0% found this document useful (0 votes)
2K views53 pages

Conway "Functions of One Complex Variable" Solutions

This document outlines the contents and topics covered in a course on complex analysis. It includes chapters on elementary properties of analytic functions, complex integration, singularities, the maximum modulus theorem, compactness and convergence in the space of analytic functions, and harmonic and entire functions.

Uploaded by

darkphysicist4
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
2K views53 pages

Conway "Functions of One Complex Variable" Solutions

This document outlines the contents and topics covered in a course on complex analysis. It includes chapters on elementary properties of analytic functions, complex integration, singularities, the maximum modulus theorem, compactness and convergence in the space of analytic functions, and harmonic and entire functions.

Uploaded by

darkphysicist4
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 53

Sadman Rahman

MATH 399 – FA’22


Complex Analysis
Instructor: Dr. David Schmitz

Main Textbook: Functions of One


Complex Variable I by Conway

April 21, 2023

Springer Nature
Contents

I Elementary Properties of Analytic Functions . . . . . . . . . . . . . . . . . . . . . . 1

II Complex Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

III Singularities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

IV The Maximum Modulus Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

V Compactness and Convergence in the Space of Analytic Functions . . . 27


Module – 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Module – 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
Module – 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

VI Harmonic and Entire Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

v
Chapter I
Elementary Properties of Analytic Functions


∑︁
Problem I.1 Find the radius of convergence of the series 𝑎 𝑛 𝑧 𝑛 ; 𝑎 ∈ C.
𝑛=0

Solution I.1 Let, radius of convergence be 𝑅. Here, we set 𝑎 𝑛 = 𝑎 𝑛 and by analogy



∑︁
to the power series 𝑎 𝑛 𝑧 𝑛 , we have
𝑛=0

𝑎𝑛
𝑅 = lim
𝑛→∞ 𝑎 𝑛+1

𝑎𝑛
= lim 𝑛+1
𝑛→∞ 𝑎

1 1
= lim =
𝑛→∞ 𝑎 |𝑎|

1
Thus, 𝑅 =
|𝑎|
if 𝑎 ≠ 0, and 𝑅 = ∞ if 𝑎 = 0. ♦

∑︁ 2
Problem I.2 Find the radius of convergence of the series 𝑎 𝑛 𝑧 𝑛 ; 𝑎 ∈ C.
𝑛=0

2
Solution I.2 Let, radius of convergence be 𝑅. Here, we set 𝑎 𝑛 = 𝑎 𝑛 and by analogy

∑︁
to the power series 𝑎 𝑛 𝑧 𝑛 , we have
𝑛=0

1
2 I Elementary Properties of Analytic Functions

𝑎𝑛
𝑅 = lim
𝑛→∞ 𝑎 𝑛+1
2
𝑎𝑛
= lim
𝑛→∞ 𝑎 (𝑛+1) 2
1 1
= lim = lim
𝑛→∞ 𝑎 2𝑛+1 𝑛→∞ |𝑎| 2𝑛+1


 0, |𝑎| > 1;
1




Thus, 𝑅 = lim = 1, |𝑎| = 1;
𝑛→∞ |𝑎| 2𝑛+1 
 ∞, |𝑎| < 1;


Problem I.3 Show that 𝑓 (𝑧) = |𝑧| 2 is differentiable only at the origin.

Solution I.3 We recall by definition, 𝑓 (𝑧) is differentiable at 𝑧 if

𝑓 (𝑧 + ℎ) − 𝑓 (𝑧)
lim
ℎ→0 ℎ

exists; Substituting 𝑓 (𝑧) = |𝑧| 2 , we have

|𝑧 + ℎ| 2 − |𝑧| 2
𝑓 ′ (𝑧) = lim
ℎ→0 ℎ
(𝑧 + ℎ) ( 𝑧¯ + ℎ) ¯ − 𝑧 𝑧¯
= lim
ℎ→0 ℎ
(𝑧 + ℎ) ( 𝑧¯ + ℎ) ¯ − 𝑧 𝑧¯
= lim
ℎ→0 ℎ
𝑧 𝑧¯ + 𝑧 ℎ¯ + ℎ 𝑧¯ + ℎ ℎ¯ − 𝑧 𝑧¯
= lim
ℎ→0 ℎ
ℎ¯ ¯
 
= lim 𝑧¯ + 𝑧 + ℎ
ℎ→0 ℎ
¯ ¯ By definition, if the limit of 𝑤 exists, then we can approach
Now, let 𝑤 := 𝑧¯ + 𝑧 ℎℎ + ℎ.
from any any direction in C and arrive at the same limit point. In particular, we
choose the real and imaginary axes, to obtain:
Along Real Axis: Since ℎ = ℎ, ¯ we have

lim (𝑤) = 𝑧¯ + 𝑧.
ℎ→0

¯ we have
Along Imaginary Axis: Since ℎ = − ℎ,

lim (𝑤) = 𝑧¯ − 𝑧.
ℎ→0

If we assume 𝑓 is differentiable at 𝑧, then we must have


I Elementary Properties of Analytic Functions 3

𝑧¯ + 𝑧 = 𝑧¯ − 𝑧 =⇒ 𝑧 = 0.

Thus, 𝑓 (𝑧) = |𝑧| 2 is differentiable only at 𝑧 = 0 and 𝑓 ′ (0) = 0. ♦


Problem I.4 Describe the set {𝑧 : 𝑒 𝑧 = 𝑖}.
Solution I.4 Consider the set {𝑧 : 𝑒 𝑧 = 𝑖}. Taking complex logarithm on both sides,
we obtain

𝑧 =: log(e𝑧 ) = log(𝑖)
:= {ln |𝑖| + 𝑖 arg(𝑖) + 2𝜋𝑖𝑘 : 𝑘 ∈ Z}
= {ln(1) + 𝑖𝜋/2 + 2𝜋𝑖𝑘 : 𝑘 ∈ Z}
n 𝜋 o
= 𝑖 + 2𝜋𝑖𝑘 : 𝑘 ∈ Z
2
n 𝜋 o
Thus, {𝑧 : 𝑒 𝑧 = 𝑖} = 𝑖 + 2𝜋𝑖𝑘 : 𝑘 ∈ Z . ♦
2

Problem I.5 Give the principal branch of 1 − 𝑧.
Solution I.5 We recall, the principal branch of log(𝑧) is given by

Log(𝑧) = ln |𝑧| + 𝑖 Arg(𝑧),

where Arg(·) denotes the principal argument. We also recall that this branch is ana-
lytic in 𝐺, where 𝐺 = C \{𝑧 ∈ R : ℜ(𝑧) ≤ 0}.
√  
Now, by definition, 1 − 𝑧 := exp 12 Log(1 − 𝑧) . Here, the principal branch of
the logarithm in the exponent is given by

Log(1 − 𝑧) = ln |1 − 𝑧| + 𝑖 Arg(1 − 𝑧)
     
1 1 𝑖
∴ exp Log(1 − 𝑧) = exp ln |1 − 𝑧| · exp Arg(1 − 𝑧)
2 2 2
√ √︁ 
∴ 1 − 𝑧 = |1 − 𝑧| · cos 𝜃 + 𝑖 sin 𝜃 ,

where 𝜃 := 12 Arg(1 − 𝑧).



We notice that 1 − 𝑧 is analytic ∀𝑧 ∈ C \{𝑧 ∈ R : ℜ(𝑧) ≥ 1}. ♦
Problem I.6 Find the image of {𝑧 : ℜ(𝑧) < 0, |ℑ(𝑧)| < 𝜋} under the exponential
function.
Solution I.6 Let 𝑧 := 𝑥 + 𝑖𝑦. Then we have

e𝑧 = e 𝑥+𝑖𝑦 = e 𝑥 e𝑖𝑦 = e 𝑥 (cos 𝑦 + 𝑖 sin 𝑦).

Now, let 𝑟 = e 𝑥 . Since ℜ(𝑧) < 0, we have 0 < 𝑟 < 1. Moreover, we have
𝑦 ∈ (−𝜋, 𝜋). For a fixed value of 𝑟 ∈ (0, 1), letting 𝑦 vary in (−𝜋, 𝜋), the ex-
pression 𝑟 (cos 𝑦 + 𝑖 sin 𝑦) defines a circle with radius 𝑟 centered at (0, 0) but without
4 I Elementary Properties of Analytic Functions

the points (0, 0) and (−𝑟, 0). This is because 𝑦 ∈ (−𝜋, 𝜋) and 𝑟 ≠ 0.

Next, we let 𝑟 vary from 0 to 1 and observe that the expression 𝑟 (cos 𝑦 + 𝑖 sin 𝑦)
defines the open unit disk without its center and also without the negative real axis.
A sketch of the region is given in the following figure.

Fig. I.1 Sketch of the set {e𝑧 : ℜ(𝑧) < 0, |ℑ(𝑧) | < 𝜋 }.

Here, the reddish area indicates the excluded negative real axis and {0} on the
unit disk (in blue). ♦

Problem I.7 Find fixed points of a dilation, a translation and the inversion on C∞ .

Solution I.7 Dilation: We recall a Möbius Transformation 𝑀 is a dilation if


𝑀 𝑧 = 𝑎𝑧. In general, we have
𝑎𝑧 + 𝑏
𝑀𝑧 = .
𝑐𝑧 + 𝑑
So, to get dilation, we set 𝑏 = 0, 𝑐 = 0, 𝑑 = 1. Now, we find its fixed points. For fixed
points, we recall
𝑀 𝑧 = 𝑧 =⇒ 𝑎𝑧 = 𝑧.
Hence, it’s clear that 𝑧 = 0 and 𝑧 = ∞ are fixed points of 𝑀. We verify this as follows:
𝑏 0
𝑀 (0) = = = 0.
𝑑 1
𝑎 𝑎
𝑀 (∞) := = = ∞.
𝑐 0
I Elementary Properties of Analytic Functions 5

Thus, for a dilation, the fixed points are 0 and ∞.

Translation: We recall a Möbius Transformation 𝑀 is a translation if 𝑀 𝑧 = 𝑧 + 𝑏.


In general, we have
𝑎𝑧 + 𝑏
𝑀𝑧 = .
𝑐𝑧 + 𝑑
So, to get translation, we set 𝑎 = 1, 𝑐 = 0, 𝑑 = 1. Now, we find its fixed points. For
fixed points, we recall
𝑀 𝑧 = 𝑧 =⇒ 𝑧 + 𝑏 = 𝑧.
Hence, it’s clear that 𝑧 = ∞ is a fixed point of 𝑀. In fact, this is the only fixed point
of 𝑀. We verify this as follows:
𝑎 1
𝑀 (∞) := = = ∞.
𝑐 0
Thus, for a translation, the only fixed point is ∞.

Inversion: We recall a Möbius Transformation 𝑀 is an inversion if


𝑀 𝑧 = 1/𝑧. In general, we have
𝑎𝑧 + 𝑏
𝑀𝑧 = .
𝑐𝑧 + 𝑑
So, to get inversion, we set 𝑎 = 0, 𝑏 = 1, 𝑐 = 1, 𝑑 = 0. Now, we find its fixed points.
For fixed points, we recall

𝑀 𝑧 = 𝑧 =⇒ 1/𝑧 = 𝑧.

Hence, it’s clear that 𝑧 = 1 and 𝑧 = −1 are fixed points of 𝑀. We verify this as
follows:
𝑎+𝑏 0+1
𝑀 (1) = = = 1.
𝑐+𝑑 1+0
−𝑎 + 𝑏 0+1
𝑀 (−1) = = = −1.
−𝑐 + 𝑑 −1 + 0
Thus, for an inversion, the fixed points are 1 and −1. ♦
𝑎𝑧 + 𝑏
Problem I.8 If 𝑇 𝑧 = find 𝑧2 , 𝑧3 , 𝑧4 (in terms of 𝑎, 𝑏, 𝑐, 𝑑) such that
𝑐𝑧 + 𝑑
𝑇 𝑧 = (𝑧, 𝑧2 , 𝑧3 , 𝑧4 ).

Solution I.8 We recall by definition, (𝑧, 𝑧2 , 𝑧3 , 𝑧4 ) is the image of the unique Möbius
Transformation which maps 𝑧 2 ↦→ 1, 𝑧 3 ↦→ 0 and 𝑧 4 ↦→ ∞.
According to the problem, since (𝑧, 𝑧2 , 𝑧3 , 𝑧4 ) is the image of 𝑇 𝑧, 𝑇 is this unique
transformation. Thus, we are done if we can find the inverse images of 𝑧2 , 𝑧3 , 𝑧4 .
Since all Möbius transformations are invertible with the inverse given by
6 I Elementary Properties of Analytic Functions

𝑑𝑧 − 𝑏
𝑇 −1 (𝑧) = ,
−𝑐𝑧 + 𝑎

we are done if we set 𝑧2 = 𝑇 −1 (1), 𝑧3 = 𝑇 −1 (0) & 𝑧 4 = 𝑇 −1 (∞). We compute these


three values as follows:
𝑑−𝑏
𝑇 −1 (1) = ;
−𝑐 + 𝑎
−𝑏
𝑇 −1 (0) = ;
𝑎
𝑑
𝑇 −1 (∞) = ;
−𝑐
These are the values for which 𝑇 𝑧 = (𝑧, 𝑧2 , 𝑧3 , 𝑧4 ). ♦

Problem I.9 If 𝑢, 𝑣 : R2 → R are C 1 and they satisfy the Cauchy-Riemann equa-


tions, then prove that 𝑓 = 𝑢 + 𝑖𝑣 is conformal, unless 𝑢, 𝑣 are constant functions.

Solution I.9 Suppose 𝑢, 𝑣 : R2 → R are C 1 and they satisfy the CR-equations. Set
𝑓 = 𝑢 + 𝑖𝑣, a complex function in 𝐺. Now, if 𝑢, 𝑣 are constant functions, then the
Jacobian 𝐽 𝑓 of 𝑓 satisfies det 𝐽 𝑓 = 0 ≯ 0 and so 𝑓 doesn’t preserve orientation. So,
we assume 𝑢, 𝑣 are non-constant functions. Using the CR-equations, we can rewrite
the Jacobian of 𝑓 as follows:

©𝑢 𝑥 𝑢 𝑦 ª © 𝑢 𝑥 𝑢 𝑦 ª
𝐽𝑓 = ­ ®=­ ®.
« 𝑣 𝑥 𝑣 𝑦 ¬ « −𝑢 𝑦 𝑢 𝑥 ¬
   
𝑢𝑥 𝑢𝑦
Now, let, 𝛼 := and 𝛽 := be the tangent vectors of two smooth curves
𝑣𝑥 𝑣𝑦
Γ1 , Γ2 respectively. It follows from a direct computation that ⟨𝛼, 𝛽⟩ = 0 =⇒ 𝛼 ⊥ 𝛽.
Thus, 𝐽 𝑓 ∈ 𝑂 (2) and also det(𝐽 𝑓 ) = (𝑢 2𝑥 + 𝑢 2𝑦 ) > 0. So, ∃𝑇 ∈ 𝑆𝑂 (2) such that
√︃
𝐽𝑓 = 𝑢 2𝑥 + 𝑢 2𝑦 𝑇 .
√︃
So, 𝐽 𝑓 is then a composition of rotations and dilations by 𝑢 2𝑥 + 𝑢 2𝑦 and thus preserves
angles. Thus, 𝑓 = 𝑢 + 𝑖𝑣 is conformal and the proof is complete. ♦
Chapter II
Complex Integration

Problem II.1 Let 𝛾∫:= [1, 𝑖] and


∫ 𝜎 := [1, 1 + 𝑖, 𝑖] be two polygons. Express them as
paths and calculate 𝑓 and 𝑓 for 𝑓 (𝑧) = |𝑧| 2 .
𝛾 𝜎

Solution II.1 We recall that we denote the line segment from 𝑧 to 𝑤 by

[𝑧, 𝑤] := {𝑡𝑤 + (1 − 𝑡)𝑧 : 𝑡 ∈ [0, 1]}.

Also, [𝑎, 𝑧1 , . . . , 𝑧 𝑛 , 𝑏] denotes the polygon from 𝑎 to 𝑏 and is defined by the union
of line segments as

[𝑎, 𝑧1 , . . . , 𝑧 𝑛 , 𝑏] := [𝑎, 𝑧1 ] ∪ [𝑧 1 , 𝑧2 ] ∪ · · · ∪ [𝑧 𝑛 , 𝑏].

