linear mapping from continuous functions to C in
the sense that Z b Z b Z b
(λF (t) + µG(t)) dt = λ F (t) dt + µ (t) dt (III.1.5)
a a a
for all complex-valued continuous functions F and G and all complex constants λ and
µ.
More generally, if P (x, y) and Q(x, y) are continuous functions defined in Ω and γ is
the smooth curve in Ω we had before, we define the line integral
Z Z b
P (X(t), Y (t))X 0 (t) + Q(X(t), Y (t))Y 0 (t) dt.

P dx + Q dy = (III.1.6)
γ a

Example III.1.6. Let us compute the integral of f (z) = z 2 from z0 to z1 along


the segment [z0 , z1 ]. From the definition,
γ(t) = z0 + (z1 − z0 )t, γ 0 (t) = z1 − z0

1This follows from the inverse function theorem—but we won’t prove it


1. PATHS AND LINE-INTEGRALS 57

and so Z Z 1
z 2 dz = (z0 + (z1 − z0 )t)2 (z1 − z0 ) dt.
[z0 ,z1 ] 0
By the chain rule, the integrand is
d1
(z0 + (z1 − z0 )t)3
dt 3
so Z  1
2 1 3 1 1
z dz = (z0 + (z1 − z0 )t) = z13 − z03
[z0 ,z1 ] 3 0 3 3
Note that this is the same as what you might have guessed by extrapolating from
real integration.
Example III.1.7. Consider the integral of z −1 along the two curves
γ1 (t) = eit , t ∈ [0, π]
and
γ2 (t) = e−it , t ∈ [0, π].
Both γ1 and γ2 connect z = 1 to z = −1, γ1 in the upper half-plane, γ2 in the lower
half-plane. Then Z Z π
dz
= idt = πi.
γ1 z 0
Similarly, Z Z π
dz
= (−i)dt = −πi.
γ2 z 0
Note that the ‘naive’ answer log(−1) − log(1) is ambiguous in this case (because you
have to make a choice of arg to define log). This is reflected in the fact that while
the integrals along γ1 and γ2 are perfectly well-defined, the answer does depend
on the choice of path. We shall discuss this at greater length below.

1.1. Reparameterization invariance. Although it is important to appreciate


that the by ‘curve’ we mean ‘parameterized curve’ rather than ‘one-dimensional sub-
set of C’ (whatever that means) nonetheless, the integral of a function along a curve
does not depend very strongly on the parameterization.
More precisely, we say that Γ : [A, B] → S is a reparameterization of γ : [a, b] → S if
there is a smooth bijection
φ : [A, B] → [a, b] (III.1.7)
with φ0 (τ ) > 0 for all τ ∈ [A, B] and such that
Γ = γ ◦ φ : [A, B] → S. (III.1.8)
Example III.1.8. t 7→ z0 + 12 (t + t3 )(z1 − z0 ), 0 6 t 6 1 is a reparameterization of
the segment [z0 , z1 ].
Then
58 III. COMPLEX INTEGRATION

Proposition III.1.9. If Γ is a reparameterization of γ in the sense just defined, and


f is continuous in S, then Z Z
f = f.
Γ γ

Proof. Compute both sides and use the chain rule:


Z Z B
F = f (Γ(τ )) Γ0 (τ ) dτ
Γ A
Z B
= f (γ(φ(τ ))) (γ ◦ φ)0 (τ ) dτ
A
Z B
= f (γ(φ(τ )))γ 0 (φ(τ ))φ0 (τ ) dτ
A
Now make the change of variables t = φ(τ ), so t runs from a to b as τ runs from A to B
(cf. (III.1.7)). With this change of variables,
Z Z b
f = f (γ(t))γ 0 (t) dt
Γ
Za
= f.
γ

Remark III.1.10. Note that if γ : [a, b] → C is a curve, and A < B are two real
numbers, then the (affine) linear function
(b − a)
φ(τ ) = a + (τ − A)
B−A
defines a smooth bijection [A, B] → [a, b] with φ0 (τ ) = A−B
a−b
> 0. Then Γ = γ ◦ φ :
[A, B] → C is a reparameterization of γ, defined on the interval [A, B]. This means
that we can, if we wish, always assume that a given curve has some standard parameter
interval: for example, taking A = 0, B = 1, we can assume the parameter interval for
our curve is [0, 1].
1.2. Opposite orientation.
Definition III.1.11. If γ : [a, b] → C is a smooth curve, let
γ opp : [−b, −a] → C be defined by γ opp (t) = γ(−t). (III.1.9)
Then the images γ[a, b] and γ opp [a, b] are equal but, as the name suggests, γ opp is
traversed in the opposite direction from γ.
Proposition III.1.12. With γ and γ opp defined as above, if f is continuous in a
neighbourhood of the image of γ, then
Z Z
f =− f. (III.1.10)
γ γ opp

Proof. Easy from the definitions. 


1. PATHS AND LINE-INTEGRALS 59

1.3. Length, triangle inequality, and M L Lemma.


Proposition III.1.13. (Triangle inequality for integrals). If F (t) = F1 (t) + iF2 (t)
is continuous on [a, b] then
Z b Z b
F (t) dt 6 |F (t)| dt. (III.1.11)
a a

Proof. By definition, the integral is defined in terms of limits of Riemann sums.


More precisely, letR a = t0 < t1 < · · · < tN < tN +1 = b be a partition. The associated
Riemann sum for F is
N
X
F (tj )(tj+1 − tj ) (III.1.12)
j=0
Rb
and this tends to a F (t) dt as we take finer and finer partitions of [a, b]. Now by the
triangle inequality,
N
X N
X
F (tj )(tj+1 − tj ) 6 |F (tj )|(tj+1 − tj ) (III.1.13)
j=0 j=0
Rb
and as we take finer and finer partitions the RHS approaches a |F (t)| dt. The inequality
in (III.1.13) is preserved as we pass to the limit, proving the result. 
Definition III.1.14. The length of a smooth curve γ : [a, b] → C is by definition
Z b
L(γ) = |γ 0 (t)| dt.
a

By the previous result, this depends on the curve and not on the parameterization.

Exercise III.1.1. Show that


2s 1 − s2
 
(X(s), Y (s)) = , , −1 6 s 6 1 (III.1.14)
1 + s2 1 + s2
is a parameterization of the part of the circle |z| = 1 lying in the closed upper
half-plane. Use this parameterization to calculate the length of this curve.
Exercise III.1.2. Calculate the length of line segment [z0 , z1 ] from the param-
eterization in Example III.1.4 (iii).

We often need to estimate integrals that we can’t evaluate exactly. Here is the main
tool.
Proposition III.1.15. (Length-sup or M L Lemma.) Let γ : [a, b] → S be a smooth
curve of length L. Let f : S → C be a continuous function such that
sup |f (γ(t))| = M (III.1.15)
a6t6b
60 III. COMPLEX INTEGRATION

Then
Z
f 6 ML (III.1.16)
γ
R
Proof. Applying the triangle inequality for γ f,
Z Z b Z b
0
f = f (γ(t))γ (t) dt 6 |f (γ(t))| |γ 0 (t)|dt. (III.1.17)
γ a a

Now use |f (γ(t))| 6 M to get


Z Z b
f 6M |γ 0 (t)|dt = M L. (III.1.18)
γ a

as required. 

Exercise III.1.3. Show that


Z
dz
→ 0 as R → ∞.
|z|=R z4+1

Exercise III.1.4. Let f (z) = p(z)


q(z) be a rational function where deg q > degp + 2.
Show that Z
f (z) dz → 0 as R → ∞.
|z|=R

Exercise III.1.5. With f as in the previous exercise, let γ(t) = Reit , 0 6 t 6 π,


be a large semicircular path in the upper half plane. Show that
Z
eiaz f (z) dz → 0 as R → ∞
γ
if a is real and positive. Is the result still true if γ is replaced by the whole circle
|z| = R?
Exercise III.1.6. Let log z = log |z| + i arg(z), where 0 < arg(z) < 2π. Show
that if f is a rational function as in Exercise III.1.4, then
Z
f (z) log z dz → 0 as R → ∞ (III.1.19)
ΓR

where ΓR (t) = Reit , θ1 6 t 6 θ2 , for any given angles 0 < θ1 < θ2 < 2π.
Exercise III.1.7. Let log z = log |z| + i arg(z), where 0 < arg(z) < 2π. Show
that if u is a continuous function in {|z| < 1}, then
Z
u(z) log z dz → 0 as δ → 0 (III.1.20)
Γδ

where Γδ (t) = δeit , θ1 6 t 6 θ2 , for any given angles 0 < θ1 < θ2 < 2π.
1. PATHS AND LINE-INTEGRALS 61

1.4. Piecewise smooth paths. Consider the semicircular path Γ shown below. It
is not smooth because of the corners at −1 and 1.

γ2

0 γ1
z = −1 z=1

Figure 1. A piecewise smooth curve consisting of two smooth paths

If f is defined in a neighbourhood of this path, we can naturally define an integral of


f along Γ by parameterizing each piece separately and then adding. That is, we would
parameterize the diameter as
γ1 (t) = t for − 1 6 t 6 1 (III.1.21)
and
γ2 (t) = eit for 0 6 t 6 π. (III.1.22)
In this situation, usually where the initial point of γ2 is the final point of γ1 (in this case,
γ1 (1) = γ2 (0)) we write Γ = γ1 + γ2 and define
Z Z Z
f= f+ f. (III.1.23)
Γ γ1 γ2

It is worth noting that there is a consistency check to be carried out here: namely if
γ : [a, b] → C is a smooth curve and c ∈ (a, b), suppose we choose to think of γ as the
sum γ1 + γ2 where
γ1 (t) = γ(t) for a 6 t 6 c (III.1.24)
and
γ2 (t) = γ(t) for c 6 t 6 b (III.1.25)
then we’d better have Z Z Z
f= f+ f. (III.1.26)
γ γ1 γ2
This however is easily verifed and follows from the standard rule
Z c Z b Z b
F (t) dt + F (t) dt = F (t) dt (III.1.27)
a c a
for integration of continuous functions over real intervals.
We also apply the notation Γ = γ1 + γ2 if γ1 and γ2 are closed curves. We shall also
write
Γ = γ1 − γ2 = γ1 + γ2opp . (III.1.28)
Thus
62 III. COMPLEX INTEGRATION

Proposition III.1.16. Suppose that γ1 and γ2 are two smooth curves,


γj : [0, 1] → S
both of which connect z0 to z1 , that is:
γ1 (0) = γ2 (0) = z0 , γ1 (1) = γ2 (1) = z1 . (III.1.29)
Let Γ = γ1 − γ2 be the the piecewise smooth closed curve which starts and ends at z0 .
Then Z Z
f (z) dz = f (z) dz (III.1.30)
γ1 γ2
if and only if Z
f (z) dz = 0. (III.1.31)
Γ

Proof. Trivial reformulation. 


Remark III.1.17. If γ is a curve, we now have two notations for the curve traversed
in the opposite direction, γ opp and −γ. That is fine, but there is some risk of confusion
when we write −γ with the curve t 7→ −γ(t). We believe that the risks of this confusion
are fairly small and shall try to be very clear about this whenever risk of confusion seems
high!
So far we have given some examples. Here is a definition:
Definition III.1.18. If S ⊂ C is a subset, then a piecewise smooth curve γ in S is
a continuous map
γ : [a, b] → S (III.1.32)
together with a partition
a = c0 < c1 < · · · < cn = b (III.1.33)
and a smooth curve
γj : [cj−1 , cj ] → S (j = 1, . . . , n) (III.1.34)
such that γ(t) = γj (t) for t ∈ [cj−1 , cj ].
We allow ourselves to write γ = γ1 + · · · + γn .
As for smooth curves, we say that our piecewise smooth curve γ is simple if a < t1 <
t2 < b implies γ(t1 ) 6= γ(t2 ). γ is a closed curve if γ(a) = γ(b).
The point of this definition is that each γj has γj0 6= 0. Thus each γj is a smooth
path in S; γ is continuous, so the initial point of γj+1 is equal to the final point of γj for
every j; thus the curve γ can be drawn ‘without lifting the pencil from the paper’ but
the direction of travel can jump at the points cj .
We also allow ourselves to write γ = γ1 + · · · + γn if γj : [aj , bj ] → C is a smooth
curve so that the collection of γj satisfy the joining conditions
γj (bj ) = γj+1 (aj+1 )
for all j. This is just a matter of convenience, since it is often a nuisance (and unnec-
essary) to arrange the collection of γj into a single continuous map as in the definition.
1. PATHS AND LINE-INTEGRALS 63

The integral is extended from smooth curves to piecewise smooth curves in the obvious
way: if γ = γ1 + · · · + γn , then
Z n Z
X
f (z) dz = f (z) dz.
γ j=1 γj

Notation III.1.19. A piecewise smooth curve is also often informally referred to as


a contour.

Exercise III.1.8. Calculate the integrals of the following functions around the
perimeter of the square with vertices at 0, 1, 1 + i and i: f (z) = z 2 , f (z) = z,
f (z) = |z|2 .

1.5. The complex fundamental theorem of calculus.


Theorem III.1.20. Let Ω be an open subset of C and let F : Ω → C be holomorphic.
Let γ : [a, b] → Ω be a smooth curve. Then
Z
F 0 (z) dz = F (γ(b)) − F (γ(a)). (III.1.35)
γ

F 0 = 0.
R
In particular, if γ is a closed curve, then γ

Proof. By direct calculation


Z Z b Z b
0 0 0 d
F (z) dz = F (γ(t))γ (t) dt = F (γ(t)) dt = F (γ(b)) − F (γ(a)).
γ a a dt


Example III.1.21. Since the derivative of z n+1 /(n+1) is equal to z n (if n 6= −1),
it follows that if γ is a curve which connects z0 to z1 , then
z n+1 z n+1
Z
z n dz = 1 − 0 .
γ n+1 n+1
Example III.1.22. On the other hand, by direct computation,
Z
dz
= 2πi.
|z|=1 z
It follows that there does not exist a holomorphic function F (z) defined in an open
set containing the circle |z| = 1, with F 0 (z) = 1/z. In other words, there is no way
to define log z as a function holomorphic in C \ {0}.

Theorem III.1.23. Suppose that f : Ω → C is continuous, R where Ω is a domain.


Suppose that for every piecewise smooth closed curve γ in Ω, Γ f (z) dz = 0. Then there
exists a holomorphic function F (z) defined in Ω, such that F 0 (z) = f (z).
64 III. COMPLEX INTEGRATION

Proof. Pick z0 ∈ Ω. For every point z ∈ Ω let γ be a smooth (or piecewise smooth)
path connecting z0 to z, and define
Z
F [γ, z] = f (w) dw. (III.1.36)
γ

If γ1 and γ2 are two curves connecting


R z0 to z, then Γ = γ2 − γ1 is a piecewise smooth
closed curve and by hypothesis, Γ f = 0. It follows that (III.1.36) depends only on z0
and z. Thus Fγ (z) may be written F (z), a well-defined function from Ω into C.
Next we claim that F 0 (z) = f (z). Pick a point z1 ∈ Ω, and choose δ > 0 so that
D(z1 , δ) ⊂ Ω. Let γ connect z0 to z1 . Then γ + [z1 , z1 + h] is a path from z0 to z1 + h
if |h| < δ and we can use this path to define F (z1 + h) (since we can use any path to
define F (z1 + h)). Then
Z
F (z1 + h) = F (z1 ) + f (w) dw (III.1.37)
[z1 ,z1 +h]

and so Z
F (z1 + h) − F (z1 ) − hf (z1 ) = (f (w) − f (z1 )) dw.
[z1 ,z1 +h]

Rearranging,

F (z1 + h) − F (z1 )
Z
1
− f (z1 ) = (f (w) − f (z1 )) dw . (III.1.38)
h |h| [z1 ,z1 +h]

We need to show that the RHS tends to zero as |h| → 0. For this we use the continuity
of f at z1 . Given ε > 0, there exists δ > 0 so that D(z1 , δ) ⊂ Ω and
w ∈ D(z1 , δ) ⇒ |f (w) − f (z1 )| < ε.
If |h| < δ, |f (w) − f (z1 )| is bounded by ε over the segment [z1 , z1 + h] and so by the ML
lemma,
Z
1 1
|h| < δ =⇒ (f (w) − f (z1 )) dw < |h| × ε = ε, (III.1.39)
|h| [z1 ,z1 +h] |h|

since the length of the segment is |h|. Combining this with (III.1.38) we have proved
that F 0 (z1 ) exists and is equal to f (z1 ), as required. 

