MAT342-notes
MAT342-notes
2020
ii
1
Introduction
Denote the set of complex numbers by
C = {x + iy | x, y ∈ R}
√
where i = −1 is defined such that i2 = −1.
Complex analysis is the study of functions of a complex variable. In the
first few chapters, we shall explore some introductory concepts, such as basic
properties of complex numbers and continuity of complex-valued functions.
The main emphasis is the concept of holomorphic functions, i.e. complex-
valued functions which are differentiable in a complex sense, and the many
applications of their somewhat magical properties. I used the word ’magical’
because holomorphicity is such a rigid condition that many of the results you
will see are somewhat unintuitive yet true.
We will start with some motivation. Basic algebra tells us that the num-
ber of roots of a polynomial with real coefficients is at most its degree. For
example, x2 + c has two real roots if c < 0, one root if c = 0, and no roots if
c > 0. Introducing the imaginary number i provides us with a more elegant
way of formulating this idea.
Complex Numbers
In this chapter, we will go through the basic algebraic and geometric prop-
erties of complex numbers.
Let’s denote by C∗ the set of non-zero complex numbers C\{0}. This set
is equipped with an additional operator:
x−iy
• inversion of a non-zero number: (x + iy)−1 = x2 +y 2
.
3
4 CHAPTER 1. COMPLEX NUMBERS
3. × is distributive over +.
The set C of complex numbers can be identified with the real vector space
2
R by the vector space isomorphism:
C → R2 ,
z 7→ (Re z, Im z),
x + iy 7→ (x, y).
Re z = x, Im z = y.
• z z̄ = |z|2 ,
x = r cos θ, y = r sin θ.
dr dθ̂
ieiθ = (cos θ̂ + i sin θ̂) + r (− sin θ̂ + i cos θ̂)
dθ dθ
iθ
dr e dθ̂
= + i eiθ
dθ r dθ!
dr dθ̂
= +i eiθ ,
dθ dθ
where the second equality above is obtained from (1.1). From above, we see
dr
that dθ = 0 and ddθθ̂ = 1. By our initial conditions, we obtain r(θ) ≡ 1 and
θ̂ ≡ θ.
(iθ)n
Alternatively, recall the following Maclaurin series: eiθ = ∞
P
n=0 n! . Us-
ing the fact that in = (−1)n/2 if n is even, and in = (−1)n−1/2 i if n is odd,
X (−1)n θn X (−1)n−1/2 θn
eiθ = +
even n
n! odd n
n!
θ2 θ4 θ3 θ5
= 1− + − ... + i θ − + − . . . = cos θ + i sin θ.
2! 4! 3! 5!
sn eint = reiθ
2. The complex plane C is both open and closed, but not bounded.
3. The punctured plane C∗ = C\{0} is open, but not closed nor bounded.
4. The unit disk D := D(0, 1) is open and bounded, but not closed.
Figure 1.3: Four connected subsets of C. Solid boundary lines are included
in the colored set, whereas dashed boundary lines are not included. The
first (from the left) is a simply connected domain. The second is closed
and simply connected. The third is a punctured disk, which is a multiply
connected domain. The last is closed and multiply connected.
Example 7.
1. ∅, C, D, D and R are simply connected.
2. The punctured plane C∗ , the punctured unit disk D∗ := D\{0}, and
the unit circle ∂D are multiply connected.
3. The annulus {z ∈ C | r < |z| < R} of inner radius r and outer radius
R is multiply connected.
Example 8. Consider the set S = {z ∈ C | |Im z1 | < 1}. In polar form
Short Quiz 1
1+i
1. Simplify i−1
.
Complex Functions
for all > 0, there exists N > 0 such that for all n ≥ N, |zn − z| < .
• zn wn → zw.
Proof. Let’s pick > 0. There are some high N1 , N2 ∈ N such that |zn − z| <
/2 if n ≥ N1 , and |wn − w| < /2 if n ≥ N2 . By triangle inequality, when
n ≥ max{N1 , N2 },
|zn + wn − z − w| ≤ |zn − z| + |wn − w| < + = .
2 2
13
14 CHAPTER 2. COMPLEX FUNCTIONS
but the value of this limit is not the same if we choose different values
of θ. For example, the limit is 1 when θ = 0, but it is −1 if θ = π2 .
Proposition 2.7. Every holomorphic function is continuous.
Proof. Let f : U → V be holomorphic. If a ∈ U ,
f (z) − f (a)
lim f (z) − f (a) = lim (z − a)
z→a z→a z−a
f (z) − f (a)
= lim (z − a) · lim
z→a z→a z−a
0
= 0 · f (a) = 0,
where the second equality follows from Theorem 2.2. Therefore, f is contin-
uous at a. As a is arbitrary, f is continuous on U .
2.2. HOLOMORPHIC FUNCTIONS 17
f (z)g(z) − f (a)g(a)
(f · g)0 (a) = lim
z→a z−a
f (z) g(z) − g(a) g(a) f (z) − f (a)
= lim +
z→a z−a z−a
0 0
= f (a)g (a) + g(a)f (a).
Example 13.
1. Every polynomial p(z) = dn=0 an z n with complex coefficients an ∈ C is
P
an entire function. We can show this inductively by product rule above
that every monomial an z n is holomorphic with derivative an nz n−1 and
by the addition rule, p is holomorphic.