By these definitions, we have for 0 ≤ 𝑡 ≤ 1,

𝛾 = [1, 𝑖] = {𝑡𝑖 + (1 − 𝑡)1} ;


𝜎 = [1, 1 + 𝑖, 𝑖] = [1, 1 + 𝑖] ∪ [1 + 𝑖, 𝑖]
= {𝑡 (𝑖 + 1) + (1 − 𝑡)1} ∪ {𝑡𝑖 + (1 − 𝑡) (1 + 𝑖)}
= {1 + 𝑖𝑡} ∪ {𝑖 + (1 − 𝑡)}
= 𝜎1 ∪ 𝜎2 ;

where we let 𝜎1 := {𝑡 : 1 + 𝑖𝑡}, 𝜎2 := {𝑡 : 𝑖 + (1 − 𝑡)} with 𝑡 ∈ [0, 2𝜋]. Now, by


definition of line integrals,
∫ ∫ 1
𝑓 := ( 𝑓 ◦ 𝛾)𝛾 ′ (𝑡) 𝑑𝑡.
𝛾 0

7
8 II Complex Integration
∫ ∫ 1
𝑡 2 + (1 − 𝑡) 2 𝑑𝑡
 
∴ 𝑓 = (𝑖 − 1)
𝛾 0
  1
2 2
= (𝑖 − 1) 𝑡 3 − 𝑡 2 + 𝑡 = (𝑖 − 1).
3 3
0
∫ ∫ ∫ ∫
We compute 𝜎 𝑓 similarly by recalling that 𝜎1 ∪𝜎2
𝑓 = 𝜎1
𝑓+ 𝜎2
𝑓 since 𝜎(𝑡) is
of bounded variation.
∫ ∫ ∫
∴ 𝑓 = 𝑓+ 𝑓
𝜎 𝜎1 𝜎2
∫ 1 ∫ 1
2
1 + (1 − 𝑡) 2 𝑑𝑡
   
=𝑖 1+𝑡 𝑑𝑡 −
0 0
1 1
𝑡3 𝑡3
 
4
=𝑖 𝑡+
3
− 2𝑡 − 𝑡 2 +
3
=
3
(𝑖 − 1). ♦
0 0

Problem II.2 Give the power series expansion of log 𝑧 about 𝑧 = 𝑖 and find its radius
of convergence.
Solution II.2 We recall that the power series expansion of a function analytic at 𝑧 0
is given by

∑︁ 𝑓 (𝑛) (𝑧0 )
𝑓 (𝑧) = (𝑧 − 𝑧 0 ) 𝑛 ;
𝑛=0
𝑛!

In this problem, 𝑓 (𝑧) = log 𝑧 and 𝑧 0 = 𝑖. We calculate the derivatives of the complex
logarithm as follows:
1
𝑓 ′ (𝑧) = ;
𝑧
1
𝑓 ′′ (𝑧) = − 2 ;
𝑧
2
𝑓 ′′′ (𝑧) = 3 ;
𝑧
(4) 6
𝑓 (𝑧) = − 4 ;
𝑧
..
.
(𝑛 − 1)!
𝑓 (𝑛) (𝑧) = (−1) 𝑛−1 ;
𝑧𝑛
(𝑛 − 1)!
∴ 𝑓 (𝑛) (𝑖) = (−1) 𝑛−1 .
𝑖𝑛

Substituting this in the formula for the power series expansion, for 𝑘 ∈ Z, we obtain
II Complex Integration 9

∑︁ (−1) 𝑛−1 (𝑛 − 1)!
log 𝑧 = log(𝑖) + 𝑛
(𝑧 − 𝑖) 𝑛
𝑛=1
𝑖 𝑛!

𝑖𝜋 ∑︁ (−1) 𝑛−1
= 2𝜋𝑖𝑘 + + (𝑧 − 𝑖) 𝑛 .
2 𝑛=1 𝑛 𝑖 𝑛

To find the Radius of Convergence, 𝑅, we recall the formula

𝑎𝑛
𝑅 := lim ,
𝑛→∞ 𝑎 𝑛+1

(−1) 𝑛−1
where 𝑎 𝑛 = . Thus,
𝑛 𝑖𝑛
(−1) 𝑛−1 (𝑛 + 1) 𝑖 𝑛+1
𝑅 = lim ·
𝑛→∞ 𝑛 𝑖𝑛 (−1) 𝑛
𝑛+1
= lim (−𝑖) ·
𝑛→∞ 𝑛
𝑛+1
= lim
𝑛→∞ 𝑛
= 1. ♦


sin 𝑧
Problem II.3 Evaluate 𝑑𝑧, where 𝛾(𝑡) = e𝑖𝑡 and 𝑡 ∈ [0, 2𝜋].
𝛾 𝑧3

Solution II.3 Let 𝑓 (𝑧) = sin 𝑧. We recall that sin 𝑧 is analytic on the whole plane C.
¯ 0) ⊂ C and 𝑓 ′′ (𝑧) = − sin 𝑧. Thus,
Now, it is obvious that 𝐵(1;

2! 𝑓 (𝑧)
𝑓 ′′ (0) = 𝑑𝑧
2𝜋𝑖 𝛾 (𝑧 − 0) 2+1

sin 𝑧

𝛾 𝑧
3
𝑑𝑧 = −𝜋𝑖 · sin 0 = 0. ♦

 
1 1 + 𝑖𝑧
Problem II.4 Let 𝑓 (𝑧) = log ; Show that
2𝑖 1 − 𝑖𝑧

∑︁ 𝑧2𝑘+1
𝑓 (𝑧) = (−1) 𝑘
𝑘=0
2𝑘 + 1

for |𝑧| < 1. Note that tan 𝑓 (𝑧) = 𝑧 (i.e. 𝑓 is a branch of arctan 𝑧).

Solution II.4 We consider the principal branch of the logarithm. Then we find the
power series expansion of 𝑓 at 𝑧 = 0 as follows:
10 II Complex Integration

1
𝑓 (𝑧) = [log(1 + 𝑖𝑧) − log(1 − 𝑖𝑧)]
2𝑖
1 h i
= log(𝑖) + log(𝑧 − 𝑖) − log(−𝑖) − log(𝑧 + 𝑖) ;
2𝑖
′ 1 1 − 𝑖𝑧 𝑖(1 − 𝑖𝑧) + 𝑖(1 + 𝑖𝑧) 1
𝑓 (𝑧) = · · 2
= ;
2𝑖 1 + 𝑖𝑧 (1 − 𝑖𝑧) 1 + 𝑧2
2𝑧
𝑓 ′′ (𝑧) = − 2 ;
1 + 𝑧2
2(3𝑧 2 − 1)
𝑓 ′′′ (𝑧) = 3 ;
1 + 𝑧2
..
.
 
(−1) 𝑛−1 1 1
𝑓 (𝑛) (𝑧) = (𝑛 − 1)! − ;
2𝑖 (𝑧 − 𝑖) 𝑛 (𝑧 + 𝑖) 𝑛

Substituting 𝑧 = 0 in the last equation, we obtain

 (𝑛 − 1)! ,
  
(−1) 𝑛−1 1 1 if 𝑛 is odd;


(𝑛)
𝑓 (0) = (𝑛 − 1)! − = 𝑖 𝑛+1
2𝑖 (−𝑖) 𝑛 𝑖 𝑛 
 0, if 𝑛 is even;

Now, we recall that if 𝑛 is odd then 𝑛 + 1 is even and thus 𝑖 𝑛+1 is always ±1. Thus,
for |𝑧| < 1, we obtain
∑︁ (𝑘 − 1)!
𝑓 (𝑧) = 𝑘+1 𝑘!
𝑧𝑘
𝑘 is odd
𝑖
∑︁ 𝑧𝑘
=
𝑘 is odd
𝑖 𝑘+1 · 𝑘

𝑧 2𝑘+1

∑︁
∴ 𝑓 (𝑧) = (−1) 𝑘 .
𝑘=0
2𝑘 + 1

Problem II.5 Let 𝐺 be a region and 𝑓 : 𝐺 → C is an analytic function such that


there is a point 𝛼 ∈ 𝐺 with | 𝑓 (𝛼)| ≤ | 𝑓 (𝑧)| for all 𝑧 ∈ 𝐺. Show that either 𝑓 is
constant or 𝑓 (𝛼) = 0.

Solution II.5 Let 𝑓 be an analytic function in 𝐺 satisfying the above hypotheses and
assume that 𝑓 (𝛼) ≠ 0 for each 𝛼 ∈ 𝐺. It is enough to show that this implies 𝑓 is
constant in 𝐺.
1
Now, consider the function 𝑔 : 𝐺 → C defined by 𝑔(𝑧) = . Since 𝑓 is analytic
𝑓 (𝑧)
and non-zero everywhere in 𝐺, it follows that 𝑔 is analytic in 𝐺. Next, we notice that
for any 𝛼 ∈ 𝐺,
II Complex Integration 11

1
|𝑔(𝑧)| =
𝑓 (𝑧)
1
=
| 𝑓 (𝑧)|
1
≤ = |𝑔(𝛼)|.
| 𝑓 (𝛼)|

Thus, |𝑔(𝑧)| ≤ |𝑔(𝛼)|, ∀𝛼 ∈ 𝐺. But by the Maximum Modulus Principle, this means
that 𝑔 and hence 𝑓 is a constant and we are done. ♦

Problem II.6 Let 𝑓 be an entire function and suppose there is a constant 𝑀, an


𝑅 > 0, and an integer 𝑛 ≥ 1 such that | 𝑓 (𝑧)| ≤ 𝑀 |𝑧| 𝑛 for |𝑧| > 𝑅. Show that 𝑓 is a
polynomial of degree ≤ 𝑛.

Solution II.6 Since 𝑓 is entire, it has a power series around 𝑧 = 0,



∑︁ 𝑓 (𝑘 ) (0) 𝑘
𝑓 (𝑧) = 𝑧 .
𝑘=0
𝑘!

Now, for some 𝑟 > 0, since 𝐵(0; 𝑟) ∈ C, we can apply Corollary – IV.2.13 to get

𝑘! 𝑓 (𝑧)
𝑓 (𝑘 ) (0) = 𝑑𝑧 ,
2𝜋𝑖 𝛾 𝑧 𝑘+1

where 𝛾(𝑡) = 𝑟e𝑖𝑡 for 𝑡 ∈ [0, 2𝜋]. Thus, it is enough to show that 𝑓 (𝑘 ) (0) = 0 for
all 𝑘 ≥ 𝑛 + 1. We note that

𝑘! 𝑓 (𝑧)
𝑓 (𝑘 ) (0) = 𝑑𝑧
2𝜋 𝛾 𝑧 𝑘+1

𝑘! | 𝑓 (𝑧)|
≤ |𝑑𝑧|
2𝜋 𝛾 |𝑧| 𝑘+1

𝑀 𝑘! |𝑧| 𝑛
≤ |𝑑𝑧|
2𝜋 𝛾 |𝑧| 𝑘+1

𝑀 𝑘!
≤ |𝑧| 𝑛−𝑘−1 |𝑑𝑧|
2𝜋 𝛾

𝑀 𝑘! 𝑛−𝑘−1
= ·𝑟 |𝑑𝑧|
2𝜋 𝛾
𝑀 𝑘! 𝑛−𝑘−1 𝑖𝑡 2 𝜋
= ·𝑟 𝑟e 0
2𝜋
𝑀 𝑘! 𝑛−𝑘
= ·𝑟 · 2𝜋 .
2𝜋
12 II Complex Integration

Thus, 𝑓 (𝑘 ) (0) = 𝑀 𝑘! 𝑟 𝑛−𝑘 , which implies that 𝑓 (𝑘 ) (0) → 0 for 𝑘 ≥ 𝑛 + 1 since


𝑟 𝑛−𝑘 → 0. So, 𝑓 (𝑘 ) (0) = 0 for all 𝑘 ≥ 𝑛 + 1 and we are done. ♦

Problem II.7 Give an elementary proof of the Maximum Modulus Principle for
polynomials.

Solution II.7 Since all polynomials are entire, by Liouville’s Theorem, if 𝑓 (𝑧) is a
polynomial satisfying | 𝑓 (𝑧)| ≤ | 𝑓 (𝑎)| for all 𝑧 ∈ C then 𝑓 is constant in C . ♦

Problem II.8 Show that Cauchy’s Integral Formula follows from Cauchy’s Theorem.

Solution II.8 Cauchy’s Theorem states that if 𝛾 ≈ 0 (homologous) then 𝛾 𝑓 = 0
for any analytic function 𝑓 defined in an open connected set 𝐺 ⊆ C. To derive the
Integral Formula from Cauchy’s Theorem, we consider the function Φ : 𝐺 → C
defined by

 𝑓 (𝑧) − 𝑓 (𝑎) ,

for 𝑧 ≠ 𝑎;


Φ(𝑧) = 𝑧−𝑎


 𝑓 ′ (𝑎), for 𝑧 = 𝑎;

where 𝑎 ∉ {𝛾}. If we can show this function Φ(𝑧) is analytic in 𝐺, then we have that
∫ ∫
𝑓 (𝑧) − 𝑓 (𝑎)
0= Φ= 𝑑𝑧
𝛾 𝛾 𝑧−𝑎
∫ ∫
𝑓 (𝑧) 𝑓 (𝑎)
= 𝑑𝑧 − 𝑑𝑧
𝛾 𝑧−𝑎 𝛾 𝑧−𝑎
∫ ∫
𝑓 (𝑧) 𝑑𝑧
= 𝑑𝑧 − 𝑓 (𝑎)
𝛾 𝑧−𝑎 𝛾 𝑧−𝑎

𝑓 (𝑧)
= 𝑑𝑧 − 𝑛(𝛾; 𝑎) 2𝜋𝑖 𝑓 (𝑎)
𝛾 𝑧−𝑎

1 𝑓 (𝑧)
∴ 𝑛(𝛾; 𝑎) 𝑓 (𝑎) = 𝑑𝑧 ,
2𝜋𝑖 𝛾 𝑧 − 𝑎

which is the Cauchy Integral Formula. So, the proof will be complete if we can show
that Φ(𝑧), defined as above, is an analytic function in 𝐺.
If 𝑧 ≠ 𝑎, then Φ(𝑧) is clearly analytic (hence also continuous). Next, we note that
 
𝑓 (𝑧) − 𝑓 (𝑎)
lim [Φ(𝑧)] = lim = 𝑓 ′ (𝑎),
𝑧→𝑎 𝑧→𝑎 𝑧−𝑎

which shows that Φ is continuous at 𝑧 = 𝑎. For showing Φ is differentiable at 𝑧 = 𝑎,


we recall that since 𝑓 is analytic in 𝐺, this implies that 𝑓 is infinitely differentiable
in 𝐺. In particular, 𝑓 ′′ exists and is also analytic in 𝐺. Now,
II Complex Integration 13
 
Φ(𝑧) − Φ(𝑎)
Φ′ (𝑎) = lim
𝑧→𝑎 𝑧−𝑎
" 𝑓 (𝑧) − 𝑓 (𝑎) #
𝑧−𝑎 − 𝑓 ′ (𝑎)
= lim
𝑧→𝑎 𝑧−𝑎
𝑓 (𝑧) − 𝑓 (𝑎) − 𝑓 ′ (𝑎) (𝑧 − 𝑎)
 
= lim
𝑧→𝑎 (𝑧 − 𝑎) 2
′′
= 𝑓 (𝑎)/2,

where the last equality follows from the fact that the analyticity of 𝑓 implies a power
series expansion of 𝑓 around 𝑧 = 𝑎. So, we have Φ′ (𝑎) = 𝑓 ′′ (𝑎)/2, implying that Φ
is differentiable at 𝑧 = 𝑎 and the proof is complete. ♦
Chapter III
Singularities

Problem III.1 The following functions have an isolated singularity at 𝑧 = 0. De-


termine its nature; if it is a removable singularity define 𝑓 (0) so that is analytic at
𝑧 = 0; if it is a pole find the singular part.
sin 𝑧
Solution III.1 Part-(a): 𝑓 (𝑧) = . Using the power series for sin 𝑧 around 𝑧0 = 0,
𝑧
we obtain the following Laurent Series:

sin 𝑧 𝑧2 𝑧4 𝑧6
𝑓 (𝑧) = =1− + − +··· ;
𝑧 3! 5! 7!
Since 𝑎 𝑛 = 0 for all 𝑛 ≤ −1, by Corollary – 1.18, we know that 𝑓 has a removable
singularity at 𝑧 = 0. Next, we define a function 𝑔 : C → C that agrees with 𝑓 in
some 𝐵(0; 𝑅) but is also analytic at 𝑧 = 0. We note that 𝑔 defined as
(
𝑓 (𝑧), 𝑧 ≠ 0;
𝑔(𝑧) =
1, 𝑧 = 0.

is analytic at 𝑧 = 0 and we are done. ♦


log(1 + 𝑧)
Part-(e): 𝑓 (𝑧) = . Using the power series for log(1 + 𝑧) around 𝑧0 = 0,
𝑧2
we obtain the following Laurent Series:

log(1 + 𝑧) 1 1 𝑧 𝑧 2
𝑓 (𝑧) = = − + − + · · · + 2𝜋𝑖𝑘,
𝑧2 𝑧 2 3 4
where 𝑘 ∈ Z. Since 𝑎 𝑛 = 0 for all 𝑛 ≤ −2, by Corollary – 1.18, we know that 𝑓 has
a pole of order 1 at 𝑧 = 0. Next, we compute the singular part S 𝑓 (𝑧) of 𝑓 . From the
above Laurent Series, we have S 𝑓 (𝑧) = 𝑧 −1 . ♦

15
16 III Singularities
 
1
Problem III.2 Find the type of singularity of 𝑓 (𝑧) = exp at 𝑧 = 0. Determine
𝑧
𝑓 ({𝑧 : 0 < |𝑧| < 𝛿}) for arbitrarily small values of 𝛿.