Exercise III.1.9. Show that if Γ is any piecewise smooth closed curve in C, then
Z
sin z dz = 0.
Γ
Exercise III.1.10. Show that if f : D(0, R) → C is given by a convergent
power
R series in D(0, R) and Γ is any piecewise smooth closed curve in D(0, R), then
Γ f (z) dz = 0. [Can you apply Theorem III.1.20?]
2. GREEN’S THEOREM, CAUCHY’S THEOREM AND CAUCHY’S INTEGRAL FORMULA 65

2. Green’s Theorem, Cauchy’s Theorem and Cauchy’s Integral Formula


This section contains two results that are the cornerstones of complex analysis,
Cauchy’s Theorem and Cauchy’s Integral Formula. We do not attempt to give the
most general versions. The reason is that to do so would take us too far afield with
topological issues. These are quite easily dealt with using technology developed in the
20th Century but are beyond the scope of this course.
An informal statement of Cauchy’s Theorem is:
If f isR holomorphic inside and on a piecewise smooth closed curve γ,
then γ f (z) dz = 0.
The difficulty with this statement is: what does ‘inside and on’ mean? If you have a
closed curve in the plane, how do you define its inside and outside, for example?
Let us restrict the statement to simple, closed, piecewise-smooth curves γ. Then it
is still not obvious, but experience suggests that the image of γ will decompose C into
precisely two disjoint pieces, one of which is unbounded, and the other of which should
be called the ‘inside’. If this is taken on trust and f is holomorphic
R in an open set
containing the ‘inside’ of γ, then Cauchy’s Theorem states that γ f (z) dz = 0.
Let’s explore, informally, some consequences. First, path-independence of definite
integrals. Suppose that γ1 and γ2 are two piecewise smooth simple paths connecting z0
to z1 . Then γ1 − γ2 = Γ is a closed curve. If f is holomorphic ‘inside and on’ Γ, then
Cauchy states that Z Z
f (z) dz = f (z) dz. (III.2.1)
γ1 γ2
Informally, if you are given f and you can ‘deform’ γ1 to γ2 in such a way that the
intermediate curves are all contained in the domain where f is holomorphic, then (III.2.1)
will hold.
There is a corresponding result about being able to move closed contours without
changing the result. For this, suppose that Γ1 and Γ2 are two disjoint piecewise smooth
simple closed curves and that Γ1 is ‘inside’ Γ2 . Then there is a set D ‘between’ Γ1 and
Γ2 , and if f is holomorphic on an open subset containing D, Cauchy’s Theorem states
that Z Z
f= f.
Γ1 Γ2
We shall get around the problem of what is the inside of a curve by starting with
the ‘inside’ and deriving the curve as its boundary!
2.1. Domains with piecewise smooth boundary.
Definition III.2.1. We say that a bounded domain D has piecewise smooth bound-
ary ∂D = Γ1 + · · · + Γn if:
(i) D is a bounded domain, that is, D is a bounded, open, path-connected subset
of C;
(ii) The boundary ∂D is a disjoint union of path-connected subsets S1 , . . . , Sn
(iii) There is a collection of n piecewise-smooth simple closed curves Γj : [aj , bj ] → C
such that the restriction of Γj to [aj , bj ) is a bijection of [aj , bj ) with Sj ;
(iv) The domain D is on the left relative to the direction in which Γj is traversed.
66 III. COMPLEX INTEGRATION

Remark III.2.2. Recall that if z is thought of as a vector, then iz represents that


vector rotated through 90 degrees anti-clockwise. Thus we can define ‘D is on the left
relative to the direction in which Γj is traversed’ by saying that γ(t) + isγ 0 (t) ∈ D for
all small enough positive s and all t in the parameter interval (for each smooth part γ
of the boundary).
Example III.2.3. The bounded domain D = D(z0 , r) has piecewise smooth bound-
ary consisting of the single simple closed curve Γ,
Γ(t) = z0 + reit , 0 6 t 6 2π.
Note that Γ has to be traversed this way for condition (iv) of the definition to be satisfied.
Example III.2.4. The bounded domain A = {z : r1 < |z − z0 | < r2 } has piecewise
smooth boundary Γ
e2 + Γ
e 1 , where
e 1 (t) = z0 + r1 e−it , 0 6 t 6 2π.
Γ
and
e 2 (t) = z0 + r2 eit , 0 6 t 6 2π.
Γ
Note that the inner circle Γ
e 1 is traversed clockwise in order to satisfy part (iv) of the
definition. We also write ∂A = Γ2 − Γ1 , where
Γ1 (t) = z0 + r1 eit , 0 6 t 6 2π.
and
Γ2 (t) = z0 + r2 eit , 0 6 t 6 2π.
Both Γ1 and Γ2 are traversed anticlockwise. The sign in front of Γ1 reminds us that we
need to traverse it in the opposite sense in order to satisfy part (iv) of the definition.
Example III.2.5. The rectangle
R = {z : a < x < b, c < y < d}
has piecewise smooth boundary ∂R = Γ, where
Γ = γ1 + γ2 + γ3 + γ4
where
γ1 = [a + ic, b + ic], γ2 = [b + ic, b + id], γ3 = [b + id, a + id], γ4 = [a + id, a + ic].
so that the piecewise smooth curve Γ is the perimeter of the rectangle, traversed anti-
clockwise.
Example III.2.6. Continuing this example, if z0 ∈ R and δ is so small that D(z0 , δ) ⊂
R, then
Rδ = R \ D(z0 , δ)
is a bounded connected open set with boundary
∂Rδ = Γ − Γδ
where
Γδ (t) = z0 + δeit , 0 6 t 6 2π.
2. GREEN’S THEOREM, CAUCHY’S THEOREM AND CAUCHY’S INTEGRAL FORMULA 67

Exercise III.2.11. Sketch the domain


D = {z : −10 < x < 10, −10 < y < 10} ∩ {z : |z − 5| > 1} ∩ {z : |z + 5| > 1}.
Does it have piecewise smooth boundary? Draw arrows on your diagram to show
the direction in which the boundary components are to be traversed.
Exercise III.2.12.
(a) Is the open upper half space a bounded domain with piecewise smooth
boundary?
(b) Is the punctured disc D∗ (a, R) a bounded domain with piecewise smooth
boundary?

We are now going to embark on Cauchy’s Theorem. This is a theorem with different
proofs and approaches, each with their advantages and disadvantages.
The first proof uses Green’s Theorem in the plane:
Theorem III.2.7. Let D be a bounded domain with piecewise smooth boundary ∂D
(see Definition III.2.1 for notation). Suppose that P and Q are two real- or complex-
valued functions which are C 1 in an open neighbourhood of D ∪ ∂D. Then we have
Z   Z
∂Q ∂P
− dxdy = P dx + Q dy (III.2.2)
D ∂x ∂y ∂D

Remark III.2.8. The above theorem is true as stated, but you may not have seen
a rigorous proof of this. Such a rigorous proof would certainly need to define the LHS
(namely the integral over the domain D). We shall leave this for now. See Chapter IV
for discussion.
Theorem III.2.9. (Cauchy’s Theorem) Let D be a bounded domain with piecewise
smooth boundary ∂D. Suppose that f : Ω → C is holomorphic in an open set Ω which
contains D ∪ ∂D. Then Z
f (z) dz = 0. (III.2.3)
∂D

Proof. We suppose in addition that f is C 1 in a neighbourhood of D. Since


Z Z
f (z) dz = f (z)(dx + idy)
∂D ∂D
we take P = f , Q = if in Green’s Theorem. We obtain
Z Z Z
f (z) dz = ((if )x − fy ) dxdy = i (fx + ify ) dxdy.
∂D D D
But this is zero by the Cauchy–Riemann equations. 
Notation III.2.10. If Γ is a disjoint union of simple, piecewise smooth curves, we
shall allow ourselves to write ‘f is holomorphic inside and on Γ’ to mean that:
• There is a bounded domain D such that ∂D = Γ;
68 III. COMPLEX INTEGRATION

• There is an open set Ω such that D ∪ ∂D ⊂ Ω, and


• f : Ω → C is holomorphic.

Using this shorthand, Cauchy’s Theorem can be stated as follows: if Γ is a disjoint


union of piecewise smooth curves and f is holomorphic inside and on Γ, then
Z
f (z) dz = 0.
Γ

Here is an application which ties in with the complex fundamental theorem of cal-
culus.
Theorem III.2.11. Let Ω be convex open set and let f : Ω → C be holomorphic.
Then there exists F : Ω → C, holomorphic, such that F 0 = f in Ω.
Proof. We proceed as in the proof of Theorem III.1.23. We allow ourselves to
assume that f is C 1 as well. Select any point z0 ∈ Ω and for z ∈ Ω, define
Z
F (z) = f (w) dw. (III.2.4)
[z0 ,z]

Then F (z) is defined using a specific path from z0 to z, so definitely a well-defined


function in Ω. To show that F 0 = f , choose z1 ∈ Ω and a small disc D(z1 , δ) ⊂ Ω. The
triangle T with vertices z0 , z1 and z1 + h is entirely contained in Ω by convexity of Ω,
and it is a bounded domain with piecewise Rsmooth boundary; moreover f is holomorphic
on T ∪ ∂T . Hence by Cauchy’s Theorem, ∂T f = 0. Spelling this out,
Z
F (z1 + h) − F (z1 ) = f (z) dz. (III.2.5)
[z1 ,z1 +h]

This is identical to (III.1.37) in the proof of Theorem III.1.23. That F 0 (z1 ) = f (z1 )
follows in exactly the same way as in the proof of Theorem III.1.23. 
Remark III.2.12. The main idea is that the hypothesis ‘f holomorphic’ implies that
the integral of f around the boundary of any triangle ∂T is zero, and this is sufficent to
run the argument of Theorem III.1.23.
R
Corollary III.2.13. If f is holomorphic in a convex set Ω, then γ f (z) dz = 0 for
any piecewise smooth closed curve γ contained in Ω.
Proof. By Theorem III.2.11, there is a holomorphic function F : Ω → C such that
F0= f . Then the result follows from the complex fundamental theorem of calculus,
Theorem III.1.20. 

2.2. The Logarithm. Consider the cut plane Ω = C \ (−∞, 0]. Define
Z z
dw
L(z) = . (III.2.6)
1 w
where the integral is along the unique straight line which connects 1 to z. It is clear that
1/w is holomorphic in any triangle contained in Ω, so the integral around the perimeter
2. GREEN’S THEOREM, CAUCHY’S THEOREM AND CAUCHY’S INTEGRAL FORMULA 69

of such a triangle is zero, and so by the same argument as in Theorem III.1.20, L(z) is
holomorphic in Ω and L0 (z) = 1/z, L(1) = 0.
Let K(z) = z exp(−L(z)). Then K is holomorphic in Ω and
K 0 (z) = exp(−L(z)) − zL0 (z) exp(−L(z)) = 0 since zL0 (z) = 1.
Hence K(z) is constant and since K(1) = 1 this constant value is 1. So
exp(L(z)) = z
and L(z) is an inverse of the exponential function, also known as the logarithm.
Now we know that L0 (z) = 1/z in Ω, we have path-independence of the integral (as
long as the path is contained in Ω, of course). In particular, we compute
Z Z
dw dw
L(z) = +
γ1 w γ2 w

where γ1 is the segment [1, |z|] and γ2 is the circular arc from |z| to z, then we have
Z
dw
= log |z| (III.2.7)
γ1 w
and Z Z arg(z)
dw
= idt = i arg(z). (III.2.8)
γ2 w 0
Hence L(z) = log |z| + i arg(z), where
−π < arg(z) < π for z ∈ Ω
is the principal value of arg(z).
L(z) is called a holomorphic branch of log z. Any other holomorphic branch of log z
defined in Ω is obtained by adding 2nπi to L(z), where n ∈ Z. Holomorphic branches
of log z can be defined in other open sets of C. For example, if Ω0 = C \ [0, ∞), we have
a holomorphic branch of log z,
log(z)
f = log |z| + iarg(z),
f 0 < arg(z)
f < 2π.
And we can add 2nπi to any holomorphic branch to define another branch holomorphic
in the same domain.
Using the holomorphic branch L(z) of log z in Ω, we may also define a holomorphic
branch of the complex power z a , by defining
z a := exp(aL(z)) for z ∈ Ω. (III.2.9)
This is a holomorphic function in the cut plane Ω. In general, different holomorphic
branches of z a are determined by different holomorphic branches of log z. z a can only
be defined holomorphically in the whole of C if a is a non-negative integer; in this case
it is independent of the choice of holomorphic branch of log z.

Exercise III.2.13. Show that there are precisely two distinct holomorphic
branches of z 1/2 defined in the cut plane Ω. How many distinct holomorphic branches
of z 1/3 are there? Write down all holomorphic branches of z a defined in Ω.
70 III. COMPLEX INTEGRATION

Exercise III.2.14. Suggest a domain U , containing the part of the real axis with
x > 1, on which a holomorphic branch of log(z − 1) can be defined.

2.3. Cauchy’s Integral Formula.


Theorem III.2.14. (Cauchy’s Integral Formula) Let D be a bounded domain with
piecewise smooth boundary ∂D. Suppose that f is holomorphic in an open set Ω which
contains D ∪ ∂D. Let z0 be a point of D. Then
Z
1 f (z)
f (z0 ) = dz. (III.2.10)
2πi ∂D z − z0

f (z)
Proof. Let F (z) = z−z 0
and let Dδ = D \ D(z0 , δ), where δ > 0 is so small that
D(z0 , δ) ⊂ D. Then F is holomorphic in Ω \ {z0 } which contains Dδ . Thus Cauchy’s
Theorem applies to F and the domain Dδ . We have
∂Dδ = ∂D − γ
where γ(t) = z0 + δeit , 0 6 t 6 2π. Thus Cauchy gives
Z Z
F (z) dz = F (z) dz. (III.2.11)
∂D γ
The integral on the RHS is computed explicitly as
Z 2π Z 2π
f (z0 + δeit ) it
it
iδe dt = i f (z0 + δeit ) dt.
0 δe 0
Now we argue as in the proof of Theorem II.5.8.
Z 2π Z 2π Z 2π
f (z0 + δeit ) dt = f (z0 ) dt + (f (z0 + δeit ) − f (z0 )) dt. (III.2.12)
0 0 0
Given ε > 0, there exists δ > 0 so that
|z − z0 | < 2δ ⇒ |f (z) − f (z0 )| < ε/2π.
For such δ,
Z 2π Z 2π
it
(f (z0 + δe ) − f (z0 )) dt 6 |f (z0 + δeit ) − f (z0 )| dt < ε.
0 0
R 2π
Since 0 f (z0 ) dt = 2πf (z0 ), the result follows by combining this estimate with (III.2.12).