Comparing (2.1) and (2.2), it is clear by taking the real and imaginary parts
of f 0 (a) that the partial derivatives of u and v at a exist and satisfy ux = vy
and vx = −uy . To show that these partial derivatives are continuous, we
need continuity of f 0 . We will obtain this for granted in Corollary 4.4.
Conversely, suppose u and v are continuously differentiable satisfying
ux = vy and vx = −uy . The Taylor series of u and v at a can be expressed
as:
1. | exp(z)| = ex ,
We would now like to find the inverse of exp. If we denote the inverse by
log, then
log(z) := ln |z| + i arg(z)
for any z 6= 0. However, as arg is multivalued, log is multi-valued and
therefore it is not a well-defined function. This is consistent with the fact
that exp is not injective due to property 3.
This problem can be fixed by using the principle argument Arg. Doing so
will introduce a ray of discontinuity (−∞, 0] corresponding to the points in C
with argument π. We can replace arg with Arg and define the principal value
of log(z), denoted by pv log(z). If we choose the codomain to be Arg(z) ∈
(−π, π], we then have the principal branch of the logarithmic function
The ray (−∞, 0] is often called a branch cut. Using this choice of branch cut,
Log is holomorphic on C\(−∞, 0], with image {x + iy | x ∈ R, −π < y < π}
and derivative z1 . (To verify this, compute the Wirtinger derivatives in polar
coordinates.)
Remark. In general, the branch cut can be taken to be any unbounded curve
from 0 which does not intersect itself. Straight rays of different angles are
often used if necessary.
For a non-integer c ∈ C, we define the power function z c to be the mul-
tivalued function on C∗ given by
z c := exp(c log(z)).
Similar to the logarithm, we can take the principal value of z c to be
pv z c := exp(c Log(z)).
Again, this becomes a holomorphic function outside a chosen branch cut
(−∞, 0].
Example 16. ii is (perhaps surprisingly) real and multivalued since
ii = e−π/2−2πk , where k ∈ Z.
Its principal value is pv ii = e−π/2 .
As exp is entire, so are cos and sin. However, the function tan is holomorphic
everywhere except at points z such that cos(z) = 0. You may check that the
usual trigonometric identities still hold.
The generalisation of hyperbolic functions are also clear:
The functions cosh and sinh can be viewed as the even part and the odd
part of the exponential function respectively. Both are entire functions, but
tanh is only holomorphic everywhere except at points z such that cosh(z) = 0.
The following property can be easily deduced by definition.
{z ∈ C | sin(z) = 0} = {nπ}n∈Z ,
nπ o
{z ∈ C | cos(z) = 0} = + nπ ,
2 n∈Z
{z ∈ C | sinh(z) = 0} = {inπ}n∈Z ,
n π o
{z ∈ C | cosh(z) = 0} = i + nπ .
2 n∈Z
Proof. By addition rule and Proposition 2.10, sin(x + iy) = sin x cosh y +
i sinh y cos x. Thus, if z = x + iy,
As such, sin(z) = 0 if and only if sin x = 0 and sinh y = 0, and the latter
occurs exactly when x ∈ {nπ}n∈Z , and y = 0.
The zeros of cos can be obtained from those of sin using the identity
cos(z) = sin(π/2 − z), and the zeros of hyperbolic functions can be obtained
from those of trigonometric functions by Proposition 2.10.
2.4. TRIGONOMETRIC FUNCTIONS 23
i i+z
tan−1 (z) = log ,
2 i−z
sinh−1 (z) = log z + [z 2 + 1]1/2 ,
1 1+z
tanh−1 (z) = log .
2 1−z
iw −iw
Proof. Let sin−1 (z) = w, then z = e −e 2i
. This can be written in quadratic
form:
(eiw )2 − 2izeiw − 1 = 0.
The quadratic formula gives us eiw = iz + [1 − z 2 ]1/2 , and using logarithm,
w = −i log iz + [1 − z 2 ]1/2 .
The resulting function is multivalued since the square root and the logarithm
are multivalued. Similar algebraic methods can be applied to obtain the
inverses of other functions and will be left to the reader as an exercise.
For each of the functions above, we can pick a branch cut of the square
root and the logarithm in order to obtain a holomorphic branch of the func-
tion. However, finding a nice branch cut for the inverse of the function
requires a more involved argument and we shall not attempt to find it.
Example 17. Let’s find the solution for the equation sin(2πz) = 2. By the
proposition above,
i i √
z = − log 2i + (−3)1/2 = − log i(2 ± 3)
2π 2π
i √ πi
=− ln(2 ± 3) + + 2πik ,
2π 2
√
i ln(2 ± 3) 1
=− + + k, k ∈ Z,
2π 4
24 CHAPTER 2. COMPLEX FUNCTIONS
Short Quiz 2
1, i, −1, −i, 0
Contour Integration
3.1 Curves in C
Definition 7. An arc / curve / path in C is subset of C parametrized by a
continuous function γ : [a, b] → C defined on a closed interval [a, b] ⊂ R.
Example 19. The function σ(t) = eit sin(2t), 0 ≤ t ≤ 2π, parametrises the
locus of the equation r = sin(2θ). This curve is a closed and smooth but
clearly not simple. See the leftmost curve in Figure 3.1.