Solution III.2 Using the power  series


 of exp(𝑧) around 𝑧0 = 0, we obtain the
1
following Laurent series for exp :
𝑧

𝑧 −𝑘
  ∑︁
1 1 1
𝑓 (𝑧) = exp = = 1+ + 2 +··· ;
𝑧 𝑘=0
𝑘! 𝑧 2𝑧

Since 𝑎 𝑛 ≠ 0 for infinitely many 𝑛 ≤ −1, by Corollary – 1.18, we know that 𝑓 has
an essential singularity at 𝑧 = 0.
Next, we find 𝑓 ({𝑧 : 0 < |𝑧| < 𝛿}) for arbitrarily small values of 𝛿. Let 𝐴 := {𝑧 :
0 < |𝑧| < 𝛿}. Being motivated by the Casorati-Weierstrass Theorem, we claim that

𝑓 ({𝑧 : 0 < |𝑧| < 𝛿}) = C \{0}.

We now proceed to prove our claim. Let 𝜔, 𝑧 ∈ C such that 𝜔 = e1/𝑧 and 0 < |𝑧| < 𝛿.
If 1/𝑧 = 𝑥 + 𝑖𝑦 (𝑥, 𝑦 ∈ R), then we have |𝜔| = e 𝑥 |e𝑖𝑦 |. This implies that |𝜔| > 0 and
hence 0 ∉ 𝐴. So, 𝐴 ⊆ C \{0}.
Next, we show that any nonzero 𝜔 ∈ C can be expressed as 𝜔 = e1/𝑧 . This will then
imply that C \{0} ⊆ 𝐴. We notice that we can express 𝜔 in polar form as follows:

𝜔 = 𝑟e𝑖 𝜃 = eln 𝑟 e𝑖 𝜃 = e𝑖 𝜃+ln 𝑟 ,

where 𝑟, 𝜃 ∈ R. Now, if we choose a positive integer 𝑘 ∈ Z+ such that


1
| ln 𝑟 + 𝑖𝜃 + 2𝜋𝑖𝑘 | < ,
𝛿
and then set 𝑧 = ln 𝑟 + 𝑖𝜃 + 2𝜋𝑖𝑘, we have 𝜔 = e1/𝑧 for 0 < |𝑧| < 𝛿. Thus, 𝜔 ∈ 𝐴 and
we must have 𝐴 = C \{0} and we are done. ♦

∫ 𝜋
𝑑𝜃
Problem III.3 Compute .
0 (𝑎 + cos 𝜃) 2

Solution III.3 We observe that if 𝑧 = e𝑖 𝜃 , then 𝑧¯ = 1/𝑧. Using this, we obtain


 2
2 𝑧 + 𝑧¯
(𝑎 + cos 𝜃) = 𝑎 +
2
 2
1 1
= 2𝑎 + 𝑧 +
4 𝑧
2
2
𝑧 + 2𝑎𝑧 + 1
= .
4𝑧2
III Singularities 17

We also notice that since the integrand is symmetric about 𝜃 = 𝜋,


∫ 𝜋 ∫ 2𝜋
𝑑𝜃 1 𝑑𝜃
2
=
0 (𝑎 + cos 𝜃) 2 0 (𝑎 + cos 𝜃) 2

−2𝑖𝑧 𝑑𝑧
= 2 ,
𝛾 𝑧2 + 2𝑎𝑧 + 1
√ √
where 𝛾 is the circle |𝑧| = 1. Moreover, 𝛼 := −𝑎 + 𝑎 2 − 1 and 𝛽 := −𝑎 − 𝑎 2 − 1
are each poles of order 2 of the above integrand. Since 𝑎 > 1, this gives |𝛼| < 1
and |𝛽| > 1. Hence 𝑛(𝛾; 𝛼) = 1 and 𝑛(𝛾; 𝛽) = 0. Next, we compute Res( 𝑓 ; 𝛼). By
Proposition – V.2.4, since 𝛼 is a pole of order 2, if 𝑔(𝑧) = (𝑧 − 𝛼) 2 𝑓 (𝑧), we have

𝑔 ′ (𝛼) (𝛼 − 𝛽) 2 − 2𝛼 2(𝑎 2 − 1) + 𝑎 − 𝑎 2 − 1
Res( 𝑓 ; 𝛼) = = = .
(2 − 1)! (𝛼 − 𝛽) 4 8(𝑎 2 − 1) 2
By the Residue Theorem, we have

2(𝑎 2 − 1) + 𝑎 − 𝑎 2 − 1

−2𝑖𝑧 𝑑𝑧
 2 = (𝜋𝑖) · .
𝛾 𝑧2 + 2𝑎𝑧 + 1 4(𝑎 2 − 1) 2

Combining this with the previous result, we finally obtain



2(𝑎 2 − 1) + 𝑎 − 𝑎 2 − 1
∫ 𝜋

0
𝑑𝜃
(𝑎 + cos 𝜃) 2
=𝜋·
2(𝑎 2 − 1) 2
. ♦


Problem III.4 Find all possible values of exp(1/𝑧) 𝑑𝑧 for any closed curve 𝛾 not
𝛾
passing through 𝑧 = 0.

Solution III.4 From Problem – III.2, we know that the Laurent series of exp(1/𝑧)
about 𝑧 = 0 is given by

𝑧 −𝑘
  ∑︁
1 1 1
𝑓 (𝑧) = exp = = 1+ + 2 +··· ;
𝑧 𝑘=0
𝑘! 𝑧 2𝑧

Thus, Res( 𝑓 ; 0) = 𝑎 −1 = 1. Using the Residue Theorem, we obtain



exp(1/𝑧) 𝑑𝑧 = 𝑛(𝛾; 0).
𝛾

18 III Singularities

Problem III.5 Suppose that 𝑓 has a simple pole at 𝑧 = 𝑎 and let 𝑔 be analytic in an
open set containing 𝑎. Show that Res( 𝑓 𝑔; 𝑎) = 𝑔(𝑎) Res( 𝑓 ; 𝑎).
Solution III.5 Let 𝑓 , 𝑔 be functions as above.
First assume 𝑔 has a simple zero at 𝑧 = 𝑎 i.e. 𝑔(𝑎) = 0. Then 𝑓 𝑔 has a removable
singularity at 𝑧 = 𝑎. But this implies that Res( 𝑓 𝑔; 𝑎) = 0. We also have that
𝑔(𝑎) Res( 𝑓 ; 𝑎) = 0, since 𝑔(𝑎) = 0. Hence, the identity is true for this case.
Next, assume that 𝑔(𝑎) ≠ 0. This implies that 𝑓 𝑔 has a simple pole at 𝑧 = 𝑎 and we
have

Res( 𝑓 𝑔; 𝑎) = lim [(𝑧 − 𝑎) ( 𝑓 𝑔)] = lim [(𝑧 − 𝑎) 𝑓 (𝑧)] · lim [𝑔(𝑧)] = 𝑔(𝑎) Res( 𝑓 ; 𝑎).
𝑧→𝑎 𝑧→𝑎 𝑧→𝑎

So, we are done. ♦

Problem III.6 Use Problem – III.5 to show that if 𝐺 is a region and 𝑓 is analytic in
𝐺 except for simple poles at 𝑎 1 , . . . , 𝑎 𝑛 ; and if 𝑔 is analytic in 𝐺 then
∫ 𝑛
1 ∑︁
𝑓𝑔 = 𝑛(𝛾; 𝑎 𝑘 ) 𝑔(𝑎 𝑘 ) Res( 𝑓 ; 𝑎 𝑘 ).
2𝜋𝑖 𝛾 𝑘=1

Solution III.6 Since all the 𝑎 𝑖 ’s are simple poles of 𝑓 , using the Residue Theorem
and the formula from the previous Exercise, we have
∫ 𝑛
1 ∑︁
𝑓𝑔 = 𝑛(𝛾; 𝑎 𝑘 ) Res( 𝑓 𝑔; 𝑎 𝑘 )
2𝜋𝑖 𝛾 𝑘=1
𝑛
∑︁
= 𝑛(𝛾; 𝑎 𝑘 ) 𝑔(𝑎 𝑘 ) Res( 𝑓 ; 𝑎 𝑘 ).
𝑘=1

So, we are done. ♦

Problem III.7 Let 𝛾 be the rectangle 𝑛 + 12 + 𝑛𝑖, −𝑛 − 21 + 𝑛𝑖, −𝑛 − 21 − 𝑛𝑖, 𝑛 + 12 − 𝑛𝑖, 𝑛 + 1


 
2 + 𝑛𝑖 .
Evaluate 𝛾 𝜋 𝑧cot(
∫ 𝜋 𝑧)
2 −𝑎 2 𝑑𝑧 and show that the limit goes to zero as 𝑛 → ∞. Consequently,
show that

1 ∑︁ 2𝑎
𝜋 cot(𝜋𝑎) = + .
𝑎 𝑛=1 𝑎 2 − 𝑛2
Solution III.7 The above integrand has simple poles at 𝑧 = 𝑘 for 𝑘 ∈ Z, and it has
two additional simple poles at 𝑧 = ±𝑎. Using the Residue Theorem, we then have
∫ 𝑛
1 ∑︁
𝑓 = Res( 𝑓 ; 𝑎) + Res( 𝑓 ; −𝑎) + Res( 𝑓 ; 𝑘),
2𝜋𝑖 𝛾 𝑘=−𝑛

𝜋 cot(𝜋𝑧)
where 𝑓 (𝑧) := . Now, by definition,
𝑧2 − 𝑎2
III Singularities 19
 
𝜋 cot(𝜋𝑧) 1
Res( 𝑓 ; 𝑘) = lim (𝑧 − 𝑘) 2 2
= 2 .
𝑧→𝑘 𝑧 −𝑎 𝑘 − 𝑎2

Also,  
𝜋 cot(𝜋𝑧) 𝜋 cot(𝜋𝑎)
Res( 𝑓 ; ±𝑎) = lim (𝑧 ± 𝑎) · = ,
𝑧→±𝑎 𝑧2 − 𝑎2 2𝑎

for both cases. Next, we compute lim 𝑓 . But as 𝑛 → ∞, | cot(𝜋𝑧)| ≤ 2 and
𝑛→∞ 𝛾
1
→ 0 for 𝑧 ∈ {𝛾}. Thus,
𝑧2 − 𝑎2

𝜋 cot(𝜋𝑧)
lim 𝑑𝑧 = 0.
𝑛→∞ 𝛾 𝑧2 − 𝑎2

Combining this fact with our previous results, we obtain



𝜋 cot(𝜋𝑎) ∑︁ 1
=−
𝑎 𝑘=−∞
𝑘 − 𝑎2
2


∑︁ 1
= 2 − 𝑘2
𝑘=−∞
𝑎

1 ∑︁ 2
= 2
+
𝑎 𝑘=1
𝑎 − 𝑘2
2


1 ∑︁ 2𝑎
=⇒ 𝜋 cot(𝜋𝑎) = +
𝑎 𝑘=1 𝑎 2 − 𝑘 2
. ♦

Problem III.8 Let 𝑓 be an entire function and let ∫ 𝑎, 𝑏 ∈ C such that |𝑎| < 𝑅 and
|𝑎| < 𝑅. If 𝛾(𝑡) = 𝑅e𝑖𝑡 for 𝑡 ∈ [0, 2𝜋], evaluate 𝛾 [(𝑧 − 𝑎) (𝑧 − 𝑏)] −1 𝑓 (𝑧) 𝑑𝑧. Use
this result to give another proof of Liouville’s Theorem.

Solution III.8 Let 𝑓 be a bounded entire function, say for 𝑀 > 0, | 𝑓 (𝑧)| ≤ 𝑀 for
1
each 𝑧 ∈ C. Now, since 𝑔(𝑧) = has two simple poles at 𝑧 = 𝑎, 𝑏 and
(𝑧 − 𝑎) (𝑧 − 𝑏)
𝑓 is analytic everywhere, using Problem – III.6, we obtain

1
𝑓 𝑔 = 𝑓 (𝑎) Res(𝑔; 𝑎) + 𝑓 (𝑏) Res(𝑔; 𝑏)
2𝜋𝑖 𝛾
𝑓 (𝑎) 𝑓 (𝑏)
= +
𝑎−𝑏 𝑏−𝑎
𝑓 (𝑎) − 𝑓 (𝑏)
= .
𝑎−𝑏
Next, using the fact that 𝑓 is bounded and 𝛾 is a circle, we have
20 III Singularities
∫ ∫ 2𝜋
1 1 𝑓 (𝑅e𝑖𝑡 ) 𝑑𝑡
𝑓𝑔 = 𝑖𝑅e𝑖𝑡 ·
2𝜋𝑖 𝛾 2𝜋 0 (𝑅e𝑖𝑡 − 𝑎) (𝑅e𝑖𝑡 − 𝑏)
∫ 2𝜋
1 | 𝑓 (𝑅e𝑖𝑡 )| 𝑑𝑡
≤ 𝑖𝑅e𝑖𝑡 ·
2𝜋 0 |𝑅e𝑖𝑡 − 𝑎||𝑅e𝑖𝑡 − 𝑏|
∫ 2𝜋
𝑅 𝑀 𝑑𝑡

2𝜋 0 (𝑅 − |𝑎|) (𝑅 − |𝑏|)
𝑀 𝑅(2𝜋 − 0)
=
2𝜋(𝑅 − |𝑎|) (𝑅 − |𝑏|)
𝑀𝑅
= .
(𝑅 − |𝑎|) (𝑅 − |𝑏|)

1
Letting 𝑅 → ∞, we have 𝑓 𝑔 → 0. Combining this with our evaluated
∫ 2𝜋𝑖 𝛾
expression for 2 1𝜋𝑖 𝛾 𝑓 𝑔, we must have that for sufficiently large 𝑅,

𝑓 (𝑎) − 𝑓 (𝑏)
= 0,
𝑎−𝑏

which implies 𝑓 (𝑎) = 𝑓 (𝑏). But since 𝑎, 𝑏, 𝑅 were all arbitrary, we have that 𝑓 is a
constant function. ♦


𝑥 2 𝑑𝑥

Problem III.9 Compute .
0 𝑥4 + 𝑥2 + 1

𝑧2
Solution III.9 We consider the function 𝑓 (𝑧) = 4 . We observe that this
𝑧 + 𝑧2 + 1
1 √ 1 √ 1 √ 1 √
function has simple poles at 𝑧 = (1 + 3𝑖), (−1 + 3𝑖), (−1 − 3𝑖), (1 − 3𝑖).
2 2 2 2
We call them 𝑎 1 , . . . , 𝑎 4 respectively. Since each 𝑎 𝑖 is a simple pole, we have

𝑎 2𝑖
Res( 𝑓 ; 𝑎 𝑖 ) = Î ,
(𝑎 𝑖 − 𝑎 𝑗 )
𝑖≠ 𝑗

for 𝑖, 𝑗 = 1, 2, 3, 4. Here, we consider the curve 𝛾 := {𝑅e𝑖𝑡 : 𝑡 ∈ [0, 𝜋]} ∪ [−𝑅, 𝑅]


for some 𝑅 > 2. Since the region enclosed by 𝛾 only contains the poles 𝑎 1 , 𝑎 2 then
winding number 𝑛(𝛾; 𝑎 3 ) = 0 = 𝑛(𝛾; 𝑎 4 ). Combining these results using the Residue
Theorem, we obtain

1 −𝑖
𝑓 = Res( 𝑓 ; 𝑎 1 ) + Res( 𝑓 ; 𝑎 2 ) = √ .
2𝜋𝑖 𝛾 2 3
But using the definition of line integrals, we also have
III Singularities 21