Proposition III.2.15. If |a| =
6 R, then
Z 
dz 2πi if |a| < R;
= (III.2.13)
|z|=R z − a 0 if |a| > R
Proof. If |a| > R then a is outside the disc D(0, R) and so 1/(z − a) is holomorphic
in a neighbourhood of this disc. Hence by Cauchy’s Theorem, the integral is zero.
If |a| < R, then we can apply Cauchy’s Integral formula with f (z) = 1. 
2. GREEN’S THEOREM, CAUCHY’S THEOREM AND CAUCHY’S INTEGRAL FORMULA 71

Exercise III.2.15. Calculate


Z
dz
|z|=R (z − z1 )(z − z2 )
for any z1 6= z2 such that |z1 | =
6 R, |z2 | =
6 R.
Exercise III.2.16. Complete the proof of the Cauchy Integral Formula along
the following lines. Starting from the RHS of (III.2.11), show that
f (z) − f (z0 )
Z Z
F (z) dz − 2πif (z0 ) = dz
γ γ z − z0
and use the ML lemma (rather than parameterizing γ) to show that this tends to
zero as δ → 0.

2.4. Comments. Once we have Green’s Theorem for bounded domains with piece-
wise smooth boundary, the proof of Cauchy’s Theorem for a holomorphic function f is
very simple, at least if we assume that f is continuously differentiable. Here Green’s
Theorem is being used as a ‘black box’, and in the interest of rigour and conceptual
foundations of the subject it is worth thinking about how ‘easy’ a theorem Green’s
Theorem really is. I give a sketch proof of Green’s Theorem in Chapter IV, so if you
are interested, you can read about it and decide for yourself. The two main difficulties
are defining the two-dimensional integral over D, and dealing with a general piecewise
smooth boundary. It requires a fair amount of spadework to deal with these issues in
a fully rigorous fashion. I should emphasise, however, that Green’s Theorem and its
higher-dimensional cousins (the divergence theorem in 3D, for example) can be made
entirely rigorous and are fundamental tools in other areas of geometry and the analysis
of PDE (as well as applied mathematics) so I think it is good to see Cauchy’s Theorem
and the Cauchy Integral Formula derived using this tool.
Apart from the difficulties involved in giving a rigorous account of Green’s Theorem,
a possible defect in the proof which uses it is the need to assume that our holomorphic
function is C 1 . The Goursat2 or Cauchy–Goursat
R Theorem avoids this assumption.
Goursat’s Theorem establishes that ∂R f (z) dz = 0 if f is holomorphic in an open
set containing a closed rectangle3 R. For the detailed proof, see any reputable text on
complex analysis. The idea is as follows.
Suppose
Z
f (z) dz 6= 0.
∂R
By multiplying f by a non-zero complex number, suppose then that
Z
f (z) dz = 1. (III.2.14)
∂R

2Edouard Goursat, French mathematician, 1858–1936.


3or triangle
72 III. COMPLEX INTEGRATION

Also, let L = L(∂R) be the length of the perimeter of the rectangle. Divide R into four
equal rectangles by cutting R into four quarters. Denote these S1 , S2 , S3 , S4 . Then
Z X4 Z
f= f (III.2.15)
∂R j=1 ∂Sj

(there is cancellation along the ‘interior’ sides of the Sj ) and by the triangle inequality
Z X 4 Z
1= f (z) dz 6 f . (III.2.16)
∂R j=1 ∂Sj
R
It follows that for at least one of the four Sj , | ∂Sj f | > 1/4. Denote by R1 a choice of
R
Sj such that | ∂Sj f | > 1/4. Note that L(∂R1 ) = L/2.
R
Continue the argument: cut R1 into quarters, pick a quarter R2 on which | ∂R2 f | >
4−2 and so on. We obtain a nested sequence of closed rectangles
R ⊃ R1 ⊃ R2 ⊃ . . . ⊃ Rn ⊃ . . .
with
• L(∂R −n
R n ) = 2 L, −n
• | ∂Rn f (z) dz| > 4 .
Because the Rn are nested rectangles and L(∂Rn ) → 0, there is a unique point z0
common to all Rn . The proof is finished
R off by using the fact that f is holomorphic
at z0 . This allows us to estimate | ∂Rn f | for large n, and in particular to obtain an
estimate which contradicts the lower bound 4−n above.
Once we have the Cauchy–Goursat Theorem, it is possible to prove directly the
Cauchy Integral Formula in the form
Z
1 f (z) dz
f (w) = (III.2.17)
2πi |z−z0 |=r z − w
for w ∈ D(z0 , r). From this we obtain that f is analytic in D(z0 , r) and in particular
that f is C 1 . At this point we could use Green’s Theorem to obtain the more general
form of Cauchy’s Theorem stated in Theorem III.2.9. There are also variants of this line
of argument that avoid Green’s Theorem completely.

3. Holomorphic functions are analytic; higher derivatives


In this section we shall prove
Theorem III.3.1. (Cauchy’s Formula for derivatives) Let D be a bounded domain
with piecewise smooth boundary ∂D. Suppose that f is holomorphic in an open set Ω
which contains D ∪ ∂D. Let z0 be a point of D. Then f is infinitely differentiable and
Z
n! f (z)
f (n) (z0 ) = dz. (III.3.1)
2πi ∂D (z − z0 )n+1

We shall also prove that H(Ω) = A (Ω), more precisely:


3. HOLOMORPHIC FUNCTIONS ARE ANALYTIC; HIGHER DERIVATIVES 73

Theorem III.3.2. Let f ∈ H(Ω) and let z0 ∈ Ω. Let D(z0 , R) ⊂ Ω. Then the Taylor
expansion

X
f (z) = an (z − z0 )n , (III.3.2)
n=0
where
f (n) (z0 )
an = , (III.3.3)
n!
is absolutely convergent in D(z0 , R). Moreover, for any r < R, we have Cauchy’s
Inequalities
|an | 6 r−n sup |f (z)|. (III.3.4)
|z−z0 |=r

Remark III.3.3. Formally, if (III.3.1) is true for n then we might imagine being
able to differentiate both sides with respect to z0 . Assuming it is valid to pass the
differentiation through the integral, we obtain (III.3.1) with n replaced by n + 1, since
d n! (n + 1)!
n+1
= .
dz0 (z − z0 ) (z − z0 )n+2
The base case n = 0 is the Cauchy Integral Formula, Theorem III.2.14. This argument
can be made rigorous.
We shall prove Theorem III.3.2 first. Cauchy’s formula for derivatives will follow
from this proof.

Proof. (Of Theorem III.3.2.) Pick any positive r, r < R. It is enough to show that
(III.3.2) is valid in D(0, r) for every r < R.
To simplify notation, suppose that z0 = 0.
For |z| < r, Cauchy’s Integral Formula gives
Z
1 f (w)
f (z) = dw. (III.3.5)
2πi |w|=r w − z
The idea is to expand 1/(w − z) using the binomial theorem. For |w| = r, |z| < r,
we have

1 1 X  z n
= , since |z/w| < 1. (III.3.6)
w−z w w
n=0
If we naively subsitute this into (III.3.5) and interchange the sum and the integral, we
obtain
X
f (z) = an z n for |z| < r (III.3.7)
where Z
1 f (w)
an = dw. (III.3.8)
2πi |w|=r wn+1
Then (III.3.7) gives f as a convergent power series.
74 III. COMPLEX INTEGRATION

In order to justify this argument, i.e. the interchange of the infinite sum and the
integral, let us replace (III.3.6) by a finite sum plus an error term. If |u| < 1, and N > 0,
we have

X N
X ∞
X
un = un + un (III.3.9)
n=0 n=0 n=N +1
N
X uN +1
= un + . (III.3.10)
1−u
n=0

Hence in place of (III.3.6) we write


N
1 1 X (z/w)N +1
= (z/w)n + . (III.3.11)
w−z w w−z
n=0
and this gives
N
X
f (z) = an z n + EN , (III.3.12)
n=0
where an is given by (III.3.8) and
f (w)(z/w)N +1
Z
1
EN = dw (III.3.13)
2πi |w|=r w−z
Now let |z|/r = α < 1. We use the ML lemma to estimate EN . We have
|w − z| > |w| − |z| = r(1 − α).
So by ML,
αN +1 αN +1
|EN | 6 r × M × =M . (III.3.14)
r(1 − α) 1−α
where M = sup|w|=r |f (w)|. Thus if we fix z with |z| < r so that α = |z|/r < 1, then
N
X
f (z) − an z n → 0 as N → ∞.
n=0

zn
P
Thus the power series an converges to f (z) in D(0, r).
Cauchy’s Inequalities follow by applying the ML lemma to (III.3.8).

Proof. (Of Theorem III.3.1.) In the course of the previous proof we showed that
f is analytic in A (Ω), so all derivatives exist. We already know that if f is analytic,
the coefficients in its power series expansion centred at 0 are the Taylor coefficients
f (n) (0)/n!. But we have the integral formula (III.3.8) for an , so
f (n) (0)
Z
1 f (w)
= an = dw. (III.3.15)
n! 2πi |w|=r wn+1
To prove the stated result, with the more general contour, suppose that D is as given in
the statement of Theorem III.3.1; we continue to assume z0 = 0. Pick r so small that
3. HOLOMORPHIC FUNCTIONS ARE ANALYTIC; HIGHER DERIVATIVES 75

D(0, r) ⊂ D and let Dr = D \ D(0, r) Then z 7→ z −n−1 f (z) is holomorphic inside and
on ∂Dr so Cauchy’s Theorem applies to give
Z Z
f (z) dz f (z) dz
n+1
= n+1
∂D z |z|=r z

and we have already shown that the RHS is equal to f (n) (0)/n!. 
3.1. Liouville’s Theorem.
Theorem III.3.4. Suppose that f : C → C is holomorphic. If f is bounded then f
is constant.
Proof. Suppose that f is bounded, |f (z)| 6 M for all z ∈ C. We use the formula
for the derivative to prove that f 0 (z0 ) = 0 for all z0 ∈ C. Let z0 ∈ C and let R > 0.
Then Z
0 1 f (z)
f (z0 ) = dz
2πi |z−z0 |=R (z − z0 )2
and using the ML lemma,
M
|f 0 (z0 )| 6 . (III.3.16)
R
Here M is fixed, but R can be chosen as we please. Taking R to ∞, we see that
f 0 (z0 ) = 0. 

Exercise III.3.17. Prove Liouville’s Theorem by using Cauchy’s Integral for-


mula to show that
Z  
1 1 1
f (z0 ) − f (0) = − f (z) dz (III.3.17)
2πi |z|=R z − z0 z
for any R > |z0 |. Take R to ∞ and use ML to show that f (z0 ) = f (0).
Exercise III.3.18. Prove Liouville’s Theorem as follows. Since f ∈ A (Ω), f
has a convergent power series expansion

X
f (z) = an z n
n=0
where the an satisfy Cauchy’s Inequalities for any r > 0. Show that the boundedness
of |f | implies that an = 0 for n > 1.
Exercise III.3.19. Prove the following generalization of Liouville: if f ∈ H(C)
and there exists M > 0, N > 0 such that
|f (z)| 6 M (1 + |z|N ) for all z ∈ C
then f is a polynomial of degree 6 N .
Exercise III.3.20. Suppose that f ∈ H(C) and for some a ∈ C and δ > 0,
f (z) 6∈ D(a, δ) for all z ∈ C.
Show that f is constant.
76 III. COMPLEX INTEGRATION

3.2. Fundamental Theorem of Algebra.


Theorem III.3.5. Let p(z) be a non-constant complex polynomial. Then p has a
zero in C.
Proof. If not, then f (z) = 1/p(z) is holomorphic in C. We claim that f (z) is
bounded, so Liouville’s Theorem then says that f (z) is constant, a contradiction.
f (z) is certainly bounded on any disc |z| 6 R. So we need to be sure that f is also
bounded for |z| → ∞.
Suppose that
p(z) = a0 + a1 z + · · · + an−1 z n−1 + an z n , an 6= 0.
Then |p(z)| → ∞ as |z| → ∞ and so by standard results about limits, there exists R > 0
so that |z| > R implies that |p(z)| > 1. Thus |f (z)| < 1 for |z| > R. Let M be an upper
bound for |f | on the closed disc D(0, R). Then
|f (z)| 6 max(1, M ) for all z ∈ C.
Liouville gives that f is constant. Hence p = 1/f is constant, contradiction. 
Of course, once p has one complex zero, it has a complete factorization, by induction.
3.3. Morera’s Theorem.
Theorem III.3.6. Suppose that Ω is a convex open subset of C and that f is con-
tinuous in Ω. Then if the integral of f around the perimeter of any triangle in Ω is 0,
it follows that f ∈ H(Ω).
Proof. Pick z0 ∈ Ω and define
Z
F (z) = f (z) dz.
[z0 ,z]
Using the vanishing of the integrals around perimeter triangles, we argue as in the proof
of Theorem III.2.11 that F is holomorphic and F 0 = f . But now we know that if F
is holomorphic, then F 0 is also holomorphic (by Theorem III.3.2, all derivatives of F
exist!). So f ∈ H(Ω). 
CHAPTER IV

Zeros, poles, and the residue theorem

1. Zeros of analytic functions


1.1. Polynomials. If p(z) = a0 + a1 z + · · · + an z n is a non-constant polynomial
(with an 6= 0, n > 1) then by repeated application of the fundamental theorem of algebra,
we find distinct complex numbers α1 , . . . , αm such that
p(z) = an (z − α1 )k1 (z − α2 )k2 · · · (z − αm )km (IV.1.1)
such that k1 + · · · + km = n. The αj are the zeros or roots of p, and kj > 0 is the order
of the zero αj . A zero of order 1 is also called a simple zero.
Notice that the zeros are isolated in the sense that for each j, there exists δ > 0 such
that
p(z) 6= 0 if 0 < |z − αj | < δ. (IV.1.2)
In words, near any zero αj of p, p(z) is non-zero for all z sufficiently close to αj (but
different from αj ).
Thus non-constant polynomials have isolated zeros of finite order, and the only other
possibility is p(z) = a0 which has no zeros at all if a0 6= 0 or every complex number a
zero if a0 = 0.

Exercise IV.1.1. Show that a zero α of p(z) has order k if and only if
p(α) = p0 (α) = · · · = p(k−1) (α) = 0, p(k) (α) 6= 0. (IV.1.3)
Exercise IV.1.2. If p(z) is a polynomial of degree 6 10 and there are 17 distinct
complex numbers α1 , . . . , α17 for which p(αj ) = 0, what can be said about p(z)?
Generalize.

We claim that the above story is very nearly the same for analytic functions.
1.2. Zeros of analytic functions. Let Ω be a domain and let f ∈ A (Ω). Suppose
that z0 is a zero of f , so z0 ∈ Ω and f (z0 ) = 0.
Since f is analytic, we have a disc D(z0 , δ) such that we have a power series expansion,

X
f (z) = an (z − z0 )n , for z ∈ D(z0 , δ). (IV.1.4)
n=0
Since f (z0 ) = 0, we have a0 = 0. There are precisely two possibilities:
(a) z0 is a zero of finite order k, say: this happens by definition if
a0 = · · · = ak−1 = 0, ak 6= 0;
77
78 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

(b) z0 is a zero of infinite order, that is, an = 0 for all n.