25
26 CHAPTER 3. CONTOUR INTEGRATION
Figure 3.2: A simple closed curve splits its complement into a bounded do-
main U and an unbounded domain V .
Remark. Due to this theorem, simple closed curves are often called Jordan
curves. You may think that the theorem seems very intuitive, but the proof
is rather involved. The very first proof was out in 1910s (not by Camille
Jordan), relying on heavy machineries such as the theory of homology groups
3.2. INTEGRATION ALONG A CONTOUR 27
Notice that the value I is independent of the radius r and the center z0 .
3.2. INTEGRATION ALONG A CONTOUR 29
The second inequality follows from the fact thatRfor any continuous function
b Rb
h : [a, b] → C, we always have the inequality | a h(t)dt| ≤ a |h(t)|dt. We
Rb
shall prove this below. Let reiθ be the polar form of a h(t)dt. Then,
Z b Z b Z b
−iθ
h(t)dt = e h(t)dt = Re e−iθ h(t)dt
a a a
Z b Z b Z b
−iθ −iθ
= Re[e h(t)]dt ≤ |e h(t)|dt = |h(t)|dt,
a a a
3.3 Primitives
Definition 11. An antiderivative / primitive of a continuous function f on
a domain U ⊂ C is a holomorphic function F : U → C such that F 0 = f .
The existence of primitives makes computation of contour integrals much
easier. Regardless of the shape of the contour, it turns out that the integral
only depends on the endpoints of the contour.
Lemma 3.4. Suppose f is continuous on a domain U and has a primitive
F . Then, for any contour γ : [a, b] → U ,
Z
f (z)dz = F (γ(b)) − F (γ(a)),
γ
If γ is piecewise smooth, we can sum up the integral over all smooth parts
to obtain a similar result.
Corollary 3.5. If F is holomorphic on a domain U and F 0 ≡ 0, then F is
a constant function.
Proof. For any two points u and v in U , we can apply the lemma above by
setting f = 0 to obtain f (u) = f (v) because the integral over f over any
curve is 0.
Theorem 3.6. Suppose f : U → C is continuous on a domain U . The
following are equivalent:
(a) f has a primitive F : U → C,
H
(b) For any closed contour γ in U , σ
f (z)dz = 0.
Proof. Suppose (a) is true. Let γ : [a, c] → U be a closed contour, then
F (γ(b)) = F (γ(b)) because γ(b) = γ(a). Then, Lemma 3.4 immediately
gives us (a) ⇒ (b).
Suppose (b) is true. Pick a basepoint z0 ∈ U and for each z ∈ U , let
γz : [0, 1] → U be a smooth curve from γz (0) = z0 to γz (1) = z. Let’s define
F by Z
F (z) = f (w)dw.
γz
3.3. PRIMITIVES 31
F (z + h) − F (z)
f (z) −
h
Z Z
1 1
= f (z) − f (w)dw = f (z) − f (w) dw
h αh h αh
1
≤ L(αh ) max |f (z) − f (w)| = max |f (z) − f (w)|.
|h| w∈αh w∈αh
Figure 3.4: Curve α joining γ and σ. Figure 3.5: The square γ can be re-
placed by the smaller circle.
Since the function 1/(z + 2) is holomorphic on and inside γ, the second inte-
1
gral is zero by Cauchy-Goursat. The function z−2 is holomorphic everywhere
34 CHAPTER 3. CONTOUR INTEGRATION
Figure 3.6: On the gray annular domain above, curves 1, 2 and 3 have the
same endpoints but only 1 and 2 are homotopic relative to their endpoints.
Among closed curves 4 - 7, the only pair of homotopic curves is 5 and 6.
Short Quiz 3
1. What is the integral of 1/z along the circle C(0, 2)?
2. Compute the length of the curve γ(t) = cos(t)e( it) where 0 ≤ t ≤ π.
3. Compute the contour integral of the function 8z 3 along an L-shaped
contour which starts from 1 to 2i and passes through 0.
Answers: 1. 2πi, 2. π, 3. 30.
Chapter 4
Integration Formulas
35
36 CHAPTER 4. INTEGRATION FORMULAS
The case where the closed contour is chosen to be a circle yields an inter-
esting property of holomorphic functions.
Corollary 4.2 (Mean Value Property). Let f be holomorphic on a domain
U . For any closed disk D(z0 , r) in U ,
Z 2π
1
f (z0 ) = f (z0 + reit )dt.
2π 0
Proof. Let γ(t) = z0 + reit , 0 ≤ t ≤ 2π parametrize the circle C(z0 , r). Since
f is holomorphic on a simply connected open neighbourhood of D(z0 , r), by
Cauchy’s integral formula,
I
1 f (z)
f (z0 ) = dz
2πi γ z − z0
Z 2π
1 f (z0 + reit ) it
= ire dt
2πi 0 reit
Z 2π
1
= f (z0 + reit )dt.
2π 0
The reason why the corollary above deserves its name is clear if you view
the integral as the average value of f along the circle γ centered at z0 . Recall
that the length element ds in polar coordinates is ds2 = dr2 + r2 dt2 . When
r is constant, ds = dt and the equation can be rewritten as:
Z
1
f (z0 ) = f (z)ds.