𝑥 2 𝑑𝑥 𝑅 2 e2𝑖𝑡 𝑑𝑡
∫ ∫ 𝑅 ∫ 𝜋 
𝑓 = + 𝑖𝑅e𝑖𝑡 · .
𝛾 −𝑅 𝑥 + 𝑥2 + 1
4
0 𝑅 4 e4𝑖𝑡+ 𝑅 2 e2𝑖𝑡 + 1

This implies that

𝑥 2 𝑑𝑥 𝑖𝑅 3 e3𝑖𝑡 𝑑𝑡
∫ 𝑅 ∫ ∫ 𝜋
= 𝑓− .
−𝑅 𝑥4 + 𝑥2 + 1 𝛾 0 𝑅 4 e4𝑖𝑡 + 𝑅 2 e2𝑖𝑡 + 1
∫ 𝜋
From our previous calculation, we have 𝑓 = √ . Moreover, we notice that
𝛾
3

𝑖𝑅 3 e3𝑖𝑡 𝑑𝑡 𝑖𝑅 3 e3𝑖𝑡 𝑑𝑡
∫ 𝜋 ∫ 𝜋

0 𝑅 4 e4𝑖𝑡 + 𝑅 2 e2𝑖𝑡 + 1 0 𝑅 4 e4𝑖𝑡 + 𝑅 2 e2𝑖𝑡 + 1
𝑅 3 |e3𝑖𝑡 | 𝑑𝑡
∫ 𝜋
=
0 |𝑅 4 e4𝑖𝑡 + 𝑅 2 e2𝑖𝑡 + 1|
𝑅 3 𝑑𝑡
∫ 𝜋

0 𝑅 4 e4𝑖𝑡 + 𝑅 2 e2𝑖𝑡 − 1
𝑅 3 𝑑𝑡
∫ 𝜋

0 𝑅4 − 𝑅2 − 1
𝜋𝑅 3
= 4 .
𝑅 − 𝑅2 − 1
Letting 𝑅 → ∞, we finally obtain that

𝑖𝑅 3 e3𝑖𝑡 𝑑𝑡
∫ 𝜋
→ 0.
0 𝑅 4 e4𝑖𝑡 + 𝑅 2 e2𝑖𝑡 + 1

Thus, we have
𝑥 2 𝑑𝑥
∫ 𝑅
𝜋
lim =√ .
𝑅→∞ −𝑅 𝑥4 + 𝑥2 + 1 3
𝑥2
But since is an even function, we obtain
𝑥4 + 𝑥2 + 1
∫ ∞
𝑥 2 𝑑𝑥 1 ∞ 𝑥 2 𝑑𝑥

0 𝑥 4 + 𝑥2 + 1
=
2 −∞ 𝑥 4 + 𝑥2 + 1
𝜋
= √ .
2 3

22 III Singularities

Problem III.10 State and prove a more general version of Rouche’s Theorem for
curves other than circles in 𝐺.

Solution III.10 We state the following general version of Rouche’s Theorem:

Suppose 𝑓 and 𝑔 are meromorphic in a region 𝐺 with no zeros or poles on


𝑓 𝑓 𝑔 𝑔
a closed rectifiable curve 𝛾 with 𝛾 ∼ 0(𝐺). If 𝑧 𝑘 , 𝑎 𝑘 (𝑧 𝑘 , 𝑎 𝑘 ) are respectively
the zeros and poles of 𝑓 (or 𝑔), and 𝑍 𝑓 , 𝑍 𝑔 (𝑃 𝑓 , 𝑃𝑔 ) are the number of zeros
(poles) of 𝑓 and 𝑔 inside 𝛾 counted according to their multiplicities and if

| 𝑓 (𝑧) + 𝑔(𝑧)| < | 𝑓 (𝑧)| + |𝑔(𝑧)|

on 𝛾, then
𝑍𝑓
∑︁  𝑃𝑓
 ∑︁  𝑍𝑔
 ∑︁  𝑃𝑔
 ∑︁  
𝑓 𝑓 𝑔 𝑔
𝑛 𝛾; 𝑧 𝑘 − 𝑛 𝛾; 𝑎 𝑘 = 𝑛 𝛾; 𝑧 𝑘 − 𝑛 𝛾; 𝑎 𝑘 .
𝑘=1 𝑘=1 𝑘=1 𝑘=1

We now proceed with the proof of our generalization noting most of the original
arguments need no change.
From the hypothesis,
𝑓 𝑓
+1 < +1
𝑔 𝑔
on 𝛾. If 𝜆 = 𝑓 /𝑔 and 𝜆 is a positive real number, then we have 𝜆 + 1 < 𝜆 + 1, which
is false. Hence, the meromorphic function 𝑓 /𝑔 maps 𝛾 into Ω = C \[0, ∞]. We can
thus define an analytic branch of the complex logarithm in Ω, say, ℓ(𝑧). This implies
that ℓ( 𝑓 (𝑧)/𝑔(𝑧)) is a well-defined primitive for ( 𝑓 /𝑔) ′ (𝑔/ 𝑓 ) near {𝛾}. Thus,

1
0= ( 𝑓 /𝑔) ′ (𝑔/ 𝑓 )
2𝜋𝑖 𝛾
∫  ′
𝑔′

1 𝑓
= −
2𝜋𝑖 𝛾 𝑓 𝑔
𝑍𝑓
∑︁ 𝑃𝑓  ∑︁ 𝑍 𝑃𝑔 
 𝑓
∑︁
𝑓   𝑔 𝑔
∑︁
𝑔 
=  𝑛(𝛾; 𝑧 𝑘 ) −
 𝑛(𝛾; 𝑎 𝑘 )  −  𝑛(𝛾; 𝑧 𝑘 ) −
 𝑛(𝛾; 𝑎 𝑘 )  ,
 𝑘=1 𝑘=1   𝑘=1 𝑘=1 
   
where the last equality directly follows from the Argument Principle. ♦
Chapter IV
The Maximum Modulus Theorem

Problem IV.1 Prove the following Minimum Principle. If 𝑓 is a non-constant analytic


function on a bounded open set 𝐺 and is continuous on 𝐺, then either 𝑓 has a zero
in 𝐺 or | 𝑓 | assumes its minimum value on 𝜕𝐺.

Solution IV.1 Since 𝑓 is continuous on 𝐺, we must have | 𝑓 | is continuous on 𝐺.


Then there exists an 𝑎 ∈ 𝐺 such that | 𝑓 (𝑎)| ≤ | 𝑓 (𝑧)| for all 𝑧 ∈ 𝐺. If 𝑎 ∈ 𝜕𝐺, then
| 𝑓 | assumes its minimum value on 𝜕𝐺 and we are done. If 𝑎 ∉ 𝜕𝐺, then 𝑎 ∈ 𝐺 and
Ð
we can write 𝐺 = 𝑋𝑖 , where the 𝑋𝑖 ’s are the components of 𝐺. Then 𝑎 ∈ 𝑋𝑖 for
some 𝑖.
But each 𝑋𝑖 is a region, so using Problem – II.5, either 𝑓 (𝑎) = 0 or 𝑓 is constant in
each 𝑋𝑖 (and hence in 𝐺). But 𝑓 was assumed to be non-constant, hence 𝑓 has to
have a zero in 𝐺. Therefore, either 𝑓 has a zero in 𝐺 or | 𝑓 | assumes its minimum
value on 𝜕𝐺 and we are done. ♦

Problem IV.2 Let 𝐺 be a bounded region and suppose 𝑓 is continuous on 𝐺 and


analytic on G. Show that if there is a constant 𝑐 ≥ 0 such that | 𝑓 (𝑧)| = 𝑐 for all 𝑧 on
the boundary of 𝐺 then either 𝑓 is a constant function or 𝑓 has a zero in 𝐺.

Solution IV.2 Let 𝑐 > 0 be a real constant such that | 𝑓 (𝑧)| = 𝑐 for all 𝑧 ∈ 𝜕𝐺. By
the Maximum Modulus Theorem (Second Version), we obtain

max{| 𝑓 (𝑧)| : 𝑧 ∈ 𝐺} = max{| 𝑓 (𝑧)| : 𝑧 ∈ 𝜕𝐺} = 𝑐.

Thus, there exists a complex number 𝜔 ∈ 𝐺 such that

| 𝑓 (𝑧)| ≤ | 𝑓 (𝜔)| = 𝑐

for all 𝑧 ∈ 𝐺. Then by the Maximum Modulus Theorem (First Version), 𝑓 must be
constant in 𝐺.
Now, if 𝑐 = 0, then by the same arguments, there must exist an 𝛼 ∈ 𝐺 such that
| 𝑓 (𝛼)| = 𝑐 = 0. But since 𝑓 is continuous on 𝐺, this would imply that 𝑓 has a zero
in 𝐺 and we are done. ♦

23
24 IV The Maximum Modulus Theorem

Problem IV.3 Suppose 𝐺 is a region, 𝑓 : 𝐺 → C is analytic, and 𝑀 is a constant


such that whenever 𝑧 ∈ 𝜕∞ 𝐺 and {𝑧 𝑛 } is a sequence in 𝐺 with 𝑧 = lim 𝑧 𝑛 , we have
lim sup | 𝑓 (𝑧 𝑛 )| ≤ 𝑀. Show that | 𝑓 (𝑧)| ≤ 𝑀, for each 𝑧 in 𝐺.

Solution IV.3 The proof is by contrapositive. Let 𝑎 ∈ 𝜕∞ 𝐺 and lim sup | 𝑓 (𝑧)| > 𝑀.
𝑧→𝑎
We will show that this implies lim sup | 𝑓 (𝑧 𝑛 )| > 𝑀.
Now, since 𝑧 𝑛 → 𝑧 as 𝑧 → ∞, and since 𝑓 is analytic (continuous), we must have
𝑓 (𝑧 𝑛 ) → 𝑓 (𝑧). Thus,

lim sup | 𝑓 (𝑧)| > 𝑀 =⇒ lim sup | 𝑓 (𝑧 𝑛 )| > 𝑀.


𝑧→𝑎

Then by negation of this statement, we obtain that

lim sup | 𝑓 (𝑧 𝑛 )| ≤ 𝑀 =⇒ lim sup | 𝑓 (𝑧)| ≤ 𝑀


𝑧→𝑎

for all 𝑎 ∈ 𝜕∞ 𝐺, and we are done. ♦

Problem IV.4 Suppose | 𝑓 (𝑧)| ≤ 1 for |𝑧| < 1 and 𝑓 is a non-constant analytic
function. Considering the function 𝑔 : D → D defined by

𝑓 (𝑧) − 𝑎
𝑔(𝑧) = ,
1 − 𝑎¯ 𝑓 (𝑧)

where 𝑎 = 𝑓 (0), prove that

| 𝑓 (0)| − |𝑧| | 𝑓 (0)| + |𝑧|


≤ | 𝑓 (𝑧)| ≤
1 + | 𝑓 (0)||𝑧| 1 − | 𝑓 (0)||𝑧|

for |𝑧| < 1.

Solution IV.4 Considering the given function 𝑔(𝑧) above, we obtain

𝑓 (0) − 𝑎
𝑔(0) =
1 − 𝑎¯ 𝑓 (0)
𝑎−𝑎
= = 0.
1 − 𝑎𝑎
¯
Again, using the triangle and reverse triangle inequalities, we obtain

𝑓 (𝑧) − 𝑎
|𝑔(𝑧)| =
1 − 𝑎¯ 𝑓 (𝑧)
| 𝑓 (𝑧) − 𝑎|
=
|1 − 𝑎¯ 𝑓 (𝑧)|
| 𝑓 (𝑧)| + |𝑎|
≤ ≤ 1,
|1 − | 𝑎||
¯ 𝑓 (𝑧)||
IV The Maximum Modulus Theorem 25

since | 𝑓 | ≤ 1 on D. So, 𝑔 satisfies the hypotheses of Schwarz’s Lemma, and we have


|𝑔(𝑧)| ≤ |𝑧| for |𝑧| < 1. Using this we can write

=⇒ | 𝑓 (𝑧) − 𝑎| ≤ |𝑧| · |1 − 𝑎¯ 𝑓 (𝑧)|


=⇒ | 𝑓 (𝑧)| − |𝑎| ≤ | 𝑓 (𝑧) − 𝑎| ≤ |𝑧| · |1 − 𝑎¯ 𝑓 (𝑧)| ≤ |𝑧| + |𝑧||𝑎| · | 𝑓 (𝑧)|

=⇒ | 𝑓 (𝑧)| − |𝑎| ≤ |𝑧| + |𝑧||𝑎| · | 𝑓 (𝑧)|


=⇒ −|𝑧| − |𝑧||𝑎| · | 𝑓 (𝑧)| ≤ | 𝑓 (𝑧)| − |𝑎| ≤ |𝑧| + |𝑧||𝑎| · | 𝑓 (𝑧)|
|𝑎| − |𝑧| |𝑎| + |𝑧|
=⇒ ≤ | 𝑓 (𝑧)| ≤ .
1 + |𝑎||𝑧| 1 − |𝑎||𝑧|

Since 𝑎 := 𝑓 (0), we are done. ♦


 
1 3
Problem IV.5 Does there exist an analytic function 𝑓 : D → D such that 𝑓 2 = 4
 
and 𝑓 ′ 12 = 23 ?

Solution IV.5 The proof is by contradiction. Assume such an analytic function 𝑓


exists. By Proposition – V.2.2, we know that if 𝛼 = 𝑓 (𝑎), then

1 − |𝛼| 2
| 𝑓 ′ (𝑎)| ≤ .
1 + |𝑎| 2
3
Setting 𝛼 = 4 and 𝑎 = 12 , we obtain

3 2
1−
 
1 4 7 2
𝑓′ ≤ = < .
2 1 2 12 3
1+ 2
 
This is a contradiction since we assumed 𝑓 ′ 1
2 = 23 . ♦

Problem IV.6 Let 𝑓 : 𝐺 → C be analytic and suppose 𝑀 is a constant such that


lim sup | 𝑓 (𝑧 𝑛 )| ≤ 𝑀 for each sequence {𝑧 𝑛 } in 𝐺 which converges to a point in
𝜕∞ 𝐺. Show that | 𝑓 (𝑧)| ≤ 𝑀. (See Problem – IV.3).

Solution IV.6 We have already proved this result for open and connected regions in
Problem – IV.3. We claim that the result holds even if the region is not required to
be connected. The proof, as before, is by contrapositive.
Here, 𝐺 is assumed to be an open set. But every open set can Ð be written as a
union of its components, which are regions in C. That is, 𝐺 = 𝑋𝑖 , where the
𝑋𝑖 areÐcomponents (maximal open connected subsets) of 𝐺. This also gives that
𝜕𝐺 = (𝜕 𝑋𝑖 ).
Now, assume that there exists an 𝑎 ∈ 𝜕∞ 𝐺 such that for every 𝑀 > 0 we have
26 IV The Maximum Modulus Theorem
h i
lim sup | 𝑓 (𝑧)| := lim+ sup{| 𝑓 (𝑧)| : 𝑧 ∈ 𝐺 ∩ 𝐵(𝑎; 𝑟)} > 𝑀.
𝑧→𝑎 𝑟→0

But then there exists a 𝛿 > 0 such that


h i
lim+ sup{| 𝑓 (𝑧)| : 𝑧 ∈ 𝐺 ∩ 𝐵(𝑎; 𝑟)} > 𝑀 + 𝛿.
𝑟→0

Next, if we have a sequence {𝑧 𝑛 } with lim 𝑧 𝑛 = 𝑎 and let 𝑎 𝑛 := 2|𝑧 𝑛 − 𝑎|, then we
have lim 𝑎 𝑛 = 0. Now consider the sequence

𝑟 𝑛 := sup{| 𝑓 (𝑧)| : 𝑧 ∈ 𝐺 ∩ 𝐵(𝑎; 𝑎 𝑛 )}.

Then lim 𝑟 𝑛 > 𝑀 +𝛿. But since {𝑟 𝑛 } is non-increasing, this implies that 𝑟 𝑛 > 𝑀 +𝛿 for
each 𝑛. Thus, there exist {𝑦 𝑛 } such that each 𝑦 𝑛 ∈ 𝐺 ∩ 𝐵(𝑎; 𝑎 𝑛 ) and | 𝑓 (𝑦 𝑛 )| > 𝑀 + 𝛿
for every 𝑛. Then we finally have

lim sup | 𝑓 (𝑦 𝑛 )| ≥ 𝑀 + 𝛿 ≥ 𝑀,
𝑛→∞

where lim 𝑦 𝑛 = 𝑎. By negation, we obtain the desired result. ♦


Chapter V
Compactness and Convergence in the Space of
Analytic Functions

Module – 1

Problem V.1 If (𝑆, 𝑑) is a metric space and if

𝑑 (𝑠, 𝑡)
𝜇(𝑠, 𝑡) = (V.1)
1 + 𝑑 (𝑠, 𝑡)

then prove that (𝑆, 𝜇) is also a metric space.