By Taylor’s Theorem, the condition that z0 be a zero of order k is equivalent to
f (z0 ) = f 0 (z0 ) = · · · = f (k−1) (z0 ) = 0, f (k) (z0 ) 6= 0. (IV.1.5)
Note that if z0 is a zero of finite order k, then we have
f (z) = ak (z − z0 )k + ak+1 (z − z0 )k+1 + · · · (IV.1.6)
X∞
k
= (z − z0 ) an+k (z − z0 )n (IV.1.7)
n=0
The sum is clearly convergent so we have a factorization
f (k) (z0 )
f (z) = (z − z0 )k g(z), g(z0 ) 6= 0, g(z0 ) = . (IV.1.8)
k!
Since g(z0 ) 6= 0, g(z) 6= 0 for all z with |z − z0 | sufficiently small. Hence if z0 is a zero
of finite order, then it is an isolated zero: f (z) 6= 0 for all z sufficiently near but
not equal to z0 .
Proposition IV.1.1. Let z0 be a zero of f ∈ A (Ω). Then EITHER
• z0 is a zero of finite order, or equivalently
• z0 is an isolated zero of f ;
OR
• z0 is a zero of infinite order, or equivalently
• z0 is a non-isolated zero, or equivalently
• f (z) is identically zero in some small disc D(z0 , δ).

Proof. We’ve seen that finite-order is equivalent to isolated. Thus non-isolated


must be equivalent to infinite-order since these are mutually exclusive and exhaustive
possibilities! So the only thing left is to see why infinite order is equivalent to the third
condition, that f is identically zero in some small disc. But infinite-order means that
all coefficients in the power-series expansion of f centred at z0 vanish from which it is
immediate that f (z) = 0 for all z in some D(z0 , δ). 
In fact, using the path-connectedness of Ω we have a stronger statement:
Theorem IV.1.2. Let Ω be a domain and let f ∈ A (Ω). Then the following are
equivalent:
(a) f has a non-isolated zero z0 in Ω;
(b) f has a zero z0 of infinite order in Ω;
(c) f is identically zero in some disc D(z0 , δ) ⊂ Ω;
(d) f is identically zero in Ω.

Proof. We have already seen that conditions (a)–(c) are equivalent. So we have to
prove that any one of them is equivalent (d). Clearly (d) is the strongest of the lot and
implies the other three. So it is enough to prove that (c)⇒(d). For this, we follow a
line which we have encountered at least twice before (Theorem II.4.5 and the Maximum
Principle, Theorem II.5.12), using the path-connectedness of Ω.
1. ZEROS OF ANALYTIC FUNCTIONS 79

Let z1 ∈ Ω and let γ : [0, 1] → Ω be a curve connecting z0 to z1 , so that γ(0) = z0 ,


γ(1) = z1 . Let
T = sup{t ∈ [0, 1] : f (γ(t)) = 0} (IV.1.9)
The set of which we are taking the sup is bounded and non-empty (it contains t = 0)
and so T is a real number 6 1. By continuity of f at γ(T ), f (γ(T )) = 0, for f (γ(t)) = 0
for all t < T and γ(t) → γ(T ) as t → T . So if T = 1, we are done, because this means
that f (z1 ) = 0.
Suppose, for a contradiction, that T < 1. Let γ(T ) = z 0 . We have seen that z 0 is a
zero of f and it must be a non-isolated zero, because again by definition of T , f (γ(t)) = 0
for all t < T and so there are zeros of f arbitrarily close to z 0 —namely any point γ(t)
with t < T . By Proposition IV.1.1, f is identically zero in some D(z 0 , δ) and such a disc
must necessarily contain some points γ(t) with t > T . This contradicts the definition of
T.
Since z1 was an arbitrary point of Ω we conclude that f (z) = 0 for all z ∈ Ω. 
Remark IV.1.3. In some sense, this shows that, as far as zeros are concerned,
analytic functions and polynomials behave in quite similar fashion. A polynomial is
either identically zero or it has some positive degree d. If it has positive degree d, then
it has a finite number of isolated zeros.
If f is now an analytic function, then either it is identically zero or every zero is of
finite order and isolated. The difference is that there may be infinitely many zeros (or
none); moreover, if f has infinitely many zeros, there need not be any upper bound on
the orders of these zeros.
In practice, the following summarizes the essential part of the previous Theorem: it
is often referred to as the the identity theorem.
Theorem IV.1.4 (Identity Theorem). Suppose that Ω is a domain of C and that
f ∈ A (Ω). Let Z be the set of zeros of f in Ω. Then if Z has an accumulation point in
Ω, f is identically zero in Ω.
Proof. Let z0 ∈ Ω be a limit point of Z. Then by continuity, z0 is a zero of f , and
by assumption it is non-isolated. Now apply Theorem IV.1.2 

Exercise IV.1.3. What are the orders of the zeros of ez −1, of sin πz, of z −sin z
(at z = 0)?
Exercise IV.1.4. Write down three further simple analytic functions and find
the order of their zeros.
Exercise IV.1.5. A function f ∈ A (C) is identically 0 on the real axis. Show
that f (z) = 0 in C.
Exercise IV.1.6. A function f (z) is holomorphic in C and is real on the real
axis. Show that g(z), where
g(z) = f (z̄)
is also holomorphic in C. Show that f (z) = g(z) for all z.
80 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

2. Poles and Laurent series


In this section we discuss and classify isolated singularities of holomorphic functions
and we define the residue at an isolated singular point.
Definition IV.2.1. We say that z0 is an isolated singularity of the function f if f is
holomorphic in an open set Ω which contains a punctured disc D∗ (z0 , δ), for some δ > 0.
Example IV.2.2. For the function f (z) = sin z, in Ω = C \ {0}, 0 is an isolated
singularity. For the function g(z) = 1/(1−z), for Ω = C\{1}, 1 is an isolated singularity.
For the function h(z) = exp(1/z), for Ω = C \ {0}, 0 is an isolated singularity.
The three examples illustrate three different types of behaviour consistent with the
definition. In the first case, if we define f (0) = 0, then we have extended the definition
of f to the ‘singularity’ z = 0 so that f is still holomorphic. Such an isolated singularity
is called ‘removable’. In the case of g(z), we see that g(z) → ∞ as z → 1, so there is no
way to extend the definition of g so as to be holomorphic in C1. This type of isolated
singularity is called a pole.
Finally, the singularity of h(z) is known as an essential singularity. The limiting
behaviour of h(z) as z → 0 is highly direction-dependent in this case: see the following
exercise.

Exercise IV.2.7. Show that h(z) remains bounded as z → 0 along the imaginary
axis. What happens to h(z) as z → 0 along the real axis a) from the left? and b)
from the right?
Show that for every non-zero w ∈ C, there is a sequence zn → 0 (depending on
w) such that f (zn ) = w as n → ∞.
Thus in any neighbourhood of 0, exp(1/z) takes every complex value apart from
0. Is that shocking?

We shall now give a precise and general definition of these three types of singularity.
We shall do this by the use of Laurent Series.
By an annulus A, we mean a set of the form
A = {z ∈ C : a < |z| < b}. (IV.2.1)
where a and b are real numbers. We allow the possibility that a = 0, in which case A is
the punctured disc centred at 0 of radius b, also denoted D∗ (0, b).
Laurent’s Theorem gives an expansion in positive and negative powers of z of any
function holomorphic in A.
Theorem IV.2.3. Let f : A → C be holomorphic. Then there exist coefficients
cn ∈ C such that

X
f (z) = cn z n for all z ∈ A. (IV.2.2)
n=−∞

1However, it can be regarded as a holomorphic mapping from C into the Riemann sphere S
2. POLES AND LAURENT SERIES 81

Moreover, we have the formula


Z
1
cn = z −n−1 f (z) dz (IV.2.3)
2πi γ

where γ(t) = reit , 0 6 t 6 2π, for any fixed radius r between a and b.
Proof. For a < a0 < b0 < b, the smaller annulus A0 = {a0 < |z| < b0 } is a bounded
domain whose closure is contained in A. We have ∂A0 = γ1 − γ2 , where
γ1 (t) = b0 eit , 0 6 t 6 2π,
is the outer circle traversed anti-clockwise, and
γ2 (t) = a0 eit , 0 6 t 6 2π.
is the inner circle traversed anticlockwise. The formula ∂A0 = γ1 − γ2 takes care of the
orientation of the two components of ∂A0 . If z ∈ A0 , we have, by Cauchy’s Integral
Formula, Theorem III.2.14,
Z Z Z
1 f (w) 1 f (w) 1 f (w)
f (z) = dw = dw − dw. (IV.2.4)
2πi ∂A0 w − z 2πi γ1 w − z 2πi γ2 w − z
Now we expand (w − z)−1 in powers, noting that on γ1 , |w| = b0 > |z|, while on γ2 ,
|w| = a0 < |z|. Thus we have different expansions on the two curves. Indeed, for the
γ1 -integral, we must use

1 1 z −1 X z n
= 1− = , for |z| < |w| (IV.2.5)
w−z w w wn+1
n=0
while for the γ2 - integral we must use
X wn ∞
1
− = , for |w| < |z|. (IV.2.6)
w−z z n+1
n=0
By the same argument as in the proof of CIF, we can interchange the sum and integral
in each case, leaving us with

X −1
X
f (z) = Cn z n + Dn z n for z ∈ A0 (IV.2.7)
n=0 n=−∞
where Z Z
1 −n−1 1
Cn = f (z)z dz, Dn = f (z)z −n−1 dz, (IV.2.8)
2πi γ1 2πi γ2
We have now almost proved the theorem, for we have the expansion (IV.2.7) for z ∈ A0
and we have formulae for the coefficients. Note that in fact the coefficients given in
(IV.2.8) agree with the formula (IV.2.3) by Cauchy’s Theorem. Indeed f (z)z −n−1 is
holomorphic in A for each integer n. So we have
Z Z
f (z)z −n−1 dz = f (z)z −n−1 dz (IV.2.9)
|z|=r |z|=ρ
for any two radii a < r < ρ < b by Cauchy’s Theorem. Thus the coefficients in the
expansion of f , (IV.2.7) are independent of the choice of annulus A0 . Since any point z
of A is in some annulus A0 , the proof is finished. 
82 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

Of course, there is a version of this theorem for functions holomorphic in an annulus


centred at z0 rather than 0. In this case, the expansion is in powers (positive and
negative) of (z − z0 ). If the expansion (IV.2.2) holds, then by multiplying by z −n−1 and
integrating over |z| = r, we obtain the formula (IV.2.3) for cn . Thus the coefficients cn
are uniquely determined by f .
Remark IV.2.4. If f is holomorphic in 1 − δ < |z| < 1 + δ, for some positive δ,
then we may write F (θ) = f (eiθ ), so F is a complex-valued 2π-periodic function of θ.
Restricting the Laurent expansion of f to the unit circle |z| = 1, we obtain the (complex
form of the) Fourier expansion of F ,

X
F (θ) = cn einθ
n=−∞

where, taking r = 1 in the formula for the cn , we have the standard formula for the
Fourier coefficients
Z 2π
1
cn = e−inθ F (θ) dθ.
2π 0
Since the Laurent coefficients are unique, any method which yields a valid expansion
in positive and negative powers must give the Laurent expansion.
Example IV.2.5. The power series expansion of the function
z3
f (z) = sin z = z − + ···
3!
is also its Laurent series expansion. So the cn for n < 0, all vanish for this function.
For the function 1/(1 − z), we have c−1 = −1 and all other terms in the expansion
in powers of (z − 1) are zero.
For exp(1/z), we have the standard power series for the exponential function, which
implies

X 1 −n
exp(1/z) = z
n!
n=1
for all z 6= 0. Thus in the Laurent expansion about z = 0, we have c−n = 1/n! for all
n > 0, cn = 0 for all n > 0.
2.1. Classification of isolated singularities.
Definition IV.2.6. Let f be holomorphic in D∗ (z0 , δ) so that z0 is an isolated
singularity of f . Let the Laurent expansion of f in D∗ (z0 , δ) be

X
f (z) = cn (z − z0 )n
n=−∞

Then z0 is a removable singularity if cn = 0 for all n < 0; it is a pole of order m if c−m 6= 0


but c−n = 0 for all n > m; otherwise, it is an essential singularity (i.e. infinitely many
of the cn , for n < 0, are non-zero.)
2. POLES AND LAURENT SERIES 83

Definition IV.2.7. If z0 is an isolated singularity then the principal or singular part


of f at z0 is the part of the Laurent series involving the c−n for n > 0:

X
c−n (z − z0 )−n (IV.2.10)
n=1

The residue Resz0 (f ) at z0 is the coefficient c−1 .

Note that: the principal part is zero if the singularity is removable; it is a finite sum
of negative powers (z − z0 )−n if the singularity is a pole; and it is an infinite sum of
negative powers in the case of an essential singularity.

Proposition IV.2.8. If z0 is a removable singularity of f , then it can be removed


in the sense that there is a unique holomorphic function fe : Ω ∪ {z0 } → C with f = fe in
Ω.

Proof. There can be at most one holomorphic function with this property, for fe(z0 )
must be defined as limz→z0 f (z). To see that fe exists, note that by definition, the Laurent
expansion of f near z0 is:

X
f (z) = cn (z − z0 )n , 0 < |z − z0 | < δ.
n>0

Thus we define fe(z0 ) = c0 and then fe(z) is a convergent power series centred at z0 , and
this gives the needed holomorphic extension of f . 

Once we have found a removable singularity of a holomorphic function, it is usual


not to distinguish between it and its unique holomorphic extension. In other words, we
tend not to bother to distinguish between f and fe.
sin z
Example IV.2.9. f (z) = z has an isolated singularity at z = 0. However,

z3 z2
sin z = z − + O(z 5 ) so f (z) = 1 − + O(z 4 ).
3! 3!
This shows that z = 0 is a removable singularity if we define f (0) = 1. Having removed
the singularity, f is now holomorphic in the entire complex plane.

Definition IV.2.10. A function holomorphic in the whole of C is called entire.


A function with isolated singularities, all of which are poles, is called meromorphic.

Exercise IV.2.8. Suppose that z0 is an isolated singularity of f and that |f (z)| 6


M for all |z − z0 | < δ, where M and δ are some constants. (Informally, f is bounded
near z = z0 .) Show that z0 must be a removable singularity. [Hint: you will need to
use the formulae for the c−n to show that they are all zero.]
84 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

3. The Residue Theorem


We state the residue theorem for meromorphic functions.
Theorem IV.3.1. Let Ω be an open subset of C and let f be a meromorphic function
in Ω. Let D be a bounded domain with piecewise smooth boundary, D∪∂D ⊂ Ω. Suppose
that no pole of f lies on ∂D. Then the number of poles of f contained in D is finite. If
they are denoted z1 , . . . , zn ,
Z n
X
f (z) dz = 2πi Reszj (f ). (IV.3.1)
∂D j=1

Proof. Suppose, for a contradiction, that there were an infinite number {z1 , z2 , . . .}
of poles in D. Now D is closed and bounded, so {zj } has a limit point w, say, in D.
Then w cannot possibly be an isolated singularity, contradicting the assumption that f
is meromorphic.
Let the (distinct) poles be {z1 , . . . , zn }.
Pick δ > 0 so small that Dj = D(zj , δ) ⊂ D for each j, and so that these discs are
mutually disjoint. Let Dδ = D \ nj=1 Dj . Then by definition f is holomorphic inside
S
and on Dδ , so Cauchy’s Theorem gives:
Z Z n Z
X
f (z) dz = 0 and so f (z) dz = f (z) dz. (IV.3.2)
∂Dδ ∂D j=1 |z−zj |=δ

We claim that the integral around |z − zj | = δ is 2πi Reszj (f ).


Consider the case j = 1. Let the Laurent expansion of f near z1 be
X
cn (z − z1 )n . (IV.3.3)
n>−k

)n
P
Now the sum n>0 cn (z − zz converges to a holomorphic function in a neighbourhood
of z = z1 and so the integral around |z − z1 | = δ is zero. Thus the integral is equal to
the integral of the principal part:
−1 Z
X
cn (z − z1 )n dz. (IV.3.4)
n=−k |z−z1 |=δ

Now (z − z1 )n is the derivative of a holomorphic function unless n = −1. So by the


complex fundamental theorem of calculus, only the term c−1 (z − z1 )−1 contributes to
the integral around |z − z1 | = δ. By explicit computation, this contribution is 2πic−1 =
2πi Resz1 (f ) by definition. (The whole computation of (IV.3.4) can be done by explicit
computation, but it is neater to invoke the complex fundamental theorem of calculus.)
Repeating the argument at the other singular points gives the result.

This result has many theoretical applications (in the complex function theory) and
practical applications (evaluation of definite integrals and sums).
In the next section we shall give some typical applications of the residue theorem in
the evaluation of integrals.
The explicit calculation of residues is easy for simple poles, but harder in general:
3. THE RESIDUE THEOREM 85

Proposition IV.3.2. If f (z) = h(z)


g(z) where h and g are holomorphic near z0 , suppose
that
g(z0 ) = 0, g 0 (z0 ) 6= 0 and h(z0 ) 6= 0. (IV.3.5)
h(z0 )
Then f has a simple pole at z = z0 and Resz0 (f ) = g 0 (z0 ) .

Proof. Without loss of generality, suppose that z0 = 0. Near 0, the power-series


expansion of f and g take the form
h(z) = h0 + h1 z + O(z 2 ), g(z) = g1 z + O(z 2 )
where h0 = h(0) and g1 = g 0 (0). So
h(z) h0 + O(z) h0
f (z) = = = + O(1) for z → 0.
g(z) g1 z(1 + O(z)) g1 z
Thus the residue is equal to h0 /g1 = h(0)/g 0 (0). 
Remark IV.3.3. The extension to double-poles is given in Exercise IV.3.11. I do
not recommend that you memorize the result of that exercise!
The following result is also sometimes useful:
g(z)
Proposition IV.3.4. Suppose that f (z) = (z−z 0)
m where g is holomorphic near

z = z0 and m is a positive integer. Then f has a pole of order at most m at z0 and


g (m−1) (z0 )
Resz0 (f ) = . (IV.3.6)
(m − 1)!
Proof. Since it is holomorphic, g has a Taylor expansion centred at z = z0 , and
dividing this by (z − z0 )m we obtain the Laurent expansion
g(z0 ) g 0 (z0 ) g (m−1) (z0 ) 1 g (m) (z0 )
f (z) = + +· · ·+ + +O(z−z0 ). (IV.3.7)
(z − z0 )m (z − z0 )m−1 (m − 1)! z − z0 m!
This shows that the pole must be at most of order m (it will be strictly less if g(z0 ) = 0)
and the residue, the coefficient of (z − z0 )−1 , is as claimed. (Note the singularity is
removable (and so the residue is zero) if g has a zero of order > m at z = z0 . 
Remark IV.3.5. (or exercise). The previous result shows that the Cauchy Integral
Formula and the Formula for derivatives are special cases of the residue theorem. The
main difference is that the residue theorem explicitly applies to holomorphic functions
with more than one singularity inside the contour.

Exercise IV.3.9. Suppose that f is holomorphic with an isolated singularity at


z = z0 , which is a pole of order m. If σ(z) is the principal part of f at z0 , show that
f − σ has a removable singularity at z = z0 .
Exercise IV.3.10. Suppose that f is holomorphic with an isolated singularity
at z = z0 . Let σ(z) be the principal part of f at z0 . Show that σ is holomorphic in
C \ {z0 }. [Hint: the hard case is that z0 is an essential singularity. In this case, you
may find it helpful to reinterpret the problem in terms of the variable w = 1/(z − z0 )
86 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

and use the integral formulae for the c−n from Theorem IV.2.3 to get bounds on
these coefficients (analogous to the Cauchy Inequalities) to prove the result.]
Exercise IV.3.11. Let p(z) and q(z) be holomorphic in the disc D(z0 , δ), δ > 0.
Assume that
p(z0 ) 6= 0, q(z0 ) = q 0 (z0 ) = 0, q 00 (z0 ) 6= 0. (IV.3.8)
Show that f (z) = p(z)/q(z) has a double pole at z = z0 and that the residue is equal
to
p0 (z0 ) 2 p(z0 )q 000 (z0 )
Resz0 f = 2 00 − . (IV.3.9)
q (z0 ) 3 q 00 (z0 )2

Exercise IV.3.12. *For enthusiasts. But worth doing this problem for the
intuition you will get about the relation between zeros and poles and just how bad
essential singularities are.
Suppose that f (z) is a holomorphic function in a punctured disc D∗ (z0 , r) =
{z ∈ C : 0 < |z − z0 | < r}.
(a) Show that f (z) has a pole of order m > 0 at z = z0 if and only if there
exists a holomorphic function g : D(z0 , r) → C, with g(z0 ) 6= 0, and such
that
g(z)
f (z) = . (IV.3.10)
(z − z0 )m
(b) Show that f (z) has a pole of order m > 0 at z = z0 if and only if 1/f (z)
has a zero of order m at z = z0 .
(c) Use (b) to show that the isolated singularity of f at z0 is a pole if and only
if |f (z)| → ∞ as z → z0 .
(d) Prove the theorem of Casorati and Weierstrass: if z0 is an (isolated) es-
sential singularity of f , then given w ∈ C, there exists a sequence zn → 0
(depending upon w), such that f (zn ) → w as zn → ∞. Thus near an es-
sential singularity f comes arbitrarily close to every complex value. [Hint:
suppose that for some w, there is no such sequence. Write down carefully
what this means and then figure out what kind of isolated singularity the
function z 7→ (f (z) − w)−1 must have at z0 . If you succeed in this, you
should end up with a contradiction.]

4. Definite integrals
4.1. Rational functions.

Example IV.4.1. Let a be a positive real number. Calculate


Z ∞
dx
4 + x4
. (IV.4.1)
−∞ a
4. DEFINITE INTEGRALS 87

Solution: Let Γ = γ1 + γ2 be the semicircular contour where γ1 is the interval [−R, R]


of the real axis and γ2 is the part of |z| = R in the upper half plane. Let
1
f (z) = 4
z + a4
Then f is meromorphic in C, with simple poles at the four roots of z 4 + a4 = 0, namely
z1 = aeπi/4 , z2 = ae3πi/4 , z3 = ae5πi/4 , z4 = ae7πi/4 .

√ points can also be written (a/ 2)(±1 ± i) and are the four corners of a square of
These
side 2a. Only z1 and z2 are inside of Γ. Hence
Z
dz
4 4
= 2πi(Resz1 (f ) + Resz2 (f ))
Γ z +a
where z1 = zeπi/4 and z2 = ae3πi/4 . To calculate the residues, we can use Prop IV.3.2
with h(z) = 1 and g(z) = z 4 + a4 . Then
1 1 1
Resz1 (f ) = 3 = 3 3πi/4 = 3 e−3πi/4
4z1 4a e 4a
and
1 1 1
Resz2 (f ) = 3 = 3 9πi/4 = 3 e−πi/4
4z2 4a e 4a
Hence Z
2πi 2π π π
f (z) dz = 3 (e−3πi/4 + e−πi/4 ) = 3 cos = √ ,
Γ 4a 2a 4 2a3
using i = eπi/2 to simplify.
By the ML Lemma, and the fact that the rational function is O(|z|−2 ) for |z| → ∞,
the contribution from γ2 tends to zero as R → ∞. (In an exam context, you’d have to
write more than this for full credit. See Homework 4). Taking the limit R → ∞, we find
Z ∞
dx π
4 + x4
=√ . (IV.4.2)
−∞ a 2a3
Always good to do a sanity check: the integrand is positive, so probably the answer
should be positive. Check! The integrand is real, so probably the answer should be real.
Check!

The above method works for any rational function f (z) which is O(|z|−2 ) as
|z| → ∞ having no poles on the real axis. Integration around the above semicircular
contour gives
Z ∞
f (x) dx = 2πi × sum of residues of f in the upper half-plane.
−∞

Exercise IV.4.13. Calculate


Z ∞
(x2 + 3) dx
.
−∞ (x2 + 1)(x2 + 4)
88 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

Exercise IV.4.14. If a > 0, calculate


Z ∞
x2 dx
2 2 2
.
−∞ (x + a )
(This one is harder because you have to calculate the residue at a double-pole.)

Note that if the rational function is even, you get the integral from 0 to ∞ as well,
just by dividing the answer for the integral from −∞ to ∞ by 2.
4.2. Rational functions of cos and sin. Consider the integral
Z 2π

J=
0 a + b cos θ
where a > b > 0 (so that a + b cos θ is never zero). Such an integral can be computed
directly, but also using contour integral methods. I like to think of this as reverse en-
gineering. We interpret the integral as the parameterized version of an integral around
|z| = 1 of a holomorphic function. To do this, we set z = eiθ (the standard parameteri-
zation of the unit circle) and use the formulae
1 1
cos θ = (z + z −1 ), sin θ = (z − z −1 )
2 2i
dz
along with dθ = iz . In the case of the integral J, we obtain
Z Z Z
1 dz dz 2i dz
J= −1
= −2i 2
=−
|z|=1 a + b(z + z )/2 iz |z|=1 b + 2az + bz b |z|=1 1 + 2λz + z 2
where λ = a/b > 1 by assumption.
Let f (z) = (z 2 + 2λz + 1)−1 . Then f is holomorphic in C away from the two roots
z1 and z2 of the quadratic z 2 + λz + 1 = 0. Solving,
p p
z1 = −λ + λ2 − 1, z2 = −λ − λ2 − 1.

Since λ > 1, |z2 | = λ + λ2 − 1 > λ > 1 but z1 z2 = 1 so |z1 | < 1. Hence
2i
J =− × 2πi × Resz1 (f ).
b
We can use Proposition IV.3.2 to calculate this residue as the pole at z = z1 is simple.
We obtain
1 1
Resz1 (f ) = = √ .
z1 − z2 2 λ2 − 1
Hence

J=√ .
a2 − b2
Integrals from 0 to π of rational functions can sometimes be obtained from the full
integral from 0 to 2π by using the usual properties
sin(θ + π) = − sin θ, cos(θ + π) = − cos θ
and/or the even/odd properties of cos and sin.
4. DEFINITE INTEGRALS 89

Exercise IV.4.15. Let a be a real number, 0 < a < 1. By evaluating the integral
Z
1 1
dz
2πi |z|=1 (z − a)(z − a−1 )
prove that Z 2π
1 2π
dt = .
0 1 + a2 − 2a cos t 1 − a2
Exercise IV.4.16.
(a) By considering a suitable complex integral, show that if 0 < r < 1,
Z π
cos nθ dθ πrn
2
= .
0 1 − 2r cos θ + r 1 − r2
(b) By considering a suitable complex integral, show that
Z 2π
cos(cos θ) cosh(sin θ) dθ = 2π.
0
For (a), you will have to use symmetry to get an integral from [0, 2π]. There are a
couple of ways to do the reverse engineering, and if you choose the right one, you
will save yourself work. For (b) we do not have a rational function of sin and cos.
The reverse-engineering idea still works, though, but may require a bit of trial and
error.

4.3. Integrals of a rational times a trig function. A typical problem is to


calculate
Z ∞
cos ax dx
2 2
−∞ x + b

where a and b are real numbers. This one appears on workshop problem set 8 (Nov
27), so I will not discuss it in detail here. In that solution, we used the Cauchy Integral
Formula instead of the residue theorem.
The main thing to remember with these types of integrals is that the holomorphic
function to be used is not
cos az
(IV.4.3)
z 2 + b2
but
eiaz
(IV.4.4)
z 2 + b2
The reason is that whatever the sign of a, cos az blows up as Im(z) → ±∞. So if you try
to integrate (IV.4.3) around a semicircular contour, you will never get the contribution
from the curved part of the contour to go to zero as R → ∞. On the other hand, if
a > 0, eiaz is bounded in the upper half-plane and if a < 0, eiaz is bounded in the lower
half-plane. The function eiaz is also simpler than cos az or sin az!
Here is the general story of integrals of a trig function times a rational function:
90 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

If f (z) is a rational function with no pole on the real axis and satisfying |f (z)| =
O(|z|−2 ) for |z| → ∞, then if a is a positive real number,
Z ∞
f (x)eiax dx = 2πi × sum of residues of f (z)eiaz in the upper half-space.
−∞
(IV.4.5)
If f is real on the real axis then the integrals of f (x) cos ax and f (x) sin ax can be
deduced by taking real and imaginary parts.
Note that f need not be a rational function as long as it has no pole on the real
axis and satisfies the given decay condition.

Exercise IV.4.17. Show that if a is real and positive and f is a rational function
as in the box, then
Z ∞
f (x)e−iax dx = −2πi × sum of residues of f (z)e−iaz in the lower half-space.
−∞
(IV.4.6)
Exercise IV.4.18. Let a be a positive real number. Calculate
Z ∞ iξx
e dx
x2 + a2
−∞
making sure your answer is valid for all real values of ξ. (This is the Fourier transform
of the function (x2 + a2 )−1 .)

We consider two extensions of the above technology. The first concerns functions
which do have a singularity on the real axis.
Example IV.4.2. Find the value of
Z ∞
1 − cos x
I= dx. (IV.4.7)
0 x2

Solution: Let
1 − cos x
u(x) = .
x2
Because 1 − cos x = x2 /2 + O(x4 ) for small x,
1
u(x) = + O(x2 ) for small x
2
so if we define u(0) = 1/2, the integrand becomes a continuous function near 0. Thus
despite the apparent singularity at x = 0, the integral
Z R
u(x) dx
0
4. DEFINITE INTEGRALS 91

exists for any finite R. Technically, I is defined as the limit


Z R
1 − cos x
I = lim dx
R→∞ 0 x2
and since u(x) = O(x−2 ) for x → ∞, this is absolutely convergent. Having satisfied
ourselves that the integral exists, let’s evaluate it by contour integral methods. Introduce
1 − eiz
f (z) =
z2
so that Re(f ) = u on the real axis. Let Γδ be the standard semicircular contour with
diameter [−R, R], but indented as shown below near z = 0 to avoid having to integrate
through the singularity of f (z) at z = 0.

Γδ

Figure 1. Indented semicircular contour. The small semicircle has ra-


dius δ, the large one has radius R.

Denote by γ the small semicircular arc of radius of δ, traversed anticlockwise. Since


f is holomorphic inside and on Γ, Cauchy’s Theorem gives
Z −δ Z R Z
f (x) dx + f (x) dx = f (z) dz + O(1/R). (IV.4.8)
−R δ γ
Here we have used the ML lemma in the usual way to estimate the integral of f along
the large semi-circular arc: this gives the O(1/R) error term.
Now the imaginary part of f along the real axis is −x−2 sin x, which is odd. Therefore
from −R to −δ of this imaginary part cancels with the integral from δ to R. The real
part of f along the real axis is even and equal to u(x), so (IV.4.8) can be simplified to
Z R
1 − cos x
Z
2 dx = f (z) dz + O(1/R). (IV.4.9)
δ x2 γ
In order to calculate the integral along γ, note first that
1 − (1 + iz + O(z 2 )) i
f (z) = = − + φ(z), (IV.4.10)
z2 z
where φ(z) is holomorphic, hence O(1), near z = 0. This calculation can be thought of
as splitting f into its principal (or singular) part −i/z and a φ(z) which is holomorphic
near z = 0.
By direct calculation Z
i
− dz = π (IV.4.11)
γ z
92 IV. ZEROS, POLES, AND THE RESIDUE THEOREM
R
whereas γ φ(z) dz = O(δ) by the ML lemma.
Putting all this information together,
Z R
1 − cos x
2 dx = π + O(δ) + O(1/R). (IV.4.12)
δ x2
Taking the limit as δ → 0 and R → ∞, we obtain
Z ∞
1 − cos x π
2
dx = . (IV.4.13)
0 x 2

Notation IV.4.3. Here we wrote O(z 2 ) in (IV.4.10) for the sum of the higher
order terms in the expansion of eiz near z = 0. This is an extension of our previous
use of the ‘big Oh’ notation. It is convenient to allow ourselves to write O(z n ) to
mean a function of the form z n ψ(z), where ψ is holomorphic (hence bounded) in a
neighbourhood of z = 0.

Remark IV.4.4. The residue of f (z) in this example at z = 0 is −i by (IV.4.10).


Thus the integral around the semicircle γ picks up πi times the residue, exactly half the
answer you’d get from integrating all the way around the pole. This pleasingly simple
behaviour only works for simple poles.

The other direction in which we wish to extend (IV.4.5) is to weaken the decay
condition |f (z)| = O(|z|−2 ) for |z| → ∞. Let us note first that we can certainly relax
this condition to |f (z)| = O(|z|−a ) for a > 1 without any additional work. For if
|f (z)| 6 C|z|−a for all sufficiently large |z| (IV.4.14)
then the ML lemma gives that the integral of f along the large semicircular contour of
radius R is 6 2πR1−a and if a > 1 this still tends to 0 as R → ∞. This situation can
arise not for rational functions but for functions involving fractional powers
R ∞(or log).
−a
Note also that the condition |f (z)| = O(|z| ) with a > 1 also implies that 0 f (x) dx
is absolutely convergent.
The more delicate case is |f (z)| = O(|z|−1 ). So suppose that f (z) is a rational
function which is O(1/z) for large |z|. Even if f (z) has no poles on the real axis, it is
not clear that Z ∞ Z M
f (x)eix dx = lim f (x)eix dx (IV.4.15)
0 M →∞ 0
exists in this case. To see that such integrals do exist, let’s consider the simplest case,
f (x) = 1/x (but where we integrate from 1 to ∞ to avoid the singularity at x = 0—it is
the behaviour at ∞ that is the issue here).
Lemma IV.4.5. The integral
Z ∞ M
eix eix
Z
dx = lim dx (IV.4.16)
1 x M →∞ 1 x
is convergent.
4. DEFINITE INTEGRALS 93

Proof. Proving the result is equivalent to proving that


Z N ix
e
IM,N := dx → 0 as M, N → ∞. (IV.4.17)
M x
This can be established by integration by parts. Indeed, integrating eix , we obtain
 ix N Z N ix
e e
IM,N = −i 2
dx (IV.4.18)
ix M M x
Now Z N
1 1 dx 1 1 1 1 2
|IM,N | 6 + + 2
= + + − = .
N M M x N M M N M
Hence the result. 
It follows that if f (z) is rational and f (z) = O(1/z) for large |z|, then
Z M
lim f (x)eix dx exists (IV.4.19)
M →∞ 0

provided that f has no poles on the real axis. Indeed, if f is such a rational function,
write
f (z) = a/z + g(z) where |g(z)| = O(|z|−2 ) for large |z|.
Then Z M Z 1 Z M ix Z M
ix ix e
f (x)e dx = f (x)e dx + a dx + g(x)eix dx.
0 0 1 x 1
The integral from 0 to 1 definitely exists and is independent of M . If we let M tend to
∞, then the third integral on the RHS is absolutely convergent because g(x) = O(x−2 )
for large x. The middle integral on the RHS converges by the Lemma. So
Z ∞ Z M
f (x)eix dx = lim f (x)eix dx
0 M →∞ 0

also exists.

Exercise IV.4.19. Use the same argument as in the proof of Lemma IV.4.5 to
prove the following generalization: if f is continuously differentiable on [0, ∞) and
satisfies the two conditions
f (x) → 0, f 0 (x) = O(x−2 ) as x → ∞ (IV.4.20)
then Z M
lim f (x)eix dx
M →∞ 0
exists. Can you weaken the hypotheses on f still further?

Suppose (to be definite) that f (z) is a rational function such that |f (z)|R = O(|z|−1 ) as

|z| → ∞ and that f has no pole on the real axis. Then by the Proposition, −∞ f (x)eix dx
exists (as an improper integral) and it is reasonable to ask if integrating f (z)eiz around
the usual semicircular contour allows us to evaluate this integral. For this to work,
94 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

we need to show that the contribution over the large semicircular arc goes to zero as
R → ∞.
This is achieved by Jordan’s Lemma
Proposition IV.4.6. Suppose that f (z) is continuous in the upper half-plane and
that |f (z)| → 0 for all z in the upper half-plane as |z| → ∞. Let γR be the contour

γR (t) = Reit , 0 6 t 6 π.
and let a > 0. Then Z
lim f (z)eiz dz = 0.
R→∞ γR

Proof. It important to understand that the condition on f means, precisely: given


ε > 0, there exists M > 0 so that
|z| > M and Im z > 0 =⇒ |f (z)| < ε. (IV.4.21)
The argument is the same for all positive a, so let us simplify to a = 1. Using the
parameterization,
Z Z π
f (z)eiz dz = exp(iR cos t − R sin t)f (Reit )iReit dt.
γR 0

Given ε > 0, choose M as in (IV.4.21) and R > M . Then by the triangle inequality and
the given bound on |f |
Z Z π
iz
f (z)e dz 6 εR exp(−R sin t) dt (IV.4.22)
γR 0

We can’t do this integral explicitly, but note that


Z π Z π/2
exp(−R sin t) dt = 2 exp(−R sin t) dt
0 0

(because sin t = sin(π − t)) and that sin t > 2t/π for 0 6 t 6 π/2. (This is the chord
which connects (0, 0) to (π/2, 1), and if you draw the graph, you’ll see that it is entirely
below sin t on the interval [0, π/2]. Hence
Z π Z π/2
π
exp(−R sin t) dt 6 2 exp(−2Rt/π) dt = (1 − e−2R )
0 0 R
Combining this with (IV.4.22) the factors of R cancel out, leaving us with
Z
eiz f (z) dz 6 πε.
γR

if R > M . This proves the result. 

To sum up
4. DEFINITE INTEGRALS 95

Suppose that f (z) is a holomorphic function apart from a finite number of poles
in a neighbourhood of the closed upper half-plane and with no pole on the real axis.
Suppose that |f (z)| → 0 for z in the upper half plane as |z| → ∞. Then if a is a
positive real number,
Z ∞
f (x)eiax dx = 2πi × sum of residues of f (z)eiaz in the upper half-space.
−∞
(IV.4.23)
If f is real on the real axis then the integrals of f (x) cos ax and f (x) sin ax can be
deduced by taking real and imaginary parts.
The hypotheses apply in particular if f is a rational function p/q with deg q >
deg p.

R∞ R∞
Exercise IV.4.20. Calculate −∞ f (x) cos x dx and −∞ f (x) sin x dx for the fol-
lowing functions f :
1
(a) f (x) = x2 +2x+5 ;
1
(b) f (x) = (x2 +2x+5)2 ;
(c) f (x) = xx+1
2 +1 .

For which of these do you need to invoke Jordan’s Lemma?


Exercise IV.4.21. Show that
Z ∞
sin x π
dx = (IV.4.24)
0 x 2
by contour-integral methods. This example is in almost every textbook. It combines
the idea of having to avoid a singularity on the real axis that we saw in Exam-
ple IV.4.2 with the application of Jordan’s Lemma.

4.4. Integrals involving log x or complex powers. Now we turn to definite


integrals involving non-integral powers of z or log z.
Example IV.4.7. Evaluate ∞
xa dx
Z
. (IV.4.25)
0 1 + x2
Solution: We are going to integrate an appropriately chosen holomorphic function over
the indented semicircular contour. The most obvious holomorphic function is
za
f (z) = (IV.4.26)
1 + z2
but we need to make a choice of holomorphic branch of z a . The domain of definition
of z a must contain our contour (and its inside) so we cut the plane along the negative
imaginary axis and define
π 3π
log z = log |z| + i arg z, − < arg(z) < (IV.4.27)
2 2
96 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

and define
z a = exp(a log z) = |z|a × exp(ia arg z)
for the choice of log and arg given in (IV.4.27) that if z = x+i0 with x > 0, then z a = xa
is real and positive, but if x < 0, we have x = |x|eπi so xa = |x|a eπia . So defined, f (z)
is holomorphic inside and on the indented semicircular contour Γδ . See Figure 2.

arg(z) = π/2

Γδ

arg(z) = π arg(z) = 0
0

Figure 2. Indented semicircular contour and cut plane for (IV.4.26).


The small semicircle has radius δ, the large one has radius R. The cut
needed to define a holomorphic branch of z a is shown in red.

By the residue theorem,


Z
f (z) dz = 2πi Resz=i (f ) = πeπia/2 . (IV.4.28)
Γ
Moreover, the contribution from the large semicircle is O(Ra−1 ) so if a < 1, this contri-
bution goes to zero as R → ∞.
On the small semicircular indentation, |f (z)| = O(δ a ) and the length of this semicir-
cle is πδ. So the contribution from this small semicircle is O(δ a+1 ) and this also tends
to zero if a + 1 > 0.
To sum up,
Z −δ Z R
f (z) dz + f (z) dz = πeπia/2 + O(δ a+1 ) + O(Ra−1 ). (IV.4.29)
−R δ
To make the error terms go away as δ → 0 and R → ∞, assume from now on
− 1 < a < 1. (IV.4.30)
(These conditions are also needed for the convergence of the original integral.)
Substituting the value xa = |x|a eπia for x ∈ [−R, −δ] and taking the limit, we obtain
Z 0 Z ∞ a
πia |x|a dx x dx
e 2
+ = πeπia/2 (IV.4.31)
−∞ 1 + x 0 1 + x2
But the integral from −∞ to 0 is equal to the integral from 0 to ∞, (make the substitution
x0 = −x) so we obtain Z ∞ a
πia x dx
(1 + e ) 2
= πeπia/2 (IV.4.32)
0 1 + x
4. DEFINITE INTEGRALS 97

and so

xa dx πeπia/2
Z
π
2
= πia
= . (IV.4.33)
0 1+x e +1 2 cos(πa/2)
As a sanity check, the answer is positive for −1 < a < 1, which is good because the
function being integrated is everywhere positive. Also as a → ±1, the answer blows up,
which is consistent with the failure of convergence of the original integral if a 6 −1 or
a > 1.

Exercise IV.4.22. Use the same method to evaluate


Z ∞
log x dx
. (IV.4.34)
0 a2 + x2

In this example it was essential that the function multiplying xa was even. If this
had not been the case, we would not have been able to relate the contribution from the
interval [−R, −δ] to the contribution from [δ, R]. If we have to calculate
Z ∞
K= f (t)ta dt (IV.4.35)
0

where f is not necessarily even, then the substitution t = x2 gives


Z ∞
K=2 f (x2 )x2a+1 dx. (IV.4.36)
0
Now the function f is explicitly even and the previous method can be applied—provided
that f (t) is the restriction to the real axis of some suitable meromorphic function defined
in the upper half plane.
The simplest possible example is
Z ∞ −b
t
K= dt, 0 < b < 1. (IV.4.37)
0 1+t
The conditions on b are needed for the integral to converge. Applying the above trick,
Z ∞ 1−2b
x dx
K=2
0 1 + x2
which is the integral we’ve just been considering with a = 1 − 2b. Hence
π π
K= = .
cos(π/2 − πb) sin πb

Exercise IV.4.23. Evaluate


Z ∞ Z ∞ a
log t dt t dt
3
and
0 1 + t 0 1 + t3
using the methods just discussed. For what values of a does the second integral
converge?
98 IV. ZEROS, POLES, AND THE RESIDUE THEOREM

Remark IV.4.8. An alternative to the squaring trick is to integrate the original


version of the function around a so-called keyhole contour as shown:

Figure 3. Keyhole contour

To use this approach, z a or log z are defined with 0 < arg z < 2π, yielding a holo-
morphic branch that is holomorphic inside and on the contour.
As part of the limiting process, not only does the radius of the small circle go to zero
and that of the large circle to ∞, but also the angle between the two straight segments
and the positive x-axis has to be taken to zero as well.
On balance, the squaring trick has a lot to recommend it!
CHAPTER V

The argument principle

1. Counting zeros and poles


Suppose that f is meromorphic in an open subset Ω ⊂ C. Recall that this means
that f is holomorphic apart possibly from isolated singularities at points of Ω, and that
these isolated singularities are poles.
Suppose that D = D(z0 , r) is a disc whose closure is contained in Ω,
{z : |z − z0 | 6 r} ⊂ Ω.
Suppose further that no zero or pole of f lies on the boundary of D. Then there is a
finite number of zeros of f in D, z1 , z2 , . . . , zn , say, and similarly a finite number of poles
p1 , . . . , pm . Each pole has an order or multiplicity as does each zero. Define N to be the
total number of zeros, counted with multiplicity, and P to be the total number of poles,
counted with multiplicity. Explicitly, if the order of the zero zj is mj , then
N = m1 + · · · + mn
and similarly for P .
Then there is a remarkable formula for N − P as an integral:
Theorem V.1.1. (Argument Principle) With the above notation and definitions,
f 0 (z)
Z
1
N −P = dz. (V.1.1)
2πi |z−z0 |=r f (z)

Proof. This follows from the residue theorem. Let F (z) = f 0 (z)/f (z). Then F is
holomorphic apart from at the poles and zeros of f . If a is a zero of order m, then we
have seen that we can write
f (z) = (z − a)m g(z), g(a) 6= 0
where g(z) is holomorphic in a small disc centred at a. Then
f 0 (z) = m(z − a)m−1 g(z) + (z − a)m g 0 (z)
and so
m g 0 (z)
F (z) = +
z−a g(z)
Thus F has a simple pole with residue equal to m at z = a. Similarly if b is a pole of f
order k, then a precisely similar calculation shows that F has a simple pole at z = b of
order −k.
99
100 V. THE ARGUMENT PRINCIPLE

Applying the residue theorem


Z
1 X X X
F (z) dz = Resw (f ) = orders of zeros− orders of poles = N −P.
2πi |z−z0 |=r
w∈D

Remark V.1.2. We counted zeros and poles in a disc for simplicity. If D is a more
general bounded domain with piecewise smooth boundary such that D ∪ ∂D contained
in Ω (and no zero or pole of f on ∂D) the the analogous result is true: if N is the total
number of zeros in D counted with multiplicity and P is the total number of poles in D
counted with multiplicity, then
f 0 (z)
Z
1
N −P = dz.
2πi ∂D f (z)
Remark V.1.3. Note that on the RHS we have the integral of a complex-valued
function and so you might expect to get pretty much any complex number as an answer.
However, one of the remarkable features of this result is that the answer is always an
integer.

2. Winding number and argument principle


The integral in the argument principle,
Z 0
1 f (z)
I= dz (V.2.1)
2πi γ f (z)
where γ : [t0 , t1 ] → C is a smooth (or piecewise smooth) closed curve can be written in
the following form, using the chain rule. Let
Γ(t) = f (γ(t)).
Then Z t1 0
f (γ(t))γ 0 (t) dt
Z t1 0 Z
1 1 Γ (t) dt 1 dz
I= = = . (V.2.2)
2πi t0 f (γ(t)) 2πi t0 Γ(t) 2πi Γ z
Here the second equality is the chain rule.
We make a brief study of integrals of this form where Γ is any smooth or piecewise
smooth curve, not necessarily of the form f ◦ γ.
So let Γ : [t0 , t1 ] → C . Suppose further that a ∈ C does not lie on Γ (i.e. pedantically,
Γ(t) 6= a for all t0 6 t 6 t1 .)
Definition V.2.1. The winding number n(Γ, a) (or index of a with respect to Γ) is
defined to be the integral Z
1 dz
n(Γ, a) = . (V.2.3)
2πi Γ z − a
We shall see that this is always an integer, and in simple examples has the interpre-
tation of the number of times that Γ ‘winds around’ a. Of course ‘winds around’ has
not been defined here. The second interpretation of (V.2.3) is as 1/2π× the change in
argument of z − a as z runs around Γ. (This is where the term ‘argument principle’
comes from.
2. WINDING NUMBER AND ARGUMENT PRINCIPLE 101

Example V.2.2. Let a = 0, let γ(t) = eit , 0 6 t 6 2π, and let f (z) = z n . Then
Γ(t) = eint . This has the interpretation of the unit circle, traversed n times anticlockwise
if n > 0 and clockwise if n < 0. So the total change of argument ∆Γ arg(z) = 2πn in
this case, which is 2π times the number of times Γ(t) winds around 0.
Proposition V.2.3. With the above definition, n(Γ, a) is an integer.
Proof. For simplicity, suppose that Γ is smooth.
Define Z
dz
L(t) = (V.2.4)
Γt z
where Γt is the restriction of Γ to the interval [t0 , t]. Then
Z t 0
Γ (s) ds
L(t) = (V.2.5)
t0 Γ(s) − a

and L(t1 ) = 2πin(Γ, a). By the fundamental theorem of calculus,


Γ0 (t)
L0 (t) = . (V.2.6)
Γ(t) − a
and so
d
(exp(−L(t))(Γ(t) − a)) = exp(−L(t)) −L0 (t)(Γ(t) − a) + Γ0 (t) = 0.

(V.2.7)
dt
It follows that
exp(L(t)) = C(Γ(t) − a) (V.2.8)
for a constant C, but since L(t0 ) = 0 we must have
1
C=
Γ(t0 ) − a
and so
Γ(t) − a
exp(L(t)) = . (V.2.9)
Γ(t0 ) − a
Evaluating at t = t1 we have
Γ(t1 ) − a
exp(L(t1 )) = =1 (V.2.10)
Γ(t0 ) − a
since Γ(t1 ) = Γ(t0 ). It follows that L(t1 ) is an integer multiple of 2πi and the proposition
is proved. 
From the formula, we have defined a continuous function L(t) such that
Γ(t) − a
exp(L(t)) = . (V.2.11)
Γ(t0 ) − a
Thus for each t, L(t) is thus a continuous choice of
log(Γ(t) − a) − log(Γ(t0 ) − a).
(Other continuous choices of log(Γ(t) − a) differ from this one by adding on 2πin, where
n ∈ Z.)
102 V. THE ARGUMENT PRINCIPLE

Recalling that log z = log |z| + i arg z, it follows that t 7→ Im L(t) is a continuous
choice of arg(Γ(t) − a) − arg(Γ(t0 ) − a) and so
L(t1 ) = i × total change in the argument of z − a as z goes around Γ. (V.2.12)
So
Z
1 dz
= ∆Γ arg(z − a) = total change in arg(z − a) along Γ. (V.2.13)
i Γ z−a

Notation V.2.4. For any (piecewise) smooth closed curve Γ, denote by


∆Γ arg(z − a) the total change in (a continuous choice of) arg(z − a) as z goes
once around Γ.

If now Γ happens to have the form f ◦ γ, and a = 0, we can interpret


Z 0
1 f (z)
dz (V.2.14)
2πi γ f (z)
equally as the winding number of the composite curve f ◦ γ around 0, as 1/2π times
1
∆Γ arg(z) along this curve, or as 2π ∆γ (arg(f )) according to taste.
In particular, we have the following restatement of the argument principle (Theo-
rem V.1.1) whicht give the Theorem its name:

Let D = {z : |z − z0 | < R} be a disc so that D ∪ ∂D is contained in the domain


Ω of a meromorphic function. Suppose no zero or pole of f lies on ∂D. Then if N
is the number of zeros of f in D and P is the number of poles of f in D, we have
1 1
N −P = ∆Γ arg(z) = ∆γ (f (z)) (V.2.15)
2π 2π
where γ(t) = z0 + Reit , 0 6 t 6 2π is ∂D traversed once anticlockwise, and Γ(t) =
f ◦ γ(t). Here zeros and poles are counted, as always, with their multiplicity.
The analogous result remains true if D is replaced by any bounded domain with
piecewise smooth boundary such that D ∪ ∂D ⊂ Ω.

2.1. The Fundamental Theorem of Algebra revisited. We apply (V.2.15) to


give a second proof of the FTA. This proof is more conceptual than the one via Liouville’s
Theorem (in my opinion).
Let
p(z) = z n + an−1 z n−1 + · · · + a0 = z n + g(z) (V.2.16)
where g(z) is the sum of the ‘lower-order terms’. We use the argument principle (V.2.15)
to show that p has n zeros in D(0, R) if R is chosen large enough. The boundary ∂D is
our friend |z| = R, parameterized by as γ(t) = Reit , 0 6 t 6 2π. By Example V.2.2, if
f (z) = z n , then n(f ◦ γ, 0) = n. We argue that if R is large enough, then adding
on g cannot change the winding number!
This is illustrated in the geogebra for the example
p(z) = z 4 + 3iz + 2 − 2i, (V.2.17)
2. WINDING NUMBER AND ARGUMENT PRINCIPLE 103

see https://www.geogebra.org/m/XFEFFdYz. Taking R = 2, we see explicitly that


n(p ◦ γ, 0) = n(f ◦ γ, 0) = 4.
The example illustrates the general idea: we view p(z) as a small perturbation of
f (z), for |z| = R. This means that p(γ(t)) is much closer to f (γ(t)) than it is to 0—
and it follows that both must wind the same number of times around 0. (Think of an
astronomical analogy: as the Earth goes once around the Sun, the Moon also goes once
around the Sun. The Moon can do exactly what it wants but it is so much closer to the
Earth than it is to the Sun, it cannot help but go once around.)
Let us define
Γ(t) = R−n f (γ(t)) = eint and ∆(t) = R−n p(γ(t)) (V.2.18)
so that
n−1
X
|Γ(t) − ∆(t)| = | aj Rj−n eijt | 6 C/R (V.2.19)
j=0
for some constant C once R > R0 . (We have seen this type of estimate already many
times before in this course.) The rescaling of the curves by R−n does not affect the
winding number so
n(f ◦ γ, 0) = n(Γ, 0) = n (V.2.20)
and
n(p ◦ γ, 0) = n(∆, 0) = number of zeros of p in D(0, R). (V.2.21)
By construction |Γ(t)| = 1 so by taking R to be large, we can make ∆(t) as close to Γ(t)
as we please, uniformly for t ∈ [0, 2π]. Thus Γ(t) is the Earth going n times around the
Sun and ∆(t) is the Earth’s moon: it must also go n times around the Sun. It follows
that
n(∆, 0) = n (V.2.22)
and p(z) has n zeros in D(0, R). This completes our proof of the Fundamental Theorem
of Algebra.

Figure 1. Proof of the fundamental theorem of algebra from the argu-


ment principle. The arrow points to Γ(t), a point of the unit circle. ∆(t)
is somewhere in the grey disc. So if Γ(t) goes n times around the origin,
then so does ∆(t).
104 V. THE ARGUMENT PRINCIPLE

Remark V.2.5. Note that this proof is ‘quantitative’: we learn not only that p(z)
has n zeros in C, but that they are all in D(0, R), where R is any radius which makes
the above argument work. In fact, Rouché’s Theorem, proved in the next section, makes
this even more concrete.

3. Rouché’s Theorem
The essential idea of the proof we have just given of the FTA is generalized and
made more precise in Rouché’s Theorem:
Theorem V.3.1. Suppose that Ω ⊂ C is an open set, that f : Ω → C is holomorphic
and that D = D(z0 , R) is a disc such that D ∪ ∂D ⊂ Ω, as before. Suppose further that
g is holomorphic in Ω and
|f (z)| > |g(z)| for all z ∈ ∂D. (V.3.1)
Then f and f + g have the same number of zeros (counted with multiplicity) in D.
We give first a geometric proof:
Proof. Let h(z) = f (z) + g(z). Note first that the hypothesis (V.3.1) implies that
f has no zeros on ∂D (because of the inequality being strict). It also implies that h has
no zero, for
|h(z)| = |f (z) + g(z)| > |f (z)| − |g(z)| > 0
also by (V.3.1).
Let us write h = f + g and let γ : [a, b] → C be any piecewise smooth closed curve.
Let
Γ(t) = f (γ(t)), ∆(t) = h(γ(t)). (V.3.2)
If we assume that |f | > |g| on γ, then we have
|Γ(t) − ∆(t)| < |Γ(t)| (V.3.3)
for t ∈ [a, b]. From Figure 2, this means that ∆(t) is in the open disc centred at Γ(t),
with radius |Γ(t)| and so the angle at 0 between Γ(t) and ∆(t) lies in (−π/2, π/2).

Γ(t)

∆(t)

Figure 2. Proof of Rouché’s Theorem: if |Γ(t)−∆(t)| < |Γ(t)|, the angle


at 0 between Γ(t) and ∆(t) lies in (−π/2, π/2).
3. ROUCHÉ’S THEOREM 105

Using this geometric fact, we can prove that


n(Γ, 0) = n(∆, 0). (V.3.4)
If now γ is the boundary of D, it follows from (V.2.15) that f and f + g have the same
number of zeros in D.
So why is (V.3.4) true? To give a precise argument, suppose that A(t) is a continuous
choice of arg Γ(t), for t ∈ [a, b]. (A(t) can be defined by integration as in the proof of
proposition V.2.3.) From the geometry of Figure 2, it follows that there is a unique
choice B(t) of arg ∆(t) which satisfies
|A(t) − B(t)| < π/2 (V.3.5)
for all t ∈ [a, b]. Such a choice is automatically continuous1 (since A(t) is), and so we
have
2πn(Γ, 0) = B(b) − B(a), 2πn(∆, 0) = A(b) − A(a). (V.3.6)
But then
2π |n(Γ, 0) − 2πn(∆, 0)| = |B(b) − B(a) − A(b) + A(a)|
6 |B(b) − A(b)| + |B(a) − A(a)|
< π (V.3.7)
by (V.3.5). On the other hand the LHS is 2π× an integer, so that integer must be zero
and the result follows. 
Example V.3.2. How many zeros has p(z) = z 7 − 4z 3 + z − 1 inside |z| = 1?
We look for the biggest term (term with largest modulus) in the sum and it is −4z 3 .
So if we put f (z) = −4z 3 and g(z) = z 7 + z − 1, can we apply Rouché? Yes, because
|g(z)| 6 1 + 1 + 1 = 3, |f (z)| = 4
if |z| = 1. Thus our polynomial has the same number of zeros as −4z 3 inside |z| = 1,
namely 3—remember to count with multiplicity!
See https://www.geogebra.org/m/ySk796fV for Geogebra illustration.
Example V.3.3. How many further zeros of this polynomial lie in |z| = 2? On this
circle, the first term is dominant:
|z 7 | = 27 = 128, |4z 3 | = 32, |z| = 2,
so if we now take f (z) = z 7 , g(z) = −4z 3 + z − 1, we can apply Rouché to conclude that
p(z) has the same number of zeros as z 7 inside |z| = 2, that is, 7. So the other 4 zeros
of p lie in the annulus {1 < |z| < 2}.
Remark V.3.4. We could also have counted the zeros inside |z| = 1 by m taking
f (z) = −4z 3 + 1 and g(z) = z 7 + z, remarkably getting the same answer!
Remark V.3.5. Rouché gives a sharpening of the above proof of the FTA. Indeed
if p and f are as in §2.1, p has n zeros in D(0, R) if |z|n > |g(z)| for all |z| = R, where g
is the sum of the lower order terms in p(z) as in (V.2.16).
1If you are not satisfied with this claim, see the analytic proof in §3.1
106 V. THE ARGUMENT PRINCIPLE

3.1. Analytic proof of Rouché. Let N be the total number of zeros of f and let
N 0 be the total number of zeros of f + g in D.
By the argument principle, Theorem V.1.1,
f 0 (z) f 0 (z) + g 0 (z)
Z Z
1 0 1
N= dz, N = dz. (V.3.8)
2πi ∂D f (z) 2πi ∂D f (z) + g(z)
Now
f 0 + g0 f 0 f g 0 − gf 0 d
− = = log(1 + g/f ). (V.3.9)
f +g f f (f + g) dz
The point is that the hypothesis of Rouché means that Re(1 + g(z)/f (z)) > 0 for z ∈ ∂D
and so log(1+g/f ) can be defined uniquely by taking the argument to be in (−π/2, π/2).
By the complex FTC, Z
d
log(1 + g/f ) dz = 0 (V.3.10)
∂D dz
and this completes our second proof of Rouché.

Exercise V.3.1. Let C be the unit circle |z| = 1 traversed anticlockwise. De-
termine the variation of the argument ∆C arg f (z) for the functions
z3 + 2
(a) f (z) = z 2 , (b) f (z) =.
z
Exercise V.3.2. Let λ be real and λ > 1, Show that the equation
zeλ−z = 1
has exactly one solution in the disc |z| = 1, which is real and positive.
Exercise V.3.3. (a) How many roots does the polynomial f (z) = z 4 −2z−2
have inside the annulus
1 3
< |z| < ?
2 2
Explain your answer.
(b) Find the number of the roots of the equation
z 6 − 5z 4 + 8z − 1 = 0
in the annulus {z : 1 < |z| < 2}.
(c) How many roots does the polynomial z 6 −5z 4 +8z −1 have inside the square
with vertices ±1/2 ± i1/2?
Exercise V.3.4. Suppose that p(z) = a0 + a1 z + · · · + an−1 z n−1 + z n is a
polynomial with |aj | 6 1 for all j. Show that all zeros of p(z) lie in D(0, 2).
CHAPTER VI

Green’s Theorem

As a historical note, the story of George Green (1793–1841) is a pretty remarkable


one. His career was far from traditional and his achievements are stunning given that
he was almost entirely self-taught and left school when he was 9. He was a baker’s
son, lived in Nottingham for most of his life, but somehow gained access to some of the
most important works on electromagnetism and what we now call potential theory by
for example Coulomb, Laplace and Poisson. He wrote an essay1 in 1821, in which he
developed many remarkable new ideas, including what we now call Stokes’s Theorem and
Green’s functions. He published this essay himself (rather than in a scientific journal),
and it took a while for his ideas to gain the prominence they deserved. However, a fellow
subscriber to the Nottingham library, Sir Edward Bromhead, recognized his genius and
encouraged him to pursue his education. Green enrolled in Cambridge when he was
nearly 40 and excelled there as an undergraduate. After graduating, he pursued his
research until his early death in 1841.
In order for Green’s Theorem in the plane (Theorem III.2.7) to be meaningful, we
need a definition of the LHS, that is to say,
Z
f (x, y) dxdy, (VI.0.1)
D
where f is a continuous function of (x, y) and D is a bounded domain with piecewise
smooth boundary ∂D. To do this ‘properly’ is beyond the scope of this course (see
Measure Theory (Math 3101).) The main complication is dealing with the fact that D
can have a pretty complicated ‘shape’. We build up to the general case by splitting D into
a union of simpler shapes, define (VI.0.1) for these shapes and verify Green’s Theorem
in the plane for these shapes as we go. The general case then follows by adding up the
answers for all the simple shapes and verifying that the sum gives Green’s Theorem in
the plane for the original domain D.

1. Green’s Theorem for a rectangle


Suppose that f is continuous on the rectangle
R = {z = x + iy : a 6 x 6 b, c 6 y 6 d}. (VI.1.1)
We mean continuous in the 2D sense: for every z0 ∈ R, given ε > 0, there exists δ > 0
such that
z ∈ R ∩ D(z0 , δ) ⇒ |f (z) − f (z0 )| < ε. (VI.1.2)

1An essay on the application of mathematical analysis to the theories of electricity and magnetism

107
108 VI. GREEN’S THEOREM

As in the one-variable case, R being compact (closed and bounded) implies that f is
actually uniformly continuous in R, so given ε > 0, the same δ > 0 will guaranteee
(VI.1.2) for all z0 ∈ R. R
It is pretty plausible that one can make a definition of R f (x, y) dxdy in terms of
2D partitions of R: that is, we cover R by a grid of tiny rectangles of the form
Rjk = {sj < x < sj+1 , tk < y < tk+1 }
where
a = s0 < · · · < sN = b, c = t0 < · · · < tM = d.
and try to define the integral as the limit of sums of the form
X
fjk (sj+1 − sj )(tk+1 − tk ) (VI.1.3)
j,k

where we take
fjk = sup f or fjk = inf f
Rjk Rjk

for the ‘upper’ and ‘lower’ sums respectively. It is not hard to adapt the one-variable
proof to prove that if f is continuous then the limit over finer and finer partitions of R
of both upper and
R lower sums exist and that they are equal. Thus such a limit serves as
a definition of R f (x, y) dxdy.
From this definition one can prove various properties, in particular that the integral
over R of f can be computed by integration first with respect to x and then with respect
to y.
Proposition VI.1.1. If f is continuous on the closed rectangle R of (VI.1.1), then
Z d
F (x) := f (x, y) dy (VI.1.4)
c
is continuous for x ∈ [a, b] and
Z b
G(y) := f (x, y) dx (VI.1.5)
a
is continuous for y ∈ [c, d]. Moreover,
Z Z b Z d
f (x, y) dxdy = F (x) dx = G(y) dy. (VI.1.6)
R a c
One can also prove versions of the fundamental theorem of calculus, such as
Theorem VI.1.2. Let f (x, y) be continuous in R. Then if we define
Z x
F (x, y) = f (t, y) dt,
a
the partial derivative Fx exists in R and is equal to f . Conversely, if g is continuous
and gx exists and is also continuous, then
Z b
g(b, y) − g(a, y) = gx (x, y) dx.
a
There are analogous statements with the roles of x and y switched.
2. GREEN’S THEOREM FOR A REGION BOUNDED BY A GRAPH 109

With all this in hand, we can prove Green’s Theorem for a rectangle.
Theorem VI.1.3. Let R be our closed rectangle (VI.1.1) and let P and Q be two C 1
functions in R. Then
Z Z
(Qx − Py ) dxdy = P dx + Q dy. (VI.1.7)
R ∂R
Here, as always, ∂R is traversed anticlockwise.
Proof. It is clearly enough to prove the separate identities
Z Z Z Z
Qx dxdy = Q dy, Py dxdy = − P dx. (VI.1.8)
R ∂R R ∂R
Consider the first of these. By the results above, we can calculate the LHS by doing the
integral with respect to x first, then the integral with respect to y. For the first integral,
we can use the FTC, giving
Z Z d
Qx dxdy = (Q(b, y) − Q(a, y)) dy (VI.1.9)
R y=c
On the other hand, the RHS of the first of (VI.1.8) is the sum of the integrals along the
line segments from (a, c) to (b, c), from (b, c) to (b, d), from (b, d) to (a, d) and finally
from (a, d) to (a, c). Only the contibutions along the vertical segments are non-zero
for the integral of Qdy (for dy = 0 along the horizontal segments) and the sum of the
integrals along the segments is precisely the RHS of (VI.1.9). This proves the first of
(VI.1.8). The second is proved in precisely the same way, switching the roles of x and y
(and in particular computing the LHS by integrating first with respect to y). 

2. Green’s Theorem for a region bounded by a graph


To make the transition from rectangles to more complicated domains D, we consider
regions of the following kind.
Definition VI.2.1. We say that D is a region bounded by a graph y = φ(x), if φ is
a smooth function of x ∈ [a, b] and
D = {z = x + iy : a 6 x 6 b, c 6 y 6 φ(x)}. (VI.2.1)
It is part of the definition that φ(x) > c for a 6 x 6 b (and is positive for at least one
point, so that D is the closure of an open set). See Figure 1.

(b, φ(b))
(a, φ(a))
D

(a, c) (b, c)

Figure 1. A region bounded by the graph y = φ(x).

If φ(a) > 0, our region D is bounded on the left-hand side by the segment from (a, c)
to (a, φ(a)). If φ(b) > 0, then it is bounded on the right-hand side by the segment from
110 VI. GREEN’S THEOREM

(b, c) to (b, φ(b)). Either or both of these segments may collapse to points if φ(a) = 0 or
φ(b) = 0. So D is bounded by one, two, or three straight-line segments and the graph
of y = φ(x). R
One can prove that there is a good definition (via partitions) of D f (x, y) dx for
such domains D, and moreover,
Z Z b Z φ(x) !
f (x, y) dx = f (x, y) dy dx. (VI.2.2)
D x=a y=c

Green’s Theorem in the plane holds for regions bounded by a graph:


Proposition VI.2.2. Let D be the region defined in (VI.2.1) and let P and Q be
two C 1 functions in D. Then
Z Z
(Qx − Py ) dxdy = P dx + Q dy. (VI.2.3)
D ∂D

Proof. Consider terms in P and Q separately. The P -term is the easier. Starting
from (VI.2.2), putting f = Py , we have
Z Z b
− Py dxdy = (−P (x, φ(x)) dx + P (x, c)) dx. (VI.2.4)
D x=a
Now we turn to the integral over ∂D of P dx. The integrals over the vertical sides (if
they are present) of ∂D vanish because dx = 0 along any vertical line. Thus the integral
over ∂D is reduced to the integral along the segment from (a, c) to (b, c), together with
the integral along the upper, curved part of the boundary. The integral along the bottom
segment from (a, c) to (b, c) is
Z b
P (x, c) dx. (VI.2.5)
a
The integral along the boundary y = φ(x) is computed via the parameterization
γ(x) = (x, φ(x)) for a 6 x 6 b (VI.2.6)
and changing the sign, since this curve is traversed from right to left! Thus the contri-
bution from this curve is Z b
− P (x, φ(x))dx. (VI.2.7)
a
Combining (VI.2.4), (VI.2.5) and(VI.2.7), we obtain the Green’s theorem for D in the
case that Q = 0.
Now let’s do the case that P = 0. It is convenient to introduce a function Q(x,
e y)
whose derivative with respect to y is equal to Q. By the FTC, we may take
Z y
Q(x, y) =
e Q(x, t) dt. (VI.2.8)
c

Then Q
e y = Q, Q
e xy = Qx and we have
Z Z Z b
Qx dxdy = Qxy dxdy =
e e x (x, φ(x)) − Q(x,
(Q e c)) dx (VI.2.9)
D D x=a
2. GREEN’S THEOREM FOR A REGION BOUNDED BY A GRAPH 111

doing the integral with respect to y first. Now by the FTC yet again, we have
Z b
− Q(x, e c) − Q(b,
e c) dx = Q(a, e c). (VI.2.10)
a

The other term, Qe x (x, φ(x)) is trickier. We’d like it to be the derivative with respect to
x of something, so that we can use the FTC to simplify the integral. Now
d e e y (x, φ(x))φ0 (x) = Qe x (x, φ(x)) + Q(x, φ(x))φ0 (x).
Q(x, φ(x)) = Q e x (x, φ(x)) + Q
dx
(VI.2.11)
Hence
Z b Z b
Qx (x, φ(x)) dx = Q(b, φ(b)) − Q(a, φ(a)) −
e e e Q(x, φ(x))φ0 (x) dx. (VI.2.12)
a a

Combining, we obtain
Z
e c) − Q(a,
Qx (x, y) dxdy = Q(a, e φ(a)) (VI.2.13)
D
e φ(b)) − Q(a,
+ Q(b, e c) (VI.2.14)
Z b
− Q(x, φ(x))φ0 (x) dx. (VI.2.15)
a

By the FTC, (VI.2.13) is equal to the integral of Q dy along the left-hand edge, the
segment joining (a, φ(a)) to (a, c). Similarly, (VI.2.14) is the integral of Q dy along the
right-hand vertical segment and (VI.2.15) is the integral of Q dy along the edge y = φ(x),
traversed from right to left. Since the integral of Q dy along the lower horizontal edge
is zero, (VI.2.13–VI.2.15) are equivalent to
Z Z
Qx (x, y) dxdy = Q(x, y) dy.
D ∂D

This completes the proof. 

A precisely similar result is true for regions bounded below by a graph, that is

D = D = {z = x + iy : a 6 x 6 b, φ(x) 6 y 6 c} (VI.2.16)

(where φ(x) 6 c for x ∈ (a, b)) and for regions bounded on the right or left by a graph:

D = {z = x + iy : a 6 x 6 ψ(y), c 6 y 6 d} or (VI.2.17)
D = {z = x + iy : ψ(y) 6 x 6 b, c 6 y 6 d} (VI.2.18)

where ψ is in each case a suitable smooth function of y.


The idea now is that if D is a bounded domain with smooth boundary, we can draw
a grid of tiny squares Q with the property that every Q is either contained in D or Q ∩ D
is one of the four types of region bounded by a graph.
112 VI. GREEN’S THEOREM

3. Green’s Theorem for bounded domains with smooth boundary


Theorem VI.3.1. Let D be a bounded domain with smooth boundary ∂D (which may
consist of several disjoint path-connected components). If P and Q are two C 1 functions
defined on D ∪ ∂D, then Green’s Theorem holds:
Z Z
(Qx − Py ) dxdy = P dx + Q dy. (VI.3.1)
D ∂D
Proof. Consider Figure 2.

Figure 2. Two abutting domains

It shows two regions with a common vertical edge. Let the one on the left be D1 ,
the one on the right D2 . Each of D1 and D2 is a region bounded by a graph, so Green’s
Theorem holds for them. Write
∂D1 = γ + ∂D10 , ∂D2 = −γ + ∂D20 .
Here γ is the vertical segment, traversed upwards; ∂D10 and ∂D20 denote the ‘rest’ of the
boundaries respectively of D1 and D2 .. Let D = D1 ∪ D2 . Then clearly
Z Z Z
(Qx − Py ) dxdy = (Qx − Py ) dxdy + (Qx − Py ) dxdy.
D D1 D2
Furthermore,
Z Z Z
P dx + Q dy = P dx + Q dy + P dx + Q dy
∂D1 ∂D10 γ

and Z Z Z
P dx + Q dy = P dx + Q dy − P dx + Q dy
∂D2 ∂D20 γ
Thus Z Z Z
P dx + Q dy = P dx + Q dy + P dx + Q dy.
∂(D1 ∪D2 ) ∂D1 ∂D2
It follows that Green’s Theorem holds for the larger domain D = D1 ∪D2 . This argument
is simple, but the key point is that the common part of the boundary is traversed once
in each direction, and the contributions to the boundary integral cancel.
Now cover the entire complex plane by a grid of squares with side δ > 0 and sides
parallel to the x- and y-axes. (The grid can be defined by the vertical lines x = x0 + mδ
and the horizontal lines y = y0 + nδ, where (x0 , y0 ) is some given point and m and n
run over the integers.) See Figure 3.
The squares are supposed to be so small that for any square S, either the interior
of S does not meet the boundary at all, or it does so in a simple arc whose tangent
doesn’t vary too much over its intersection with S. Examples of the types of possible
intersections of D with a square are shown in Figure 4.
3. GREEN’S THEOREM FOR BOUNDED DOMAINS WITH SMOOTH BOUNDARY 113

Figure 3. Region D (light shading) with a superimposed grid of small


squares. The squares entirely inside D are shaded darker.

Figure 4. Examples of allowable intersections of D with squares from


the grid. In the first example, we have a region bounded above by a
graph. In the third, we have a region bounded to the right by a graph.
In the fourth and fifth, we have a region bounded to the left by a graph.
The second is bounded above by a graph, but not the graph of a smooth
function. The proof of Proposition VI.2.2 is easily adapted for this kind
of region, though.

A rigorous proof that this is possible would use fact that the tangent vector is
continuously varying (and never zero). Combined with the fact that the boundary is
compact, this means that we can choose the squares so small that the tangent vector
varies only by, say, 10◦ over the square. If the square meets the boundary in two or more
pieces, it is pretty plausible that by repeated subdivision, we can reduce to the case that
the boundary meets each square in a single arc. A key point is the compactness of the
boundary, which implies that it meets at most finitely many squares.
Granted that this is possible, in Figure 3, Green’s Green’s Theorem is true for each
of the pale grey squares which meet ∂D and for all the dark grey squares which are
interior to D. Thus D is a finite union of abutting regions, each of which is either a
square or a region bounded by a graph. Since Green’s Theorem is true for such regions,
and we have seen that if it is true separately on abutting regions, it is also true for the
union, it follows that it is true for D itself. 
114 VI. GREEN’S THEOREM

4. Piecewise smooth boundaries

Figure 5. Corners

Essentially the same argument can be used to prove Green’s Theorem for bounded
domains with piecewise smooth boundary as well.
Suppose that the non-smooth points of ∂D are z1 , . . . , zN . Cover D with a grid of
rectangles as before, by drawing a large number of horizontal and vertical lines, including
the lines passing through the zj ,
Re(z) = Re(zj ), Im(z) = Im(zj ).
Once again we make this mesh so small that when the interior of a rectangle meets
the boundary, it does so as the graph of a function. Since we have arranged for each
corner to be on the intersection of two grid-lines, near each corner we shall see figures
like the ones shown in Figure 5. In the example on the left in Figure 5, there are two
squares, each of which contains one boundary arc. And we already know that Green’s
Theorem applies to such domains. Thus if this is the only thing that ever happens with
the corners, the previous argument goes through as before. Unfortunately, the picture
on the right can also happen, where the two boundary arcs lie in a single rectangle. We
do not yet know Green’s Theorem for regions with the shape of the one on the right.
However, denote by S the top-right square and by W1 and W2 the two white parts
of the square. Denote by G the region shaded grey. Then
S = G ∪ W1 ∪ W2 and ∂G = ∂S − ∂W1 − ∂W2 . (VI.4.1)
And the point is that we know Green’s Theorem for S and for W1 and W2 because W1
and W2 are regions bounded by a graph. So if P and Q are defined over the whole of
S, we obtain Green’s Theorem for G from Green’s Theorem for G, W1 and W2 by use
of (VI.4.1) and subtraction! In this argument, we needed to assume that P and Q are
defined and C 1 over the whole square. This is OK, because the assumption about P
and Q is that they are defined and C 1 in a slightly larger open set Ω containing D and
its boundary. Thus we can suppose that the squares in the above pictures are so small
that the closure of the top right square is contained in Ω, so that P and Q are defined
there.
Thus the idea of covering a domain with piecewise smooth boundary by very small
rectangles and obtaining Green’s Theorem for the complicated domain by summing over
the small rectangles can be made to work in this case as well.
This completes our discusssion of Green’s Theorem in the plane.

You might also like