L(γ) γ
Example 25. Let’s evaluate the integral I along γ parametrizing C(0, 2),
given by
ez
I
I= 2
dz.
γ z −1
By partial fractions decomposition, we can split I into I1 + I2 where
ez ez
I I
1 1
I1 = dz, I2 = dz.
2 γ z−1 2 γ z+1
By Cauchy’s formula, we immediately obtain I1 = πie and I2 = πie−1 . Thus,
I = 2πi cosh(1).
Theorem 4.3 (Cauchy’s Differentiation Formula). Let f : U → V be a
holomorphic function, γ be a simple closed contour in U , and W be the
domain enclosed by γ such that U ⊂ W . Then, the nth derivative f (n) (z0 ) of
f at a point z0 in W satisfies the following formula:
I
(n) n! f (z)
f (z0 ) = dz.
2πi γ (z − z0 )n+1
4.2. APPLICATIONS OF CAUCHY’S FORMULAS 37
Proof. We can again assume without loss of generality that z0 = 0. The base
case where n = 0 is exactly the previous theorem. Suppose the formula holds
for some natural number n. Then, for some small non-zero a,
Example 27. All polynomials of degree ≥ 1, exp, sin, cos, sinh, and cosh
are all unbounded.
Remark. There is no such analogue of Liouville’s theorem for real functions.
For example, the functions tanh and x21+1 are bounded and infinitely differ-
entiable (i.e. real analytic) in the whole R.
One consequence of Liouville’s theorem is the following.
Corollary 4.8. The image of a non-constant entire function is dense in C,
i.e. it intersects every open disk in the plane.
Proof. Suppose for a contradiction that the image avoids some open disk
1
D(z0 , r), i.e. |f (z) − z0 | ≥ r for all z ∈ C. Then, f (z)−z 0
is an entire function
whose modulus is bounded above by r. By Liouville’s theorem, this fraction
must be a constant function, but then this raises a contradiction because f
is not constant.
Liouville’s theorem also gives us a standard proof of the fundamental
theorem of algebra.
Theorem 4.9 (Fundamental Theorem of Algebra). Every complex polyno-
mial p(z) of degree d ≥ 1 has exactly d roots (counting multiplicity).
Proof. Let p(z) = dn=0 an z n where ad 6= 0 and suppose for a contradiction
P
that p has no roots. Then, 1/p(z) is an entire function. As |z| → ∞,
1 1 1 1
lim = lim d ad−1 a0 =0· = 0.
|z|→∞ p(z) |z|→∞ |z | |ad + + ... + | |ad |
z zd
Let’s pick any small > 0. By the definition of continuity, there is some
1
R > 0 such that p(z) ≤ whenever |z| > R. Since the closed disk D(0, R)
is a compact disk, the maximum value
1
M := max
|z|≤R p(z)
1
exists and is finite. Clearly, 1/p(z) is bounded because p(z) ≤ max{, M }.
By Liouville’s theorem, 1/p is constant, but this contradicts the fact that
p(z) is a non-constant polynomial.
Therefore, p has some root z1 ∈ C. This allows us to express p as a
product p(z) = (z − z1 )q1 (z) for some polynomial q1 (z) of degree d − 1. By
the same reasoning, q1 has a root z2 ∈ C and q1 (z) = (z − z2 )q2 (z) for some
polynomial q2 (z) of degree d − 2. Inductively, we can find d roots of p(z),
namely z1 , z2 , . . . zd (may be repeated).
Above is one of the many ways for us to prove the theorem. We will
present a much shorter one in section 5.5.
40 CHAPTER 4. INTEGRATION FORMULAS
Example 28. Let’s find the maximum and minimum of sin z on the closed
square S with vertices πi, π + πi, π + 2πi and 2πi. Recall from 2.3 that
| sin z|2 = sin2 x + sinh2 y. Observe the following.
Example 29. Let’s find the maximum and minimum of |f |, where f (z) =
z 2 − 7z + 12, on the closed unit disk D̄. Since f only attains 0 outside of D̄
at 3 and 4, it is sufficient to check maxima and minima along the boundary,
which is the unit circle. When |z| = 1,
Short Quiz 4
sin z
1. What is the integral of z−π
along the circle C(0, 5)?
2z 5
2. What is the integral of (2z−1)3
along the circle C(0, 5)?
45
46 CHAPTER 5. SERIES, ZEROS, AND POLES
When z0 = 0, the expansion is often called the MacLaurin series for f . The
radius of convergence of the Taylor series is the largest possible radius R > 0
such that the series converges for |z − z0 | < r. If there is no such maximum
value (e.g. when f is entire), we say that R = ∞.
1
Example 32. Examples 30 and 31 show the MacLaurin series of 1−z and
Log(1 − z). Both series cannot be extended beyond the unit disk since imme-
diately at z = 1, both functions are not well-defined. Therefore, both have
radius of convergence 1.
Example 33. The exponential function ez−z0 is an entire function. Since its
nth derivative at z0 is 1 for all n, it has a Taylor series about z0 with infinite
radius of convergence given below:
∞
X (z − z0 )n
ez−z0 = .
n=0
n!
Using this series, we can also derive the corresponding series for the functions
sin, cos, sinh and cosh.
1
Example 34. Let us compute the Taylor series for f (z) = 2i−z
about 2.
1 1 1 1
= = · z−2
2i − z (2i − 2) − (z − 2) 2i − 2 1 − 2i−2
∞ n X ∞
1 X z−2 (z − 2)n
= =
2i − 2 n=0 2i − 2 n=0
(2i − 2)n+1
∞ n+1
X 1+i
= − (z − 2)n .
n=0
4
√ z−2
The radius of convergence is 2 2 because the series converges when | 2i−2 |<
√
1, or equivalently, |z − 2| < 2 2.
5.2 Zeros
Definition 14. Suppose f is a holomorphic function on a domain U . We say
that f has a zero at a point z0 ∈ U order k if f (k) (z0 ) 6= 0 and f (n) (z0 ) = 0
for all n < k. If k = 1, we say that the zero is simple.
(c) there is some holomorphic function g on U such that g(z0 ) 6= 0 and for
each z ∈ U , f (z) = (z − z0 )k g(z).
Proof. (a) ⇐⇒ (b) follows immediately from Taylor’s theorem, and (c) ⇒ (b)
follows from direct computation of derivatives using Leibniz’s rule.
f (z)
Assume (b) holds. The function g(z) = (z−z k is holomorphic on U \{z0 }.
0)
About z0 , the Taylor series of f gives us a well-defined power series repre-
sentation of g:
P∞ ∞
an (z − z0 )n X
g(z) = n=k
= an+k (z − z0 )n .
(z − z0 )k n=0
Lemma 5.4. Let f be holomorphic on some disk D(z0 , r) and let {zn }n∈N be
an infinite sequence of distinct zeros of f such that zn → z0 . Then, f ≡ 0.
This is a contradiction.
Example 35. The only holomorphic function f (z) which has zeros on the
set of rational numbers Q is the zero function.
for any simple closed contour γ in A enclosing the disk D(z0 , r).
50 CHAPTER 5. SERIES, ZEROS, AND POLES
∞ ∞
1 1 1 1 X z n X z n
= · z = = n+1
. (5.1)
w−z w 1− w
w n=0 w n=0
w
w
However, if w ∈ γ2 , then z
< 1 and thus,
∞ −1
1 1 1 1 X w n X zm
= · w = = m+1
. (5.2)
z−w z 1− z
z n=0 z m=−∞
w
By (5.1) and (5.2), we can convert the two integrals into the desired series.
∞
! −1
!
z n f (w) z m f (w)
I I
1 X 1 X
f (z) = dw + dw
2πi γ1 n=0 wn+1 2πi γ2 m=−∞ wm+1
∞ I −1 I
X 1 f (w) n
X 1 f (w)
= n+1
dw z + m+1
dw z m
n=0
2πi γ1 w m=−∞
2πi γ2 w
∞
X −1
X ∞
X
= an z n + am z m = an z n .
n=0 m=−∞ n=−∞
Definition 15. The bi-infinite series in the theorem above is called the Lau-
rent series of f about the point z0 .
∞ n −1
1 1 1 1X 1 X
= · 1 = = zn.
z−1 z 1− z
z n=0 z n=−∞
This series converges when | z1 | < 1, or equivalently, when 1 < |z| < ∞.
3z
Example 39. Let’s find a Laurent series for f (z) = z2 +z−2 converging on
1 2
the annulus {1 < |z + 1| < 2}. By partial fractions, f (z) = z−1 + z+2 has
1
singularities at −1 and 2. The Taylor series of z−1 about −1 is
∞ n ∞
1 1 1 1X z+1 X
=− · z+1 = − =− 2−n−1 (z + 1)n .
z−1 2 1− 2 2 n=0 2 n=0
52 CHAPTER 5. SERIES, ZEROS, AND POLES
1
convergent when | z+1 | < 1, i.e. in the domain U1 = {z | 1 < |z + 1|}.
Combining the two series, we deduce that the Laurent series of f about −1
is
∞
(
X −2−n−1 , if n ≥ 0,
f (z) = an (z + 1)n , where an = n+1
n=−∞
2(−1) , if n < 0,
convergent in the annulus U1 ∩ U2 = {z | 1 < |z + 1| < 2}.
5.4 Singularities
We will use an asterisk on a disk to introduce a puncture at its centre:
(b) there is a holomorphic function g on U such that g(z0 ) 6= 0 and for all
z 6= z0 ,
g(z) = (z − z0 )k f (z).
z3 z5 z2 z4
sin z 1
= · z− + − ... = 1 − + − ...
z z 3! 5! 3! 5!
sin z sin 0
Since limz→0 z
= 1, we may set 0
= 1 so that it becomes an entire
function.
Theorem 5.8 (Riemann’s Removable Singularity). Suppose U is a domain
and f is a holomorphic function on U \{z0 } with a singularity at z0 . The
following are equivalent.
(a) z0 is a removable singularity.
Example 44. Let γ(t) = e2πit , 0 ≤ t ≤ 1 parametrise the unit circle. The
image of γ under the power map f (z) = z n , where n ≥ 1, is f ◦ γ(t) = e2πint .
The winding number of f ◦ γ is n, which coincides with the order of the zero
of f at 0.
Therefore, we have the first part of the equation. Suppose {z1 , . . . zm } and
{w1 , . . . wn } are the sets of zeros and poles in V respectively.
Pick any zero zj and let kj be its order. There is some meromorphic
function gj such that f (z) = (z −zj )kj gj (z) and gj is holomorphic at zj where
gj (zj ) 6= 0. Pick a small radius j > 0 so that inside the closed disk D(zj , j )
gj have no poles nor zeros aside from zj . Let γj be the circle C(z1 , j ), then
where the last equality follows from the fact that gj0 (z)/gj (z) is holomorphic
on D(zj , j ) and Cauchy-Goursat.
Pick any pole wj and let lj be its order. There is some meromorphic
function hj such that f (z) = hj (z)(z − wj )−lj and hj is holomorphic at wj
where hj (zj ) 6= 0. Pick a small radius δj > 0 so that inside the closed
disk D(wj , δj ) g have no poles nor zeros aside from wj . Let σj be the circle
C(wj , δj ), then
where the last inequality follows from the fact that h0j (z)/hj (z) is holomorphic
on D(wj , δj ).
The curve γ can be split into some m + n simple closed contours each of
which encloses exactly one zero or pole. By deformation theorem, the integral
along γ is equal to the sum of the integrals along each of the contours γj for
j = 1 . . . m and σj for j = 1 . . . n.
5.5. COUNTING ZEROS AND POLES 57
Figure 5.3: The contour γ can be split into a collection of m+n simple closed
contours each of which encloses exactly one zero or pole and deformed into
γj ’s and σj ’s.
Example 45. Let’s compute the integral of sec z along a square γ of side
length 7 centered at 0 by using the fact that sec0 z = sec z tan z. There are
four simple poles of sec enclosed by γ, namely ± π2 , ± 3π
2
. sec has no zeros.
By the argument principle,
I I
sec z tan z
tan z dz = dz = 2πi · (−4) = −8πi.
γ γ sec z
g(z)
Proof. Define a meromorphic function h(z) = f (z)
+ 1 on U . Along γ, h is
holomorphic and |h(z) − 1| = fg(z)
(z)
< 1. Therefore, the contour h ◦ γ lies in
the disk D(1, 1) disjoint from 0 and consequently has zero winding number.
Denote by Zf +g and Zf the respective numbers of zeros of f + g and f inside
58 CHAPTER 5. SERIES, ZEROS, AND POLES
Short Quiz 5
sin z cos z −2
, , e(z−π) , tan((z − π)−1 )
z−π z−π
3. Which of the four functions above have a removable singularity at π?
Evaluation of Integrals
To extract out a−1 from the series, we can take the (k − 1)th derivative of g
evaluated at z0 :
g (k−1) (z0 ) = (k − 1)!a−1 .
Therefore, when we are given the order k of the pole, the residue can be
computed as follows:
1 dk−1 h i
Resf (z0 ) = lim k−1 (z − z0 )k f (z) .
(k − 1)! z→z0 dz
61
62 CHAPTER 6. EVALUATION OF INTEGRALS
Proof. Pick a pole wj and let kj be its order. Let j > 0 be small enough such
that the only pole of f inside the closed disk D(wj , j ) is wj , and parametrize
the boundary of this disk by
γj (t) = wj + j e2πit , 0 ≤ t ≤ 1.
If the Laurent series of f about wj is ∞ n
P
n=−kj an (z − wj ) , then
I ∞
X I
f (z)dz = an (z − wj )n dz
γj n=−kj γj
∞
X Z 1
= an nj e2πitn · γj0 (t)dt
n=−kj 0
X Z 1 Z 1
= an 2πin+1
j e2πit(n+1) dt + a−1 2πidt
n≥−kj ,n6=−1 0 0
X e2πit(n+1) 1
= an n+1
j + 2πia−1
n≥−kj ,n6=−1
n+1 0
circle C(0, 1/2). The only pole of f enclosed by γ is 0 and its order is 3. By
Residue Theorem,
d2
2πi 1
I = 2πi Resf (0) = lim
2! z→0 dz 2 z 2 − 1
−6z 2 + 2
= πi lim 2 = −2πi.
z→0 (z − 1)3
for some generic rational function F . With the substitution z = eix where
0 ≤ x ≤ 2π, we can transform I into a contour integral
z + z −1 z − z −1 dz
I
I= F , ,
C(0,1) 2 2i iz
We will explore in detail different types of integrals you can solve accord-
ing to the type of integrand f and the shape of the contour γ used.
6.3. DEFINITE INTEGRALS 65
Semicircular Contour
Consider integrals of the type
Z ∞
I= f (x)dx,
−∞
I = lim I1 = lim I0 − I2 = I0 .
R→∞ R→∞
Example 49. Suppose the integrand is f (x) = (x6 + 1)−1 . This function
can be extended to a meromorphic function on C and it has 6 simple poles
forming the set {z | z 6 = −1} = {±i, e±πi/6 , e±5πi/6 }. The ones on the upper
half plane are i, eπi/6 , e5πi/6 .
By L’Hopital’s rule, the residue of f at any pole c is
z−c 1 c c
Resf (c) = lim 6
= lim 5 = 5 = − .
z→c z + 1 z→c 6z 6c 6
Let’s pick a large R > 0 and define γ as outlined before. By residue
theorem,
I
I0 = f (z)dz = 2πi Resf (i) + Resf (eπi/6 ) + Resf (e5πi/6 )
γ
eπi/6 e5π/6
i πi 2π
= 2πi − − − =− i + eπi/6 + e5πi/6 = .
6 6 6 3 3
66 CHAPTER 6. EVALUATION OF INTEGRALS
eiz
f (z) =
z2 + 1
and then take its real and imaginary parts (since cos z = Re eiz and sin z =
Im eiz ). The function f has simple poles at ±i. With the same setup as
before,
e−1
I
I0 = f (z)dz = 2πi Resf (i) = 2πi = πe−1 .
γ i+i
Since
1 1
lim ≤ lim 2 = 0,
R→∞ z 2+1 R→∞ R − 1
where the function f has a finite number of poles throughout C, with 0 being
the only real pole. Due to the presence of singularity at 0, we need to add
a little dent to our semicircular contour. We pick an arbitrarily small ρ > 0
and construct γ by gluing together the following curves:
I= lim I1 + I3 = I0 − lim I4 .
R→∞,ρ→0 ρ→0
Therefore, I = πi. Taking the real and imaginary parts, we obtain the
following two integrals as well:
Z ∞ Z ∞
cos x sin x
dx = 0, and dx = π.
−∞ x −∞ x
Sector Contour
Consider integrals of the type
Z ∞
I= f (x)dx,
0
Rotational symmetry will imply that there is some constant C such that
Z Z
f (z)dz = C f (z)dz.
γ3 γ1
After showing by ML inequality that I2 → 0 as R → ∞,
I = lim I1 = lim I0 − I2 − I3 = I0 − CI
R→∞ R→∞
−1
Therefore, I = (1 + C) I0 .
Example 52. Let’s evaluate I when f (x) = (1 + x3 )−1 . This integrand has
rotational symmetry of angle θ = 2π 3
since
1 1
f (e2πi/3 z) = = = f (z).
1 + e2πi z 3 1 + z3
The poles of f are −1 and e±iπ/3 , each of which is simple. Pick a large R
and define a sector contour γ of angle 2π 3
as outlined above. Since eiπ/3 is the
only pole enclosed by γ,
z − eiπ/3 1 2πeπi/3
I0 = 2πi Resf (eiπ/3 ) = 2πi lim = 2πi lim = .
z→eiπ/3 z3 + 1 z→eiπ/3 3z
2 3
Let’s parametrize γ3 by γ3 (r) = re2πi/3 as r varies from R to 0. Then,
Z 0 2πi/3 Z R
e 2πi/3 1
I3 = 3
dr = −e 3
dr = −e2πi/3 I1 .
R 1 + r 0 1 + r
By ML inequality, we have that I2 → 0 because
Z
2πR/3 2πR
f (z)dz ≤ L(γ2 ) max |f (z)| = 3
≤ → 0.
γ2 z∈γ2 min|z|=R,Imz≥0 |z + 1| 3(R3 − 1)
Thus,
2π
I = (1 − e2πi/3 )−1 I0 = √ .
3 3
6.3. DEFINITE INTEGRALS 69
Keyhole Contour
Consider integrals of the type
Z ∞
I= f (x)dx.
0
Additional parameters ρ and are needed and they are taken to be arbitrarily
small.
over (0, ∞). Taking the branch cut of the square root to be arg z = 0, we
complexify f to a meromorphic function on C\[0, ∞) with a pole at −1 of
order 2. By residue theorem,
d√
I0 = 2πi Resf (−1) = 2πi lim z = π.
z→−1 dz
70 CHAPTER 6. EVALUATION OF INTEGRALS
I1 + I2 + I3 + I4 = I0
I +0+I +0=π
π
∴I= .
2
Chapter 7
Harmonic Functions
7.1 Harmonicity
Definition 21. Let U ⊂ R2 be a non-empty open subset. A real-valued
function u : U → R is harmonic if it is continuously twice differentiable and
it satisfies the Laplace’s equation:
∆u := uxx + uyy = 0.
2
∂ 2
∂
The operator ∆ = ∂x 2 + ∂y 2 is known as the Laplacian / Laplace operator.
There are many instances in which polar coordinates are used. With the usual
change of variables (x, y) = (r cos θ, r sin θ), this operator becomes
∂2 1 ∂ 1 ∂2
∆= + +
∂r2 r ∂r r2 ∂θ2
in polar coordinates.
Example 56. Denote the real and imaginary parts of the principal logarithm
Log : C\(−∞, 0] as u and v accordingly. In Cartesian coordinates,
p y
u(x, y) = ln x2 + y 2 , v(x, y) = tan−1 .
x
It may not seem clear at first that both u and v are harmonic, but in polar
coordinates, u(r, θ) = ln r and v(r, θ) = θ and it is much easier to show that
these harmonic whenever r > 0 and −π < θ < π.
71
72 CHAPTER 7. HARMONIC FUNCTIONS
It is worth noting that the only ingredients of the proof of the maximum
modulus principle of holomorphic functions in Lemma 4.11 and Theorem 4.12
are continuity, mean value property and connectivity of the domain. Since
these three properties are satisfied by every harmonic function on a domain,
harmonic functions automatically satisfy the maximum modulus principle.
Nonetheless, we shall give a slicker proof using harmonic conjugates.
Theorem 7.5. Harmonic functions satisfy the maximum modulus principle.
Proof. Let u be a harmonic function on a domain U ⊂ C. Pick a har-
monic conjugate v of u on U and define f (x + iy) = u(x, y) + iv(x, y) on U .
The function is a non-constant holomorphic function ef (z) with modulus is
|ef (z) | = eu (x, y). Applying the maximum modulus principle to ef (z) , then
eu and therefore u(x, y) do not attain maximum on U . We can apply the
same argument on −u to conclude that u does not attain minimum on U .
Therefore, |u| does not attain maximum on U , and if U is bounded, maxima
are achieved only on the boundary.
Consequently,
Z 2π
1
u(r, θ) − F (θ) = P (r, t − θ) (F (t) − F (θ))dt.
2π 0
Since P is positive on D,
Z 2π
1
|u(r, θ) − F (θ)| ≤ P (r, t − θ) |F (t) − F (θ)|dt. (7.4)
2π 0
1 − r2
Z Z
1 1
P (r, t − θ) |F (t) − F (θ)|dt = |F (t) − F (θ)|dt
2π J 2π J |1 − rei(t−θ) |2
1 − r2
Z
≤ |F (t) − F (θ)|dt
2π sin2 δ J
D(1 − r2 )
≤ .
2π sin2 δ
R 2π
where D = 0 |F (t)|dt. This upper bound converges to 0 as r → 1. Combin-
ing this with (7.5), we have shown that the upper bound in (7.4) converges
to 0 as r → 1 and subsequently δ → 0. In other words, u(r, θ) → F (θ) as
r → 1. Hence, u extends continuously to the boundary condition F .
Example 60. Let F (θ) = 1 when 0 ≤ θ < π and F (θ) = 0 when π ≤ θ < 2π.
The unique harmonic function on the unit disk with boundary value F (θ) is
Z π π
1 1 −1 1+r t−θ
u(r, θ) = P (r, t − θ)dt = tan tan
2π 0 π 1−r 2 0
1 −1 1 + r π − θ 1 −1 1 + r θ
= tan tan + tan tan
π 1−r 2 π 1−r 2
2
1 r −1
= tan−1 .
π 2r sin θ
Proof. It is clear from Proposition 7.3 that harmonicity implies mean value
property. Assume now that u satisfies the MVP. Let’s pick a point z0 ∈ U
in complex coordinates. By openness, there is some r > 0 such that U
contains the closed disk D(z0 , r). By the Dirichlet principle, there is a unique
continuous function v on D(z0 , r) that coincides with u on the circle C(z0 , r)
and is harmonic on the open disk D(z0 , r).
The function u − v on D(z0 , r) satisfies the MVP because both u and v
satisfy the MVP. Therefore, u − v satisfies the maximum modulus principle
too. Since u − v ≡ 0 on the boundary, u − v ≡ 0 on D(z0 , r). In particular,
u is harmonic on the open disk D(z0 , r). Since z0 is an arbitrary point in U ,
u must be harmonic everywhere on U .
Definition 23. The flow is incompressible if and only if the vector field V
has zero divergence:
div V = ∇ · V = px + qy ≡ 0.
curl V = ∇ × V = qx − py ≡ 0.
Notice that the integrand coincides with the curl of V . Irrotationality essen-
tially ensures that the circulation along γ is zero because this integral always
vanishes. (Alternatively, you may say that the vector field is conservative.)
When both assumptions are met, we say that the fluid has an ideal fluid
flow. Observe that they very much resemble the Cauchy-Riemann equations.
is holomorphic on Ω ⊂ C.
For an ideal fluid flow, the function f above is often called the complex
velocity of the fluid flow.
One implication of irrotationality is that f admits a primitive F (x+iy) =
φ(x, y) + iψ(x, y) on Ω. By Cauchy-Riemann equations and equation 2.1,
80 CHAPTER 7. HARMONIC FUNCTIONS
dF
p − iq = f = = φx + iψx = φx − iφy .
dz
In particular, φ satisfies ∇φ = (φx , φy ) = (p, q) = V . Moreover, both
φ and ψ are necessarily harmonic because they are the real and imaginary
parts of the holomorphic function F .
Definition 25. The level sets of the velocity potential {(x, y) ∈ Ω | φ(x, y) =
c} for some real constant c are called the equipotentials of the flow. The level
sets of the stream function {(x, y) ∈ Ω | ψ(x, y) = d} for some real constant
d are called the streamlines of the flow.
∇φ · ∇ψ = φx ψx + φy ψy = ψy ψx + (−ψx )ψy = 0.
The proof above also shows that streamlines have tangent vectors ∇φ.
Given any fluid particle (x0 , y0 ) in Ω, since the vector field V coincides
with ∇φ and determines its flow trajectory, the particle always flows along
a unique streamline {(x, y) | ψ(x, y) = ψ(x0 , y0 )}.
Example 62. The vector field V (x, y) = (x, −y) induces an ideal fluid flow
2
on the plane. The complex potential can be chosen to be F (z) = z2 =
x2 −y 2
2
+ ixy. V induces hyperbolic streamlines {xy = d}. The equipotentials
of V are hyperbolas of the form {x2 − y 2 = 2c}. At the origin, the flow has
a unique fixed point, which is a saddle point of the flow. Perpendicularity
of the equipotentials and the streamlines fails here because the vector field
vanishes.
Short Quiz 7