Solution V.1 Now, since (𝑆, 𝑑) is a metric space, we have

𝑑 (𝑠, 𝑡) ≥ 0
𝑑 (𝑠, 𝑡) = 𝑑 (𝑡, 𝑠)
𝑑 (𝑠, 𝑡) = 0 ⇐⇒ 𝑠 = 𝑡
𝑑 (𝑠, 𝑡) ≤ 𝑑 (𝑠, 𝑢) + 𝑑 (𝑢, 𝑡),

for all 𝑠, 𝑡, 𝑢 ∈ 𝑆. To show (𝑆, 𝜇) is a metric space we must show that 𝜇(𝑠, 𝑡) satisfies
these properties as well. Now, since 𝑑 (𝑠, 𝑡) ≥ 0, we must have

𝑑 (𝑠, 𝑡)
𝜇(𝑠, 𝑡) = ≥ 0.
1 + 𝑑 (𝑠, 𝑡)

Moreover, if 𝜇(𝑠, 𝑡) = 0, then

𝑑 (𝑠, 𝑡)
= 0 =⇒ 𝑑 (𝑠, 𝑡) = 0 ⇐⇒ 𝑠 = 𝑡.
1 + 𝑑 (𝑠, 𝑡)
Again, for any 𝑠, 𝑡 ∈ 𝑆, we have

𝑑 (𝑠, 𝑡) 𝑑 (𝑡, 𝑠)
𝜇(𝑠, 𝑡) = = = 𝜇(𝑡, 𝑠).
1 + 𝑑 (𝑠, 𝑡) 1 + 𝑑 (𝑡, 𝑠)

27
28 V Compactness and Convergence in the Space of Analytic Functions

We are done if we can show that 𝜇(𝑠, 𝑡) satisfies the triangle inequality. Let 𝑠, 𝑡, 𝑢 ∈ 𝑆.
Now, since 𝜇(𝑠, 𝑡) ≥ 0, if 𝑑 (𝑠, 𝑡) ≤ max{𝑑 (𝑠, 𝑢), 𝑑 (𝑢, 𝑡)} then

𝜇(𝑠, 𝑡) ≤ max{𝜇(𝑠, 𝑢), 𝜇(𝑢, 𝑡)}.

This would then imply that

𝜇(𝑠, 𝑡) ≤ 𝜇(𝑠, 𝑢) + 𝜇(𝑢, 𝑡).

Again, if 𝑑 (𝑠, 𝑡) > max{𝑑 (𝑠, 𝑢), 𝑑 (𝑢, 𝑡)} then


 
1 1 1
< min , .
1 + 𝑑 (𝑠, 𝑡) 1 + 𝑑 (𝑠, 𝑢) 1 + 𝑑 (𝑢, 𝑡)

Since 𝑑 (𝑠, 𝑡) also satisfies the triangle inequality, this implies

𝑑 (𝑠, 𝑡) 𝑑 (𝑠, 𝑢) + 𝑑 (𝑢, 𝑡)


𝜇(𝑠, 𝑡) = ≤
1 + 𝑑 (𝑠, 𝑡) 1 + 𝑑 (𝑠, 𝑡)
𝑑 (𝑠, 𝑢) 𝑑 (𝑢, 𝑡)
= +
1 + 𝑑 (𝑠, 𝑡) 1 + 𝑑 (𝑠, 𝑡)
𝑑 (𝑠, 𝑢) 𝑑 (𝑢, 𝑡)
≤ +
1 + 𝑑 (𝑠, 𝑢) 1 + 𝑑 (𝑢, 𝑡)
= 𝜇(𝑠, 𝑢) + 𝜇(𝑢, 𝑡).

So, (𝑆, 𝜇) is a metric space. ♦

Problem V.2 Recall that Proposition – VII.1.2 states that if 𝐺 is an open set then
Ð
there exists a sequence {𝐾𝑛 } of compact subsets of 𝐺 such that 𝐾𝑛 = 𝐺. Moreover,
the sets 𝐾𝑛 can be chosen to satisfy some other properties. Find such sets 𝐾𝑛 for:
(a) 𝐺 an open disk;
(d) 𝐺 an infinite strip.

Solution V.2 Part-(a): Let 𝑎 ∈ C and 𝑟 > 0 such that 𝐺 = 𝐵(𝑎; 𝑟). Consider the
sets  
1
𝐾𝑛 := 𝐵 𝑎; 𝑟 − .
𝑛
1
Here, if 𝑟 − ≤ 0, then we set 𝐾𝑛 = ∅. Then each 𝐾𝑛 is compact and satisfies
𝑛 Ð
𝐾𝑛 ⊂ int 𝐾𝑛+1 . Moreover, we trivially have that 𝐾𝑛 = 𝐺 and so we are done.

Part-(d): Without loss of generality, we assume 𝐺 = {𝑧 ∈ C : |ℑ(𝑧)| < 1}. Let


𝑧 ∈ C is of the form 𝑧 = 𝑥 + 𝑖𝑦, where 𝑥, 𝑦 ∈ R. Consider
 
1
𝐾𝑛 := {𝑧 ∈ C : |𝑥| ≤ 𝑛} ∩ 𝑧 ∈ C : |𝑦| ≤ 1 − .
𝑛
Module – 1 29

Now, since the intersection of two closed sets is closed, each 𝐾𝑛 is closed. Moreover,
Ð 𝐾𝑛 is bounded and hence 𝐾𝑛 is compact for all 𝑛 ∈ N. We also have that
each
𝐾𝑛 = 𝐺 and so we are done. ♦

Problem V.3 (Dini’s Theorem) Consider C(𝐺, R) and suppose that { 𝑓𝑛 } is a se-
quence in C(𝐺, R) which is monotonically increasing (i.e., 𝑓𝑛 (𝑧) ≤ 𝑓𝑛+1 (𝑧) for all
𝑧 ∈ 𝐺) and lim 𝑓𝑛 (𝑧) = 𝑓 (𝑧) for all 𝑧 in 𝐺, where 𝑓 ∈ C(𝐺, R). Show that 𝑓𝑛 → 𝑓 .

Ð we know that there exists a sequence {𝐾𝑛 }


Solution V.3 By Proposition – VII.1.2,
of compact subsets of 𝐺 such that 𝐾𝑛 = 𝐺. Without loss of generality, we can
thus assume that 𝐺 itself is compact. We consider the sequence of functions

𝑔𝑛 := 𝑓 − 𝑓𝑛 .

Since 𝑓𝑛 is monotonically increasing and lim 𝑓𝑛 (𝑥) = 𝑓 (𝑥), 𝑔𝑛 is monotonically


decreasing and lim 𝑔𝑛 (𝑥) = 0. That is, for every 𝜀 > 0, there exists an 𝑁 ∈ N such
that
𝜀
0 ≤ 𝑔 𝑁 (𝑥) <
2
for all 𝑥 ∈ 𝐺. Moreover, each 𝑔𝑛 is continuous since each 𝑓𝑛 and 𝑓 are in C(𝐺, R).
By the continuity of 𝑔𝑛 , we have that for every 𝜀 > 0, there exists an 𝑁 ∈ N such
that
𝑔 𝑁 (𝑦) < 𝜀
for all 𝑦 ∈ 𝐵(𝑥; 𝑟) with 𝑟 < ∞ and any 𝑥 ∈ 𝐺. Since 𝑔 𝑁 is monotonically decreasing,
we must have 𝑔𝑛 < 𝜀 for every 𝑛 ≥ 𝑁 and 𝑦 ∈ 𝐵(𝑥; 𝑟). These open balls form an open
cover for 𝐺. But 𝐺 is compact and hence we must have finitely many 𝑥 1 , 𝑥2 , . . . , 𝑥 𝑘
such that for every 𝜀 > 0, 𝐵(𝑥 𝑖 ; 𝑟) forms a subcover that covers 𝐺 whenever 𝑛 ≥ 𝑁𝑖 .
Thus, we finally have that 𝑔𝑛 < 𝜀 for every 𝑛 ≥ max{𝑁1 , . . . , 𝑁 𝑘 }. This implies that

| 𝑓𝑛 (𝑥) − 𝑓 (𝑥)| < 𝜀

for every 𝜀 > 0 and thus, 𝑓𝑛 → 𝑓 . ♦

Problem V.4 Let { 𝑓𝑛 } ⊂ C(𝐺, Ω) and suppose that { 𝑓𝑛 } is equicontinuous at each


point of 𝐺. If 𝑓 ∈ C(𝐺, Ω) and 𝑓 (𝑧) = lim 𝑓𝑛 (𝑧) for each 𝑧 then show that 𝑓𝑛 → 𝑓 .
Solution V.4 Using the same argument as in the previous exercise, we can assume
without loss of generality that 𝐺 is compact. Since 𝑓𝑛 is equicontinuous at each
𝑧 ∈ 𝐺, we have that for every 𝜀 > 0 there exists a 𝛿 > 0 such that
𝜀
𝑑 ( 𝑓𝑛 (𝑥), 𝑓𝑛 (𝑦)) <
3
whenever 𝑑 (𝑥, 𝑦) < 𝛿. Since 𝑓 (𝑧) = lim 𝑓𝑛 (𝑧), we also have
𝜀
𝑑 ( 𝑓 (𝑥), 𝑓 (𝑦)) <
3
30 V Compactness and Convergence in the Space of Analytic Functions

whenever 𝑑 (𝑥, 𝑦) < 𝛿. As in the previous exercise, we can find finitely many points
𝑦 1 , . . . , 𝑦 𝑘 ∈ 𝐺 such that the open balls 𝐵(𝑦 𝑖 ; 𝛿) cover 𝐺, since 𝐺 is compact. Since
𝑓𝑛 → 𝑓 (pointwise), we thus have for every 𝑛 ≥ 𝑁𝑖
𝜀
𝑑 ( 𝑓𝑛 (𝑦 𝑖 ), 𝑓 (𝑦 𝑖 )) < .
3
Finally, for an arbitrary 𝑧 ∈ 𝐺 and 𝑛 ≥ max{𝑁1 , . . . , 𝑁 𝑘 }, we have
𝜀
𝑑 ( 𝑓𝑛 (𝑧), 𝑓 (𝑧)) ≤ 𝑑 ( 𝑓𝑛 (𝑧), 𝑓𝑛 (𝑦 𝑖 )) + 𝑑 ( 𝑓𝑛 (𝑦 𝑖 ), 𝑓 (𝑦 𝑖 )) + 𝑑 ( 𝑓 (𝑦 𝑖 ), 𝑓 (𝑧)) < 3 · = 𝜀.
3
This then means 𝑓𝑛 → 𝑓 and we are done. ♦

Problem V.5 Prove Vitali’s Theorem: If 𝐺 is a region and { 𝑓𝑛 } ⊂ H (𝐺) is locally


bounded and 𝑓 ∈ H (𝐺) that has the property that 𝐴 = {𝑧 ∈ 𝐺 : lim 𝑓𝑛 (𝑧) = 𝑓 (𝑧)}
has a limit point in 𝐺 then 𝑓𝑛 → 𝑓 . [H (𝐺) is the space of all analytic functions
from 𝐺 to C.]

Solution V.5 The proof is by contradiction. We assume 𝑓𝑛 ̸→ 𝑓 in H (𝐺). The


sequence { 𝑓𝑛 } is a locally bounded sequence in H (𝐺). Then by Montel’s Theorem,
{ 𝑓𝑛 } is normal. So, there is a subsequence { 𝑓𝑛𝑘 } ⊂ { 𝑓𝑛 } that converges to 𝑓 in
H (𝐺). Now, since 𝑓𝑛 ̸→ 𝑓 , there must be a compact set 𝐾 ⊂ 𝐺 such that the
convergence is not uniform on 𝐾. Thus, there must be an 𝜀 > 0 such that for all
𝑛 ∈ N there is 𝑧 𝑛 ∈ 𝐾 with

| 𝑓𝑛 (𝑧 𝑛 ) − 𝑓 (𝑧 𝑛 )| ≥ 𝜀.

By compactness of 𝐾 we choose a subsequence of 𝑧 𝑛 , say, 𝑧 𝑛𝑚 so that it converges


to a point 𝑧 0 ∈ 𝐾.
Here, since { 𝑓𝑛 } is locally bounded, the subsequence { 𝑓𝑛𝑚 } is locally bounded. Again
by Montel’s Theorem, there is an analytic function in H (𝐺) such that 𝑓𝑛𝑚 → 𝑔 in
H (𝐺). If 𝐴 is the set of points of pointwise convergence 𝑓𝑛 (𝑧) → 𝑓 and 𝑓𝑛 (𝑧) → 𝑔.
Now since 𝐺 is a region and 𝐴 has a limit point in 𝐺, by Corollary – IV.3.8, we have
𝑓 ≡ 𝑔 on 𝐺, which gives a contradiction on the set 𝐾 and the point 𝑧0 because

| 𝑓𝑛𝑚 (𝑧 𝑛𝑚 ) − 𝑓 (𝑧 𝑛𝑚 )| → |𝑔(𝑧0 ) − 𝑓 (𝑧 0 )| ≥ 𝜀.

So, we must have that 𝑓𝑛 → 𝑓 in H (𝐺). ♦

Problem V.6 Let D = 𝐵(0; 1) and for 0 < 𝑟 < 1, let 𝛾𝑟 (𝑡)∫= 𝑟e2 𝜋𝑖𝑡 , 𝑡 ∈ [0, 1]. Show
that a sequence { 𝑓𝑛 } in H (D) converges to 𝑓 if and only if 𝛾 | 𝑓 (𝑧)− 𝑓𝑛 (𝑧)| |𝑑𝑧| → 0
𝑟
as 𝑛 → ∞ for each 𝑟, 0 < 𝑟 < 1.

Solution V.6 We first assume lim 𝑓𝑛 (𝑧) = 𝑓 (𝑧) for all 𝑧 ∈ D. Then we have
Module – 1 31
∫ ∫
| 𝑓 (𝑧) − 𝑓𝑛 (𝑧)| |𝑑𝑧| ≤ sup {| 𝑓 (𝑧) − 𝑓𝑛 (𝑧)|} |𝑑𝑧|
𝛾𝑟 𝑧 ∈𝛾𝑟 𝛾𝑟
= 2𝜋𝑟 · sup {| 𝑓 (𝑧) − 𝑓𝑛 (𝑧)|}.
𝑧 ∈𝛾𝑟

But sup {| 𝑓 (𝑧) − 𝑓𝑛 (𝑧)|} → 0 and thus, we have


𝑧 ∈𝛾𝑟

| 𝑓 (𝑧) − 𝑓𝑛 (𝑧)| |𝑑𝑧| → 0.
𝛾𝑟

Next, we assume 𝛾 | 𝑓 (𝑧) − 𝑓𝑛 (𝑧)| |𝑑𝑧| → 0 as 𝑛 → ∞ for each 𝑟, 0 < 𝑟 < 1. Let
𝑟
𝐾 be some compact subset of D and choose 0 < 𝑟 < 1 such that 𝐾 ⊂ 𝐵(0; 𝑟) and
define 𝛿 > 0 by
𝛿 = min{𝑑 (𝑥, 𝑦) : 𝑥 ∈ 𝐾, 𝑦 ∈ 𝐵(0; 𝑟)}.
By the assumption, we have

| 𝑓 (𝑧) − 𝑓𝑛 (𝑧)| |𝑑𝑧| < 2𝜋𝜀𝛿.
𝜕𝐵

If 𝑎 ∈ 𝐾, by Cauchy’s Integral Formula,



1 𝑓 (𝑧) − 𝑓𝑛 (𝑧)
𝑓 (𝑎) − 𝑓𝑛 (𝑎) = 𝑑𝑧.
2𝜋𝑖 𝜕𝐵 𝑧−𝑎
Thus, we have

1 | 𝑓 (𝑧) − 𝑓𝑛 (𝑧)|
| 𝑓 (𝑎) − 𝑓𝑛 (𝑎)| ≤ |𝑑𝑧|
2𝜋 𝜕𝐵 |𝑧 − 𝑎|

1
≤ | 𝑓 (𝑧) − 𝑓𝑛 (𝑧)||𝑑𝑧|
2𝜋𝛿 𝜕𝐵
2𝜋𝜀𝛿
< = 𝜀.
2𝜋𝛿
Thus, 𝑓𝑛 (𝑎) → 𝑓 (𝑎) for all 𝑎 ∈ 𝐾. Since 𝐾 was arbitrary, we must have that
{ 𝑓𝑛 (𝑧)} → 𝑓 (𝑧) for all 𝑧 ∈ D . ♦

Problem V.7 Let { 𝑓𝑛 } ⊂ H (𝐺) be a sequence of one-one functions which converge


to 𝑓 . If 𝐺 is a region, show that either 𝑓 is one-one or 𝑓 is a constant function.

Solution V.7 Assume { 𝑓𝑛 } ⊂ H (𝐺) is a sequence of one-one functions which


converge to 𝑓 where 𝐺 is a region. Suppose 𝑓 is not the constant function. Then,
we have to show that 𝑓 is one-one. Choose an arbitrary point 𝑧 0 ∈ 𝐺 and define the
sequence
𝑔𝑛 (𝑧) = 𝑓𝑛 (𝑧) − 𝑓𝑛 (𝑧0 ).
32 V Compactness and Convergence in the Space of Analytic Functions

It is clear that the sequence {𝑔𝑛 } converges to 𝑔(𝑧) = 𝑓 (𝑧) − 𝑓 (𝑧0 ) on the open
connected set 𝐺 \ {𝑧 0 }, which is again a region. We also have that 𝑔𝑛 never vanishes
on 𝐺 \ {𝑧0 }, since 𝑓𝑛 are one-to-one functions. Therefore, the sequence {𝑔𝑛 } satisfies
the hypotheses of Corollary – VII.2.6, where the region is 𝐺 \ {𝑧 0 }. Thus, either
𝑔 ≡ 0 or 𝑔 never vanishes. Since 𝑔(𝑧) = 𝑓 (𝑧) − 𝑓 (𝑧0 ) is not identically zero on
𝐺 \ {𝑧0 }, we must have 𝑔(𝑧) = 𝑓 (𝑧) − 𝑓 (𝑧 0 ) has no zeros in 𝐺 \ {𝑧0 }. This implies

𝑓 (𝑧) ≠ 𝑓 (𝑧 0 )

for all 𝑧 ∈ 𝐺 \ {𝑧 0 }. But this in turn implies that 𝑓 is one-to-one on 𝐺. ♦

Problem V.8 Prove Proposition – 3.3(b): If 𝜌 > 0 is given and 𝑎 ∈ C then there is a
number 𝑟 > 0 such that 𝐵(𝑎; 𝑟) ⊂ 𝐵∞ (𝑎; 𝜌).
𝛿
Solution V.8 Let 𝑎 ∈ C. Assume 𝛿 > 0 is given. We pick an 𝑟 > 0 such that 𝑟 < .
2
Now, if 𝑧 ∈ 𝐵(𝑎; 𝑟) we have

|𝑧 − 𝑎| < 𝑟
𝛿
< ·1
2
𝛿 √︁ √︁
≤ 1 + |𝑧| 2 1 + |𝑎| 2
2
2|𝑧 − 𝑎|
=⇒ 𝛿 > √︁ √︁
1 + |𝑧| 2 1 + |𝑎| 2
∴ 𝑑 (𝑧, 𝑎) < 𝛿.

This implies that 𝑧 ∈ 𝐵∞ (𝑎; 𝜌), which implies that

𝐵(𝑎; 𝑟) ⊂ 𝐵∞ (𝑎; 𝜌). ♦

Problem V.9 Prove Proposition – 3.3(d): If a compact set 𝐾 ⊂ C is given, there is a


number 𝑝 > 0 such that 𝐵∞ (∞; 𝜌) ⊂ C∞ \𝐾.

Solution V.9 Let 𝑟 > 0 and 𝐾 ⊂ C, where 𝐾 is compact with 𝐾 ⊂ 𝐵(0; 𝑟). We
trivially have
C∞ \𝐵(0; 𝑟) ⊂ C∞ \𝐾.
So, it is enough to show that for a given 𝑟 > 0, there exists a 𝛿 > 0 such that

𝐵∞ (∞; 𝛿) ⊂ C∞ \𝐵(0; 𝑟).


2
To show this we choose 𝛿 < √ . We then have
1 + 𝑟2
Module – 2 33

𝐵∞ (∞; 𝛿) = {𝑧 ∈ C : 𝑑 (∞, 𝑧) < 𝛿} ∪ {∞}


( )
2
= 𝑧 ∈ C : √︁ < 𝛿 ∪ {∞}
1 + |𝑧| 2
( )
2 2
⊂ 𝑧 ∈ C : √︁ < √ ∪ {∞}
1 + |𝑧| 2 1 + 𝑟2
= {𝑧 ∈ C : 𝑟 ≤ |𝑧|} ∪ {∞}
= C ∪{𝑧 ∈ C : |𝑧| < 𝑟} 𝑐 ∪ {∞}
= C∞ \{𝑧 ∈ C : |𝑧| < 𝑟}
= C∞ \𝐵(0; 𝑟).

We then have
𝐵∞ (∞; 𝛿) ⊂ C∞ \𝐵(0; 𝑟)
and so we are done. ♦

Module – 2

Problem V.10 Let 𝑓 be analytic on 𝐺 = {𝑧 : ℜ(𝑧) > 0}, one-one, with ℜ[ 𝑓 (𝑧)] > 0
for all 𝑧 in 𝐺, and 𝑓 (𝑎) = 𝑎 for some real number 𝑎. Show that | 𝑓 ′ (𝑎)| ≤ 1.

Solution V.10 Here, we notice 𝐺 is a ’proper’ simply connected region in C. If


𝑎 ∈ 𝐺 then by the Riemann Mapping Theorem, there is a unique analytic function
𝑔 : 𝐺 → D with the properties:
(i) 𝑔(𝑎) = 0 and 𝑔 ′ (𝑎) > 0;
(ii) 𝑔 is one-to-one;
(iii) 𝑔(𝐺) = D.
Now, since 𝑔 : 𝐺 → D is bijective, 𝑔 −1 : D → 𝐺 is also bijective, and since
𝑓 : 𝐺 → 𝐺 is also one-to-one, we can define a one-to-one analytic function ℎ(𝑧) by

ℎ(𝑧) = 𝑔 ◦ 𝑓 ◦ 𝑔 −1 (𝑧).

We note ℎ(𝑧) is also bijective with ℎ(D) = D. Thus, we have

ℎ(0) = 𝑔 ◦ 𝑓 ◦ 𝑔 −1 (0) = 𝑔 ◦ 𝑓 (𝑎) = 𝑔(𝑎) = 0,

where we have used the facts that 𝑔 −1 (0) = 𝑓 (𝑎) = 𝑎. By definition, we also have

|ℎ(𝑧)| ≤ 1

for all 𝑧 ∈ D. Applying Schwarz’s Lemma to ℎ(𝑧), we now have

|ℎ′ (0)| ≤ 1.
34 V Compactness and Convergence in the Space of Analytic Functions

Taking the derivative of ℎ(𝑧), using the Chain Rule, we obtain


 ′
ℎ′ (𝑧) = 𝑔 ◦ 𝑓 ◦ 𝑔 −1 (𝑧)
h i h i h i′
= 𝑔 ′ ◦ 𝑓 ◦ 𝑔 −1 (𝑧) · 𝑓 ′ ◦ 𝑔 −1 (𝑧) · 𝑔 −1 (𝑧)
h i h i 1
= 𝑔 ′ ◦ 𝑓 ◦ 𝑔 −1 (𝑧) · 𝑓 ′ ◦ 𝑔 −1 (𝑧) · ′ −1 ,
𝑔 (𝑔 (𝑧))

where the last equality follows from Proposition – III.2.20 when 𝑔 ′ (𝑔 −1 (𝑧)) ≠ 0.
Next, we evaluate ℎ′ at 𝑧 = 0 as follows
h i h i 1
ℎ′ (0) = 𝑔 ′ ◦ 𝑓 ◦ 𝑔 −1 (0) · 𝑓 ′ ◦ 𝑔 −1 (0) · ′ −1
𝑔 (𝑔 (0))
h i h i
𝑔 ′ ◦ 𝑓 (𝑎) · 𝑓 ′ (𝑎)
=
𝑔 ′ (𝑎)
𝑔 ′ (𝑎) · 𝑓 ′ (𝑎)
= = 𝑓 ′ (𝑎).
𝑔 ′ (𝑎)

We note the last equality is well-defined since 𝑔 ′ (𝑎) > 0. But combining this result
with our earlier result, we have |ℎ′ (𝑎)| = | 𝑓 ′ (𝑎)| ≤ 1 and we are done. ♦

Problem V.11 Let 𝐺 1 and 𝐺 2 be simply connected regions neither of which is the
whole plane. Let 𝑓 be a one-one analytic mapping of 𝐺 1 onto 𝐺 2 . Let 𝑎 ∈ 𝐺 1 and put
𝛼 = 𝑓 (𝑎). Prove that for any one-one analytic map ℎ of 𝐺 1 into 𝐺 2 with ℎ(𝑎) = 𝛼 it
follows that |ℎ′ (𝑎)| ≤ | 𝑓 ′ (𝑎)|. What can be said if ℎ is not assumed to be one-one?

Solution V.11 First, we define a one-to-one analytic function 𝐹 (𝑧) : 𝐺 1 → 𝐺 1 by

𝐹 (𝑧) = 𝑓 −1 ◦ ℎ(𝑧).

Here, 𝐹 (𝑧) well-defined since 𝑓 : 𝐺 1 → 𝐺 2 is bijective and ℎ : 𝐺 1 → 𝐺 2 is


one-to-one. We also have for any 𝑎 ∈ 𝐺 1

𝐹 (𝑎) = 𝑓 −1 ◦ ℎ(𝑎) = 𝑓 −1 (𝛼) = 𝑎.

Thus, by the previous exercise (Problem – V.10), we obtain |𝐹 (𝑎)| ≤ 1. But we also
have by differentiating 𝐹 (𝑧),
 ′
𝐹 ′ (𝑧) = 𝑓 −1 ◦ ℎ(𝑧)
h i′
= 𝑓 −1 (ℎ(𝑧)) · ℎ′ (𝑧)
ℎ′ (𝑧)
= ,
𝑓 ′ ( 𝑓 −1◦ ℎ(𝑧))
Module – 2 35

where the last equality follows from Proposition – III.2.20. Evaluating 𝐹 ′ at 𝑧 = 𝑎,


we obtain
ℎ′ (𝑎)
𝐹 ′ (𝑎) =
𝑓 ′ ( 𝑓 −1
◦ ℎ(𝑎))
ℎ′ (𝑎)
= ′ −1
𝑓 ( 𝑓 (𝛼))
ℎ′ (𝑎)
= ′ ,
𝑓 (𝑎)

which is well-defined since 𝑓 ′ (𝑎) ≠ 0 (by injectivity). Finally, combining this with
our previous result, we directly have |ℎ′ (𝑎)| ≤ | 𝑓 ′ (𝑎)|.
If ℎ(𝑧) is not injective, we still must have the same result since we do not require the
assumption that 𝐹 (𝑧) is injective in Problem – V.10. ♦

Problem V.12 Let 𝑟 1 , 𝑟 2 , 𝑅1 , 𝑅2 be positive numbers such that 𝑅1 /𝑟 1 = 𝑅2 /𝑟 2 ; Show


that ann(0; 𝑟 1 , 𝑅1 ) and ann(0; 𝑟 2 , 𝑅2 ) are conformally equivalent.

Solution V.12 We notice that the function 𝑓 : ann(0; 𝑟 1 , 𝑅1 ) → ann(0; 𝑟 2 , 𝑅2 ) de-


fined by
𝑟2
𝑓 (𝑧) = 𝑧,
𝑟1
is analytic and maps the annulus ann(0; 𝑟 1 , 𝑅1 ) bijectively onto the annulus
ann(0; 𝑟 2 , 𝑅2 ). In particular, 𝑓 maps the inner boundary of the domain to the in-
ner boundary of the range, that is,

𝑓 ({𝑧 : |𝑧| = 𝑟 1 }) = {𝑧 : |𝑧| = 𝑟 2 }.


𝑅2 𝑟 1
As for the outer boundaries, we note that since 𝑅1 = , if |𝑧| = 𝑅1 , then
𝑟2
𝑟 2 𝑅2 𝑟 1
| 𝑓 (𝑧)| = · = 𝑅2 .
𝑟1 𝑟2
So, 𝑓 maps the outer boundary of the domain onto the outer boundary of the range.
Now, the inverse function of 𝑓 is
𝑟1
𝑓 −1 (𝑧) = 𝑧.
𝑟2

We observe that 𝑓 −1 is also a non-constant analytic function. Therefore, the annulus


ann(0; 𝑟 1 , 𝑅1 ) is conformally equivalent to the annulus ann(0; 𝑟 2 , 𝑅2 ). ♦
36 V Compactness and Convergence in the Space of Analytic Functions

log(1 + 𝑧)
Problem V.13 Prove that lim = 1.
𝑧→0 𝑧
log(1 + 𝑧)
Solution V.13 We first note that the function 𝑓 (𝑧) := has a removable
𝑧
singularity at 𝑧 = 0 since

lim 𝑧 𝑓 (𝑧) = lim log(1 + 𝑧) = log(1) = 0.


𝑧→0 𝑧→0

So, 𝑓 (𝑧) can be defined to be analytic at 𝑧 = 0. We know



∑︁ 𝑧𝑘
log(1 + 𝑧) = (−1) 𝑘+1 ,
𝑘=1
𝑘

where |𝑧| < 1. Then we have


∞ ∞ ∞
log(1 + 𝑧) ∑︁ 𝑧 𝑘−1 ∑︁ 𝑧 𝑘−1 ∑︁ 𝑧𝑘
= (−1) 𝑘+1 =1+ (−1) 𝑘+2 =1+ (−1) 𝑘 .
𝑧 𝑘=1
𝑘 𝑘=2
𝑘 𝑘=1
𝑘 +1

We note that the above sum is an analytic function. Thus,


∞ ∞
𝑧𝑘 0

∑︁ ∑︁
lim 𝑓 (𝑧) = 1 + lim (−1) 𝑘 =1+ (−1) 𝑘 = 1.
𝑧→0 𝑧→0
𝑘=1
𝑘 +1 𝑘=1
𝑘 +1

∞  
Ö 1 1
Problem V.14 Prove that 1− 2 = .
𝑛=2
𝑛 2

Solution V.14 First, we consider the finite product,


𝑛   Ö 𝑛   
Ö 1 1 1
𝑃𝑛 = 1− 2 = 1− 1+
𝑘=2
𝑘 𝑘=2
𝑘 𝑘
" 𝑛   #
Ö 1 1
= exp log 1− 1+
𝑘=2
𝑘 𝑘
" 𝑛     #
∑︁ 1 1
= exp log 1 − 1+
𝑘=2
𝑘 𝑘
" 𝑛   #
∑︁ 𝑘 −1 𝑘 +1
= exp log
𝑘=2
𝑘 𝑘
" 𝑛 # " 𝑛 #
∑︁ ∑︁
= exp [log(𝑘 − 1) − log(𝑘)] + exp [log(𝑘 + 1) − log(𝑘)]
𝑘=2 𝑘=2
Module – 2 37
   
= exp log(1) − log(𝑛) + exp log(𝑛 + 1) − log(2)
 
= exp log(𝑛 + 1) − log(2𝑛)
  
𝑛+1 𝑛+1
= exp log = .
2𝑛 2𝑛

Now, by definition, the infinite product is the limit of 𝑃𝑛 as 𝑛 → ∞. Taking the limit,
we obtain
∞  
1 𝑛+1 1

Ö
1 − 2 = lim 𝑃𝑛 = lim = .
𝑘=2
𝑘 𝑛→∞ 𝑛→∞ 2𝑛 2

For the next few exercises, we first state a few definitions. Let 𝐺 be a region and
I ⊆ H (𝐺).
(i) I is called an ideal iff : (a) 𝑓 , 𝑔 ∈ I implies 𝑎 𝑓 + 𝑏𝑔 ∈ I for all 𝑎, 𝑏 ∈ C;
(b) 𝑓 ∈ I and 𝑔 ∈ H (𝐺) implies 𝑓 𝑔 ∈ I .
(ii) I is called a proper ideal iff I ≠ (0) and I ≠ H (𝐺).
(iii) I is a maximal ideal iff I is a proper ideal and whenever J (fancy version
of the letter ’J’) is an ideal with I ⊆ J , this implies I = J or J = H (𝐺).
(iv) I is a prime ideal if whenever 𝑓 , 𝑔 ∈ H (𝐺) and 𝑓 𝑔 ∈ I then either 𝑓 ∈ I
or 𝑔 ∈ I .
(v) If 𝑓 ∈ H (𝐺) then Z ( 𝑓 ) denotes the set of zeros of 𝑓 counted according to
multiplicity.
(vi) If S ⊂ H (𝐺) then Z (S ) := {Z ( 𝑓 ) : 𝑓 ∈ S }, where the zeros are again
Ñ
counted according to multiplicity.
(vii) If 𝐴 ⊂ 𝐺, then I ( 𝐴) := { 𝑓 ∈ H (𝐺) : 𝐴 ⊂ Z ( 𝑓 )}.
(viii) If 𝑓 , 𝑔 ∈ H (𝐺) then 𝑓 divides 𝑔 (in symbols, 𝑓 |𝑔) if there is an ℎ in H (𝐺)
such that 𝑔 = 𝑓 ℎ.
(ix) If S ⊂ H (𝐺) and S contains a non-zero function then 𝑓 is the greatest
common divisor of S if: (a) 𝑓 |𝑔 for each 𝑔 ∈ S ; (b) whenever ℎ|𝑔 for all
𝑔 ∈ S , we have ℎ| 𝑓 . We write 𝑓 = 𝑔.𝑐.𝑑.S .

Problem V.15 Prove that 𝑓 |𝑔 iff Z ( 𝑓 ) ⊂ Z (𝑔).

Solution V.15 First, we assume Z ( 𝑓 ) ⊂ Z (𝑔). Then every zero of 𝑓 of multiplicity


𝑚 is also a zero of 𝑔 of multiplicity at least 𝑚. We observe that to show 𝑓 |𝑔 it is
𝑔
enough to show that the function ℎ = is analytic in 𝐺. But by the above reasoning,
𝑓
ℎ is analytic in 𝐺. Hence, 𝑔 = 𝑓 ℎ and we have 𝑓 |𝑔.
Next, we assume that 𝑓 |𝑔. This implies that there exists an analytic function in 𝐺
with 𝑔 = 𝑓 ℎ. But then we have

Z (𝑔) = Z ( 𝑓 ) ∪ Z (ℎ),

which implies that


Z ( 𝑓 ) ⊂ Z (𝑔). ♦
38 V Compactness and Convergence in the Space of Analytic Functions

Problem V.16 Prove that 𝑓 = 𝑔.𝑐.𝑑.S iff Z ( 𝑓 ) = Z (S ).

Solution V.16 Let 𝑓 = 𝑔.𝑐.𝑑.S . Then by definition, 𝑓 |𝑔 for all 𝑔 ∈ S . By Problem


– V.15, this implies
Z ( 𝑓 ) ⊂ Z (𝑔)
for all 𝑔 ∈ S . But then we have
Ù
Z (𝑓) = {Z (𝑔) : 𝑔 ∈ S } =: Z (S ).

Next, assume that Z ( 𝑓 ) = Z (S ). Then Z ( 𝑓 ) ⊂ Z (𝑔) for all 𝑔 ∈ S . By Problem


– V.15, this gives 𝑓 |𝑔. Now, let ℎ|𝑔 for all 𝑔 ∈ S . We are done if we can show that
ℎ| 𝑓 . Since ℎ|𝑔, ∀𝑔 ∈ S , by Problem – V.15,

Z (ℎ) ⊂ Z (𝑔)

for all 𝑔 ∈ S . By definition, this implies


Ù
Z (ℎ) = {Z (𝑔) : 𝑔 ∈ S } =: Z (S ).

Problem V.17 Show that every maximal ideal in H (𝐺) is a prime ideal.

Solution V.17 The proof is by contrapositive. Let I be a maximal ideal in H (𝐺).


Assume that 𝑓 , 𝑔 ∈ H (𝐺) with 𝑓 , 𝑔 ∉ I . Then since I is a maximal ideal, we can
write
H (𝐺) = I + { 𝑓 ℎ1 : ℎ1 ∈ H (𝐺)},
and
H (𝐺) = I + {𝑔ℎ2 : ℎ2 ∈ H (𝐺)}.
We now notice that

1 = 𝑚 + 𝑓 ℎ1
1 = 𝑛 + 𝑔ℎ2 ,

where 𝑚, 𝑛 ∈ I . Multiplying, we obtain

1 = 𝑚𝑛 + 𝑚𝑔ℎ2 + 𝑛 𝑓 ℎ1 + 𝑓 𝑔ℎ1 ℎ2 .

Here, the analytic function 𝑚𝑛 + 𝑚𝑔ℎ2 + 𝑛 𝑓 ℎ1 is in I . Since 1 ∉ I , we must then


have that
𝑓 𝑔ℎ1 ℎ2 ∉ I .
By contraposition, we get the desired result. ♦
Module – 2 39

Problem V.18 Give an example of an ideal which is not a prime ideal.

Solution V.18 Let 𝑧1 , 𝑧2 ∈ 𝐺. Consider the set

J := { 𝑓 ∈ H (𝐺) : 𝑓 (𝑧1 ) = 0 and 𝑓 (𝑧2 ) = 0}.

We notice that if 𝑓 , 𝑔 ∈ J and 𝑎, 𝑏 ∈ C, then (𝑎 𝑓 + 𝑏𝑔) ∈ J since

(𝑎 𝑓 + 𝑏𝑔) (𝑧1 ) = 𝑎 𝑓 (𝑧1 ) + 𝑏𝑔(𝑧1 ) = 0 + 0 = 0,

and
(𝑎 𝑓 + 𝑏𝑔) (𝑧 2 ) = 𝑎 𝑓 (𝑧2 ) + 𝑏𝑔(𝑧2 ) = 0 + 0 = 0.
Again, if 𝑓 ∈ J and 𝑔 ∈ H (𝐺), then 𝑓 𝑔 ∈ J since

( 𝑓 𝑔) (𝑧 1 ) = 𝑓 (𝑧 1 )𝑔(𝑧1 ) = 0 · 0 = 0,

and
( 𝑓 𝑔) (𝑧2 ) = 𝑓 (𝑧2 )𝑔(𝑧 2 ) = 0 · 0 = 0.
Thus, J is an ideal in H (𝐺). Next, we observe that if 𝑓 𝑔 ∈ J , then none of them
is necessarily in the ideal J . For instance, consider 𝑓 (𝑧) = 𝑧 − 𝑧 1 and 𝑔(𝑧) = 𝑧 − 𝑧2 .
Then 𝑓 (𝑧)𝑔(𝑧) = (𝑧 − 𝑧1 ) (𝑧 − 𝑧2 ) is in J , but clearly neither 𝑓 nor 𝑔 is in J .
Hence, J is an ideal which is not a prime ideal. ♦

Problem V.19 Find an entire function 𝑓 such that 𝑓 (𝑛 + 𝑖𝑛) = 0 for every integer 𝑛.

Solution V.19 Let (𝑎 𝑛 ) := 𝑛 + 𝑖𝑛. First, we observe that


∞  2 ∞
∑︁ 𝑟 ∑︁ 𝑟2
= <∞
𝑛=1
|𝑛 + 𝑖𝑛| 𝑛=1
2𝑛2

for all 𝑟 > 0. Then by the Weierstrass Factorization Theorem for 𝐺 = C, we obtain
that the function
Ö∞
𝑓 (𝑧) = 𝐸 2 (𝑧/𝑎 𝑛 )
𝑛=1

is entire and have zeros at 𝑧 = 𝑎 𝑛 for each integer 𝑛. Here, 𝐸 2 (𝑧/𝑎 𝑛 ) is the second
order Elementary Factor evaluated at 𝑧 0 = 𝑧/𝑎 𝑛 . By definition,

𝑧2
   
𝑧 𝑧
𝐸 2 (𝑧/𝑎 𝑛 ) = 1 − exp + .
𝑎𝑛 𝑎 𝑛 2𝑎 2𝑛

Thus, our desired function is given by


∞ 
𝑧2
  
𝑧 𝑧

Ö
𝑓 (𝑧) = 1− exp + .
𝑛=1
𝑎𝑛 𝑎 𝑛 2𝑎 2𝑛
40 V Compactness and Convergence in the Space of Analytic Functions

Problem V.20 Find a factorization for cos(𝜋𝑧).

Solution V.20 We recall that the complex sine and cosine functions satisfy the well-
known double-angle identity

sin(2𝜋𝑧) = 2 sin(𝜋𝑧) cos(𝜋𝑧).

Since a factorization for sin(𝜋𝑧) is known, we can use the above identity to derive a
factorization for cos(𝜋𝑧). We recall
∞ 
𝑧2
Ö 
sin(𝜋𝑧) = 𝜋𝑧 1− 2 .
𝑛=1
𝑛
∞ 
4𝑧2
Ö 
∴ sin(2𝜋𝑧) = 2𝜋𝑧 1− 2 .
𝑛=1
𝑛

Substituting these in the double-angle identity, we obtain


∞  ∞ 
4𝑧 2 𝑧2
Ö  Ö 
2𝜋𝑧 1− = 2𝜋𝑧 cos(𝜋𝑧) 1 −
𝑛=1
𝑛2 𝑛=1
𝑛2
∞  ∞   ∞ 
𝑧2 4𝑧2 Ö 4𝑧2
Ö  Ö 
=⇒ cos(𝜋𝑧) 1− 2 = 1− 1 − ,
𝑛=1
𝑛 𝑚=1
(2𝑚) 2 𝑚=1 (2𝑚 − 1) 2

where the last equality holds since we know that rearrangement of terms does not
alter the convergence of an infinite product. Continuing, we further obtain
∞  ∞   ∞ 
𝑧2 𝑧2 Ö 4𝑧 2
Ö  Ö 
=⇒ cos(𝜋𝑧) 1−= 1 − 1 −
𝑛=1
𝑛2 𝑛=1
𝑛2 𝑛=1 (2𝑛 − 1) 2
∞ 
4𝑧2


Ö
⇐⇒ cos(𝜋𝑧) = 1− .
𝑛=1
(2𝑛 − 1) 2
Module – 2 41

Problem V.21 Find a factorization for sinh 𝑧.

Solution V.21 We first note that


1 𝑧
sinh 𝑧 = (𝑒 − 𝑒 −𝑧 )
2
1  −𝑖2 𝑧 2

= 𝑒 − 𝑒 − (−𝑖 ) 𝑧
2
−𝑖  𝑖 (𝑖𝑧) 
= 𝑒 − 𝑒 −𝑖 (𝑖𝑧)
2𝑖
∴ sinh 𝑧 = (−𝑖) sin(𝑖𝑧).

Now, by Equation – VII.6.2 (from the text), we know


∞ 
𝑧2
Ö 
sin(𝜋𝑧) = 𝜋𝑧 1− .
𝑛=1
𝑛2

This implies that


∞ 
𝑧2
Ö 
sin 𝑧 = 𝑧 1− 2 2 .
𝑛=1
𝜋 𝑛
Substituting 𝑖𝑧 into the above identity, we obtain
∞  ∞ 
𝑖2 𝑧2 𝑧2
 

Ö Ö
sinh 𝑧 = (−𝑖) sin(𝑖𝑧) = (−𝑖) (𝑖𝑧) 1− 2 2 = 𝑧 1+ 2 2 .
𝑛=1
𝜋 𝑛 𝑛=1
𝜋 𝑛

Problem V.22 Prove Wallis’ Formula:



𝜋 Ö 4𝑛2
= .
2 𝑛=1 4𝑛2 − 1

Solution V.22 Substituting 𝑧 = 1/2 into the factorization of the sin(𝜋𝑧) function
(Equation – VII.6.2), we obtain
42 V Compactness and Convergence in the Space of Analytic Functions
∞  
𝜋Ö 1/4
1 = sin(𝜋/2) = 1− 2
2 𝑛=1 𝑛

𝜋 Ö 𝑛2 − 1/4
 
=
2 𝑛=1 𝑛2
∞ 
𝑛2

𝜋 Ö
⇐⇒ =
2 𝑛=1 𝑛2 − 1/4
∞ 
𝑛2
Ö 
=
𝑛=1
(𝑛 + 1/2) (𝑛 − 1/2)

!
Ö 𝑛2
= (2𝑛+1) (2𝑛−1)
𝑛=1 4
∞ 
4𝑛2
Ö 
=
𝑛=1
(2𝑛 + 1) (2𝑛 − 1)

4𝑛2
 
𝜋 Ö
∴ =
2 𝑛=1 4𝑛2 − 1
. ♦

Module – 3

Problem V.23 Show that 0 < 𝛾 < 1, where 𝛾 is the Euler-Mascheroni constant.

Solution V.23 We consider the (real) integral


∫ 1/𝑘  
𝑡 1 𝑘 +1
𝑑𝑡 = − log ,
0 1+𝑡 𝑘 𝑘

where we used the substitution 𝑢 = 1 + 𝑡 to evaluate it. We note that


∫ 1/𝑘 ∫ 1/𝑘
𝑡 1
𝑑𝑡 ≤ 𝑡 𝑑𝑡 = 2 .
0 1 + 𝑡 0 2𝑘
And, ∫ 1/𝑘 ∫ 1/𝑘
𝑡 𝑡 1
𝑑𝑡 ≥ 𝑑𝑡 = .
0 1+𝑡 0 1 + 1/𝑘 2𝑘 (𝑘 + 1)
From this we have the estimates
∞  ∞
𝜋2
   ∑︁
∑︁ 1 𝑘 +1 1
𝛾= − log ≤ 2
= ≈ 0.8225 < 1.
𝑘=1
𝑘 𝑘 𝑘=1
2𝑘 12
Module – 3 43

We also have
∞    ∞
∑︁ 1 𝑘 +1 ∑︁ 1
𝛾= − log ≥ > 0.
𝑘=1
𝑘 𝑘 𝑘=1
2𝑘 (𝑘 + 1)

Combining these results, we obtain 0 < 𝛾 < 1. ♦

Problem V.24 Show that √ Γ(𝑧)Γ(1 − 𝑧) = 𝜋 csc(𝜋𝑧) for 𝑧 not an integer. Deduce
from this that Γ(1/2) = 𝜋.
Solution V.24 If 𝑧 is not an integer, then by Gauss’ Formula (VII.7.6), we obtain

𝑛! 𝑛 𝑧 𝑚! 𝑚 1−𝑧
Γ(𝑧)Γ(1 − 𝑧) = lim lim
𝑛→∞ 𝑧(𝑧 + 1) · · · (𝑧 + 𝑛) 𝑚→∞ (1 − 𝑧) (2 − 𝑧) · · · (𝑚 + 1 − 𝑧)

𝑛! 𝑛! 𝑛 𝑧 𝑛1−𝑧
= lim
𝑛→∞ 𝑧(1 + 𝑧) (1 − 𝑧) · · · (𝑛 + 𝑧) (𝑛 − 𝑧) (𝑛 + 1 − 𝑧)

(𝑛!) 2 · 𝑛
= lim
𝑛→∞ 𝑧(12 − 𝑧 2 ) · · · (𝑛2 − 𝑧 2 ) (𝑛 + 1 − 𝑧)
(𝑛!) 2 · 𝑛
= lim  2
  
𝑧2
𝑧 (𝑛!) 2 1 −
𝑛→∞ 𝑧
12
··· 1− 𝑛2
(𝑛 + 1 − 𝑧)
1
= h   i  
𝑧2 𝑧2 𝑛+1−𝑧
𝑧 · lim 1− 12
··· 1− 𝑛2
· lim 𝑛
𝑛→∞ 𝑛→∞
1
= h   i .
𝑧2 𝑧2
𝑧 · lim 1− 12
··· 1− 𝑛2
𝑛→∞

Now, using the definition of infinite products we know that


∞ 
𝑧2 𝑧2 𝑧2
    
Ö sin(𝜋𝑧)
lim 1− 2
· · · 1 − 2
:= 1 − 2
= ,
𝑛→∞ 1 𝑛 𝑛=1
𝑛 𝜋𝑧

where the last equality follows from Equation – VII.6.2. Combining this with our
previous result, we obtain
1 𝜋
Γ(𝑧)Γ(1 − 𝑧) = h   i = = 𝜋 csc(𝜋𝑧).
𝑧 · lim 1− 𝑧2
··· 1− 𝑧2 sin(𝜋𝑧)
𝑛→∞ 12 𝑛2

Now, substituting 𝑧 = 1/2, we get

[Γ(1/2)] 2 = 𝜋 csc(𝜋/2).

This implies that



Γ(1/2) = 𝜋. ♦
44 V Compactness and Convergence in the Space of Analytic Functions

Problem V.25 Use Theorem – VII.8.17 to prove that 𝑝 𝑛−1 = ∞, if {𝑝 𝑛 } is a


Í
sequence of primes. (Note that this implies there are infinitely many primes.)
Solution V.25 The proof is by contradiction. Assume that 𝑝 𝑛−1 is finite. Euler’s
Í
Theorem (VII.8.7) states that for ℜ(𝑧) > 1,
∞  
Ö 1
𝜁 (𝑧) = .
𝑛=1
1 − 𝑝 𝑛−𝑧

Here, as ℜ(𝑧) → 1, both Í sides of the equality becomes unbounded. Now, since
𝑝 𝑛 > 0 for all 𝑛, the sum 𝑝 𝑛−1 , if finite, also converges absolutely. By Proposition
– VII.5.4, we then have that the series
∞  
∑︁ 1
log 1 −
𝑛=1
𝑝𝑛

converges absolutely. But this implies


∞   ∞  
∑︁ 1 Ö 1
− log 1 − = log = log 𝜁 (1) < ∞.
𝑛=1
𝑝𝑛 𝑛=1
1 − 𝑝𝑛

This is a contradiction since 𝜁 (𝑧) has a simple pole at 𝑧 = 1. Hence, there are in-
finitely many primes. ♦


∑︁ 𝑑 (𝑛)
Problem V.26 Prove that 𝜁 2 (𝑧) = , for ℜ(𝑧) > 1, where 𝑑 (𝑛) is the number
𝑛=1
𝑛𝑧
of divisors of 𝑛.
Solution V.26 We start with the definition of the zeta function and expanding the
product as follows

! ∞ !   
2
∑︁ 1 ∑︁ 1 1 1 1 1 1 1
𝜁 (𝑧) = = 1 + + + + · · · 1 + + + · · ·
𝑛=1
𝑛 𝑧 𝑛=1 𝑛 𝑧 2𝑧 3𝑧 4𝑧 2𝑧 3𝑧 4𝑧
1 1 1 1
=1+ + 𝑧 + 𝑧 + 𝑧 +···
2 𝑧 3 4 5
1 1 1 1
+ 𝑧 + 𝑧 + 𝑧 + 𝑧 +···
2 4 6 8
1 1 1 1
+ 𝑧 + 𝑧 + 𝑧 + 𝑧 +···
3 6 9 12
1 1 1 1
+ 𝑧 + 𝑧 + 𝑧 + 𝑧 +···
4 8 12 16
..
.
2 2 3 2 4
= 1+ 𝑧 + 𝑧 + 𝑧 + 𝑧 + 𝑧 +··· ;
2 3 4 5 6
Module – 3 45

1
Here, we observe that each term of the form 𝑧 appears exactly as many times as
𝑛
there are divisors of 𝑛, counting order. For instance, since 2 = 1 × 2 = 2 × 1, it
appears twice. Same for any prime 𝑛. For composite 𝑛, for instance for 𝑛 = 4, since
4 = 1 × 4 = 2 × 2 = 4 × 1, it appears thrice. Hence, we have

𝑑 (𝑛)

∑︁
𝜁 2 (𝑧) = .
𝑛=1
𝑛𝑧

√ √
Problem V.27 Find a meromorphic function with poles √ of order 2 at 1, 2, 3, . . .
2
√ (𝑧 − 𝑚) 𝑓 (𝑧) = 1 for all 𝑚.
such that the residue at each pole is 0 and lim
𝑧→ 𝑚

Solution V.27 We claim that the function 𝑓 defined by


∞ √
∑︁ (3 𝑛 − 2𝑧)𝑧 2
𝑓 (𝑧) = √ √ 2
3
𝑛=1 𝑛 (𝑧 − 𝑛)

has the required


√ properties. To verify this we observe that the only poles of 𝑓
occur at 𝑧 = 𝑛, ∀𝑛 ∈ N. These are all poles √ of multiplicity 2. Moreover, 𝑓 is
meromorphic in√ C. Next, we verify that Res( 𝑓 ; 𝑛) = 0, ∀𝑛.√By Proposition – V.2.4,
if 𝑔(𝑧) = (𝑧 − 𝑛) 2 𝑓 (𝑧), then it is enough to show that 𝑔 ′ ( 𝑛) = 0, ∀𝑛 ∈ N.

3 𝑛 − 2𝑧 𝑧2

𝑔(𝑧) = √
𝑛3
√ 
′ 2𝑧 3 𝑛 − 2𝑧 2𝑧2
∴ 𝑔 (𝑧) = √ −√
𝑛3 𝑛3
√ 
6𝑧 𝑧 − 𝑛
=− √ ;
𝑛3
√ √
=⇒ 𝑔 ′ ( 𝑛) = Res( 𝑓 ; 𝑛) = 0.
√ 2
√ (𝑧 − 𝑚) 𝑓 (𝑧) = 1 for all 𝑚. We can write
Finally, we verify that lim
𝑧→ 𝑚

∞ √ √
√ 2 ∑︁ (𝑧 − 𝑚) 2 (3 𝑛 − 2𝑧)𝑧2
lim
√ (𝑧 − 𝑚) 𝑓 (𝑧) = lim
√ √ √
𝑧→ 𝑚 𝑧→ 𝑚
𝑛=1 𝑛3 (𝑧 − 𝑛) 2
" √ √ √ √ #
(𝑧 − 𝑚) 2 (3 𝑛 − 2𝑧)𝑧2 ∑︁ (𝑧 − 𝑚) 2 (3 𝑛 − 2𝑧)𝑧2
= lim√ √ √ + √ √
𝑧→ 𝑚 𝑛3 (𝑧 − 𝑛) 2 𝑛≠𝑚 𝑛3 (𝑧 − 𝑛) 2
√ √
3 𝑚3 − 2 𝑚3
= √ + 0 = 1.
𝑚3

Thus, our claimed function satisfies all the requirements and we are done. ♦
Chapter VI
Harmonic and Entire Functions

Problem VI.1 Show that if 𝑢 is harmonic then so are 𝑢 𝑥 and 𝑢 𝑦 .

Solution VI.1 By definition, if 𝑢 is harmonic then

𝑢 𝑥 𝑥 + 𝑢 𝑦𝑦 = 0.

Again since 𝑢 is harmonic, 𝑢 is infinitely differentiable. In particular, the third partial


derivatives of 𝑢 exist and are continuous. So, we have

𝑢 𝑥 𝑦 𝑦 = 𝑢 𝑦𝑦 𝑥 and 𝑢 𝑦 𝑥 𝑥 = 𝑢 𝑥 𝑥 𝑦 .

To show that 𝑢 𝑥 is harmonic, we notice that

(𝑢 𝑥 ) 𝑥 𝑥 + (𝑢 𝑥 ) 𝑦𝑦 = 𝑢 𝑥 𝑥 𝑥 + 𝑢 𝑥 𝑦𝑦 = (𝑢 𝑥 𝑥 + 𝑢 𝑦𝑦 ) 𝑥 = 0.

Thus, by definition, 𝑢 𝑥 is harmonic. Similarly, to show that 𝑢 𝑦 is harmonic, we note


that
(𝑢 𝑦 ) 𝑥 𝑥 + (𝑢 𝑦 ) 𝑦𝑦 = 𝑢 𝑦 𝑥 𝑥 + 𝑢 𝑦𝑦𝑦 = (𝑢 𝑥 𝑥 + 𝑢 𝑦𝑦 ) 𝑦 = 0.
Thus, 𝑢 𝑦 is also harmonic and we are done. ♦
Problem VI.2 If 𝑢 is harmonic, show that 𝑓 = 𝑢 𝑥 − 𝑖𝑢 𝑦 is analytic.

Solution VI.2 By definition, if 𝑢 is harmonic then

𝑢 𝑥 𝑥 + 𝑢 𝑦𝑦 = 0.

Again since 𝑢 is harmonic, 𝑢 is infinitely differentiable. In particular, the second


partial derivatives of 𝑢 exist and are continuous, that is, 𝑢 𝑥 𝑦 = 𝑢 𝑦 𝑥 . Now, to show
that 𝑓 = 𝑢 𝑥 − 𝑖𝑢 𝑦 is analytic, it is enough to show the following three conditions are
true:
• ℜ( 𝑓 ) = 𝑢 𝑥 is harmonic;
• ℑ( 𝑓 ) = −𝑢 𝑦 is harmonic;

47
48 VI Harmonic and Entire Functions

• 𝑢 𝑥 and −𝑢 𝑦 satisfy the Cauchy – Riemann Equations.


By Problem – VI.1, we already have shown that 𝑢 𝑥 and 𝑢 𝑦 (hence −𝑢 𝑦 ) are harmonic.
To show that 𝑢 𝑥 and −𝑢 𝑦 satisfy the Cauchy – Riemann Equations, we note that

(𝑢 𝑥 ) 𝑥 = −(𝑢 𝑦 ) 𝑦 ⇐⇒ 𝑢 𝑥 𝑥 + 𝑢 𝑦𝑦 = 0.

Again, we also notice

(𝑢 𝑥 ) 𝑦 = −(−𝑢 𝑦 ) 𝑥 ⇐⇒ 𝑢 𝑥 𝑦 = 𝑢 𝑦 𝑥 .

Thus, the Cauchy – Riemann Equations are satisfied and so we are done. ♦
Problem VI.3 Let 𝑢 be harmonic in 𝐺 and suppose 𝐵(𝑎; 𝑅) ⊂ 𝐺. Show that

1
𝑢(𝑎) = 𝑢(𝑥, 𝑦) 𝑑𝑥 𝑑𝑦 .
𝜋𝑅 2 𝐵(𝑎;𝑅)

Solution VI.3 Let D be a disk such that 𝐵(𝑎; 𝑅) ⊂ D ⊂ 𝐺. There exist an 𝑓 ∈ H (D)
such that ℜ( 𝑓 ) = 𝑢 on D. From Cauchy’s Integral Formula, we have
∫ ∫ 2𝜋
1 𝑓 (𝑤) 1
𝑓 (𝑎) = 𝑑𝑤 = 𝑓 (𝑎 + 𝑟e𝑖 𝜃 ) 𝑑𝜃,
2𝜋𝑖 𝛾 𝑤−𝑎 2𝜋 0

where 𝛾(𝑡) = 𝑎 + 𝑟e𝑖 𝜃 and 𝑡 ∈ [0, 2𝜋]. Now, multiplying by 𝑟 and then integrating
both sides w.r.t. 𝑟, we obtain
∫ 𝑅 ∫ 𝑅 ∫ 2𝜋
1
𝑟 𝑓 (𝑎) 𝑑𝑟 = 𝑟 𝑓 (𝑎 + 𝑟e𝑖 𝜃 ) 𝑑𝜃 𝑑𝑟
0 2𝜋 0 0
∫ 𝑅 ∫ 2𝜋
1 𝑅 1
⇐⇒ 𝑟 2 𝑓 (𝑎) = 𝑟 𝑓 (𝑎 + 𝑟e𝑖 𝜃 ) 𝑑𝜃 𝑑𝑟
2 0 2𝜋 0 0
∫ 𝑅 ∫ 2𝜋
1
⇐⇒ 𝑓 (𝑎) = 𝑟 𝑓 (𝑎 + 𝑟e𝑖 𝜃 ) 𝑑𝜃 𝑑𝑟
𝜋𝑅 2 0 0

1
⇐⇒ 𝑓 (𝑎) = 𝑓 (𝑥, 𝑦) 𝑑𝑥 𝑑𝑦,
𝜋𝑅 2 𝐵(𝑎;𝑅)

where the last equality follows by a change of coordinates from polar to rectangular.
Now, we note
∬ ∬
1 1
𝑢(𝑎) = ℜ( 𝑓 (𝑎)) =
𝜋𝑅 2 𝐵(𝑎;𝑅)
ℜ( 𝑓 (𝑥, 𝑦)) 𝑑𝑥 𝑑𝑦 =
𝜋𝑅 2 𝐵(𝑎;𝑅)
𝑢(𝑥, 𝑦) 𝑑𝑥 𝑑𝑦 . ♦
Problem VI.4 Let 𝑢 : 𝐺 → R be harmonic and let 𝐴 = {𝑧 ∈ 𝐺 : 𝑢 𝑥 (𝑧) = 𝑢 𝑦 (𝑧) =
0}. Can 𝐴 have a limit point in 𝐺 ?
Solution VI.4 Let 𝑢 be a nonconstant harmonic function. By Problem – VI.1, we
know that 𝑢 𝑥 and 𝑢 𝑦 are harmonic, and by Problem – VI.2, we know that 𝑓 = 𝑢 𝑥 −𝑖𝑢 𝑦
is analytic. Define
VI Harmonic and Entire Functions 49

𝐵 = {𝑧 ∈ 𝐺 : 𝑓 (𝑧) = 0} = {𝑧 ∈ 𝐺 : 𝑢 𝑥 (𝑧) − 𝑖𝑢 𝑦 (𝑧) = 0}.

Then we note that 𝐴 = 𝐵 since 𝑢 𝑥 (𝑧) = 𝑢 𝑦 (𝑧) = 0 ⇐⇒ 𝑓 (𝑧) = 0. Since 𝑓 is


analytic, the zeros are isolated and therefore 𝐵 cannot have a limit point in 𝐺. This
implies that 𝐴 cannot have a limit point in 𝐺.
But if 𝑢 is a constant harmonic function, then we must have 𝐴 = 𝐵 = 𝐺, and 𝑓 ≡ 0.
Thus, 𝐴 must have a limit point in 𝐺. ♦
Problem VI.5 In the hypothesis of Jensen’s Formula, do not suppose that 𝑓 (0) ≠ 0.
Show that if 𝑓 has a zero at 𝑧 = 0 of multiplicity 𝑚 then
∫ 2𝜋
𝑓 (𝑚) (0)
𝑛  
∑︁ 𝑟 1  
𝑚 log 𝑟 + log =− log + log 𝑓 𝑟e𝑖 𝜃 𝑑𝜃 .
𝑚! 𝑘=1
|𝑎 𝑘 | 2𝜋 0

𝑟 (𝑧 − 𝑎 𝑘 )
Solution VI.5 We know that the map maps 𝐵(0; 𝑟) onto itself. We let
𝑟 2 − 𝑧 𝑎¯ 𝑘
𝑛
Ö 𝑟 2 − 𝑧 𝑎¯ 𝑘
𝑟𝑚
𝐹 (𝑧) = 𝑓 (𝑧) .
𝑧𝑚 𝑘=1
𝑟 (𝑧 − 𝑎 𝑘 )

Then 𝐹 ∈ H (𝐺), 𝐹 (𝑧) ≠ 0 ∀𝑧 ∈ 𝐵(0; 𝑟), and |𝐹 (𝑧)| = | 𝑓 (𝑧)| whenever |𝑧| = 𝑟.
Thus, by Jensen’s Formula, we have
∫ 2𝜋 ∫ 2𝜋
1 
𝑖𝜃
 1  
log |𝐹 (0)| = log 𝐹 𝑟e 𝑑𝜃 = log 𝑓 𝑟e𝑖 𝜃 𝑑𝜃 .
2𝜋 0 2𝜋 0

Since 𝑓 has a zero at 𝑧 = 0 of multiplicity 𝑚, we also notice that



𝑓 (𝑧) 1 ∑︁ 𝑓 (𝑘 ) (0) 𝑘
𝑔(𝑧) ≡ = 𝑧
𝑧𝑚 𝑧 𝑚 𝑘=0 𝑘!

1 ∑︁ 𝑓 (𝑘 ) (0) 𝑘
= 𝑧
𝑧 𝑚 𝑘=𝑚 𝑘!

∑︁ 𝑓 (𝑘 ) (0) 𝑘−𝑚
= 𝑧 .
𝑘=𝑚
𝑘!
𝑛  
𝑚
Ö 𝑟
∴ |𝐹 (0)| = |𝑔(0)| 𝑟 .
𝑘=1
|𝑎 𝑘 |

Substituting this into our previous equation for log |𝐹 (0)|, we obtain
∫ 2𝜋
𝑓 (𝑚) (0) ∑︁
𝑛  
𝑟 1  
𝑚 log 𝑟 + log + log = log 𝑓 𝑟e𝑖 𝜃 𝑑𝜃
𝑚! 𝑘=1
|𝑎 𝑘 | 2𝜋 0
∫ 2𝜋
𝑓 (𝑚) (0)
𝑛  
1

∑︁ 𝑟  
⇐⇒ 𝑚 log 𝑟 + log =− log + log 𝑓 𝑟e𝑖 𝜃 𝑑𝜃 .
𝑚! 𝑘=1
|𝑎 𝑘 | 2𝜋 0

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy