Taylor JL Complex Variables
Taylor JL Complex Variables
Sally
The UNDERGRADUATE TEXTS 16
SERIES
Complex
Variables
Joseph L. Taylor
Complex
Variables
Joseph L. Taylor
2010 Mathematics Subject Classification. Primary 30–01, 30Axx, 30Bxx, 30Dxx, 30Exx.
QA331.7.T389 2011
515.9—dc23
2011019541
Copying and reprinting. Individual readers of this publication, and nonprofit libraries
acting for them, are permitted to make fair use of the material, such as to copy a chapter for use
in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Requests for such
permission should be addressed to the Acquisitions Department, American Mathematical Society,
201 Charles Street, Providence, Rhode Island 02904-2294 USA. Requests can also be made by
e-mail to [email protected].
c 2011 by the American Mathematical Society. All rights reserved.
Printed in the United States of America.
∞ The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at http://www.ams.org/
10 9 8 7 6 5 4 3 2 1 16 15 14 13 12 11
Contents
Preface vii
v
vi Contents
Basic complex variables is a very popular subject among mathematics faculty and
students. There is one big theorem (the Cauchy Integral Theorem) with a somewhat
difficult proof, but this is followed by a host of easy to prove consequences that are
both surprising and powerful. Furthermore, for undergraduates, the subject is
refreshingly new and different from the analysis courses which precede it.
Undergraduate complex variables is our favorite course to teach, and we teach
it as often as we can. Over a period of years we developed notes for use in such a
course. The course is a one-semester undergraduate course on complex variables at
the University of Utah. It is designed for junior level students who have completed
three semesters of calculus and have also had some linear algebra and at least one
semester of foundations of analysis.
Over the past several years, these notes have been expanded and modified to
serve other audiences as well. We have used the expanded notes to teach a course
in applied complex variables for engineering students as well a course at the first
year graduate level for mathematics graduate students. The topics covered, the
emphasis given to topics, and the choice of exercises were different for each of these
audiences.
When taught as a one-semester junior level course for mathematics students,
the course is a transitional course between freshman and sophomore level calculus,
linear algebra, and differential equations and the much more sophisticated senior
level mathematics courses taught at Utah. The students are expected to understand
definitions and proofs, and the exercises assigned will include proofs as well as
computations. The course moves at a leisurely pace, and the material covered
includes only Chapters 1, 2, and 3 and selected sections from Chapters 4, 5, and 6.
A full year course could easily cover the entire text.
When we teach a one-semester undergraduate course for engineers using these
notes, we cover essentially the same material as in the course for mathematics
majors, but not all the proofs are done in detail, and there is more emphasis on
vii
viii Preface
The Real Numbers are Insufficient. The complex number system was devel-
oped in response to the need for solutions to polynomial equations. The simplest
polynomial equation that does not have a solution in the real number system is the
equation
x2 + 1 = 0,
which has no real solution because −1 has no real square root. More generally, a
quadratic equation
(1.1.1) ax2 + bx + c = 0,
1
2 1. The Complex Numbers
where a, b and c are real numbers, formally has two solutions given by the quadratic
formula
√
−b ± b2 − 4ac
(1.1.2) x= ,
2a
but these will not be real numbers if b2 − 4ac is negative. If we could take square
roots of negative numbers, then the quadratic formula would give us solutions to
(1.1.1) for all choices of real coefficients a, b, c. To make this possible, we expand
the real number system in the following way, thus creating the complex number
system C.
2i . 3+2i
−1 0 1 2 3 4
−i
The properties described in the above theorem are some of the properties that
must hold in a field. A field must also have additive and multiplicative identities –
that is, elements 0 and 1 which satisfy
(1.1.3) z+0=z
and
(1.1.4) 1·z =z
for every element z in the field. That this holds for C follows immediately from the
fact that C is a vector space over R.
A field must also have the properties that every element z has an additive
inverse, that is, an element −z such that
(1.1.5) z + (−z) = 0,
1.1. Definition and Simple Properties 5
and every non-zero element z has a mutiplicative inverse, that is, an element z −1
such that
(1.1.6) z · z −1 = 1.
The first of these follows immediately from the fact that C is a vector space
over R. The second is nearly as easy. If z = x + iy = 0, then a direct calculation
shows that
x − iy x y
z −1 = 2 2
= 2 2
− 2 i
x +y x +y x + y2
satisfies (1.1.6). We conclude:
Theorem 1.1.5. With addition and multiplication defined as in Definition 1.1.1,
the complex numbers form a field.
Note that the modulus, as defined above, is just the usual Euclidean norm in
the vector space R2 . Thus, if z1 , z2 ∈ C, then |z1 − z2 | is the Euclidean distance
from z1 to z2 . The term modulus is traditional, but the terms norm and absolute
value are also commonly used to mean the same thing. We will use all three.
Note also that the two solutions of a quadratic equation with real coefficients
given in Example 1.1.3 are complex conjugates of each other. Thus, the solutions
to a quadratic equation with real coefficients occur in conjugate pairs. Quadratic
equations with complex coefficients also have roots and they are also given by the
quadratic formula. However, we cannot prove this until we prove that every complex
number has a square root. In fact, in Section 1.4 we will prove that every complex
number has roots of all orders.
For a complex number z = x + iy, the real number x is called the real part
of z and is denoted Re(z), while the number y is called the imaginary part of z
and is denoted Im(z). In graphing complex numbers using a rectilinear coordinate
system, x determines the coordinate on the horizontal axis, while y determines the
coordinate on the vertical axis.
Note that a complex number z is real if and only if z = z, and it is purely
imaginary if and only if z = −z. Note also, that if z = x + iy, then
z+z z−z
Re(z) = x = and Im(z) = y = .
2 2i
The elementary properties of conjugation and modulus are gathered together in the
next theorem.
Theorem 1.1.7. If z and w are complex numbers, then
(a) z = z;
6 1. The Complex Numbers
−z = −x+iy
. . z = x+iy
. .
−z = −x−iy z = x−iy
(b) zz = |z|2 ;
(c) z + w = z + w;
(d) zw = z w;
(e) |zw| = |z||w| and |z| = |z|;
(f) |z| is a non-negative real number and is 0 if and only if z = 0;
(g) | Re(zw)| ≤ |z||w|;
(h) |z + w| ≤ |z| + |w|.
Proof. We will prove (g) and (h). The other parts are elementary computations
or observations and will be left as exercises.
Parts (g) and (h) are the Cauchy-Schwarz inequality and the triangle inequality
for the vector space R2 . Versions of these inequlities hold in general Euclidian space
Rn . The proofs we give here are specializations to C of the standard proofs of these
inequalities in Rn .
To prove (g), we begin with the observation that (a) and (d) imply that
zw + zw = 2 Re(zw).
We then let t be an arbitrary real number and note that, by Parts (c), (d), and (f),
0 ≤ |zt + w|2 = (zt + w)(zt + w) = |z|2 t2 + 2 Re(zw)t + |w|2
for all values of t. This implies that the quadratic polynomial in t given by
|z|2 t2 + 2 Re(zw)t + |w|2
is never negative and, therefore, has at most one real root. This is only possible if
the expression under the radical in the quadratic formula is negative or zero. Thus,
4(Re(zw))2 − 4|z|2 |w|2 ≤ 0.
Part (g) follows immediately from this.
1.1. Definition and Simple Properties 7
Part (h) follows directly from Part (g) and the other parts of the theorem. In
fact,
|z + w|2 = (z + w)(z + w)
= |z|2 + 2 Re(zw) + |w|2
≤ |z|2 + 2|z||w| + |w|2 = (|z| + |w|)2 .
Inversion and Division. Recall that the inverse of a non-zero complex number
z = x + iy is
x − iy z z
z −1 = 2 = 2 = .
x +y 2 |z| zz
Stating this in the last form makes the identity zz −1 = 1 obvious.
This also suggests the right way to do complex division problems in general: to
express w/z as a complex number in standard form (as a real number plus i times
a real number), simply multiply both numerator and denominator by z. That is,
w wz wz
= = 2.
z zz |z|
The number wz is then easily put in standard form and the problem is finished by
dividing by the real number |z|2 .
1
Example 1.1.9. Express in the standard form x + yi.
2 + 3i
Solution:
1 2 − 3i 2 − 3i 2 3
= = = − i.
2 + 3i (2 + 3i)(2 − 3i) |2 + 3i|2 13 13
3 + 4i 3 − 4i
Example 1.1.10. Express and in standard form.
3 − 4i 3 + 4i
Solution:
3 + 4i (3 + 4i)2 7 24
= = − + i,
3 − 4i |3 + 4i|2 25 25
3 − 4i (3 − 4i)2 7 24
= = − − i.
3 + 4i |3 − 4i|2 25 25
8 1. The Complex Numbers
1.2. Convergence in C
We assume the reader is familiar with the basics concerning convergent sequences
and series of real numbers – particularly those results which follow from the com-
pleteness of the real number system, such as the fact that bounded monotone
sequences converge and the various convergence tests for series. We also assume
a familiarity with the basics of power series in a real variable. The purpose of
1.2. Convergence in C 9
this section is to extend these ideas and results to sequences and series of complex
numbers and power series in a complex variable. This is an introductory section.
A deeper study of complex power series will come later in the text.
There is no natural order relation on the complex numbers. The statement
z < w makes no sense for complex numbers z and w. However, if these numbers
happen to be real, then the inequality does make sense, because R is an ordered
field. We make heavy use of inequalities in this section, but note they are always
inequalities between real numbers. Thus, if an inequality of the form a < b occurs
in this text the numbers a and b are assumed to be real numbers, even if there is
no explicit statement to that effect.
The first of these observations reduces the problem of showing that a sequence
of complex numbers converges to a given complex number to showing that a certain
sequence of non-negative real numbers converges to 0. This is useful because we
have many tools at our disposal to show that a sequence of non-negative numbers
converges to 0. One of the most useful of these tools is the second observation
above. The next example and the proof of the following theorem are excellent
examples of how this works.
Example 1.2.3. Prove that lim z n = 0 if |z| < 1.
Solution: Note that |z n | = |z|n follows from Part (e) of Theorem 1.1.7. If
|z| < 1 then lim |z|n = 0. That lim z n = 0, as well, follows from (1) of Remark
1.2.2.
Theorem 1.2.4. A sequence of complex numbers {zn } converges to a complex
number w if and only if {Re(zn )} converges to Re(w) and {Im(zn )} converges to
Im(w).
10 1. The Complex Numbers
divergence is called the term test. Its proof for complex series is the same as the
proof for real series. We leave it as an exercise (Exercise 1.2.9).
Example 1.2.7. For what values of z does the complex geometric series
∞
zk
k=0
converge and what does it converge to?
Solution: We first note that if |z| ≥ 1, then |z|k ≥ 1 for all k and the series
diverges by the term test. For |z| < 1 we use the same trick that is used to study
the real geometric series. The nth partial sum of the series is
n
sn = zk .
k=0
∞
k=0 zk are convergent and, by Theorem 1.2.4, so is the sequence {sn } itself. Thus,
the series converges.
It follows from the triangle inequality that
n
n
|sn | = zk ≤ |zk |.
k=0 k=0
The inequality of the theorem follows when we pass to the limit as n → ∞ in this
inequality and use the result of Exercise 1.2.7.
∞ 2 −1
Example 1.2.9. Prove that the complex series k=1 (k + k i) converges abso-
lutely.
Solution: By the second form of the triangle inequality (see Exercise 1.2.2)
we have
k2
|k + k2 i| ≥ k2 − k ≥ if k ≥ 2.
2
Thus,
|k + k2 i|−1 ≤ 2k−2 if k ≥ 2.
∞ −2 ∞ −2
Since
∞ the p-series k=1 k
converges, so does k=1 2k and, by comparison,
2 −1 ∞ 2 −1
k=1 |k + k i| . Thus, k=1 (k + k i) converges absolutely.
there are elementary ways to show that this result holds and to find R. One simply
uses the standard convergence tests from calculus – particularly the ratio test.
Example 1.2.11. Find the radius of convergence of the series
∞
zn
(1.2.3) .
n=1
n
Solution: If we apply the ratio test to the series ∞ n=1 |z| /n, we conclude
n
that this series converges for |z| < 1 and diverges for |z| > 1. Hence, by Theorem
1.2.8, the series (1.2.3) also converges for |z| < 1. If |z| > 1, then the sequence of
terms of this series fails to converge to zero (since |z|n /n → +∞) and so the series
diverges by the term test. Thus our series converges on D1 (0) and diverges outside
of D1 (0). We conclude that it has radius of convergence 1.
Example 1.2.12. Show that the radius of convergence of the series
∞
zn
(1.2.4)
n=0
n!
is +∞ – that is, the series converges for all z.
∞
Solution: We apply the ratio test to the series n=0 |z|n /n!. We have
n −1
|z|n+1 |z| |z|
=
(n + 1)! n! n+1
which converges to 0 for all z. Hence, by the ratio test, the power series (1.2.4)
converges for all z ∈ C. The radius of convergence is, thus, +∞.
Example 1.2.13. Find the radius of convergence of the power series
∞
an z n
n=0
where an = 2n if n is prime and an = 0 if n is not prime.
Solution: We cannot use the ratio test on this onebecause most of the terms
∞
of
∞ series
the are 0. However, if we compare the series n=0 an |z|n with the series
n=0 2 |z| – a series to which we can apply the ratio test – we conclude that the
n n
series converges for |z| < 1/2. If |z| > 1/2, then the terms of our series do not tend
to 0 and so the series diverges by the term test. We conclude that the radius of
convergence is 1/2.
Proof. The proof uses the complex form of the binomial formula:
n
n!
(1.3.1) (z + w)n = z j wn−j .
j=0
j!(n − j)!
and collect terms of degree n. Provided this operation is valid, we conclude that
ez+w = ez ew . It turns out that it is valid to expand the product of two infinite
series in this fashion if they are both absolutely convergent. We prove this in the
following lemma, which will complete the proof of the theorem.
∞ ∞
Lemma 1.3.3. Let j=0 aj and k=0 bk be two absolutely convergent series of
complex numbers. Then
⎛ ⎞ ⎛ ⎞
∞ ∞ ∞
(1.3.3) ⎝ aj ⎠ bk = ⎝ aj bk ⎠ .
j=0 k=0 n=0 j+k=n
The left side of (1.3.3) is, by definition, (lim sJ )(lim tK ), while the right side is
lim uN . We know lim sJ and lim tK both exist since the series defining them con-
verge absolutely. We must prove that lim uN exists and equals (lim sJ )(lim tK ).
16 1. The Complex Numbers
This shows that, not only is eiy a point on the unit circle, it is the point which
is reached by rotating through an angle y (measured in radians) from the initial
point (1, 0).
Example 1.3.5. Express the complex numbers e2πi , eπi and eπi/2 in standard
form.
1.3. The Exponential Function 17
There are a number of properties of the exponential function which follow easily
from this characterization of ez . We collect them together in the following theorem
whose proof is left to the exercises.
Theorem 1.3.7. The exponential function has the following properties:
(a) ez is never 0;
(b) | ez | = eRe(z) ;
(c) | ez | ≤ e|z| ;
(d) ez is periodic of period 2πi, meaning ez+2πi = ez for every z ∈ C;
(e) ez = 1 if and only if z = 2πni for some integer n.
Example 1.3.9. How would you define a complex version of the arctan function?
Solution: The power series which converges to arctan x on (−1, 1) is
∞
x2n+1
arctan x = (−1)n .
n=0
2n + 1
If we replace x by the complex variable z, we obtain a series which converges on
the disc D1 (0). We then use it to define arctan z on this disc:
∞
z 2n+1
arctan z = (−1)n .
n=0
2n + 1
15. How would you define a complex version of the function log(1 + x)?
16. Verify that the power series defining arctan in Example 1.3.9 converges on the
disc D1 (0).
Example 1.4.1. Find the polar form for the complex number z = 2 + 2i.
Solution: We have
√
r = 22 + 22 = 2 2 and θ = arctan(1) = π/4.
√
Thus, z = 2 2 eπi/4 is the polar form of z.
Example 1.4.2. Put the complex number z = 2 eπi/6 in standard form x + iy.
Solution: We have
√
x = 2 cos π/6 = 3 and y = 2 sin π/6 = 1
√
and so z = 3 + i.
20 1. The Complex Numbers
r e iθ
r
θ
Products, Powers, and Roots. Polar form is particularly useful in dealing with
products of complex numbers, since the product has such a simple expression if the
numbers are given in polar form. If z1 = r1 eiθ1 and z2 = r2 eiθ2 , then the law of
exponents implies that
(1.4.1) z1 z2 = r1 r2 ei(θ1 +θ2 ) .
Thus, z1 z2 is the complex number whose norm (distance from the origin) |z1 z2 | is
the product of |z1 | and |z2 | and whose argument is the sum of the arguments of z1
and z2 .
Similarly, the quotient of z1 and z2 is given by
(1.4.2) z1 /z2 = (r1 /r2 ) ei(θ1 −θ2 ) .
Example 1.4.3. Find z1 z2 and z1 /z2 if z1 = 2 eπi/3 and z2 = 3 e2πi/3 .
Solution: By (1.4.1) and (1.4.2), we have
√
z1 z2 = 6 eπi = −6 and z1 /z2 = (2/3) e−πi/3 = 1/3 − i 3/3.
It is evident from (1.4.1) that the nth power of a complex number z = r eiθ is
(1.4.3) z n = r n einθ .
From this we conclude that if z = r eiθ , and we choose w = r 1/n eiθ/n , then
n
w = z. Thus, w is an nth root of z. It is not the only one, however. Since z can
also be written as z = ei(θ+2πk) for any integer k, each of the numbers
wk = r 1/n ei(θ/n+2πk/n) ,
where k is an integer, is also an nth root of z. Of course, these numbers are not all
different. Those whose arguments differ by an integral multiple of 2π are the same.
The numbers w0 , w1 , w2 , · · · , wn−1 are all distinct, but every other wk is equal to
one of these. This proves the following theorem.
Theorem 1.4.4. If z = r eiθ is a non-zero complex number, then z has exactly n
nth roots. They are the numbers
r 1/n ei(θ/n+2πk/n) for k = 0, 1, 2, · · · , n − 1.
1.4. Polar Form for Complex Numbers 21
−1/2+i 3/2
−1/2−i 3/2
Note that the n numbers ei(θ/n+2πk/n) that appear in this result are evenly
spaced around the unit circle, with each successive pair separated by an angle of
2π/n.
The nth roots of the number 1 play a special role. They are called the nth roots
of unity. If we apply the above theorem in the special case where r = 1 and θ = 0,
it tells us that the nth roots of unity are the numbers
(1.4.4) e2πki/n for k = 0, 1, 2, · · · , n − 1.
Example 1.4.5. Find the cube roots of unity.
Solution: In the particular case where n = 3, (1.4.4) tells us that the roots of
unity are
e0 = 1,
√
e2πi/3 = −1/2 + ( 3/2)i, and
√
e4πi/3 = −1/2 − ( 3/2)i.
Example 1.4.6. Find the cube roots of 2i.
Solution: Since 2i = 2 eπi/2 , Theorem 1.4.4 implies that the cube roots of 2i
are
√
21/3 eπi/6 = 21/3 ( 3/2 + i/2),
√
21/3 ei(π/6+2π/3) = 21/3 e5πi/6 = 21/3 (− 3/2 + i/2), and
21/3 ei(π/6+4π/3) = 21/3 e3πi/2 = −21/3 i.
The Logarithm. If z = r eiθ , and log r is the natural logarithm of the positive
number r, then the law of exponents implies that
z = elog r+iθ .
Thus, it would make sense to define log z to be log r + iθ = log |z| + i arg z. There is
a problem with this, however. There are infinitely many possible choices for arg z,
and so log z is not well defined, just as arg z is not well defined.
The solution to this problem is to restrict θ = arg z to lie in a specific half-open
interval of length 2π.
22 1. The Complex Numbers
z iπ
iπ/2
z log(z)
0 1 2 0 1 2
−iπ/2
−iπ
The following properties of the various branches of the log function follow easily
from the definition and the properties of the exponential function. The proofs are
left to the exercises.
Theorem 1.4.8. If log is the branch of the the log function determined by an
interval I, then
(a) if z = 0, then elog z = z;
(b) if z ∈ C, then log ez = z + 2πki for some integer k;
(c) if z, w ∈ C, then log zw = log z + log w + 2πki for some integer k;
(d) log 1 = 2πki for some integer k;
(e) log agrees with the ordinary natural log function on the positive real numbers
if and only if the interval I contains 0.
Suppose the interval I defining a branch of the log function has endpoints a
and b with a < b. Then, since b − a = 2π, the polar coordinate equations θ = a
and θ = b define the same ray. This ray is called the cut line for this branch of the
logarithm. Observe that if z and w are two complex numbers with |z| = |w| > 0
which are close to each other, but on opposite sides of the cut line – say with argI z
near a and argI w near b – then log w − log z is nearly 2πi. In other words, as we
cross the cut line moving in the clockwise direction, the value of log jumps by 2πi.
Thus, log is not continuous at points on the cut line. Later we will show that it is
continuous everywhere else.
1.4. Polar Form for Complex Numbers 23
Note that the nth root function, as defined by (1.4.5), is giving only one of the
nth roots of z. Which one is determined by the branch of the log function that is
used. The other nth roots are obtained by multiplying this one by the nth roots
of unity (Exercise 1.4.9). For example, given one square root of z, the other is
obtained by multipling it by −1.
√
Example 1.4.9. If z √ is defined using the principal branch of the log function,
analyze the behavior of z near the cut line for this branch.
Solution: For the principal branch of the log function the interval I is (−π, π]
and so the cut line is the line defined by θ = π in polar coordinates – that is, it is
the negative real axis. If z = r eiθ is just above the negative real axis, then log z is
nearly log r + iπ and
√
z = e(1/2) log z
√
is nearly i r. On the other hand, if z is just below the negative
√ axis, then √
log z is
nearly log r − iπ, (1/2) log z is close to √(log r − iπ)/2 and z is close to −i r. In
other words, as z crosses the cut line, z jumps from one square root of z to its
negative, which is the other square root of z.
√
Example 1.4.10. Analyze the function z 2 + 1, where the square root function
is defined, as above, using the principal branch of the log function.
√
Solution: The cut√ line for z is the same as for log – the half-line of negative
reals. The function z jumps from values on the positive imaginary axis to their
negatives as z crosses this line in the counterclockwise direction. The number z 2 +1
crosses the negative real half-line in the counterclockwise direction as z crosses either
(i, i∞) or (−i∞, −i) in the counterclockwise direction. √ Thus, these two half-lines
on the imaginary axes are where discontinuities of z 2 + 1 occur. As either of
these half-lines is crossed in the counterclockwise direction, a typical value of this
function jumps from a number it with t > 0 to −it (see Figure 1.4.4).
−it it
it −it
√
Figure 1.4.4. The Cut Line Discontinuities of z 2 + 1.
Note that for any function defined in terms of a branch log z of the log func-
tion, we can expect trouble along the cut line for log. Since, log has a jump or
discontinuity as we cross the cut line, we can expect the same for functions defined
in terms of it.
Also, we emphasize that functions defined this way depend on the choice of a
branch of the log function. If one wants such a function to agree with the standard
one on the positive real axis, then one must choose a branch of the log function
which is the ordinary natural logarithm on the positive real numbers. By Theorem
1.4.8, Part(e), this happens if and only if the interval I, in which arg(z) is required
to lie, contains 0. In particular, the principal branch has this property.
11. For the principal branch of the log function, find log(1 − i).
12. Find log(1 − i) for the branch of the log function determined by the interval
[0, 2π).
13. Prove (a) and (b) of Theorem 1.4.8.
14. Prove (c), (d), and (e) of Theorem 1.4.8.
15. Analyze the function z i defined by (1.4.7) using the principal branch of the log
function. What kind of a jump, if any, does it have as z crosses the negative
real axis?
√
16. Analyze the function 1 − z 2 , where the square root function is defined by
the principal branch of the log function. Where does it have discontinuities
(jumps)?
17. Let the square root function be √ defined by the √ principal
√ branch of the log
function. Compare the functions z 2 − 1 and z + 1 z − 1. Where are the
discontinuities of each function?
1 − z n+1
18. The identity 1 + z + z 2 + · · · + z n = was derived in Example 1.2.7.
1−z
Use this to derive Lagrange’s trigonometric identity:
1 sin (n + 1/2)θ
1 + cos θ + cos 2θ + · · · + cos nθ = + .
2 2 sin θ/2
Hint: Take the real parts of both sides in the first identity.
Chapter 2
Analytic Functions
27
28 2. Analytic Functions
already knows what it means for such a function to be continuous. Still, we will
review the basics of continuity in this context, partly in order to set terminology
and notation that is peculiar to the complex variables context.
Open and Closed Sets. Recall that the open disc Dr (z0 ) and closed disc
Dr (z0 ), centered at z0 , with radius r > 0, are defined by
Dr (z0 ) = {z ∈ C : |z − z0 | < r} and Dr (z0 ) = {z ∈ C : |z − z0 | ≤ r}.
Open intervals and closed intervals on the real line play an important part in
calculus in one variable. Open and closed discs are the direct analogues in C of
open and closed intervals on the line. However, the geometry of the plane is much
more complicated than that of the line. We will need the concepts of open and
closed for sets that are far more complicated than discs. This leads to the following
definition.
Definition 2.1.1. If W is a subset of C, we will say that W is open if, for each
point w ∈ W , there is an open disc centered at w which is contained in W . We
will say that a subset of C is closed if its complement is open. A neighborhood of
a point z ∈ C is any open set which contains z.
It might seem obvious that open discs are open sets and closed discs are closed
sets. However, that is only because we have chosen to call them open discs and
closed discs. We actually have to prove that they satisfy the conditions of the
preceding definition. We do this in the next theorem.
Theorem 2.1.2. We have:
(a) the empty set ∅ is both open and closed;
(b) the whole space C is both open and closed;
(c) each open disc is open;
(d) each closed disc is closed.
Proof. The empty set ∅ is open because it has no points, and so the condition that
a set be open, stated in Definition 2.1.1, is vacuously satisfied. The set C is open
because it contains any open disc centered at any of its points. Thus, ∅ and C are
both open. Since they are complements of one another, they are also both closed.
To prove (c), we suppose Dr (z0 ) is an open disc and w is one of its points.
Then |w − z0 | < r and so, if we set s = r − |w − z0 |, then s > 0. Also, if z ∈ Ds (w),
then |z − w| < s and so
|z − z0 | ≤ |z − w| + |w − z0 | < s + |w − z0 | = r,
and so z ∈ Dr (z0 ) (see Figure 2.1.1). Thus, we have shown that, for each w ∈
Dr (z0 ), there is an open disc, Ds (w), centered at w, which is contained in Dr (z0 ).
By definition, this means that Dr (z0 ) is open. This completes the proof of (c).
To prove (d), we consider a closed disc Dr (z0 ). To prove that it is a closed set,
we must show its complement is open. Suppose w is a point in its complement. This
means w ∈ C but w ∈ / Dr (z0 ), and so |w − z0 | > r. This time we set s = |w − z0 | − r
and we claim that the open disc Ds (w) is contained in the complement of Dr (z0 ).
2.1. Continuous Functions 29
s w
r s r
w
z0 z0
In fact, if z ∈ Ds (w), then |z − w| < s and so, by the second form of the triangle
inequality (Exercise 1.2.2),
|z − z0 | ≥ |w − z0 | − |z − w| > |w − z0 | − s = r,
which means z is in the complement of Dr (z0 ). Thus, we have proved that each
point of the complement of Dr (z0 ) is the center of an open disc contained in the
complement of Dr (z0 ). This proves that this complement is open, hence, that
Dr (z0 ) is closed.
The next theorem shows that the collection of all open subsets of C forms what
is called a topology for C. It says that the collection of open subsets is closed under
arbitrary unions and finite intersections.
Theorem 2.1.3. The union of an arbitrary collection of open sets is open, while
the intersection of any finite collection of open sets is open. On the other hand,
the intersection of an arbitrary collection of closed sets is closed, while the union
of any finite collection of closed sets is closed.
Proof. If z0 ∈ C, V is an arbitrary collection of open sets, and U = V is its
union, then z0 is in U if and only if it is in at least one of the sets in V. Suppose, it
is in V ∈ V. Then, since V is open, there is a disc Dr (z0 ), centered at z0 , which is
contained in V . Then this disc is also contained in U . This proves that U is open.
Now suppose {V1 , V2 , · · · , Vk } is a finite collection of open sets and
z0 ∈ U = V1 ∩ V2 ∩ · · · ∩ Vn .
Then, since each Vk is open, there exists for each k a radius rk such that Drk (z0 ) ⊂
Vk . If r = min{r1 , r2 , · · · , rn }, then Dr (z0 ) ⊂ Vk for every k, which implies that
Dr (z0 ) ⊂ U . It follows that U is open.
The proofs of the statements for closed sets follow from those for open sets by
taking complements. We leave the details to Exercise 2.1.1.
A consequence of the above theorem is: if U is open and K is closed, then the
set-theoretic difference U \ K is open. Similarly, K \ U is closed (Exercise 2.1.2).
Example 2.1.4. If 0 < r < 1, prove that the annulus A = {z ∈ C : r < |z| < R}
is open.
30 2. Analytic Functions
E E° E
Figure 2.1.2. The Set E of Example 2.1.7, its Interior E ◦ , and Closure E.
Solution: The disc DR (0) is open, the disc Dr (0) is closed, and A is the
set-theoretic difference DR (0) \ Dr (0). Thus, by the previous remark, A is open.
Interior, Closure, and Boundary. If E is a subset of C, then E contains a
largest open subset, meaning an open subset of E that contains all other open
subsets of E. In fact, the union of all open subsets of E is open, by Theorem 2.1.3,
and is a subset of E which contains all open subsets of E. Similarly, the intersection
of all closed sets containing E is the smallest closed set containing E. Thus, the
following definition makes sense.
Definition 2.1.5. Let E be a subset of C. Then:
(a) the largest open subset of E is called the interior of E and is denoted E ◦ ;
(b) the smallest closed set containing E is called the closure of E and is denoted
E;
(c) the set E \ E ◦ is called the boundary of E and is denoted ∂E.
Recall that a neighborhood of a point z0 ∈ C is any open set containing z0 .
The proof of the following theorem is elementary and is left to the exercises.
Theorem 2.1.6. Let E be a subset of C and z an element of C. Then:
(a) z ∈ E ◦ if and only if there is a neighborhood of z that is contained in E;
(b) z ∈ E if and only if every neighborhood of z contains a point of E;
(c) z ∈ ∂E if and only if every neighborhood of z contains points of E and points
of the complement of E.
Example 2.1.7. Find the interior, closure and boundary for the set
E = {z ∈ C : |z| < 1, Im(z) ≥ 0} ∪ {−iy : y ∈ [0, 1]}.
Solution: It is immediate from the previous theorem that
E ◦ = {z ∈ C : |z| < 1, Im(z) > 0},
E = {z ∈ C : |z| ≤ 1, Im(z) ≥ 0} ∪ {−iy : y ∈ [0, 1]},
∂E = {z ∈ C : |z| = 1, Im(z) ≥ 0} ∪ [−1, 1] ∪ {−iy : y ∈ [0, 1]}.
2.1. Continuous Functions 31
The condition “z ∈ E” in this definition means that whether or not the limit
exists may depend on E. That is, given a function f defined on a set E and a
subset D of E, we can always consider a new function f |D which is f restricted
to the new domain D. It may be that the limit as z → a does not exist for f as
defined on its original domain E, but the limit of f |D does exist.
Example 2.1.9. Show that the function f (z) = z/|z| with domain D1 (0) \ {0}
does not have a limit as z → 0, but if this function is restricted to the domain
(0, 1) ⊂ R, then its limit as z → 0 is 1.
Solution: If f is considered as a function with domain D1 (0) \ {0}, then for
each z = r eiθ in this domain, f (z) = eiθ . Thus f takes on every value of modulus
one in every open disc Dδ (0) with δ < 1. So there is certainly no one number
that f (z) is approaching as z → 0. On the other hand, suppose f is restricted to
(0, 1) ⊂ R. On this interval, f is identically 1 and so its limit as z → 0 is also 1.
there is a δ > 0 such that |f (z) − L| < whenever z ∈ E and 0 < |z − a| < δ. This
is equivalent to the statement
f (E ∩ Dδ (a) \ {a}) ⊂ D (L) ⊂ W.
Thus, if we choose V = Dδ (a)\{a}, then V is a deleted neighborhood of a satisfying
(2.1.1). This proves the “only if” part of the theorem.
To prove the “if” part, we assume that for every neighborhood W of L there
is a deleted neighborhood V of a such that f (E ∩ V ) ⊂ W . Then, given > 0,
let W = D (L) and let V be a deleted neighborhood of a satisfying (2.1.1). Since
V ∪ {a} is open, there exists a δ > 0 such that Dδ (a) ⊂ V ∪ {a}. Then
f (E ∩ Dδ (a) \ {a}) ⊂ f (E ∩ V ) ⊂ W = D (L).
This is equivalent to the statement that |f (z) − w| < whenever z ∈ E and
0 < |z − a| < δ. Thus, by definition, limz→a f (z) = L.
The definition of continuous function is also very domain dependent. For ex-
ample, if the function f (z) = z/|z| of Example 2.1.9 is given the value 1 at 0, then
it is continuous at 0 as a function with domain [0, 1), but if we consider its domain
to be D1 (0), then it is not continuous.
The next result characterizes functions which are defined and continuous on an
open set U as those functions f on U for which f −1 preserves open sets.
Theorem 2.1.13. If f is a complex-valued function defined on an open set U ⊂ C,
then f is continuous on U if and only if f −1 (W ) is open for every open set W ⊂ C.
If f and g are functions with domain a set E and they are both continuous at
a ∈ E, then f + g and f g are continuous at a, and f /g is continuous at a provided
g(a) = 0. The proofs of these facts for complex-valued functions of a complex
variable are no different than the proofs, given in calculus, of the corresponding
facts for functions of a real variable. Thus, we will accept them without further
comment.
Obviously, constants are continuous everywhere, as is the function z itself. It
follows that polynomials in z are also continuous everywhere.
If g with domain E is continuous at a and f with domain D is continuous at
b = f (a), and if g(E) ⊂ D, then the composite function f ◦ g, with domain E, is
continuous at a. This is another fact whose proof for functions of a complex variable
is no different than the proof from calculus of the analogous result for functions of
a real variable.
Example 2.1.14. Prove that if g is a continuous non-vanishing function on an
open subset U ⊂ C, then log |g| is continous on U .
Solution: The function |g(z)| is continuous on U because it is the composition
of the function g, which is continuous on U , and the function | · |, which is continous
everywhere. Also, |g(z)| has its values in the positive reals, since g is non-vanishing
on U . The function log is continuous on the positive reals and so the composition
log |g| is continuous on U .
1. Prove the second statement of Theorem 2.1.3. That is, prove that the intersec-
tion of an arbitrary collection of closed sets is closed, while the union of any
finite collection of closed sets is closed.
2. Prove that if U is open and K is closed, then the set-theoretic difference U \ K
is open, while K \ U is closed.
3. Prove Theorem 2.1.6.
4. Show that the set A = {z ∈ C : Re(z) > 0} is open. Hint: You must show that
each point w ∈ A is the center of an open disc which is entirely contained in A.
5. Tell which of the following sets are open subsets of C, which are closed, and
which are neither (no proof required):
(a) {z ∈ C : 1 < |z| < 2};
(b) {z ∈ C : Im(z) = 0, 0 < Re(z) < 1};
(c) {z ∈ C : −1 ≤ Re(z) ≤ 1}, −1 ≤ Im(z) ≤ 1}.
34 2. Analytic Functions
6. Find the interior, closure, and boundary for the set {z ∈ C : 1 ≤ |z| < 2} (no
proof required).
7. Prove that w ∈ C is in the closure of a set E ⊂ C if and only if there is a
sequence {zn } ⊂ E such that lim zn = w. Thus, a set E is closed if and only if
it contains all limits of convergent sequences of points in E.
|z − z|
8. Does limz→0 f (z) exist if f (z) = with domain C \ {0}? How about if
|z|
the domain is restricted to be just R \ {0}?
9. Prove that Re(z), Im(z), and z are continuous functions of z.
10. At which points of C is the function (1 − z 4 )−1 continuous.
11. Prove that argI is continuous except on its cut line.
12. Use the result of the preceding exercise to prove that a branch of the log function
is continuous except on its cut line.
13. Use Theorem 2.1.13 to prove that if f and g are continuous functions with open
domains Uf and Ug and if g(Ug ) ⊂ Uf , then f ◦ g is continuous on Ug .
14. Prove that if f is a continuous function defined on an open subset U of C, then
sets of the form {z ∈ U : |f (z)| < r} and {z ∈ U : Re(f (z)) < r} are open.
15. Use the result of the preceding exercise to come up with an open subset of C
that has not been previously described in this text.
16. Prove that a function f with open domain U is continuous at a point a ∈ U
if and only if whenever {zn } ⊂ U is a sequence converging to a, the sequence
{f (zn )} converges to f (a).
The first hint that there is something fundamentally different about this notion
of derivative is in the following example.
ez+λ − ez eλ −1
= ez .
λ λ
Thus, to show that the derivative of ez is ez we need only show that
eλ −1
(2.2.2) lim = 1.
λ→0 λ
However, if t = |λ|, inspection of the power series for eλ and et shows that
λ λ
e −1 e −1 − λ et −1 − t
≤
(2.2.3) λ − 1 = λ t
.
Now to show that the expression on the left has limit zero and, thus, verify (2.2.2),
we simply apply L’Hôpital’s rule to the expression on the right.
36 2. Analytic Functions
The complex derivative has all of the familiar properties in relation to sums,
products, and quotients of functions. The proofs of these are in no way different
from the proofs of the corresponding results for functions of a real variable. In the
following theorem, Part (a) is trivial and we leave Parts (b) and (c) to the exercises.
Theorem 2.2.6. If f and g are functions of a complex variable which are differ-
entiable at z ∈ C, then
(a) f + g is differentiable at z and (f + g) (z) = f (z) + g (z);
(b) f g is differentiable at z and (f g) (z) = f (z)g(z) + f (z)g (z);
(c) if g(z) = 0, 1/g is differentiable at z and (1/g) (z) = −g (z)/g 2 (z).
Parts (a) and (b) of this theorem and the fact that constant functions and the
function z are analytic on C imply that every polynomial in z is analytic on C. Of
course, since z is not analytic, we cannot expect mixed polynomials that contain
powers of both z and z to be analytic.
Parts (b) and (c) of the theorem imply that f /g is differentiable at z if f and
g are and if g(z) = 0. They also imply the quotient rule
f f (z)g(z) − g (z)f (z)
(z) = .
g g 2 (z)
The chain rule also holds for the complex derivative.
Theorem 2.2.7. If g is differentiable at a and f is differentiable at b = g(a), then
f ◦ g is differentiable at a and
(f ◦ g) (a) = f (g(a))g (a).
ux = vy ,
(2.2.10)
uy = −vx
at (x, y). Equations (2.2.10) are the Cauchy-Riemann equations. Equations (2.2.9)
also show that if f exists at z, then f (z) = C + iD = ux + ivx = −i(uy + ivy ). If
we set fx = ux + ivx and fy = uy + ivy , then this can be written as f = fx = −ify
wherever f exists.
The above discussion shows that, at any point where f has a complex derivative,
its real and imaginary parts are differentiable functions and satisfy the Cauchy-
Riemann equations. The converse is also true: If the real and imaginary parts of f
are differentiable and satisfy the Cauchy-Riemann equations at a point z = x + iy,
then f (z) exists. The proof of this is a matter of working backwards through the
above discussion, beginning with the assumption that u and v are differentiable at
(x, y), with partial derivatives that satisfy ux = vy = C and uy = −vx = −D. This
leads to (2.2.8), which eventually leads back to the conclusion that C + iD is the
derivative of f at z = x + iy. We leave the details to the exercises. The result is
the following theorem.
f = fx = −ify .
Example 2.2.11. Use the Cauchy-Riemann equations to prove that, for each
branch of the log function, log(z) is analytic everywhere except on its cut line
and has derivative 1/z.
Solution: We first prove that the principal branch of the log function is an-
alytic on the right half-plane H = {z ∈ C : Re(z) > 0}. For z ∈ H we have
z = x + iy = r eiθ where
r= x2 + y 2 and θ = tan−1 (y/x).
Harmonic Functions. In the next chapter, we will prove that analytic functions
are C∞ – that is, they have continuous complex derivatives of all orders. This, in
particular, implies that analytic functions have continuous partial derivatives of all
orders with respect to x and y. Assuming this result for the moment, we have
Theorem 2.2.12. The real and imaginary parts of an analytic function on U are
harmonic functions on U , meaning they satisfy Laplace’s equation
uxx + uyy = 0.
Solution: The function u is the real part of f (z) = ez and is, therefore,
harmonic by the previous theorem. The imaginary part of f is v(x, y) = ex sin y,
and so this function v is a harmonic conjugate of u.
1. Fill in the details in Example 2.2.4 by verifying the inequality (2.2.3) and
showing that the limit of the expression on the right is 0.
2. Prove Theorem 2.2.5.
3. Prove Part (b) of Theorem 2.2.6.
4. Prove Part (c) of Theorem 2.2.6.
5. Use induction and Theorem 2.2.6 to show that (z n ) = nz n−1 if n is a non-
negative integer.
6. Find the derivative of z 7 + 5z 4 − 2z 3 + z 2 − 1. Which results from this section
are used in this calculation?
3
7. Find the derivative of ez .
8. If we use the principal branch of the log function, at which points of C does
log z
have a complex derivative? What is its derivative at these points?
z
9. Finish the proof of Theorem 2.2.9 by showing that if f = u + iv, u and v are
differentiable at z, and u and v satisfy the Cauchy-Riemann equations at z,
then f (z) exists.
10. Use the Cauchy-Riemann equations to verify that the function f (z) = z 2 is
analytic everywhere.
11. Describe all real-valued functions which are analytic on C.
12. Derive the Cauchy-Riemann equations in polar coordinates:
ur = r −1 vθ ,
uθ = −rvr
by using the change of variable formulas x = r cos θ, y = r sin θ and the chain
rule.
13. We showed in Example 2.2.11 that each branch of the log function is analytic
on the complex plane with its cut line removed. Use the Cauchy-Riemann
equations in polar form (previous problem) to give another proof of this fact.
14. Assuming each branch of the log function is analytic, use the chain rule to give
another prove that each such function has derivative 1/z.
15. Use the Cauchy-Riemann equations to prove that if f is analytic on an open set
U , then the function g defined by g(z) = f (z) is analytic on the set {z : z ∈ U }.
16. Verify that the function log |z| is harmonic on C \ {0} and find a harmonic con-
jugate for it on the set consisting of C with the non-positive real axis removed.
2.3. Contour Integrals 41
Smooth Curves. Let I = [a, b] be a closed interval on the real line and let
γ : I → C be a complex-valued function on I. If c ∈ I, then the derivative γ (c) of
γ at c is defined in the usual way:
γ(t) − γ(c)
(2.3.1) γ (c) = lim .
t→c t−c
Of course, γ is complex-valued and so this limit should be interpreted as the type
of limit discussed in Section 2.1. It can be calculated by expressing γ in terms of
its real and imaginary parts, that is, by writing γ(t) = x(t) + iy(t), where x(t) and
y(t) are real-valued functions on I. Then γ (t) = x (t) + iy (t) (Exercise 2.3.6).
What about the endpoints a and b of the interval I? Should we either not talk
about the derivative at the endpoints or, perhaps, use one-sided derivatives defined
in terms of one-sided limits (limit from the right at a and limit from the left at b)?
Actually, there is no need to do anything special at a and b or to exclude them. If
the domain of γ is [a, b], then our domain dependent definition of limit takes care
of the problem. If c = a, the limit as t → a in (2.3.1) only involves values of t to
the right of a, since only those are in the domain of the difference quotient that
appears in this limit. Similarly, if c = b, the limit as t → b involves only points to
the left of b. Thus, the derivatives at a and b that our definition leads to are what
in calculus would be called the right derivative at a and the left derivative at b.
The curve γ is differentiable at c if the limit defining γ (c) exists. It is contin-
uously differentiable or smooth on I if it is differentiable at every point of I and if
the derivative is a continuous function on I. In this case we will write γ ∈ C1 (I).
Definition 2.3.1. A curve γ : [a, b] → C in C is called piecewise smooth if there
is a partition a = a0 < a1 < · · · < an = b of [a, b] such that the restriction of γ to
[aj−1 , aj ] is smooth for each j = 1, · · · , n. A curve which is piecewise smooth will
be called a path.
γ(b)
.
.
γ(a)
.
γ(t)
Example 2.3.3. Let z and w be two points in C. Find a path which traces the
straight line from z to w and find its derivative.
Solution: The path γ, with parameter interval [0, 1], defined by
γ(t) = (1 − t)z + tw = z + t(w − z),
satisfies γ(0) = z and γ(1) = w. It is a parametric form of a straight line in the
plane, and its derivative is γ (t) = w − z.
Example 2.3.4. Find a path that traces once around the square with vertices
0, 1, 1 + i, i in the counterclockwise direction. Find γ (t) on the subintervals where
γ is smooth.
Solution: We choose [0, 1] as the parameter interval and define a path γ as
follows (see Figure 2.3.2):
⎧
⎪
⎪ 4t, if 0 ≤ t ≤ 1/4;
⎪
⎨1 + (4t − 1)i, if 1/4 ≤ t ≤ 1/2;
γ(t) =
⎪
⎪ 3 − 4t + i, if 1/2 ≤ t ≤ 3/4;
⎪
⎩
(4 − 4t)i, if 3/4 ≤ t ≤ 1.
This is continuous on [0, 1] and smooth on each subinterval in the partition 0 <
1/4 < 1/2 < 3/4 < 1. It traces each side of the square in succession, moving in
the counterclockwise direction. On the first interval, γ is the constant 4, on the
second it is 4i, on the third it is −4, and on the fourth it is −4i.
i 1+i
0 1
This Riemann integral for complex-valued functions has the properties one
would expect given knowledge of the Riemann integral for real-valued functions.
The next three theorems cover some of these properties.
Theorem 2.3.5. Let f1 and f2 be Riemann integrable functions on [a, b] and α
and β complex numbers. Then, αf1 + βf2 is integrable on [a, b], and
b b b
(αf1 (t) + βf2 (t)) dt = α f1 (t) dt + β f2 (t) dt.
a a a
Proof. That this is true if the constants α and β are real follows directly from
expressing f1 and f2 in terms of their real and imaginary parts. Thus, to prove
b b
the theorem we just need to show that a if (t) dt = i a f (t) dt if f = g + ih is an
integrable function on [a, b]. However,
b b
i(g(t) + ih(t)) dt = (−h(t) + ig(t)) dt
a
a
b b b
=− h(t) dt + i g(t) dt = i (g(t) + ih(t)) dt .
a a a
Proof. This follows from the fact that the same things are true of the integrals of
the real and imaginary parts g and h of f .
Theorem 2.3.7. If f is an integrable function on [a, b], then
b b
f (t) dt ≤ |f (t)| dt.
a a
44 2. Analytic Functions
b
Proof. This is proved using a trick. We set w = a f (t) dt. If w = 0, there is
nothing to prove. If w = 0, let u = w/|w|. Then uw = |w| and so
b b
b
f (t) dt = u f (t) dt = uf (t) dt.
a a a
Since this is a real number, the integral of the imaginary part of uf is zero and we
have
b b b b
f (t) dt = Re(uf (t)) dt ≤ |uf (t)| dt = |f (t)| dt.
a a a a
One may think of this definition in the following way: the contour integral on
the left in (2.3.3) is defined to be the Riemann integral obtained by replacing z by
γ(t) and dz by γ (t)dt and integrating over the parameter interval for γ.
In practice, we will calculate contour integrals by breaking the path up into its
smooth sections, calculating the integrals over these sections and then adding the
results. That this is legitimate follows from the fact that the Riemann integral of
a function over the union of two contiguous intervals on the line is the sum of the
integrals over the two intervals.
Examples.
Example 2.3.9. Find γ
z dz if γ is the circular path defined in Example 2.3.2.
Solution: By Example 2.3.2, we have γ(t) = r eit for 0 ≤ t ≤ 2π and γ (t) =
it
ir e . Thus,
2π 2π
z dz = r eit ir eit dt = ir 2 e2it dt = ir 2 (cos 2t + i sin 2t) dt
γ 0 0
2π 2π
= ir 2 cos 2t dt − r 2 sin 2t dt = 0.
0 0
On [1, 2] we have (γ(t))2 = t2 − 2t + 2(t − 1)i and γ (t) = 1. Thus, the integral over
the second section of the path is
2 1
2
2
(γ(t)) γ (t) dt = (t2 − 2t + 2(t − 1)i) dt = (t3 /3 − t2 + (t2 − 2t)i)1 = −2/3 + i.
1 0
Thus, γ
z dz = −i/3 − 2/3 + i = −2/3 + 2i/3.
2
Example 2.3.11. Find a path γ which traces once around the triangle with vertices
0, 1, i in the counterclockwise direction, starting at 0. For this path γ, find γ z dz.
Solution: A path γ with the required properties has parameter interval [0, 3]
and is given by
⎧
⎪
⎨t, if 0 ≤ t ≤ 1;
γ(t) = 2 − t + (t − 1)i if 1 ≤ t ≤ 2;
⎪
⎩
(3 − t)i if 2 ≤ t ≤ 3.
46 2. Analytic Functions
On the interval [1, 2], we have γ(t) = 2 − t − (t − 1)i and γ (t) = −1 + i. Hence,
2 2
γ(t)γ (t) dt = (2t − 3 + i) dt = i.
1 1
On the interval [2, 3], we have γ(t) = t − 3 and γ (t) = −i. Hence,
1 3
γ(t)γ (t) dt = (3 − t)i dt = i/2.
0 2
Parameter Changes that Change the Integral. The following example shows
that some changes of parameterization do change the integral.
Example 2.4.1. Find γ1 1/z dz if γ1 (t) = r eit on [0, 2π] is the circular path of
Example 2.3.2. Does the answer change if the circle is traversed in the clockwise
direction instead, using the path γ2 (t) = r e−it on [0, 2π]?
Solution: From Example 2.3.2 we know that the path γ1 (t) = r eit has γ1 (t) =
ir eit and so the given integral is
2π 2π
dz (r eit )
= dt = i dt = 2πi.
γ1 z 0 r eit 0
This example shows that the integral along a path depends not only on the
geometric figure that is the image γ(I) of the path, but also on the direction the
path is traversed (at the very least).
Also, traversing a portion of the curve more than once may affect the integral.
For example, if we were to go around the circle twice in Example 2.4.1, by choosing
γ(t) = e2it on [0, 2π], the result would be 4πi instead of 2πi.
for every function f defined and continuous on a set E containing γ1 ([a, b]) =
γ2 ([c, d]).
48 2. Analytic Functions
= f (z) dz,
γ1
where the third equality follows from the substitution s = α(t). This completes the
proof.
Note that the condition that α(c) = a and α(d) = b is essential in the above
theorem. It says that α takes the endpoints of the parameter inverval [c, d] to the
endpoints of the parameter interval [a, b] in an order preserving fashion.
Example 2.4.3. Are the integrals of a continuous function over the two paths in
(2.4.1) necessarily the same?
Solution: Yes. If we set α(t) = − cos t, then α is a smooth√function√ mapping
the parameter interval [π/6, 5π/6] to the parameter interval [− 3/2, 3/2] in an
order preserving fashion. Furthermore, γ2 = γ1 ◦ α. Thus, the above theorem
insures that the integral of a continuous function over γ1 is the same as its integral
over γ2 .
Doesn’t Example 2.4.1 contradict Theorem 2.4.2? After all, if γ1 (t) = eit on
[0, 2π] and α : [0, 2π] → [0, 2π] is defined by α(t) = 2π − t, then γ2 (t) = γ1 (α(t)) =
e−it . By Example 2.4.1 the integrals of 1/z over these two curves are different.
Doesn’t Theorem 2.4.2 say they should be the same? No. The conditions α(a) = c
and α(b) = d are not satisfied by this choice of α, since α(0) = 2π and α(2π) = 0.
In other words, this choice of α reverses the order of the endpoints of the parameter
interval rather than preserving that order.
In general, the conditions α(a) = c and α(b) = d guarantee that, overall, γ2
traverses the curve in the same direction as γ1 . If α were positive on the entire
interval, then α would be increasing on this interval and γ1 and γ2 would be moving
in the same direction at each point of the curve. If α is not positive on all of [c, d],
then there may be intervals where one path reverses direction and backtracks, while
the other path does not. These things do not affect the integral, because if a curve
does backtrack for a time, it has to turn around and recover the same ground in
order to catch up to the other curve in the end. This is an intuitive explanation;
the actual proof that the integral is unaffected is in the proof of the above theorem.
Theorem 2.4.2 leads to a strategy which, for some paths γ1 and γ2 with the
same image, yields a proof that they determine the same integral: Suppose that the
parameter intervals for the two paths can each be partitioned into n subintervals
2.4. Properties of Contour Integrals 49
in such a way that for j = 1, · · · , n, γ1 on its jth subinterval and γ2 on its jth
subinterval are related by a smooth function
αj , as in Theorem 2.4.2. If this can
be done, then it clearly follows that γ1 f (z) dz = γ2 f (z) dz for any function f
which is continuous on a set containing γ1 (I). For this reason, Theorem 2.4.2 is
sometimes called the independence of parameterization theorem.
Remark 2.4.4. Since path integrals are essentially independent of the way the
path is parameterized, we will often describe a path without specifying a parame-
terization. Instead, we will just give a description of the geometric object that is
traced, the direction, and how many times. For example, we may describe a path
as tracing once around the unit circle in the counterclockwise direction, or tracing
once around the boundary ∂Δ of a given triangle Δ in the counterclockwise direc-
tion, or as tracing the straight line path from a complex number w1 to a complex
number w2 . In the first two cases we may simply write
f (z) dz or f (z) dz
|z|=1 ∂Δ
for the corresponding path integral. In the latter case, we may write
w2
f (z) dz
w1
for the path integral along the straight line from w1 to w2 .
Closed Curves. The curves in Examples 2.3.2 and 2.3.4 both have the property
that they begin and end at the same point – that is, they are closed curves. A closed
curve γ on a parameter interval [a, b] is one that satisfies γ(a) = γ(b). A closed
curve which is a path will be called a closed path.
The famous integral theorem of Cauchy states that the integral of an analytic
function f around a closed path is 0, provided there is an appropriate relationship
between the curve γ and the domain U on which f is analytic (roughly speaking,
the curve should lie in U but not go around any holes in U ). Since the function
f (z) = z is analytic on C (as is any polynomial in z), the next example illustrates
this phenomenon.
Example 2.4.5. Find γ z dz if γ is the path of Example 2.3.4.
Solution: From Example 2.3.4 we know that the path γ(t) has values 4t, 1 +
(4t−1)i, 3−4t+i, (4−4t)i and derivatives 4, 4i, −4, and −4i on the four subintervals
of the partition 0 < 1/4 < 1/2 < 3/4 < 1. Thus, the integrals over the four smooth
pieces of our curve are
1/4
1/4
4t · 4 dt = 8t2 0 = 1/2,
0
1/2 1/2
(1 + (4t − 1)i) · 4i dt = (4ti − 8t2 + 4t)1/4 = i − 1/2,
1/4
3/4 3/4
(3 − 4t + i) · (−4) dt = (−12t + 8t2 − 4ti)1/2 = −1/2 − i,
1/2
1 1
(4 − 4t)i · (−4i) dt = (+16t − 8t2 )3/4 = 1/2.
3/4
50 2. Analytic Functions
γ2 (c)
γ1
γ1 (a) γ2
γ1 (b) = γ2 (b)
a b c
Since these add up to 0, we have γ
z dz = 0.
The function 1/z is also analytic, except at z = 0. The circular path of Example
2.4.1 is closed and lies in the domain where 1/z is analytic. So why is the integral
not 0? Because the path goes around a hole in the domain of 1/z – it goes around
{0}.
Then −γ(a) = γ(b) and −γ(b) = γ(a). In fact, −γ traces the same geometric figure
as γ, but it does so in the opposite direction.
For some closed curves, such as circles, and boundaries of rectangles, triangles,
etc., there is clearly a clockwise direction around the curve and a counterclockwise
direction. If such a curve is parameterized so that it is traversed in the counter-
clockwise direction, we will say the resulting closed path has positive orientation.
If it is traversed in the clockwise direction, we will say it has negative orientation.
Clearly, if γ has positive orientation, then −γ has negative orientation.
The common starting and ending point of a closed path can be changed without
changing the integral of a function over this path. This is done by representing the
closed path as the join of two paths which connect the original starting and ending
point to the new one. One then uses part (b) of the next theorem. The details are
left to the exercises.
The next theorem states the elementary properties of path integrals having
to do with linearity and path additivity. Part (b) follows immediately from the
corresponding additivity property of the Riemann integral on the line and we have
already used it several times. We leave the proofs of (a) and (c) to the exercises
(Exercise 2.4.5).
Theorem 2.4.6. Let γ, γ1 , γ2 be paths with γ1 ending where γ2 begins, f and g two
functions which are continuous on a set E containing the images of these paths,
and a and b complex numbers. Then
(a) (af (z) + bg(z)) dz = a f (z) dz + b g(z) dz;
γ γ γ
(b) f (z) dz = f (z) dz + f (z) dz;
γ1 +γ2 γ1 γ2
(c) f (z) dz = − f (z) dz.
−γ γ
Part (a) of this theorem says that a path integral is a linear function of the
integrand, Part (b) says that it is an additive function of the path, while Part (c)
shows why the notation −γ is appropriate for the curve that is γ traversed in the
opposite direction.
Length of a Path. We define the length (γ) of a path γ in C in the same way
the length of a curve in R2 is defined in calculus.
Definition 2.4.7. If γ(t) = x(t) + iy(t) is a path in C with parameter interval
[a, b], then the length (γ) of γ is defined to be
b b
(γ) = |γ (t)| dt = x (t)2 + y (t)2 dt.
a a
Example 2.4.8. Prove that the above definition of length yields the correct length
for a path which traces once around a circle of radius r.
Solution: The path is γ(t) = r eit , with parameter interval [0, 2π]. The deriv-
2π
ative of γ is γ (t) = ir eit and so |γ (t)| = r. Thus, (γ) = 0 r dt = 2πr.
52 2. Analytic Functions
for each w ∈ C.
Solution: The statement that f is bounded means there is an upper bound
M for |f |. That is, |f (z)| ≤ M for all z ∈ C. We also have |z − w| ≥ |z| − |w|
by the second form of the triangle inequality. If z ∈ γ(I), then |z| = R and so
|z − w| ≥ R − |w|, which implies |z − w|−2 ≤ (R − |w|)−2 . Thus, for z ∈ γ(I), we
have the following bound on the integrand of (2.4.3):
f (z) M
z − w ≤ (R − |w|)2 .
Since (γR ) = 2πR, Theorem 2.4.9 implies that
f (z)
dz ≤ 2πM R .
(z − w) 2 (R − |w|)2
γR
The right side of this inequality has limit 0 as R → ∞ and this implies (2.4.3).
4. Compute the integral of the previous exercise for any smooth path γ which
begins at z0 and ends at w0 .
5. Prove Parts (a) and (c) of Theorem 2.4.6.
6. Describe a smooth, order preserving function α which takes the parameter
interval [0, 1] to the parameter interval [2, 5].
7. Prove that a parameter change γ → γ ◦ α, like the one in Theorem 2.4.2, does
not change the length of a path provided α is an non-decreasing function (has
a non-negative derivative).
cos z
8. Show that dz ≤ 2πe if γ is a path that traces the unit circle once.
γ z
Hint: Show that | cos z| ≤ e if |z| = 1.
9. Show that if Δ is a triangle in the plane of diameter d (length of its longest
side), and if f is a continuous function on Δ with |f | bounded by M on Δ,
then
f (z) dz ≤ 3M d.
∂Δ
10. Prove that γ p(z) dz = 0 if γ(t) = eit , 0 ≤ t ≤ 2π, and p(z) is any polynomial
in z (this is a special case of Cauchy’s Theorem, but do not assume Cauchy’s
Theorem in your proof).
11. Let R(z) be the remainder after n terms in the power series for ez . That is,
n ∞
zk zk
R(z) = ez − = .
k! k!
k=1 k=n+1
e−1
Prove that |R(z)| ≤ if |z| ≤ 1.
(n + 1)!
12. Prove that γ ez dz = 0 if γ(t) = eit , 0 ≤ t ≤ 2π using the previous exercise
and Exercise 10.
13. Prove that if γ is a closed path with parameter interval I = [a, b] and common
starting and ending point z = γ(a) = γ(b) and w is any other point on γ(I),
then there is another closed path γ1 with γ1 (I) = γ(I), which determines the
same integral, but has w as common starting and ending point. Hint: Use part
(b) of Theorem 2.4.6.
We take a first step toward the proof of Cauchy’s Theorem in this section by
proving it in the case where the path is the boundary of a triangle and the function
is analytic on an open set containing the triangle.
The proof of this result will make essential use of a couple of properties of
compact sets. Thus, we will precede the proof with a discussion of compact sets.
Theorem 2.5.2. A set K is compact if and only if, whenever a collection of closed
subsets of K has the property that each finite subcollection has non-empty intersec-
tion, then the full collection also has non-empty intersection.
Corollary 2.5.3. If
A1 ⊃ A2 ⊃ · · · ⊃ An ⊃ · · ·
is a nested sequence of non-empty compact subsets of Rn , then An = ∅.
This is one of the properties of compact sets that we shall need in our proof of
Cauchy’s Theorem on a triangle. The others are as follows:
Antiderivatives. There is one case in which it is very easy to prove that the
integral of a continuous function f around a closed path is 0. This is the case where
the function f has an antiderivative.
As with functions of a real variable, an antiderivative for a function f , defined
on an open set U , is a function g such that g = f on U .
Theorem 2.5.6. If f is a continuous function defined on an open set U , and if f
has an antiderivative g on U , then
f (z) dz = g(γ(b)) − g(γ(a))
γ
if γ is any path in U with parameter interval [a, b]. If γ is a closed path, then this
integral is 0.
There is a version of the chain rule which holds for the composition of an analytic
function with a path (Exercise 2.5.4). It tells us that
(g(γ(t))) = g (γ(t))γ (t).
Thus,
b
f (z) dz = (g(γ(t))) dt = g(γ(b)) − g(γ(a)),
γ a
by the complex version of the Fundamental Theorem of Calculus (Exercise 2.3.7).
If the path is closed, then γ(b) = γ(a) and the integral is 0.
Proof. We set
I= f (z) dz.
∂Δ
Our objective is to prove that I = 0. We will do this by showing that |I| < for
every positive number . Thus, let be an arbitrary positive number.
We subdivide the triangle Δ into four smaller triangles by joining the midpoints
of the sides of Δ. The resulting four triangles are all similar to Δ with sides exactly
half as long as the corresponding sides of Δ (see Figure 2.5.1).
We now apply Parts (b) and (c) of Theorem 2.4.6. The sum of the integrals
around each of these smaller triangles is a sum of integrals along their edges, with
each of the edges interior to the original triangle occurring twice – once going one
2.5. Cauchy’s Integral Theorem for a Triangle 57
direction and once going the opposite direction. Thus, the contributions of these
interior edges cancel, leaving only the contributions from the edges which lie along
the boundary of the original triangle. It follows that the sum of the integrals
of f around the boundaries of these four smaller triangles is I, and so one of
these integrals must have modulus at least |I|/4. Let Δ1 denote the corresponding
triangle. In other words, Δ1 is chosen from the four subtriangles so that
|I1 | ≥ |I|/4 where I1 = f (z) dz.
∂Δ1
Note also that, if h is the diameter of Δ (which is the length of its longest side),
then Δ1 has diameter h1 = h/2.
We now repeat the above construction with Δ replaced by Δ1 . That is, we
subdivide Δ1 into four similar triangles and choose one of them, call it Δ2 , with
the property that
|I2 | ≥ |I1 |/4 ≥ |I|/4 where I2 =
2
f (z) dz,
∂Δ2
2
and with diameter h2 = h1 /2 = h/2 .
Proceeding by induction, we may choose for each n a triangle Δn , of diameter
hn , so that
Δn ⊂ Δn−1 ,
(2.5.4) |In | ≥ |I|/4n
where In = f (z)dz
∂Δn
and
(2.5.5) hn = h/2n .
We may then choose n large enough that hn = h/2n < δ. Since hn is the
diameter of Δn and w ∈ Δn , we have
h2n
|In | < .
h2
Putting this together with (2.5.5), we conclude
|In | < n .
4
Combined with (2.5.4), this yields
|I| ≤ 4n |In | < .
Since was an arbitrary positive number, we conclude that I = 0. This completes
the proof.
We will need a slightly stronger version of this theorem in which we allow the
possibility that there is one point in Δ where f may not have a complex derivative,
but where f is continuous.
Theorem 2.5.9. Let f, U, and Δ be as in the previous theorem except that we
assume that f is continuous on U and analytic on U \ {c} for some exceptional
point c ∈ Δ. Then we still have
f (z)dz = 0.
∂Δ
(a) (b)
of each of these is zero, so the integral around ∂Δ is also 0. This completes the
proof.
U z0
z
a
Theorem 2.6.1. Let U be a convex open set and suppose f is a function which is
continuous on U and has the property that its integral around the boundary of any
triangle in U is zero. If a ∈ U is fixed and F (z) is defined for all z ∈ U by
z
F (z) = f (w) dw,
a
then F (z) = f (z) for all z ∈ U .
Proof. Let [a, z] denote the line segment joining a to z, considered as a path from
a to z. For z ∈ U , this line segment lies entirely in U and so we may define a
function F (z) by
z
F (z) = f (w) dw,
a
where by this we mean the path integral of f along the path [a, z], as in Remark
2.4.4. We will show that F (z) = f (z). To do this, we let z and z0 be points of U
and consider the triangle Δ with vertices a, z, z0 (see Figure 2.6.1). The fact that
U is convex implies that Δ ⊂ U .
Let ∂Δ denote the boundary of Δ, considered as a contour which goes from a
to z to z0 and then to a again. Since, by hypothesis, the integral of f around the
boundary of any triangle in U is 0, we have
0= f (w) dw
∂Δ
z z0 a
= f (w) dw + f (w) dw + f (w) dw
a z z0
z
= F (z) − F (z0 ) − f (w) dw.
z0
We conclude that
z
F (z) − F (z0 ) = f (w) dw.
z0
62 2. Analytic Functions
If we add and subtract the number f (z0 ) in the integrand of this integral, we obtain
z z
F (z) − F (z0 ) = f (z0 ) dw + (f (w) − f (z0 )) dw
z0 z
0 z
= f (z0 )(z − z0 ) + (f (w) − f (z0 )) dw.
z0
If we divide by z − z0 and subtract the first term on the right from both sides, we
get
z
F (z) − F (z0 ) 1
(2.6.1) − f (z0 ) = (f (w) − f (z0 )) dw.
z − z0 z − z0 z0
Thus, to finish the proof that F (z0 ) = f (z0 ), we just need to show that the
expression on the right in (2.6.1) has limit 0 as z → z0 .
Given > 0, we may choose a δ > 0 such that |f (w) − f (z0 )| < when
|w − z0 | < δ. This follows from the fact that f is continuous on U . If |z − z0 | < δ,
then |w − z0 | < δ for every w on the line segment [z0 , z] and so |f (w) − f (z0 )| <
for every w on this line segment. Then
z
(f (w) − f (z0 )) dw < |z − z0 |
z0
and so z
1
(f (w) − f (z )) dw <
z − z0 0
z0
whenever |z − z0 | < δ. This shows that
z
1
lim (f (w) − f (z0 )) dw = 0,
z→z0 z − z0 z
0
Cauchy’s Integral Theorem. We now have all the tools in place to prove
Cauchy’s Integral Theorem for convex sets. This is not the most general form
of the theorem – that will come later – but it is sufficiently general to allow us to
derive a wealth of surprising consequences.
Theorem 2.6.2. Let U be a convex open set and suppose f is a function which is
analytic on U , except possibly at one point, where it is at least continuous. Then
f (z) dz = 0
γ
for every closed path in U .
Proof. By the previous theorem, and Theorem 2.5.8, f has a complex antideriva-
tive F in U . Then Theorem 2.5.6 implies that the integral of f around any closed
path is zero.
Proof. Let z0 a point in the complement of γ(I). Since γ is a closed path, we have
γ(a) = γ(b). We define a complex-valued function λ(t) on the interval a ≤ t ≤ b by
t
γ (s)
λ(t) = ds.
a γ(s) − z0
Then λ(a) = 0 and λ(b) = 2πi Indγ (z). If we can show that eλ(b) = 1, then the
proof will be complete, since the only numbers w with ew = 1 are the numbers
2πi n where n is an integer.
By the Fundamental Theorem of Calculus,
γ (t)
λ (t) = ,
γ(t) − z0
while the derivative of eλ(t) is
γ (t)
(eλ(t) ) = eλ(t) λ (t) = eλ(t) .
γ(t) − z0
It follows that
eλ(t) 1 γ (t) λ(t)
= e λ(t)
(γ(t) − z0 ) − e γ (t) = 0.
γ(t) − z0 (γ(t) − z0 )2 γ(t) − z0
Hence, eλ(t) /(γ(t) − z0 ) is a constant. We conclude that
eλ(t) eλ(a) 1
= =
γ(t) − z0 γ(a) − z0 γ(a) − z0
for every t ∈ [a, b]. If we set t = b, this gives us
γ(b) − z0
eλ(b) = .
γ(a) − z0
Since γ(a) = γ(b), it follows that eλ(b) = 1. This completes the proof.
even if the function is not analytic at some point but is continuous there, it follows
that
f (w) f (z)
0= g(z, w)dw = dw − dw
γ γ w − z γ w −z
f (w)
= dw − 2πi Indγ (z)f (z),
γ w −z
as long as z is not on the contour γ (note that this is required in order to write
the integral in the first line above as the difference of two integrals, since otherwise
these two integrals might not exist individually). We conclude that
1 f (w)
Indγ (z)f (z) = dw,
2πi γ w − z
as required. This completes the proof.
This is a striking result, for it says that the values of an analytic function at
points “inside” a closed path are determined by its values at points on the path.
Here, a point is considered inside the path if the path has non-zero index at the
point.
Corollary 2.6.8. If U is a convex open set, z ∈ U and γ is a closed path in U
with Indγ (z) = 1, then
1 f (w)
f (z) = dw
2πi γ w − z
for every function f analytic on U .
Intuitively, the meaning of the hypothesis Indγ (z) = 1 in the above corollary is
that the closed path γ goes around z once and does so in the positive direction.
Cauchy’s Integral Theorem and Cauchy’s Integral Formula have a wealth of
applications. We will begin exploring these in the next chapter.
However, in order for Cauchy’s Integral Theorem, in the above form, to be us-
able, we need to be able to easily compute the index of a curve around a given point.
The last section of this chapter is devoted to developing the essential properties of
the index function which make this possible.
Connected Sets.
The union of a family of connected sets with a point in common is also connected
(Exercise 2.7.1). It follows that, if z ∈ E, then the union of all connected subsets
of E containing z is itself connected. This implies that each point of E is contained
in a maximal connected subset of E. A maximal connected subset of E is called
a connected component of E or simply a component of E. Two components of E
are either disjoint or identical, since, otherwise, their union would be a connected
set larger than one of them. Thus, the components of E form a pairwise disjoint
family of subsets of E whose union is E.
In this section we are primarily concerned with open sets and their components.
If E is open, then the sets A ∩ E and B ∩ E of Definition 2.7.1 are open subsets
of E. It follows that an open set E is separated if and only if it is the union of
two disjoint non-empty open subsets of itself. It is connected if this is not the case.
The next theorem states the essential facts regarding connected open sets that we
will need in this section.
An open set U is said to be path connected if every two points in U can be
connected by a path which lies entirely in U .
Proof. The set γ(I) is the image of a compact set under a continuous function and
so it is compact, hence, closed. Its complement C \ γ(I) is, therefore, open. Thus,
if z0 ∈ C \ γ(I), there is an open disc, centered at z0 , and contained in C \ γ(I). Let
R be the radius of one such disc. We will show that, on some smaller disc, centered
at z0 , Indγ (z) is constant.
Suppose r is a positive number less than R and z ∈ Dr (z0 ). Then
1 dw 1 dw
Indγ (z) − Indγ (z0 ) = −
2πi γ w − z 2πi γ w − z0
(2.7.1)
1 z − z0
= dw.
2πi γ (w − z)(w − z0 )
Furthermore, every point w of γ(I) is at least a distance R away from z0 and a
distance R − r away from z. That is,
|w − z0 | ≥ R and |w − z| ≥ R − r.
Since |z − z0 | < r, this implies that the integrand of the last integral in (2.7.1) is
less than or equal to r/R(R − r) and, hence, that
r(γ)
| Indγ (z) − Indγ (z0 )| ≤ .
2πR(R − r)
We can make the right side of this inequality as small as we want by choosing r
sufficiently small. In particular, we can make it less than 1. However, Indγ (z) and
Indγ (z0 ) are both integers. If they differ by less than 1, then they are the same.
Thus, if r is chosen small enough, Indγ (z) is constant on Dr (z0 ).
2.7. Properties of the Index Function 69
Let A be a component of C \ γ(I), and for each integer n let Vn be the set of
points of A on which Indγ (z) = n. The above argument shows that each Vn is an
open subset of A. So is the union of all the sets Vm for which m = n. These two
sets separate A unless one of them is empty. Since A is connected, one of them
must be empty. This means that if Vn is not empty, then it is all of A. Thus, only
one of the sets Vn can be non-empty and this means that Indγ (z) is constant on A.
It remains to show that Indγ (z) = 0 on the unbounded component of C \ γ(I).
Let D be an open disc containing γ(I) and let z0 be a point outside this disc. Then
z0 is in the unbounded component of C \ γ(I). Furthermore,
1 dw
Indγ (z0 ) = =0
2πi γ w − z0
1
by Cauchy’s Integral Theorem, since D is a convex set containing γ(I) and
w − z0
is analytic on D. Since Indγ (z) is constant on each component of C \ γ(I), it must
be identically 0 on the unbounded component.
Example 2.7.6. Calculate Indγ (z) for the path γ which traces n times around a
circle of radius r centered at z0 , where, if n is positive, this means γ traces the
circle in the counterclockwise direction, and, if n is negative, it means γ traces the
circle in the clockwise direction.
Solution: A parameterization for such a γ is γ(t) = z0 + r eint with parameter
interval I = [0, 2π]. Here γ (t) = irn eint and so
2π 2π
1 γ (t)dt 1
Indγ (z0 ) = = in dt = n.
2πi 0 γ(t) − z0 2πi 0
We use the previous theorem to find the index at other points z. Since the interior
of the circle traced by γ is a component of C \ γ(I), Indγ (z) must be constant on
it. Therefore, it has the value n at every z with |z − z0 | < r. The other component
of C \ γ(I) is the unbounded component {z : |z − z0 | > r}. On it, Indγ (z) = 0.
Thus, in this example, Indγ (z) is an integer n which is the number of times the
path goes around z if the path has positive orientation (goes in the counterclockwise
direction), and is the negative of this number if the path has negative orientation.
This includes the case where the path does not go around z at all, because z lies
outside the circle. Then n = 0.
Crossing a Path. For most paths, the index function can easily be computed
using the principle that if a path is crossed from right to left at a “simple” point of
the path, then the index increases by 1. We will make this statement precise below
and outline its proof. Most of the details are left to the exercises.
Definition 2.7.7. Let γ be a path with parameter interval I = [a, b] and let D be
an open disc in the plane. We will say that γ simply splits D if
(a) J = γ −1 (D) is a non-empty open subinterval of I, or (in the case where the
path is closed and γ(a) = γ(b) ∈ D) a union of two half-open subintervals
[a, c) and (d, b]; and
(b) D \ γ(J) has exactly two components.
70 2. Analytic Functions
z0 D
Roughly speaking, a disc D is simply split by a path if the path passes just
once through D and cuts it into exactly two connected components.
In this situation, there is a way to make sense of which of the two components
is to the left of the path and which is to the right (Exercise 2.7.3).
Theorem 2.7.8. Let γ be a closed path which simply splits a disc D. Then
Indγ (z) = 1 + Indγ (w)
if z is in the left and w in the right component of D \ γ(J).
Thus, if z0 is a point of γ which is the center of a disc which is simply split by
γ, then Indγ (z) increases by 1 as z crosses γ from right to left at z0 .
If γ is a path with parameter interval I = [a, b], and t0 ∈ (a, b), then the
derivatives of γ as a function on [a, t0 ] and γ as a function on [t0 , b] both exist at
t0 (see Definition 2.3.1 and the discussion preceding it). They are equal if t0 is a
smooth point of the path. If t0 is not a smooth point, then γ has two derivatives
at t0 – a left derivative D γ and a right derivative Dr γ. At the endpoint a of I, γ
has only a right derivative, while at the endpoint b it has only a left derivative.
2.7. Properties of the Index Function 71
2i
A
1+i
B
C
In other words, a point is a simple point of a path if the path passes through
the point just once and it both approaches the point from a definite direction with
a positive speed and leaves the point in a definite direction with a positive speed.
This leads to a very useful criterion for a point on a path to be the center of a disc
that is simply split by the path.
Theorem 2.7.11. If z0 is a simple point of the path γ, then there is an open disc,
centered at z0 , which is simply split by γ.
Exercises 2.7.11 through 2.7.14 are devoted to proving this theorem in the case
where the point z0 is actually a smooth simple point of γ. In Exercise 2.7.15, the
reader is asked to modify this argument so as to prove the theorem in general.
Note that in attempting to prove the above theorem, we may assume that
z0 = γ(t0 ) for some interior point t0 of the parameter interval (otherwise we can
just reparameterize to make this the case). Also, we may assume that z0 = 0 since,
otherwise, we can just translate the curve to make this the case. Both of these
assumptions are made in Exercises 2.7.11 to 2.7.15.
Theorems 2.7.11 and 2.7.8 imply that if a closed path γ is crossed from right
to left at a simple point, then Indγ increases by 1. For most of the paths γ that
we shall encounter, all but finitely many points of γ are simple points. Exceptions
to this rule are paths which retrace parts of themselves, cross themselves infinitely
often, or come to a dead stop over some segment of the parameter interval.
1. Prove that the union of a family of connected sets with a point in common is
also a connected set.
72 2. Analytic Functions
2. Prove Part (a) of Theorem 2.7.2. That is, prove that each component of an
open set is open.
3. Let D be an open disc and γ : I → C be a closed path which simply splits D.
Argue that, if we think of the positive direction along the curve through D as
being “up”, then it makes sense to think of one of the components into which
γ splits D as the “left” one and the other as the “right” one. Describe how to
tell which is which.
4. Suppose a closed path γ simply splits a disc D as in Figure 2.7.1. Define two new
paths γ1 and γ2 as follows: The curve γ1 agrees with γ until γ enters D. It then
departs from γ and instead traces the boundary of D in the counterclockwise
direction until it rejoins γ. It agrees with γ from that point on. The path
γ2 does the same thing except it traces the boundary of D in the clockwise
direction. For which points z inside D does Indγ (z) = Indγ1 (z)? For which
points z inside D does Indγ (z) = Indγ2 (z)? What is Indγ1 (z) − Indγ2 (z) if z is
any point inside D? Hint: Use Cauchy’s Integral Theorem and Example 2.7.6.
5. Use the results of the previous exercise to prove Theorem 2.7.8.
6. Prove that if γ is a closed path whose complement has just two components
and if γ has at least one simple point, then Indγ (z) = ±1 on the bounded
component.
7. In Example 2.7.9 how would the answers differ if the inner circle is traced in
the clockwise direction rather than the counterclockwise direction?
8. If a path γ traces a figure eight once, what are the possibilities for Indγ (z) in
the two bounded components of the complement of the figure eight?
9. Determine the value of Indγ (z) in each of the components of C \ γ(I) if γ is the
curve of Figure 2.7.3.
10. Suppose γ1 and γ2 are closed paths and z is not on either path. Show that
Indγ1 +γ2 (z) = Indγ1 (z) + Indγ2 (z) and Ind−γ1 = − Indγ1 (z). Hint: Use Theo-
rem 2.4.6.
11. Let γ(t) = x(t) + iy(t) be a path with parameter interval I = [a, b] and t0 a
point in (a, b) at which the path is smooth and simple. For further simplicity,
assume γ(t0 ) = 0 (we can always achieve this by translating the path). Prove
that, even though γ (t0 ) may not exist, the second derivative of the function
h(t) = |γ(t)|2 = x2 (t) + y 2 (t)
does exist at t0 and equals 2|γ (t0 )|2 .
2.7. Properties of the Index Function 73
12. Let γ and t0 be as in the previous exercise. Prove that there is an inter-
val (c, d) ⊂ (a, b), containing t0 , such that |γ(t)| is strictly decreasing on
[c, t0 ] and strictly increasing on [t0 , d]. Show that this implies that, if δ =
min{|γ(c)|, |γ(d)|}, then γ(t) crosses each circle of radius less than δ, centered
at 0, exactly once for t ∈ (c, t0 ) and exactly once for t ∈ (t0 , d).
13. Let γ, t0 and (c, d) be as in the previous exercise. Prove that there is an open
disc D, centered at 0, such that γ −1 (D) is an open subinterval J of (c, d). Hint:
Begin by showing you can choose D small enough that it contains no points
γ(t) for t ∈/ (c, d); then use the result of the previous exercise.
14. With γ, t0 , (c, d), and D as in the previous exercise, prove that D is simply
split by γ. This proves Theorem 2.7.11 in the case of a smooth simple point.
15. How would the argument outlined in the previous four exercises need to be
modified to prove Theorem 2.7.11 for a point which is a simple point of γ, but
not a smooth point? Note that, in this case, γ(t) will have two derivatives at
t0 – one from the left and one from the right.
Chapter 3
There is a subtle but crucial difference between statements (a) and (b) in the
above definition: In (b), given , there must be an N that works for all z ∈ S. In
75
76 3. Power Series Expansions
(a), for each z there must be an N , but N depends on z, in general, and there may
not be an N that works simultaneously for all z ∈ S.
The importance of uniform convergence stems primarily from two facts that
are proved below and are extensively used thereafter: (1) The limit of a uniformly
convergent sequence of continuous functions is also continuous; and (2) the integral
along a path of the limit of a uniformly convergent sequence of continuous functions
is the limit of their integrals.
Theorem 3.1.2. If E is a subset of C and {fn } is a sequence of continuous func-
tions on E which converges uniformly on E to a function f , then f is also contin-
uous on E.
.8
.6
.4
x2
.2 x4
x8
x 16
x 32
0
0 .2 .4 .6 .8 1
Figure 3.1.1. The Sequence {xn } does not Converge Uniformly on [0, 1].
The fact that the convergence is not uniform can also be seen directly: No
matter how large n is chosen, we can always find a z with |z| < 1 such that
|fn (z) − f (z)| = |z|n ≥ 1/2. In fact, (1/2)1/n is such a z. Thus, the condition for
uniform convergence fails to hold for = 1/2.
On the other hand, if z ∈ D r (0) with r < 1, then |z| ≤ r and |z|n ≤ r n . Given
> 0, if N is chosen larger than log / log r, then n ≥ N implies
|fn (z) − f (z)| = |z|n ≤ r n < .
Since N was chosen independent of z, the convergence is uniform on Dr (0).
The real analogue of the above example is the sequence {xn } on [0, 1], which is
illustrated in Figure 3.1.1.
Uniform Convergence of Series. We say that an infinite series ∞ k=0 fk (z) of
functions, defined on a set E, converges uniformly on E if the sequence of partial
sums {sn } converges uniformly on E, where we recall that
n
sn (z) = fk (z).
k=0
There is a very useful criterion which insures uniform convergence of such a se-
quence. This is the Weierstrass M -test:
Theorem 3.1.5 (Weierstrass M -Test). Let
∞
(3.1.2) fk (z)
k=0
78 3. Power Series Expansions
such that |fk (z)| ≤ Mk for all k and all z ∈ E, then (3.1.2) converges uniformly on
E.
Proof. The comparison test, comparing (3.1.2) to (3.1.3), shows that, for each
z ∈ E, the series (3.1.2) converges. Let s(z) be the number it converges to, and let
sn (z) denote the nth partial sum of (3.1.2). Then
∞
∞
(3.1.4) |s(z) − sn (z)| ≤ |fk (z)| ≤ Mk .
k=n+1 k=n+1
Since the series (3.1.3) converges, given > 0, we can choose N such that the right
side of (3.1.4) is less than for all n ≥ N . Then (3.1.4) implies that |s(z)−sn (z)| <
for all n ≥ N and all z ∈ E. Since N was chosen independently of z, this shows
that the convergence is uniform.
∞ k 2
Example 3.1.6. Show that the series k=1 z /k converges uniformly on the
closed unit disc D1 (0).
Solution: We have |z k /k2 | ≤ 1/k2 if |z| ≤ 1. Furthermore, ∞ 2
k=1 1/k con-
verges,because it is a p-series with p = 2. Hence, by the Weierstrass M -test, the
∞
series k=1 z k /k2 converges uniformly on D1 (0).
−1
R = lim sup |ck |1/k .
The above theorem has the following corollary, the proof of which is left as an
exercise.
Corollary 3.1.9. If f has a power series expansion about z0 which converges on
the disc DR (z0 ), then f is continuous on this disc.
∞ k
Example 3.1.10. Prove that if k=0 ck z is a power series with radius of con-
∞ k−1
vergence R, then the power series k=1 kck z also has radius of convergence
R.
Solution: If we multiply the second series by z, the set on which the series
converges does not change, and so its radius of convergence does not change. The
∞
resulting series is k=0 kck z k . Let R1 be its radius of convergence. By the previous
theorem,
−1
R = lim sup |ck |1/k ,
and −1 −1
R1 = lim sup |kck |1/k = lim sup k1/k |ck |1/k .
The sequence k1/k has limit 1 (Exercise 3.1.11), and so the factor k1/k does not
effect the lim sup (Exercise 3.1.12). Hence, R1 = R.
1. Show that the sequence {1/(nz)} converges uniformly to 0 on every set of the
form {z : |z| ≥ r} for fixed r > 0, but it does not converge uniformly on
{z : z = 0}.
3.1. Uniform Convergence 81
find a power series expansion for E(z) about 0. Where does this power series
converge?
82 3. Power Series Expansions
Differentiating Power Series. The next theorem and its corollaries concern a
function f defined by a convergent power series
∞
(3.2.1) f (z) = cn (z − z0 )n
n=0
Proof. If f has a power series expansion (3.2.1) with radius of convergence R, let
g be the function which we hope turns out to be the derivative of f – that is, we
set
∞
g(z) = ncn (z − z0 )n−1 .
n=1
This series has the same radius of convergence as the series for f (see Example
3.1.10) and so it converges on DR (z0 ) and converges uniformly on any smaller
closed disc. By Corollary 3.1.9, g is continous on DR (z0 ), and by Theorem 3.1.11,
z ∞
g(w) dw = cn z n = f (z) − f (z0 )
z0 n=1
on DR (z0 ). Theorem 2.6.1 tells us this function is an antiderivative for g(z). Since
f (z0 ) is a constant, this means f = g on DR (z0 ), as required.
Example 3.2.2. Find a power series expansion for the principal branch of the log
function about the point z0 = 1.
Solution: We know that the derivative of log(z) is 1/z (Exercise 2.2.14). We
also know that, since
1 1
= ,
z 1 − (1 − z)
3.2. Power Series Expansions 83
Since we can differentiate term by term, a series which has this as its derivative is
the series
∞ ∞
n (z − 1) (z − 1)n
n+1
(3.2.3) (−1) =− (−1)n−1 .
n=0
n+1 n=1
n
The series (3.2.2) and (3.2.3) both have radius of convergence 1. If f (z) denotes
the sum of series (3.2.3), then f (z) = 1/z = log (z). It follows that f and log
differ by a constant. Since they both have the value 0 at z = 1, they are the same.
Therefore, (3.2.3) is the power series expansion of log(z) about 1.
Using Theorem 3.2.1, the following can be proved in the same way as its real
variable counterpart. The details are left to the exercises.
Corollary 3.2.3. If f has a power series expansion (3.2.1) about z0 , with radius
of convergence R, then it has derivatives of all orders on DR (z0 ). Its kth derivative
is
∞
n! cn
(k)
f (z) = (z − z0 )n−k .
(n − k)!
n=k
In particular, its kth derivative at z0 is given by
f (k) (z0 ) = k! ck .
z t s
r
z0 U
Figure 3.2.1. Setup for the Proof of the Existence of Power Series Expansions.
z0 U z
z1
A function, defined on the union of two open sets and analytic on each of them, is
clearly also analytic on their union; and so g is analytic on U .
Proof. We just need to observe that if D R (z0 ) ⊂ U , with U open, then there is an
open disc Dr (z0 ) with DR (z0 ) ⊂ Dr (z0 ) ⊂ U (Exercise 3.2.4). We can then apply
Theorem 3.2.5 with s = R to obtain a power series expansion of f with coefficients
given by (3.2.4). Then Corollary 3.2.3 relates these coefficients to the derivatives
of f at z0 .
This leads to a very powerful tool. By estimating the size of the integrands
in this formula, we can get estimates on the size of the derivatives of f . These
estimates are called Cauchy’s estimates.
Theorem 3.2.9 (Cauchy’s Estimates). If f is analytic on an open set containing
the closed disc DR (z0 ), and if |f (z)| ≤ M on the boundary of this disc, then
n!M
(3.2.8) |f (n) (z0 )| ≤
Rn
for n = 0, 1, 2, · · · .
This theorem will provide the crucial step in the proof of Liouville’s Theorem
in the next section.
3.2. Power Series Expansions 87
Morera’s Theorem. This is a very handy tool for showing that a function is
analytic.
Theorem 3.2.11 (Morera’s Theorem). Let f be a continuous function defined on
an open set U . If the integral of f is 0 around the boundary of every triangle that
is contained in U , then f is analytic in U .
Proof. Theorem 2.6.1 says that a function f that is continuous on a convex open
set U , and has the property that its integral around any triangle in U is 0, has
a complex antiderivative g in U . However, the fact that g = f on U means, in
particular, that g is analytic on U . But then f is analytic on U by Corollary 3.2.6.
The hypothesis that U is convex in the above argument is not necessary, since
every open set is a union of convex open sets (open discs, in fact). Thus, we can
apply the argument of the previous paragraph to each open disc contained in U
and conclude that f is analytic on each of them. In particular, it has a derivative
at each point of U and, hence, is analytic on U .
14. Use Morera’s Theorem to show that if f is continuous on an open set U and
analytic on U \ E, where E is either a point or a line segment, then f is actually
analytic on all of U .
15. Use Cauchy’s estimates to prove that if {fn } is a sequence of analytic functions
on an open set U , converging uniformly to f on each compact subset of U , then
(k)
{fn } converges uniformly to f (k) on each compact subset of U .
16. Use Morera’s Theorem to prove that if U is an open subset of C, I = [a, b] is
an interval on the real line, and g(z, t) is a continuous function on U × I which
is analytic in z for each t ∈ I, then the function
b
f (z) = g(z, t) dt
a
is analytic in U .
Theorem 3.3.1 (Liouville’s Theorem). The only bounded entire functions are the
constant functions.
Proof. The reader who did Exercise 3.2.10 of the preceding section has nearly com-
pleted the proof of Liouville’s Theorem. The exercise states a simple consequence
of the Cauchy estimates: If a function f is analytic on a disc of radius R, centered
at z0 , and if |f (z)| ≤ M for all z in this disc, then
M
(3.3.1) |f (z0 )| ≤ .
R
If f is bounded and entire, then f (z) is analytic on the entire plane and |f (z)|
is bounded by some number M on the entire plane. This implies that (3.3.1) holds
for all positive numbers R and all z0 ∈ C. If we take the limit as R → ∞ in (3.3.1),
we conclude that |f (z0 )| = 0 for all z0 ∈ C. In other words, the derivative of f is
identically zero. This implies f is a constant (see Exercise 2.6.1).
One has to see the consequences of this theorem to appreciate its power. In
the remainder of this section and the exercises we will introduce a number of these.
Others will occur later.
90 3. Power Series Expansions
This concept satisfies the same basic rules as other kinds of limits: limit of the
sum is sum of the limits, limit of the product is product of the limits, limit of the
quotient is quotient of the limits if the denominator does not have limit zero, etc.
These facts, as well as the fact that
1
lim =0
z→∞ z
(Exercise 3.3.1), are used in the proof of the Fundamental Theorem of Algebra.
Before proving that theorem, we prove the following simple result.
Theorem 3.3.3. If a function f is defined and continuous on the entire plane and
if limz→∞ f (z) exists, then f is bounded on C.
Proof. If limz→∞ f (z) = L, there exists an R > 0 such that |f (z) − L| < 1
whenever |z| > R. By the triangle inequality, this implies
|f (z)| < |L| + 1 if |z| > R.
Since f is continuous on C and DR (0) is closed and bounded, hence compact, f is
bounded on DR (0). Since f is bounded on DR (0) and on its exterior, it is bounded
on all of C.
Theorem 3.3.4 (Fundamental Theorem of Algebra). Every non-constant complex
polynomial has a complex root.
and so
1 h(z)
lim = lim = 0.
z→∞p(z) z→∞ z n
Since the limit of 1/p exists at infinity, the previous theorem implies that 1/p is
bounded on all of C. So Liouville’s Theorem implies that it is a constant. In
fact, the constant must be zero, since 1/p(z) has limit 0 at ∞. This is clearly a
contradiction, since 1/p(z) cannot take on the value 0 on C. We conclude that p(z)
must have a root.
The above result is the essential ingredient in the proof that every complex
matrix can be put in upper triangular form (Exercise 3.3.14).
1 2
Example 3.3.7. Find the eigenvalues of the matrix .
−1 3
Solution: The characteristic polynomial of this matrix is
λ−1 −2
det = λ2 − 4λ + 5.
1 λ−3
By the Quadratic Formula, the roots of this polynomial are 2 ± 2i. Thus, the
eigenvalues of the above matrix are 2 + 2i and 2 − 2i.
If the coefficients of (3.3.3) are real numbers and we are looking only for real so-
lutions to this differential equation, then we exploit the fact that the non-real roots
of a polynomial with real coefficients occur in conjugate pairs (Exercise 3.3.10). If
λ = α + iβ and λ = α − iβ are roots of the auxilliary equation for (3.3.3), then
eλx + eλx
eαx cos βx = and
2
e λx
− eλx
eαx sin βx =
2i
both give real solutions to (3.3.3), as does any linear combination
C eαx cos βx + D eαx sin βx = eαx (C cos βx + D sin βx),
where C and D are arbitrary real constants.
Example 3.3.9. Find all solutions to the differential equation y − 2y + 5y = 0.
Then find all real solutions.
3.3. Liouville’s Theorem 93
y = A e(1+2i)x +B e(1−2i)x ,
where A and B are complex constants. The general real solution is
y = ex (C cos 2x + D sin 2x),
where C and D are real constants.
Proof. We will prove that every polynomial satisfies an inequality of the form
(3.3.4). We leave the converse as an exercise in the application of Cauchy’s esti-
mates.
Suppose p(z) = an z n + an−1 z n−1 + · · · + a1 z + a0 is a polynomial of degree at
most n. Then
p(z)
lim = an .
z→∞ z n
This gives a bound on |p| on the complement of the closed disc DR (0). A closed
disc of finite radius is a compact set and so p is bounded on DR (0) – say, |p(z)| ≤ A
on this disc. If we combine this with our bound on the complement of D R (0), we
conclude that
|p(z)| ≤ A + B|z|n
on all of C, as required.
94 3. Power Series Expansions
Proof. Theorem 3.4.1 shows that, at each point z0 of U , there are two possibilities:
(1) there is a disc centered at z0 in which f is identically 0, and (2) there is a
disc centered at z0 in which f has no zeroes except possibly at z0 itself. Let Vj ,
j = 1, 2 be the set of points z0 for which the jth possibility is the one which occurs.
Obviously, V1 and V2 are both open sets, V1 ∩ V2 = ∅, and V1 ∪ V2 = U . Thus,
either V1 or V2 is empty, since, otherwise, they would separate the connected set
U . If V2 = ∅, then f is identically 0 on U . Since f is not identically 0, we conclude
that V1 = ∅. Since the second of the possibilities in Theorem 3.4.1 is the only one
that occurs in this situation, we conclude that (a) and (b) above both hold at every
z0 ∈ U .
We can modify the disc Dr (z0 ) in statement (a) of the theorem so that it has a
center (which may no longer be z0 ) with rational coordinates and a rational radius
and still has the property that it contains z0 and has no zeroes other than possibly
z0 . We simply choose a point z0 with rational coordinates, and a positive rational
number ρ such that |z0 − z0 | < ρ < r/2. Then z0 ∈ Dρ (z0 ) ⊂ Dr (z0 ) and so Dρ (z0 )
contains z0 but no other zeroes of f . Since we may do this for each zero of f in U
and since there are only countably many discs with rational centers and rational
radii, we conclude that f can have only countably many zeroes in U .
Thus, Part (b) of the above theorem says that the zero set Z(f ) of a non-
constant analytic function on a connected open set U is a discrete subset of U .
It turns out that a subset of an open set U is discrete if and only if no sequence
of distinct points of E converges to a point of U (see Exercise 3.4.1). Of course,
there may be sequences in E which converge to points not in U .
Theorem 3.4.2 has the following easy but important consequence. We leave the
proof as an exercise (Exercise 3.4.2).
3.4. Zeroes and Singularities 97
Theorem 3.4.4 (Identity Theorem). Suppose f and g are two analytic functions
with domain a connected open set U . If f (w) = g(w) at each point w of a non-
discrete subset E of U , then f (z) = g(z) at every point of U .
If the f of Theorem 3.4.2 actually has a zero at z0 , then the integer k is positive.
We call it the order of the zero of f at z0 .
Example 3.4.5. What is the order of the zero of the function f (z) = cos z − 1 at
0? What is the function g of Part (a) of Theorem 3.4.2 in this case if z0 = 0?
Solution: If we subtract 1 from the power series expansion of cos z about 0,
we obtain
z2 z4 z 2n
cos z − 1 = − + + · · · (−1)n + ···
2!
4! 2n!
1 z2 z 2n−2
= z2 − + + · · · (−1)n + ··· .
2! 4! 2n!
We conclude that the order of the zero of cos z − 1 at 0 is 2 and the function g of
Part (a) of Theorem 3.4.2 is the function given by the power series in parentheses
above. This power series has infinite radius of convergence by the ratio test.
Theorem 3.4.6. If an analytic function g is not zero at a point z0 in its domain,
then in some neighborhood V of z0 there is an analytic function h such that g(z) =
eh(z) in V .
Proof. To find such an h, we simply choose a branch of the log function that does
not have g(z0 ) on its cut line. Then the set on which this branch of the log function
is analytic is an open set W which contains g(z0 ). If we set V = g −1 (W ) and define
h on V by
h(z) = log(g(z)),
then V is a neighborhood of z0 and g(z) = eh(z) on V .
There are two types of isolated singularities that are not removable. A function
f defined on U \ {z0 } of the form
g(z)
f (z) = ,
(z − z0 )k
where g is analytic on U , g(z0 ) = 0, and k is a positive integer, is said to have a
pole of order k. If the order of the pole is 1, then it is called a simple pole. An
isolated singularity which is not a pole and is not a removable singularity is called
an essential singularity.
Example 3.4.10. If U is an open set and z0 ∈ U , then analyze the singularities of
a function of the form f /g, where f and g are analytic on U .
Solution: The singularities of f /g in U are all isolated because the zeroes of
g are isolated. If we factor f and g as in Theorem 3.4.2, then
f (z) = (z − z0 )j p(z) and g(z) = (z − z0 )k h(z),
where j and k are the orders of the zeroes of f and g at z0 and p and h are
analytic in U and non-vanishing in some neighborhood of z0 . Then f (z)/g(z) =
(z − z0 )j−k p(z)/h(z) with p(z)/h(z) analytic and non-vanishing in a neighborhood
of z0 . The point z0 is a removable singularity for f /g if j ≥ k and, otherwise, is a
pole of order k − j.
1
Example 3.4.11. Analyze the singularities of the function f (z) = .
1 − ez
Solution: The denominator of this fraction has zeroes at the points {2πki}
for k an integer. Each of these is a zero of order 1 because the derivative of 1 − ez
3.4. Zeroes and Singularities 99
is − ez and this is non-zero for every z and, in particular, is non-zero at the points
{2πki}. It follows from the preceding example that each of these points is a simple
pole for f .
Essential singularities are quite wild. In fact, the Big Picard Theorem states
that a function with an essential singularity at z0 takes on every complex number
but one as a value in every open disc centered at z0 . We will prove the Big Picard
Theorem and its little brother – the Little Picard Theorem – in Chapter 7. The
following is a very much weaker statement than the Big Picard Theorem, but it is
still enough to show that an analytic function behaves very wildly near an essential
singularity.
Theorem 3.4.12. If f is analytic in U \{z0 } and has an essential singularity at z0 ,
then for every open disc D, centered at z0 and contained in U , the set f (D \ {z0 })
has closure equal to the entire complex plane.
The following theorem gives an easy way to identify whether a given isolated
singularity is removable, a pole, or essential.
Theorem 3.4.13. Let f be an analytic function with an isolated singularity at z0 .
Then
(a) f has a removable singularity at z0 if and only if limz→z0 f (z) exists and is
finite;
(b) f has a pole at z0 if and only if limz→z0 f (z) = ∞;
(c) f has an essential singularity at z0 if and only if limz→z0 f (z) does not exist,
even as an infinite limit.
100 3. Power Series Expansions
f (z) f (z)
lim = lim
z→z0 g(z) z→z0 g (z)
and that this limit exists if ∞ is allowed as a possible value.
10. Suppose U is a connected open set and z0 ∈ U . Prove that if f is a non-
constant analytic function on U \ {z0 } and f takes on a certain value c at least
once in every deleted neighborhood of z0 , then f has an essential singularity
at z0 .
102 3. Power Series Expansions
11. Prove that if f is a function which is analytic in the exterior of the closed disc
Dr (0) and if limz→∞ f (z) = 0, then f has a power series expansion of the form
∞
f (z) = an z −n ,
n=1
which converges on the set {z ∈ C : |z| > r}. Hint: Consider the function
g(z) = f (1/z).
1
12. f (z) = .
z − z3
13. f (z) = sin 1/z.
ez −1 − z
14. f (z) = .
z2
1 1
15. f (z) = z − .
e −1 z
log z
16. f (z) = , where log is the principal branch of the log function.
(1 − z)2
Recall that a subset of the plane is compact if it is both closed and bounded.
Suppose U is an open set which is bounded. Then its closure U is both closed and
bounded and, hence, is compact. Suppose U is connected. If f is a continuous
complex-valued function on U , then the continuous real valued function |f (z)| has
a maximum value on U . If f is also analytic on U and non-constant, then the
previous theorem implies that this maximum cannot occur at a point of U . This
means it must occur only on the boundary ∂U . This proves the following corollary
of Theorem 3.5.2.
Corollary 3.5.3. Suppose U is a connected, bounded, open subset of C. If f is
a function which is continuous on U , analytic on U , and non-constant, then the
maximum value of |f (z)| on U is attained on ∂U and nowhere else.
104 3. Power Series Expansions
The typical example of a set U of the type described in the above corollary is
an open disc D r (z0 ) of finite radius.
Example 3.5.4. Find where the function f (z) = z 2 − z attains its maximum
modulus on the closed unit disc.
Solution: By Corollary 3.5.3, the maximum occurs only on the unit circle
{z = eit : t ∈ [0, 2π]}. Thus, we need to find where the maximum modulus of
the function |f (eit )| occurs for t ∈ [0, 2π]. This is equivalent to finding where the
square of the function has a maximum. Thus, we wish to maximize the function
h(t) = | e2it − eit |2 = | eit −1|2 = 2 − 2 cos t.
Clearly, the maximum occurs at t = π and only there. Thus, the maximum of
|z 2 − z| on the closed unit disc is 2 and it occurs only at z = −1.
Proof. Since f (0) = 0, Theorem 3.4.2 implies that f (z) = zg(z), where g is also
analytic in D1 (0). Since |f (z)| ≤ 1 on D1 (0), it follows that
1
|g(z)| ≤ on the circle |z| = r
r
for each r < 1. The Maximum Modulus Theorem implies that this inequality also
holds inside the disc of radius r. Since this is true of each r < 1, we conclude that
|g(z)| ≤ 1 on all of D1 (0) and this implies (3.5.4).
The inequality |f (0)| ≤ 1 follows from (3.5.4), since
f (z)
f (0) = lim = lim g(z) = g(0).
z→0 z z→0
This theorem leads to two important results about harmonic functions: a max-
imum principle for harmonic functions and an integral formula.
Theorem 3.5.8. If u is a harmonic function on a connected open set U and u has
a local maximum at some point z0 ∈ U , then u is constant on U .
The next theorem shows that a harmonic function u has the mean value prop-
erty: the value of u at a point is equal to its mean value over any circle centered at
the point.
Theorem 3.5.9. If u is harmonic on an open set U and Dr (z0 ) ⊂ U , then
2π
1
(3.5.6) u(z0 ) = u(z0 + r eit ) dt.
2π 0
Proof. Choose R > r such that DR (z0 ) ⊂ U . Then Theorem 3.5.7 implies that u
is the real part of an analytic function f on DR (z0 ). The Cauchy Integral Formula
applied to f and the path γ(t) = z0 + r eit yields
2π
1
f (z) = f (z0 + r eit ) dt.
2π 0
Equation (3.5.6) follows from this by equating real parts.
Example 3.5.10. Find a harmonic congugate for the harmonic function x2 − y 2 .
Solution: We recognize x2 − y 2 as the real part of the analytic function z 2 =
(x + iy)2 . The imaginary part of this function is 2xy and so 2xy is a harmonic
conjugate of x2 − y 2 .
Example 3.5.11. Prove that
x
u(x, y) =
x2 + y 2
is harmonic on C \ {0} and find a harmonic conjugate for it.
Solution: We simply observe that u(x, y) is the real part of
x − iy 1
= .
x2 + y 2 z
Therefore, u is harmonic and has
−y
x2 + y 2
as a harmonic conjugate.
1. Find where the function z 2 − 1 attains its maximum modulus on the closed
unit disc.
2. Find where the function | ez | attains its maximum value on the closed unit
disc.
3. Find where the function (z − 1)2 attains its maximum modulus on the triangle
with vertices at 0, 1 + i, 1 − i.
3.5. The Maximum Modulus Principle 107
In this chapter we extend Cauchy’s theorems in two ways: (1) we remove the
condition that the open set U be convex; and (2) we replace the path γ by a more
general type of object called a cycle.
In making these improvements, we finally come to grips with an issue which
we have avoided so far. This issue concerns the topology of the plane. Specifically,
there is a phenomenon that occurs in the plane which does not occur in the line.
On the line, every connected open set is an open interval, which certainly has no
gaps or holes. However, in the plane, there are connected open sets which do have
holes. An example is an open annulus of the form
109
110 4. The General Cauchy Theorems
for coefficients {nj } and {mj } some of which may be zero. Then we define
p
Γ+Λ= (nj + mj )γj .
j=1
Example 4.1.1. Show how to use a path in U , in the sense of the previous chapter,
to produce a 1-chain in U which is a linear combination of smooth paths.
Solution: Suppose γ : [a, b] → U is a closed path. Let
a = t0 < t1 < · · · < tn = b
be a partition such that γ is smooth on each subinterval. Let zj = γ(tj ) for
j = 1, · · · , n. For j = 1, · · · , n let γj be a reparameterization of the restriction of
γ to [tj−1 , tj ] which changes the parameter interval to [0, 1]. If we define a chain Γ
to be the formal sum
n
Γ= γj ,
j=1
then this is a chain consisting of smooth paths.
Definition 4.1.3. Suppose Γ and Λ are chains with Γ(I) and Λ(I) both subsets
of a set E ⊂ C. We will say that Γ and Λ are E-equivalent if
f (z) dz = f (z) dz
Γ Λ
for every continuous function f on E. If Γ(I) = Λ(I) = E and Γ and Λ are
E-equivalent, then will will simply say that Γ and Λ are equivalent.
Example 4.1.4. Let Δ be the triangle with vertices {a, b, c} listed in counterclock-
wise order. Recall that ∂Δ denotes the path which traverses the boundary of Δ
once in the counterclockwise direction. Show that ∂Δ and Γ = [a, b] + [b, c] + [c, a]
are equivalent, as are ∂Δ and Λ = [a, b] + [b, c] − [a, c].
Solution: Clearly ∂Δ(I) = Γ(I) = Λ(I) – each of them is just the boundary of
the triangle Δ. Also, it follows from Theorem 2.4.6, Parts (b) and (c), that ∂Δ, Γ,
112 4. The General Cauchy Theorems
and Λ all determine the same integral for continuous functions on the boundary of
Δ. Thus, by definition, the three chains are equivalent.
0-Chains. Just as the group of 1-chains is the free abelian group generated
by paths, the group of 0-chains in U is the free abelian group generated by the
singleton subsets of C (sets {z} containing a single point z ∈ C) – that is, it is the
group of formal linear combinations,
p
mj {zj },
j=1
⎛ ⎞
p
p
∂⎝ m j γj ⎠ = (mj {γj (1)} − mj {γj (0)}).
j=1 j=1
This may not result in a linear combination of distinct singleton sets, but
we fix this, combining terms involving the same singleton subset by adding their
coefficients. In order for ∂ to determine a well defined map from 1-chains to 0-
chains, one must show that it preserves the identifications made in our definition
of 1-chains. This simply requires that any summand in a 1-chain that has zero
as coefficient is sent by ∂ to a 0-chain with all zeroes as coefficients, and this is
obviously true.
Once we know that ∂ is well defined, it is quite evident from the definition that it
is a group homomorphism ∂ from 1-chains to 0-chains (meaning ∂(Γ+Λ) = ∂Γ+∂Λ).
This ensures that ker ∂ = {Γ : ∂Γ = 0} is closed under addition and taking additive
inverses and, hence, is a subgroup of the set of 1-chains in U . Elements of this
4.1. Chains and Cycles 113
Proof. We will make a succession of changes to Γ, none of which will change Γ(I)
or the integral (4.1.1).
We first express Γ as a linear combination of paths in which the coefficients are
all 1 or −1. That is, each term mj γj is replaced by a sum of mj copies of γj if mj
is positive and by a sum of −mj copies of (−1)γj if mj is negative. This does not
change the chain Γ.
The next step is to replace each summand of the form (−1)γj with the path
−γj , where we recall that −γj (t) = γj (1 − t), so that −γj is γj traversed in the
reverse direction. Theorem 2.4.6, Part (c), implies that this change results in a
cycle Γ̃ which is equivalent to Γ.
At this point we have replaced Γ with a cycle Γ̃ which is a simple sum of some
number of paths. Say the number is n. We next show that if these paths are not
all closed paths, then we can replace Γ̃ with a sum of fewer than n paths without
changing the integral.
Suppose γj is a path in Γ̃ which is not closed. Then γj (0) = γj (1). Since
∂ Γ̃ = 0, γj (1) must be equal to γk (0) for some path γk in Γ̃; otherwise there would
be no term to cancel with {γj (1)} in the expression for ∂Γ. We may join γk and
γj , as in (2.4.2), to form a new path which begins at γj (0) and ends at γk (1). By
Theorem 2.4.6, Part (b), replacing γj and γk in Γ̃ by this new path does not change
the integral. Thus, if its summands are not all closed paths, we may replace Γ̃ with
114 4. The General Cauchy Theorems
a path with fewer summands, which determines the same integral. This forms the
basis for an induction argument, which is carried out below.
If a cycle consists of a single path γ, then the condition
∂γ = {γ(1)} − {γ(0)} = 0
implies γ is a closed path. Suppose we know that any cycle which can be written
as a sum of fewer than n paths may be replaced by an equivalent cycle which is a
sum of closed paths. Then the argument of the previous paragraph shows that any
cycle which can be written as a sum of n paths is equivalent to some cycle which
is a sum of closed paths. By induction, every cycle is equivalent to a sum of closed
paths.
Index of a Cycle. The notion of index of a path around a point can be extended
to cycles as follows.
Definition 4.1.8. If Γ is a 1-cycle in C and z ∈ C is a point which does not lie on
Γ(I), then we set
1 1
IndΓ (z) = dw.
2πi Γ w − z
The number IndΓ (z) is called the index of Γ around z.
By Theorem 4.1.7 the cycle Γ may be replaced by one which is a sum of closed
paths without changing Γ(I) or the integral defining IndΓ (z). Thus, IndΓ (z) is a
finite sum of numbers of the form Indγ (z), where γ is a closed path. This, together
with Theorems 2.6.6 and 2.7.5 proves the first three parts of the following theorem.
The last part follows immediately from (4.1.2).
Theorem 4.1.9. Let Γ be a 1-cycle in C. Then
(a) IndΓ is an integer-valued function defined on the complement of Γ(I);
(b) IndΓ is constant on each component of C \ Γ(I);
(c) IndΓ is 0 on the unbounded component of C \ Γ(I);
(d) if Λ is also a cycle, and z ∈
/ Γ(I) ∪ Λ(I), then
IndΓ+Λ (z) = IndΓ (z) + IndΛ (z).
Find Indγ (z), IndΓ1 (z), and IndΓ (z) for each value of z which is an integer.
4.1. Chains and Cycles 115
... ...
... ...
γ−N γ−2 γ−1 γ0 γ1 γ 2 γN
Solution: The paths γ and γj are circles traversed once in the counterclockwise
direction. For such a path we know the index is 1 inside the circle and 0 outside
the circle. Thus, if k is an integer,
1, if |k| ≤ N ;
Indγ (k) =
0, if |k| > N ;
and
1, if k = j;
Indγj (k) =
0 j.
if k =
Thus,
k
1, if |k| ≤ N ;
IndΓ1 (k) = Indγj (k) =
j=−k
0, if |k| > N ;
and so,
IndΓ (k) = Indγ (k) − IndΓ1 (k) = 0
for every integer k.
Homologous Cycles. In the general Cauchy Theorem of the next section, the
condition that U be convex is replaced by a condition on the path or cycle over which
the integration takes place. The condition is that the path or cycle be homologous
to 0 in the sense of the following definition.
Definition 4.1.11. Let U be an open subset of C and let Γ and Λ be 1-cycles in
U . We will say that Γ and Λ are homologous in U if
IndΓ (z) = IndΛ (z)
116 4. The General Cauchy Theorems
γ2 γ1
γ3 γ4
IndΓ (z) = 0
Example 4.1.12. If γ1 (t) = e2πit and γ2 (t) = 2 e2πit for t ∈ [0, 1], show that
Γ = γ2 − γ1 is a 1-cycle which is homologous to 0 in U = {z ∈ C : 1/2 < |z| < 3},
but is not homologous to 0 in the set V = U \ {3/2}.
Solution: The curve γ1 has index 1 about points inside the unit disc D1 (0),
and index 0 about points exterior to the closed unit disc, while γ2 has index 1
about points in D2 (0) and 0 about points exterior to the closure of this disc. The
complement of U is
D1/2 (0) ∪ (C \ D3 (0)).
Since
Indγ1 = Indγ2 = 1 on D1/2 (0)
and
Indγ1 = Indγ2 = 0 on C \ D3 (0),
we conclude that Γ = γ2 − γ1 is homologous to 0 in U .
However, at the point z = 3/2, γ1 has index 0 and γ2 has index 1. Since 3/2 is
in the complement of V = U \ {3/2}, Γ is not homologous to 0 in V .
4.1. Chains and Cycles 117
γ1 (t) = 4 e2πit ,
γ2 (t) = i + e2πit ,
γ3 (t) = −i + e2πit .
for all (z0 , w0 ) ∈ U × U . Here, the limit has the usual -δ definition, using the
appropriate definition of the distance between (z, w) and (z0 , w0 ), which is
|(z, w) − (z0 , w0 )| = |z − z0 |2 + |w − w0 |2 .
This is the same as the Euclidean distance between these two points, considered
as points of R4 . It is easy to see that this type of limit satisfies all the usual limit
rules regarding sums, products, quotients, etc.
Lemma 4.2.1. If g(z, w) is defined by (4.2.1), then g is a continuous function of
two complex variables on U × U .
If |(z, w) − (z0 , z0 )| < δ, then both z and w are in Dδ (z0 ), as is any point λ on the
line segment joining them. If z = w, integrating over this line segment yields
w
f (w) − f (z) = f (λ) dλ,
z
and so
1 w
|g(z, w) − f (z0 )| = (f (λ) − f (z0 )) dλ < .
w−z z
If z = w, then |g(z, w) − f (z0 )| = |f (z) − f (z0 )| < . Hence, g is continuous at
(z0 , z0 ). This completes the proof.
The Cauchy Integral Theorem. The Cauchy Integral Theorem follows easily
from Theorem 4.2.2.
Theorem 4.2.4. If f is analytic in the open set U and if Γ is a 1-cycle in U which
is homologous to 0 in U , then
f (z) dz = 0.
Γ
Then g(z) is also analytic in U and g(z0 ) = 0. Theorem 4.2.2, applied to the
function g, tells us that
g(z)
f (z) dz = dz = 2πi IndΓ (z0 )g(z0 ) = 0.
Γ Γ z − z0
This completes the proof.
1
Example 4.2.5. Find dz if Γ is the cycle of Example 4.1.10.
Γ sin(πz)
Solution: By Example 4.1.10, Γ has index 0 at every integer k and so it is
homologous to 0 in U = C \ Z, where Z is the set of integers. The function sin(πz)
vanishes exactly on the set of integers (Exercise 4.2.1) and so 1/ sin(πz) is analytic
in U . It follows from the above theorem that the integral in question is 0.
Simple Closed Paths. Although we have stated the general Cauchy theorems
in terms of integration over 1-cycles, they also hold if the integration is over an
ordinary closed path. After all, a closed path is just a particularly simple 1-cycle.
Classically, these theorems are stated and proved for integrals over a particular kind
of closed path – a simple closed path.
A simple closed curve is a closed curve which does not intersect itself except at
the endpoints of the parameter interval I = [a, b]. That is,
Definition 4.2.6. A closed curve γ, defined on [a, b], is called a simple closed curve
if γ(s) = γ(t) for a ≤ s < t ≤ b unless s = a and t = b.
The Jordan Curve Theorem says that if γ is a simple closed curve, then the
complement of γ(I) has two components – one unbounded, and one bounded. Here
we will prove a weak version of the Jordan Curve Theorem – one that is sufficient
for our purposes.
Recall from Definition 2.7.10 that a simple point of a path γ is a point that the
path passes through just once and at which the left and right derivatives of γ are
both non-zero.
Definition 4.2.7. A closed path will be called a simple closed path if each of its
points is a simple point.
Thus, a simple closed path is a simple closed curve which is piecewise smooth
with non-zero left and right derivatives at each point. Our weak version of the
Jordan Curve Theorem is the following.
Theorem 4.2.8. If γ is a simple closed path, then C \ γ(I) has exactly two compo-
nents – a bounded component on which Indγ (z) = ±1 and an unbounded component
on which Indγ (z) = 0.
Proof. By Theorems 2.7.8 and 2.7.11, we may choose for each point z ∈ γ(I) an
open disc Dz , centered at z, such that Dz \ (Dz ∩ γ(I)) has exactly two components
– a left component Lz and a right component Rz , and Indγ (z) is one unit greater
on Lz than on Rz (see Figure 4.2.2).
It is not difficult to see that two sufficiently close points of γ(I) have discs with
overlapping left components and overlapping right components. It follows that the
122 4. The General Cauchy Theorems
union of all the left components is a connected open set L, while the union of all
the right components is a connected open set R (see Exercise 4.2.11). Furthermore,
Indγ (z) = Indγ (w) + 1 if z ∈ L and w ∈ R. It follows that L and R are disjoint.
The union of all the discs Dz is an open set U containing γ(I) and, in fact, U is
the union of the disjoint sets γ(I), L, and R.
Now every component of the complement of γ(I) has a subset of γ(I) as its
boundary (Exercise 4.2.12). This implies that every such component must have
non-empty intersection with U and, hence, must meet either L or R. But since L
and R are connected, a component which meets one of them must, in fact, contain
it. It follows that there are only two components of the complement of γ(I) –
one containing L and one containing R. One of these must be the unbounded
component and Indγ is zero on it. The other one is a bounded component and Indγ
must be plus or minus one on it, depending on whether it contains L or R. This
completes the proof of the theorem.
From the proof of the above theorem, it follows that Indγ (z) = 1 on the bounded
component of C\γ(I) if this component contains L. Intuitively, this means that the
bounded component is on the left as the path is traversed in the positive direction.
In this case, we say that the path has positive orientation. The other possibility is
that Indγ (z) = −1 on the bounded component, which implies that this component
contains R – i.e., that the bounded component is on the right as the path is traversed
in the positive direction. In this case, we say that the path has negative orientation.
4.2. Cauchy’s Theorems 123
If γ is a simple closed path, we will call the bounded component of C \ γ(I) the
inside of γ and the unbounded component the outside of γ.
We may now apply Theorems 4.2.2 and 4.2.4 in the case where the cycle Γ is
a single simple closed path γ to obtain the Cauchy Theorem and Formula in their
classical forms.
Theorem 4.2.9. If γ is a simple closed path and f is a function analytic in an
open set U containing both γ(I) and its inside, then
(4.2.9) f (w) dw = 0
γ
and
1 f (w)
(4.2.10) f (z) = dw
2πi γ w−z
for each z on the inside of γ(I).
Proof. We know that Indγ (z) = 0 on the outside of γ, and this contains the
complement of U (since U contains γ(I) and its inside). Thus, the hypothesis of
Theorem 4.2.2 is satisfied, and so we know that (4.2.9) is true and that
1 f (w)
Indγ (z)f (z) = dw
2πi γ w − z
for z on the inside of γ. Then (4.2.10) follows if we just observe that Indγ (z) = 1
on the inside of γ.
1. Verify the fact, used in Example 4.2.5, that sin(πz) = 0 if and only if z is an
integer. Hint: We know this if z is real, but how do we know there are no
zeroes of sin z except those on the real line?
2. Let Γ = γ1 − γ2 − γ3 , where γ1 , γ2 , and γ3 are positively oriented circles with
radii 5, 1, and 1 and centers 0, −2, and 3, respectively. Find
1
dz
Γ (z + 2)(z − 3)
without calculating the integrals around the individual paths γi .
3. Check your answer to the previous problem by calculating
1
dz
γj (z + 2)(z − 3)
for each of the individual paths γj . Hint: The integral around γ1 may be
calculated by using a partial fraction decomposition of the integrand.
4. If γ is any simple closed path, with −2 and 3 inside γ, find
1
dz.
γ (z + 2)(z − 3)
124 4. The General Cauchy Theorems
9. For positive numbers r < R, set γ1 (t) = r e2πit and γ2 (t) = R e2πit for t ∈ [0, 1]
and set Γ = γ2 − γ1 . Prove that if f is analytic in an open set containing the
set {z : r ≤ |z| ≤ R}, then
1 f (w)
(4.2.11) f (z) = dw
2πi Γ w − z
for every z with r < |z| < R.
10. For the cycle Γ and function f of the previous problem, what is the integral
(4.2.11) if |z| < r or |z| > R?
11. Let γ be a simple closed path and, for each z ∈ γ(I), let Dz , Lz and Rz be the
sets used in the proof of Theorem 4.2.8. Prove that if z and w are two points
of γ(I) such that Dz ∩ Dw ∩ γ(I) = ∅, then Lz ∩ Lw = ∅ and Rz ∩ Rw = ∅.
Show that this implies that the union of all the Lz is a connected set as is the
union of all the Rz .
12. Prove that if K is a closed subset of C, then each component of C \ K has its
boundary contained in K.
Lemma 4.3.1. If h is a function which is analytic in the exterior of the disc Dr (z0 )
and vanishes at ∞, then the function
h(1/w + z0 ), if w = 0;
q(w) =
0, if w = 0;
is analytic in D1/r (0).
Proof. The number 1/w + z0 is in C \ Dr (z0 ) if and only if 1/|w| > r, that is, if
and only if w ∈ D1/r (0). So the domain of q is D1/r (0). Clearly q is analytic on
the complement of {0} in this disc.
The fact that limz→∞ h(z) = 0 implies that limw→0 q(w) = 0. Thus, q is
continuous at 0. It follows from the Removable Singularity Theorem (Theorem
3.4.8) that it is analytic on all of D1/r (0).
where M = sup{|f (w)| : w ∈ γs (I)}. From this bound on the integrand in (4.3.4),
it follows that
Ms
|h(z)| ≤
|z| − |z0 | − s
and so, since |z0 | and s are fixed,
lim h(z) = 0,
z→∞
as required.
The uniqueness of g and h follows from Liouville’s Theorem. In fact, suppose
we also have
f = g1 − h1
4.3. Laurent Series 127
S
z
R
r s
z0
γs γS
Example 4.3.3. Find a decomposition like that in the previous theorem if the
function f is
1
f (z) =
(z − 1)(z − 2)
and the annulus A is A = {z ∈ C : 1 < |z| < 2}.
Solution: This is just the partial fraction decomposition of f . We set
1
g(z) = for |z| < 2,
z−2
and
1
h(z) = for |z| > 1.
z−1
Then g and h are analytic, f = g − h on A, and h vanishes at infinity.
in A = D2 (0) ∩ (C \ D1 (0)).
1
Example 4.3.6. The function f = of the previous exercise is also
(z − 1)(z − 2)
analytic on the annulus
B = {z ∈ C : 0 < |z − 1| < 1},
4.3. Laurent Series 129
while h(z) is analytic in C \ {1} with limit 0 at infinity. Since it is already a power
of (z − 1), it needs no further expansion. Thus, the Laurent expansion of f in the
annulus B is
∞
∞
−1
f (z) = − (z − 1) − (z − 1)
n
=− (z − 1)n .
n=0 n=−1
Example 4.3.7. Find the Laurent expansion of e1/z in the annulus C \ {0}.
Solution: Since f (z) = e1/z is analytic except at 0 and has limit 1 at infinity,
in the decomposition f = g −h of Theorem 4.3.2 the function g is 1 and the function
h is 1 − e1/z . The Laurent expansion is
0
zn
1/z
e = .
n=−∞
|n|!
A path which traverses the circle |w − z0 | = s once in the positive direction is given
by γ(t) = z0 + s eit , for 0 ≤ t ≤ 2π. Then
This leads to the following important application of the general Cauchy Theo-
rem.
Theorem 4.4.3 (Residue Theorem). Let f be a function which is analytic on U \E,
where U is an open subset of C and E is a discrete subset of U . If γ is a closed
path in U \ E which is homologous to 0 in U , then
(a) there are only finitely many points of E at which Indγ is non-zero;
(b) if these points are {z1 , z2 , · · · , zn }, then
n
1
(4.4.2) f (z) dz = Indγ (zj ) Res(f, zj ).
2πi γ j=1
Proof. We first prove (a). Recall from the proof of Theorem 2.7.5 that if r is
chosen so that γ(I) ⊂ Dr (0), then the bounded components of C \ γ(I) are also
contained in Dr (0). Also, Indγ (z) is non-zero only on certain bounded components
of C \ γ(I). It follows that the union of γ(I) and the components of its complement
where Indγ (z) is non-zero is a bounded set K. The set K is also the complement
of the union of the components of C \ γ(I) on which Indγ = 0, and so it is closed.
Thus, K is compact.
Since γ is homologous to 0 in U , every point of the complement of U is a point
where Indγ is 0. This implies K ⊂ U . Since the singularities of f form a discrete
subset E of U , we may choose for each point of U an open disc, centered at that
point and contained in U , which either contains one singularity (the center of the
132 4. The General Cauchy Theorems
zj
γj
disc) or no singularities. This collection of open discs covers K (since its union
is U ), and so some finite subcollection also covers K. But this means there are
only finitely many singularities of f in K, and, hence, there are only finitely many
singularities of f at which Indγ is nonzero. This completes the proof of (a).
Let z1 , z2 , · · · , zn be the singularities of f at which Indγ is nonzero. For each
of these points zj , we choose an rj > 0 such that Drj is contained in the open set
U \ γ(I). We choose r > 0 such that r < min{r1 , · · · , rn } and the discs Dr (zj ) are
non-overlapping. We then set
where γj (t) = zj + r e2πit for t ∈ [0, 1] (we may assume that γ also has [0, 1] as its
parameter interval).
Now each zj is a point where Indγ is non-zero, and so is every point in the open
disc Dr (zj ). Hence the closure D r (zj ) of this disc is contained in K ⊂ U . This
means that the complement of U is contained in the complement of Dr (zj ). Thus,
Indγj (z) = 0 on the complement of U . Since this is true for every j and is also true
of γ, we have that IndΓ (z) = 0 on the complement of U – that is, Γ is homologous
to 0 in U .
Note, f is not analytic in U and so the general Cauchy Theorem does not yet
apply. However, we also have that
while
mj Indγj (zk ) = 0 for k = j.
4.4. The Residue Theorem 133
Now (4.4.2) follows from this, Theorem 4.4.2, and (4.4.3). This completes the
proof.
1
Solution: The integrand f (z) = has isolated singularities inside γ at 0
sin z
and π. The function
z
g(z) =
sin z
has a removable singularity at 0 and the value of the resulting analytic function at
0 is 1. Thus,
Res(f, 0) = 1.
Since sin(z) = − sin(z − π), the function
z−π z−π
h(z) = =−
sin z sin (z − π)
has a removable singularity at z = π. The value at π of the resulting analytic
function is −1. Thus,
Res(f, π) = −1.
It follows from the Residue Theorem that the integral of f around γ is zero.
If we combine this with the Residue Theorem, the result is the following theo-
rem.
Theorem 4.4.8. Let f be a meromorphic function defined in a open set U and let
γ be a closed path in U which is homologous to 0 in U . Assume there are no zeroes
or poles of f on γ(I) and the zeroes and poles of f at which Indγ is not zero occur
at the points z1 , z2 , · · · , zn . Then
n
1 f (z)
(4.4.6) mj kj = dz,
j=1
2πi γ f (z)
If the path γ is simple, then the previous theorem and corollary are simpler.
Corollary 4.4.10. Let f be a meromorphic function on the open set U . If γ is a
simple closed path in U , with its inside contained in U , which does not pass through
a zero or pole of f , and {z1 , · · · , zn } is the set of zeroes and poles of f inside γ,
then
n
1 f (z)
(4.4.8) kj = dz = Indf ◦γ (0),
j=1
2πi γ f (z)
where, for j = 1, · · · , n, kj is the order of the zero or minus the order of the pole
at zj .
4. If γ is a simple closed path with 0 inside, but no other multiples of 2πi inside,
find
1
dz.
γ e −1
z
path which does not pass through 1 or −1, what are the possible values for
f (z)
and what determines which value is achieved?
γ f (z)
10. Use Theorem 4.4.8 to derive a formula for γ cot z dz, where γ is any closed
path which does not pass through an integral multiple of π.
z+1
11. Prove that there is an analytic logarithm for defined in the open set
z−1
U = C \ [−1, 1].
12. Example 4.4.11 is only one of many possible theorems concerning the existence
of analytic logarithms that can be proved using Theorem 4.4.8. Invent and
prove another one.
Proof. If f has no zeroes on γ(I), then (4.5.1) implies that g also has no zeroes
on γ(I). If we set h(z) = f (z)/g(z), then h is meromorphic in U and has no zeroes
or poles on γ(I). The inequality (4.5.1) implies that the curve h ◦ γ satisfies
|h ◦ γ(t) − 1| ≤ 1
for all t ∈ I. It follows from this that 0 is in the unbounded component of C\h◦γ(I).
Hence,
Indh◦γ (0) = 0.
The path γ is homologous to 0 in U and so we may apply Corollary 4.4.9. It tells
us that the number of zeroes minus the number of poles of h inside γ, counting
order, is 0. However, this is the number of zeroes of f minus the number of zeroes
of g inside γ, counting order. Hence, these two numbers are the same.
Example 4.5.2. How many zeroes, counting order, does the polynomial
4z 5 − z 3 + z 2 − 2
have inside the unit circle?
Solution: We apply Rouché’s Theorem with f (z) = 4z 5 − z 3 + z 2 − 2 and
g(x) = 4z 5 . On the unit circle |z| = 1, we have
|f (z) − g(z)| = | − z 3 + z 2 − 2| ≤ |z|3 + |z|2 + 2 = 4 = |g(z)|.
By Rouché’s Theorem, f and g have the same number of zeroes inside the unit
circle. Since g(z) = 4z 5 has 5 zeroes, counting order, inside this circle, so does f .
138 4. The General Cauchy Theorems
f
V
U −1
f
Proof. We set w0 = f (z0 ). If f (z0 ) = 0, then the function f (z) − w0 also has non-
vanishing derivative at z0 and, hence, it has a zero of order 1 at z0 . Furthermore,
by Theorem 3.4.2 there is an r > 0 such that, in the disc Dr (z0 ), this is the only
zero of f (z) − w0 .
We choose δ such that 0 < δ < r and f (z) = 0 for all z ∈ Dδ (z0 ). This is
possible because f is continuous and so f −1 (C \ {0}) is an open set containing z0 .
We next let γ be the circle
γ(t) = z0 + δ e2πit , t ∈ [0, 1].
Then f (z)−w0 has no zero on the compact set γ(I). This means that the minimum
value of |f (z) − w0 | in γ(I) is positive. Then if |w − w0 | < and z ∈ γ(I), we
have
|(f (z) − w) − (f (z) − w0 )| = |w − w0 | < ≤ |f (z) − w0 |.
By Rouché’s Theorem, the two functions of z, f (z) − w and f (z) − w0 , have the
same number of zeros, counting order, in Dδ (z0 ). Since, f (z) − w0 has one zero in
this disc, so does f (z) − w, and this is true for each w ∈ D (w0 ).
We conclude from the above that, for each w ∈ D (w0 ), there is exactly one
z ∈ Dδ (z0 ) such that f (z) = w. In other words, if we set
W = D (w0 ) and V = f −1 (W ),
then f : V → W is one-to-one and onto, and, hence, has a well-defined inverse
function f −1 .
We claim that f −1 is continuous on W . By Theorem 2.1.13, to show this,
we only need to show that its inverse function, f , takes open sets to open sets.
However, we just showed that an analytic function f on an open set U , which has
a non-zero derivative at a point z0 , takes U to a set which contains a neighborhood
of f (z0 ). This result applies equally well to any open set containing any point z
of V , since f (z) = 0 for z ∈ V . Thus, f takes any open subset of V to an open
subset of W and this implies that f −1 : W → V is continuous.
We next show that the inverse function is analytic and has the indicated de-
rivative. If w and w1 are two points of W , with z = f −1 (w) and z1 = f −1 (w1 ) the
corresponding points of V , then w = f (z), w1 = f (z1 ), and
−1
f −1 (w) − f −1 (w1 ) f (z) − f (z1 ) 1 1
lim = lim = = −1 .
w→w1 w − w1 z→z1 z − z1 f (z1 ) f (f (w1 ))
140 4. The General Cauchy Theorems
1
This shows that (f −1 ) (w1 ) exists and equals . Of course, the fact that
f (f −1 (w 1 ))
z → z1 as w → w1 follows from the continuity of f −1 .
at g(z0 ) = 0. Then the image of V2 under g k is the open disc Dk (0), and so the
image of V2 under f is f (z0 ) + Dk (0). Since this is a neighborhood of f (z0 ) which
is contained in f (V ), the proof is complete.
4.6. Homotopy
This section is devoted to developing tools for calculating the index of a closed
path (or more generally a cycle) about a point. This is crucial, since both the
hypotheses and conclusions of the Cauchy theorems involve the index. The tools
we will develop involve ideas from algebraic topology applied in the plane.
We begin by extending the index function to closed curves which are not nec-
essarily piecewise smooth. The key to doing this is to show that all closed paths
142 4. The General Cauchy Theorems
sufficiently near a given closed curve have the same index, and then to show that
every closed curve can be approximated arbitrarily closely by closed paths. The
common value of the index of all closed paths near a given closed curve γ is then
defined to be the index of γ. The key to the first part of this program is the fol-
lowing theorem, which is closely related to Rouché’s Theorem and has a similar
proof.
Theorem 4.6.1. Suppose z ∈ C and γ1 and γ2 are two closed paths which do not
pass through z and which satisfy
|γ1 (t) − γ2 (t)| ≤ |γ2 (t) − z|
for all t ∈ I. Then Indγ1 (z) = Indγ2 (z).
Proof. By translating γ1 and γ1 by z and using the fact, obvious from the definition
of Ind, that Indγ (z) = Indγ−z (0), we may assume that z = 0. Then our hypothesis
is that γ1 and γ2 are two paths which do not pass through 0 and which satisfy
|γ1 (t) − γ2 (t)| ≤ |γ2 (t)|.
We define a new path γ by
γ1 (t)
γ(t) = .
γ2 (t)
This path does not pass through 0, and it satisfies
|γ(t) − 1| ≤ 1.
In other words, γ(I) lies in the closed unit disc of radius 1 centered at 1 and does
not contain 0. It follows from this that 0 is in the unbounded component of C\γ(I).
Hence,
Indγ (0) = 0.
To complete the proof, we will show that
(4.6.1) Indγ (0) = Indγ1 (0) − Indγ2 (0).
In fact,
b
1 1 1 γ (t)
Indγ (0) = dz = dt,
2πi γ z 2πi a γ(t)
and so (4.6.1) follows from the calculation
γ γ γ2 − γ1 γ2 γ γ
= 1 = 1 − 2.
γ γ1 γ2 γ1 γ2
This completes the proof.
Index for Closed Curves. Recall that a curve is just a continuous function
γ : I → C, where I is a closed bounded interval – no differentiability is assumed.
For simplicity in the following discussion, we will just work with curves for which
the parameter interval I is [0, 1].
There is a notion of distance between two curves parameterized on I = [0, 1].
Definition 4.6.2. The distance between curves γ1 and γ2 on I is defined to be
||γ1 − γ2 || = sup |γ1 (t) − γ2 (t)|.
t∈I
4.6. Homotopy 143
With this notion of distance, we will show that every curve can be approximated
arbitrarily closely by curves which are piecewise smooth, in fact, piecewise linear.
Here, by a linear curve, we mean a curve of the form
γ(t) = u + tv
for fixed complex numbers u and v and t ranging over some parameter interval.
Such a curve traces a straight line segment. The linear curve which traces the line
segment from w to z and has parameter interval [a, b] is given by
b−t t−a
(4.6.2) γ(t) = w+ z
b−a b−a
(see Exercise 4.6.1).
A curve on I is piecewise linear if there is a partition 0 = t0 < t1 < · · · < tn = 1
of I such that γ is linear on each subinterval [tj−1 , tj ]. Obviously, piecewise linear
curves are piecewise smooth and, hence, are paths.
Theorem 4.6.3. If γ : I → C is a curve, then for each > 0 there is a piecewise
linear curve γ̃ such that ||γ̃ − γ|| < .
The next theorem tells us that, given a closed curve γ and a point z not on
γ(I), all closed paths sufficiently near γ have the same index about z.
144 4. The General Cauchy Theorems
Theorem 4.6.4. If γ is a closed curve and z ∈ C, a point not on γ(I), then there
is a δ > 0 such that if γ1 and γ2 are paths with ||γ − γj || < δ for j = 1, 2, then
Indγ1 (z) = Indγ2 (z).
With this definition, Theorem 4.6.4 implies that Indγ (z) is a locally constant
function of the closed curve γ. That is, the following theorem holds.
Theorem 4.6.6. If z ∈ C and γ is a closed curve in C such that z ∈ / Γ(I), then
there is a δ > 0 such that if γ1 is any other closed curve in C and ||γ − γ1 || < δ,
then
Indγ (z) = Indγ1 (z).
Proof. We choose δ as in Theorem 4.6.4. Then all closed paths within a distance δ
of γ have the same index around z, and this is the index of γ around z, by definition.
If γ1 is a closed curve with ||γ − γ1 || < δ, set
δ1 = δ − ||γ − γ1 ||.
Then all closed paths within a distance δ1 of γ1 are within a distance δ of γ, and
so they all have the same index about z as γ does. By definition, this is also the
index of γ1 about z.
4.6. Homotopy 145
Homotopy. Let U be an open set in C and let γ0 and γ1 be two closed curves in
U with parameter interval [0, 1]. The curves are said to be homotopic if it is possible
to continuously deform γ0 to γ1 while remaining in U . This is made precise in the
following definition.
Definition 4.6.7. Two closed curves γ0 and γ1 in U , with parameter interval [0, 1],
are said to be homotopic in U if there is a continuous function h, defined on the
square [0, 1] × [0, 1] in R2 , with values in U , such that
(a) h(0, t) = γ0 (t) for all t ∈ [0, 1];
(b) h(1, t) = γ1 (t) for all t ∈ [0, 1]; and
(c) h(s, 0) = h(s, 1) for all s ∈ [0, 1].
In the above definition, for each fixed s, the function γs (t) = h(s, t) defines a
curve γs in U . Parts (a) and (b) say that when s = 0 and 1, these curves are the
original curves γ0 and γ1 . Part (c) says that each of the curves γs is closed.
The family of curves {γs } determined by a homotopy h, as in the above defini-
tion, is a continuous one-parameter family of curves in the sense that the following
theorem is true.
Theorem 4.6.8. If {γs } is the family of curves determined by a homotopy, as in
Definition 4.6.7, then for each s0 ∈ I and each > 0, there is a δ > 0 such that
||γs − γs0 || < whenever |s − s0 | < δ.
Proof. This proof uses the fact that a continuous function on a compact set such
as the square I × I, is uniformly continuous. For the continuous function h, this
means that given > 0 there is a δ > 0, such that
|h(s, t) − h(s0 , t0 )| < whenever |(s, t) − (s0 , t0 )| < δ,
for any pair of points (s, t), (s0 , t0 ) ∈ I × I. In particular, in the case where t = t0
this says that if |s − s0 | < δ, then
|γs (t) − γs0 (t)| < for all t ∈ I,
which implies ||γs − γs0 || < , and this completes the proof.
We are now in a position to prove the homotopy theorem for the index function.
Theorem 4.6.9. If γ0 and γ1 are two closed curves in U which are homotopic in
U , then
Indγ0 (z) = Indγ1 (z)
for every z ∈ C \ U .
We conclude that Indγs (z) is constant in the open interval (s0 − δ, s0 + δ). Since s0
is an arbitrary point of I, we conclude that the set on which Indγs (z) takes on any
given value is an open set. Since I is connected, and the union of these open sets
is I, there can be only one of them that is non-empty. Thus, Indγs (z) is constant
on I and, in particular, Indγ1 (z) = Indγ0 (z).
If γ0 and γ1 are homotopic closed paths in an open set U , then the above
theorem implies that the cycle Γ = γ1 − γ0 has index 0 about every point of the
complement of U . Hence, Γ is homologous to 0. The general Cauchy Theorem then
says that
0= f (z) dz = f (z) dz − f (z) dz
Γ γ1 γ0
for every function f which is analytic on U . This proves the homotopy version of
Cauchy’s Theorem.
Example 4.6.11. Prove that the circle of radius 2 centered at 0, with its standard
parameterization, is homotopic in C \ {0} to the piecewise linear curve which traces
once in the positive direction around the square with vertices at 1, i, −1, −i (see
Figure 4.6.1).
Solution: We parameterize the circle by γ0 (t) = 2 e2πit , t ∈ [0, 1] and the
square by
⎧
⎪
⎪ (1 − 4t) + 4ti, if 0 ≤ t ≤ 1/4;
⎪
⎨(4t − 1) + (2 − 4t)i, if 1/4 ≤ t ≤ 1/2;
(4.6.3) γ1 (t) =
⎪
⎪ (4t − 3) + (2 − 4t)i, if 1/2 ≤ t ≤ 3/4;
⎪
⎩
(4t − 3) + (4t − 4)i, if 3/4 ≤ t ≤ 1.
4.6. Homotopy 147
We define a homotopy h between the two by simply letting h(s, t) be the point
h(s, t) = (1 − s)γ0 (t) + sγ1 (t)
on the line segment joining γ0 (t) to γ1 (t). The resulting point h(s, t) is always on
or outside the boundary of the square and so it is never 0. For 0 ≤ t ≤ 1/4, the
formula for h is
h(s, t) = 2(1 − s) cos t + s(1 − 4t) + [2(1 − s) sin t + 4st]i.
The formula for h if t is in one of the other subintervals of the partition 0 < 1/4 <
1/2 < 3/4 < 1 is also easily calculated using (4.6.3).
Example 4.6.12. Show why the curves γ0 and γ1 of the previous exercise cannot
be homotopic in V = C \ {3/2}.
Solution: The point 3/2 is in the complement of V , and the curve γ0 has
index 1 about 3/2. However, the curve γ1 has index 0 about 3/2, since 3/2 is in
the unbounded component of C \ γ1 (I). By Theorem 4.6.9, the two curves cannot
be homotopic in V .
Simply Connected Sets.
Definition 4.6.13. A connected open set U is said to be simply connected if every
closed curve in U is homotopic to a point (that is, a constant curve).
Example 4.6.14. Prove that a convex open set U is simply connected.
Solution: Fix a point z0 ∈ U . If γ : I → U is a closed curve in U , we define a
homotopy between γ and the constant curve z0 as follows:
h(s, t) = (1 − s)γ(t) + sz0 .
Note that, for each t ∈ I, h(x, t) is on the line segment joining γ(t) to z0 and,
hence, is in U . Clearly, h is continuous on I × I, h(0, t) = γ(t), h(1, t) = z0 , and
h(s, 0) = h(s, 1), since γ(0) = γ(1). This establishes that h is a homotopy between
γ and the constant curve z0 . Since this can be done for every closed curve γ in U ,
the set U is simply connected.
We have talked about an open set in C possibly having a “hole”. What do we
mean by a hole in an open set? This is usually understood to mean a bounded
component of the closed set C \ U . However, we have not explored the ideas of
connectedness and connected components for closed sets so far in this text. A few
words concerning these concepts now will make the statement and proof of the next
theorem much easier to understand.
Recall that a set S is separated by a pair of open sets U , V if U ∩ S and V ∩ S
are both non-empty, have empty intersection, and have union equal to S. If there
is no pair of open sets which separate S, then S is connected. A component of S is
a maximal connected subset of S.
If the pair U , V separates S and S happens to be closed, then A = S \ (C \ V )
is closed and is equal to U ∩ S. Similarly, B = S \ (C \ U ) is closed and equal to
V ∩ S. Thus, a closed set S is separated by a pair of open sets U, V if and only
if it is separated by a pair of its closed subsets A, B in the sense that A, B is a
disjoint pair of closed, non-empty subsets of S with union equal to S. In this case,
A = U ∩ S and B = V ∩ S.
148 4. The General Cauchy Theorems
In the next theorem we will need the following lemma concerning components
and separation of a close subset of C. The proof is left to the exercises.
Lemma 4.6.15. Let S be a closed subset of C. If C is a component of S, then C
is bounded if and only if C is contained in a closed bounded subset A of S such that
B = S \ A is also closed.
Note that the set B of the above lemma could be empty. If it is empty, then S
itself is bounded. If it is not empty, then the pair A, B separates S.
The next theorem shows that simply connected open sets are connected open
sets with no holes and that they are exactly the sets on which all the things we
would like to be true in complex analysis are, in fact, true.
Theorem 4.6.16. Let U be a connected open subset of C. Then the following
statements are equivalent.
(a) U is simply connected;
(b) every cycle in U is homologous to 0;
(c) U has no holes, that is, C \ U has no bounded components;
(d) Γ f (z) dz = 0 for every cycle Γ in U and every analytic function f on U ;
(e) every analytic function on U has an antiderivative;
(f) every harmonic function on U has a harmonic conjugate;
(g) every non-vanishing analytic function f on U has an analytic logarithm;
(h) every non-vanishing analytic function f on U has an analytic square root.
Proof. We can only give an incomplete proof at this point. We will prove that each
of the statements (a) through (h) implies the next statement on the list. However,
the proof that (h) implies (a) will have to wait until we prove the Riemann Mapping
Theorem in Chapter 6.
A constant path γ(t) ≡ c in U has γ (t) ≡ 0, and so its index about any point
other than c is 0. In particular, it is homologous to 0 in U . Since each cycle is
equivalent to a sum of closed paths, it follows from Theorem 4.6.9 that (a) implies
(b).
We next prove that (b) implies (c). Suppose (c) fails, that is, suppose there is a
bounded component C of S = C \ U . Then, by the previous lemma, C is contained
in a closed bounded subset A of S such that B = S \ A is also closed.
The set A is compact. It is contained in some bounded open rectangle R as
well as in the open set C \ B. Thus, A is contained in the bounded √ open set
V = R \ (C \ B). We fix a point z0 ∈ A. Let δ > 0 be chosen so that 2 δ is smaller
than the distance from A to the complement of V . Suppose the open rectangle R
is defined by inequalities
a < Re(z) < b, c < Im(z) < d.
Let a = s0 < s1 < · · · < sn = b and c = t0 < t1 < · · · < tn = d be partitions of
[a, b] and [c, d] into subintervals of length at most δ. We are careful to choose these
so that Re(z0 ) is not one of the sj ’s and Im(z0 ) is not one of the tk ’s.
Now each of the rectangles
Rjk = {z : sj−1 ≤ Re(z) ≤ sj , tk−1 ≤ Im(z) ≤ tk }
4.6. Homotopy 149
U
A
Γ1
V
which meets A also lies in V . Let Γ be the cycle which is the sum of all the paths
∂Rjk for which Rjk ∩ A = ∅. Now if an edge of one of these rectangles meets A,
then this edge belongs to two of the rectangles whose boundaries make Γ. The edge
occurs with opposite orientation in the two paths and, hence, their contributions
will cancel out in any integral around Γ as well as in the expression for ∂Γ. This
means that if Γ1 is the chain which is left after all edges which meet A are thrown
out of Γ, then Γ1 is still a cycle, and if f is a function continuous on Γ(I), then
(4.6.4) f (z) dz = f (z) dz.
Γ Γ1
IndΓ1 (z0 ) = 1.
Since Γ1 is a cycle in U , we conclude that (b) fails. This completes the proof that
(b) implies (c).
To show that (c) implies (d), suppose γ is a closed path in U . If V is a bounded
component of C\γ(I), then A = V ∩(C\U ) is a closed set since it is the intersection
of C \ U with the complement of the union of the other components of C \ γ(I).
The set B = (C \ U ) \ A is also closed since it is equal to C \ (U ∪ V ). By Lemma
4.6.15, if A is non-empty, then it contains a bounded component of C \ U . Thus,
if (c) holds, then A must be empty – that is, V ∩ (C \ U ) = ∅ for every bounded
component V of C \ γ(I). Then z ∈ C \ U implies z is in the unbounded component
of C\γ(I) and so Indγ (z) = 0. The hypotheses of the Cauchy Theorem are satisfied
and so (d) holds. Thus, (c) implies (d).
150 4. The General Cauchy Theorems
If (d) holds, then we may define an antiderivative g for the analytic function f
on U by fixing a point z0 ∈ U and setting
g(z) = f (w) dw,
γz
As with a homotopy for closed curves, a homotopy of the above type determines
a continuous one-parameter family of curves {γs } by
γs (t) = h(s, t).
In this case, the curves in the family all begin at z0 and end at z1 . For this type
of homotopy, we have the following results, which follow easily from the preceding
material, and whose proofs are left as exercises.
Theorem 4.6.19. If γ0 and γ1 are homotopic paths in U connecting z0 to z1 , then
f (z) dz = f (z) dz
γ0 γ1
for every function f which is analytic in U .
Theorem 4.6.20. Let U be a connected open set and let z0 , z2 be points of U .
Then U is simply connected if and only if any two curves in U connecting z0 to z1
are homotopic in U .
1. Show that the path defined by (4.6.2) is, as claimed in the text, a linear curve,
with parameter interval [a, b], which traces the line segment from w to z.
2. Find ||γ − γ1 || if γ(t) = e2πit , and γ1 (t) = 2 cos(2πt) + 3i sin(2πt), for t ∈ [0, 1].
3. If γ(t) = e2πit for t ∈ [0, 1], describe a piecewise linear path γ1 such that
||γ − γ1 || < π/8.
4. Show that the curves γ and γ1 of Exercise 2 are homotopic in C\{0} by finding
an explicit homotopy in C \ {0} between them.
5. Explain why the curves γ and γ1 of the previous exercise cannot be homotopic
in the set C \ {2i}.
6. Which of the following open sets in C are simply connected:
(a) C \ {0}, (b) C \ [−∞, 0],
(c) D2 (0) \ [−1, 1], (d) D1 (0) \ [0, 1),
(e) C \ ({0} ∪ {1}), (f) D2 (0) \ D1 (0)?
7. For the set C \ {0}, show by example that (b), (c), and (d) of Theorem 4.6.16
all fail.
8. Also show by example that (e), (f), and (g) of Theorem 4.6.16 fail for the set
C \ {0}.
9. Prove Theorem 4.6.19.
10. Prove Theorem 4.6.20.
11. Prove that if U and V are homeomorphic open subsets of C, and if U is
simply connected, then so is V . Here, U and V are homeomorphic if there is a
one-to-one continuous function f of U onto V which has a continuous inverse
function.
12. Prove Lemma 4.6.15 – that is, prove that if S is a closed subset of C and C
is a component of S, then C is bounded if and only if it contains a bounded
closed subset A for which S \ A is also closed.
Chapter 5
Residue Theory
The Residue Theorem (Theorem 4.4.3) has a wide range of applications. This
chapter is devoted to exploring some of them. We begin with a section on techniques
for computing residues.
153
154 5. Residue Theory
The first couple of cases of the preceding theorem are worth highlighting. We
do this in the following corollary, which follows immediately from the theorem.
Corollary 5.1.2. Given a function g, analytic in a neighborhood of z0 ,
g(z)
(a) if f (z) = , then Res(f, z0 ) = g(z0 );
z − z0
g(z)
(b) if f (z) = , then Res(f, z0 ) = g (z0 ).
(z − z0 )2
z2 + 1
Example 5.1.3. Find Res(f, 1) if f (z) = .
(z − 1)(z 2 − 2z + 5)
Solution: The function f is of the form
g(z)
f (z) = ,
z−1
where
z2 + 1
g(z) = .
z 2 − 2z + 5
Since g is analytic in a neighborhood of 1, Corollary 5.1.2(a) implies that
1
Res(f, 1) = g(1) = .
2
sin z
Example 5.1.4. Find Res(f, 0) if f (z) = .
z2
Solution: This is clearly a situation where Corollary 5.1.2(b) applies, with
g(z) = sin z. Thus,
Res(f, 0) = g (0) = cos 0 = 1.
We could also have done this one directly, by writing out the power series expansion
for sin z about 0, and dividing by z 2 to obtain the Laurent series expansion of f .
5.1. Computing Residues 155
where the fraction on the right has the same denominator as the original fraction,
but the numerator has changed and is now a power series with lowest order term
of degree at least m + 1, whereas the original numerator had lowest order term of
degree at least m. If we write out explicitly the polynomials involved, then 5.1.4
becomes
p(z) am z m + am+1 z m+1 + am+2 z m+2 + · · ·
=
q(z) b0 + b1 z + b2 z 2 + · · ·
(5.1.5)
am m (am+1 − b1 am /b0 )z m+1 + (am+2 − b2 am /b0 )z m+2 + · · ·
= z + .
b0 b0 + b1 z + b2 z 2 + · · ·
This tells us that the first coefficient in (5.1.3) is
(5.1.6) cm = am /b0 .
The next coefficient cm+1 is obtained by repeating the procedure on the new fraction
that appears on the right in the above equation. This procedure may be repeated
as often as is needed to obtain a given coefficient cn . For example, in Exercises
5.1.16 and 5.1.17 you are asked to use this procedure to show that
am+1 am b1
(5.1.7) cm+1 = − 2
b0 b0
and
am+2 am+1 b1 am b21 am b2
(5.1.8) cm+2 = − + − 2 .
b0 b20 b30 b0
For small k, the expression for ck−1 , given by carrying out the method described
in the above theorem, is reasonably simple. However, the complexity increases
rapidly with increasing k.
When k = 1, the above theorem and (5.1.6), with m = 0, say that if
p(z) p(z)
f (z) = = ,
h(z) (z − z0 )q(z)
with q(z0 ) = 0, then
a0 p(z0 )
Res(f, z0 ) = c0 = = .
b0 q(z0 )
5.1. Computing Residues 157
This is just Corollary 5.1.2(a). We can restate it in terms of the original functions
p and h as follows: since h(z) = (z − z0 )q(z), we have
q(z0 ) = h (z0 )
and so
Corollary 5.1.6. Let p and h be analytic in a neighborhood of z0 , where h has a
zero of order 1 at z0 . Then
p(z0 )
Res(p/h, z0 ) = .
h (z0 )
Example 5.1.7. Find Res(f, 0) if
ez
f (z) = .
sin z
Solution: Since sin z has a zero of order 1 at z = 0, we may apply Corollary
5.1.6 with p(z) = ez and h(z) = sin z. Then
e0
Res(f, 0) = = 1.
cos 0
Example 5.1.8. Find Res(f, 0) if
1
f (z) = .
ez −1 − z
Solution: Here the denominator has a zero of order 2. We set p(z) = 1 and
z2 z3
h(z) = ez −1 − z = + + ··· .
2 3!
Then h(z) = z 2 q(z) where
1 z
q(z) = + + ··· .
2 3!
By Theorem 5.1.5, the residue we seek is the coefficient of the degree 1 term in the
power series expansion of 1/q, and this may be obtained by long division. In fact,
we have
1 z/3 + · · ·
=2− = 2 − (2/3)z + · · · ,
1/2 + z/3! + · · · 1/2 + z/3! + · · ·
and so the residue is −2/3.
cot z
Example 5.1.9. Find the residue at 0 of f (z) = .
z2
Solution: We proceed as in Theorem 5.1.5. The function f has a pole of order
3 at 0. We set p(z) = cos z, h(z) = z 2 sin z, and
h(z) sin z z2 z4
q(z) = 3
= =1− + − ··· .
z z 6 120
We then use long division to find the first three power series coefficients, c0 , c1 and
c2 of p/q. By Theorem 5.1.5, the residue we seek will then be the coefficient c2 . We
158 5. Residue Theory
have,
p(z) cos z 1 − z 2 /2 + · · ·
= −1 =
q(z) z sin z 1 − z 2 /6 + · · ·
(−1/2 + 1/6)z 2 + · · ·
=1+
1 − z 2 /6 + · · ·
−(1/3)z 2 + · · ·
=1+
1 − z 2 /6 + · · ·
= 1 − (1/3)z 2 + · · · .
The integral on the left is 2πi times the sum of the residues of g(z)/iz inside the
unit circle, provided g(z)/iz has no poles on the unit circle itself.
2π
dθ
Example 5.2.1. Find .
0 2 + sin θ
Solution: If we replace sin θ in the integrand by (z − z −1 )/2i, the resulting
function is
1 2iz
= .
2 + (z − z −1 )/2i 4iz + z 2 − 1
This is the g of equation (5.2.2). We divide this by iz and integrate around the
unit circle to get our answer. Thus, we need to evaluate the integral
2 2
2 + 4iz − 1
dz = √ √ dz.
|z|=1 z |z|=1 (z + (2 − 3)i)(z + (2 + 3)i)
Here we have factored z 2 + 4iz − 1 by finding its roots using the Quadratic Formula.
The only pole of the integrand that is inside the unit circle is the one at the point
160 5. Residue Theory
√
(−2 + 3)i. The residue at this pole is found using Corollary 5.1.2(a). That is, we
simply evaluate the function
2
√
z + (2 + 3)i
√
at the pole z = (−2 + 3)i. The resulting residue is
√
2 3
√ =− i.
2 3i 3
By the Residue Theorem, the integral we seek is this number times 2πi. Thus,
2π √
dθ 2 3
= π.
0 2 + sin θ 3
This technique may be used to evaluate any integral of the form (5.2.1), where
f (θ) is a rational function of sin θ and cos θ with a denominator which does not
vanish for θ ∈ [0, 2π].
can often be evaluated using residues. Here we are assuming that f is a function
which is Riemann integrable on each finite subinterval [a, b] of R.
There are two senses in which such an integral may converge. If the expression
x
f (t) dt
−y
exists, then the integral (5.2.3) is said to converge in the principal value sense and
the above limit is called the principal value of the integral. This is weaker than
ordinary convergence of the integral – that is, (5.2.3) may converge in the principal
value sense and not converge in the ordinary sense.
An improper integral (5.2.3) is said to converge absolutely if the integral of
|f | converges. This implies the convergence of (5.2.3) and, in fact, we have the
following integral analogue of the comparison test for convergence of series.
5.2. Evaluating Integrals Using Residues 161
Theorem 5.2.2. If f and g are continuous functions on the line, with g(t) ≥ 0
and |f (t)| ≤ g(t) for all t ∈ R, then the integral of f on R converges and
∞ ∞
f (t) dt ≤ g(t) dt.
−∞ −∞
provided the integral of g on R converges.
Then the hypothesis that |f (t)| ≤ g(t) implies that |ak | ≤ bk for all n. Furthermore,
for each positive integer n,
xn
n xn n
f (t) dt = ak and g(t) dt = bk .
0 k=1 0 k=1
Thus, the convergence of the ∞improper integral of g(t) implies the convergence of
the series of positive
terms k=1 bk . This, in turn, implies the absolute convergence
of the series ∞ a
k=1 k and the inequality
∞
∞
(5.2.4) ak ≤ bk .
k=1 k=1
x
This means that, if h(x) = 0 f (t) dt, then the sequence {h(xn )} has a limit for ev-
ery increasing sequence {xn } converging to infinity. By the remark in the first para-
graph, limx→∞ h(x) exists and is the common limit of all the sequences {h(xn )}.
Hence the improper integral of f exists and is, by definition, this number L.
It follows from (5.2.4) that
∞ ∞
∞
f (t) dt = |L| ≤ b = g(t) dt.
k
0 k=1 0
0
The same argument can be used to show that −∞ f (t) dt exists and satisfies
the inequality 0 0
f (t) dt ≤ g(t) dt.
−∞ −∞
The theorem follows on combining these results.
−r r
that the integral is independent of r, for r > R, and is equal to 2πi times the sum
of the residues of f at this finite set of singularities.
If f is analytic in an open set containing the lower half-plane, then all of the
above still holds with the following changes: the lower half-plane replaces the upper
half-plane, the path γr is replaced by its reflection through the x-axis, and, since
the new path has negative orientation, the resulting integral is −2πi times the sum
of the residues of f in the lower half-plane.
This proves the following theorem:
Theorem 5.2.3. Let H be either the closed upper or lower half-plane. If f is
analytic, except at a discrete set of singularities, in an open set containing H, has
no singularities on R, and there are positive numbers R, C, and p > 1 such that
(5.2.5) is satisfied for z ∈ H, then
∞
m
f (x) dx = σ(H)2πi Res(f, zj ),
−∞ j=1
√
2
The sum of these is − i and, when multiplied by 2πi, this yields
4
∞ √
x2 2
4
dx = π.
−∞ 1 + x 2
The only pole in the upper half-plane is the one at i, and this is a pole of order
2. Corollary 5.1.2(b) applies, and it tells us that the residue of f at this pole is
obtained by differentiating
1
g(z) = ,
(z + i)2
and evaluating at z = i. The result is
−2 i
g (i) = 3
=− .
(2i) 4
Thus, ∞
1 1 −i π
2 2
dx = · 2πi · = .
0 (1 + x ) 2 4 4
In the next example, we exploit the fact that the integrand is unchanged if we
rotate coordinates by an angle π/n.
∞
1
Example 5.2.6. Find dx.
0 1 + x2n
Solution: The integrand is even, so we could just proceed as in the preceding
example, replacing this integral by one over all of R and then applying Theorem
5.2.3. However, this would mean evaluating residues at each of the n poles in the
upper half-plane. There is a better way. We set
1
f (z) =
1 + z 2n
and note that
(5.2.7) f (eπi/n z) = f (z)
for each z ∈ C. This means f has the same values along the ray {eπi/n x : 0 ≤ x}
as it does along the ray [0, +∞). To exploit this, we choose r > 1 and use the path
γr which traverses the boundary of the sector indicated in Figure 5.2.2. That is,
5.2. Evaluating Integrals Using Residues 165
r e πi/n
e πi/2n
the path γr traverses the interval [0, r], then the arc r eit , 0 ≤ t ≤ π/n, and then
returns to 0 along the interval [r eπi/n , 0].
Then
r π/n r
f (z) dz = f (x) dx + f (r e )ri e dt −
it it
f (eπi/n x) eπi/n dx
γr 0 0 0
r π/n
= (1 − eπi/n ) f (x) dx + f (r eit )ri eit dt.
0 0
The second integral on the right clearly goes to 0 as r → ∞ and the integral on the
left is 2π Res(f, z1 ), where z1 = eπi/2n is the only pole of f inside γr . Thus,
∞
Res(f, z1 )
f (x) dx = 2πi .
0 1 − eπi/n
All that remains is to evaluate this residue.
Since z12n = −1, we have
z 2n + 1 = z 2n − z12n = (z − z1 )(z 2n−1 + z 2n−2 z1 + · · · + z12n−1 ).
Corollary 5.1.2(a) says that, to find the residue of f at z1 , we simply evaluate the
inverse of the second factor on the right at z = z1 . The result is
1 z1 eπi/2n
Res(f, z1 ) = 2n−1 = − 2n = − 2n ,
2nz1
and this implies
∞
1 2πi eπi/2n π/2n
dx = − = .
0 1 + x2n 2n 1 − eπi/n sin(π/2n)
2π
dθ
2. Find .
0 ∞ 10 + 6 sin θ
1
3. Find 2 + 2x + 2
dx.
x
−∞ ∞
x2
4. Find 2 2
dx.
0 ∞ (1 + x )
x
5. Find 2 + 2x + 2)2
dx
∞
−∞ (x
1
6. Find dx (see Example 5.2.6).
0∞ 1 + x6
1
7. Find dx using contour integration over the boundary of the sector
0 1 + x3
defined in polar coordinates by {r eiθ : 0 ≤ r ≤ R, 0 ≤ θ ≤ 2π/3} with R > 1
(see Example
∞ 5.2.6).
eix
8. Find dx.
1 + x2
−∞ ∞ ∞
cos(x) eix
9. Find 2 2
dx by first finding 2 2
dx and and then taking
−∞ (1 + x ) −∞ (1 + x )
the real
∞ part of the result.
cos x
10. Find dx.
0 1 + x2
11. Prove the fact, used in the proof of Theorem 5.2.2, that if h is a function on
(0, ∞) and limk→∞ h(xk ) exists for every increasing sequence {xk } of positive
numbers converging to ∞, then these limits are all the same number L and
limx→∞ h(x) = L.
12. Prove that an improper integral of the form (5.2.3) converges if f is a contin-
uous function on R which satisfies an inequality |f (x)| ≤ C|x|−p for |x| > R,
where C and R are positive constants, and p > 1.
13. Suppose f is a continuous function on (0, ∞) (which may have a singularity at
s
∞ Show how to use the substitution t = e to convert the improper integral
0).
0
f (t) dt into an improper integral on (−∞, ∞). Use this to prove a theorem
∞ to Theorem 5.2.2 for improper integrals on (0, ∞).
analogous
dt
14. Find . Hint: Use the substitution of the previous exercise.
0 t + t log2 t
ic
−a b
Proof. We give the proof in case t < 0. The case t > 0 is the same except that
the lower half-plane is used instead of the upper half-plane.
Since limz→∞ f (z) = 0, f has only finitely many singularities. For positive
numbers a, b, and c let γ be the path which traverses the boundary of the rectangle
with vertices −a, b, b + ic, −a + ic in the positive direction (see Figure 5.3.1).
If a, b, and c are large enough, all singularities of f in the upper half-plane will
be inside γ. In particular, there will be no singularities on γ(I). Then,
b c
−izt −itx
f (z) e dz = f (x) e dx + i f (b + iy) e−it(b+iy) dy
γ −a 0
(5.3.1) b c
− f (x + ic) e−it(x+ic) dx − i f (−a + iy) e−it(−a+iy) dy.
−a 0
168 5. Residue Theory
√
If t < 0, the previous theorem applies, and it tells us that fˆ(t) is 2π i times the
sum of the residues of e−itz f (z) in the upper half-plane. Now f has only one pole
in the upper half-plane and that is at i. The residue of e−itz f (z) at this point can
be computed using Corollary 5.1.2(a), since
1
f (z) = .
(z + i)(z − i)
The residue is obtained by evaluating e−itz /(z + i) at z = i. The result is
et
,
2i
and so
π t
fˆ(t) = e if t < 0.
2
−itz
√ For t > 0, we compute the residue of e f (z) at −i and then multiply by
− 2π i. The result is
π −t
fˆ(t) = e if t > 0.
2
At t = 0 we have ∞
1 1 π
fˆ(0) = √ 2
dx = ,
2π −∞ 1 + x 2
by the case n = 1 of Example 5.2.6 (or simply by noting that arctan is an anti-
derivative for the integrand).
Putting the cases t < 0, t > 0, t = 0 together, we conclude that
π −|t|
fˆ(t) = e for all t ∈ R.
2
∞
cos x
Example 5.3.4. Show that 2
dx = π/ e .
−∞ 1 + x
Solution: The integral we seek is the real part of the integral
∞ −ix
e
2
dx,
−∞ 1 + x
√ 1
which is 2π times the Fourier transform of evaluated at t = 1. By the
1 + x2
previous example, this is π/e.
1
Example 5.3.5. Find the Fourier transform of f (x) = .
x+i
Solution: We set
1
f (z) = .
z+i
This function is meromorphic, with no poles on the real axis, and it vanishes at
infinity. Thus, Theorem 5.3.2 applies.
The only pole of f occurs at z = −i. Since there are no poles in the upper
half-plane, Theorem 5.3.2 tells us that fˆ(t) = 0 if t < 0.
170 5. Residue Theory
√
For t > 0, fˆ(t) is − 2π i times the residue of f (z) e−itz at z = −i. This
residue is e−t and so
0, if t < 0;
fˆ(t) = √ −t
− 2π i e , if t > 0.
Note the jump discontinuity at t = 0.
What about the value of fˆ at t = 0? We have
∞
1
fˆ(0) = √ f (x) dx.
2π −∞
We can attempt to evaluate this using the fact that f (z) has log(z + i) as an
antiderivative, given any branch of the log function. If we use the principal branch
of log, then the line Im(z) = i is contained in its domain of definition, and so the
real line is contained in the domain on which log(z + i) is defined and has f (z) as
its derivative. We conclude that
b
f (x) dx = log(b + i) − log(−a + i)
−a
1
= (log(b2 + 1) − log(a2 + 1)) + i(arg(b + i) − arg(−a + i)).
2
The limit as (a, b) → (∞, ∞) of this expression does not exist, since its real part
approaches +∞ as b → ∞ with a fixed and approaches −∞ as a → ∞ with b
fixed. Thus, the improper integral defining fˆ(0) does not converge. However, the
symmetric integral
a
f (x) dx = i(arg(a + i) − arg(−a + i))
−a
has limit −πi as a → ∞. In this case, we say that the integral has principal value
−πi, even though it is not a convergent improper integral. If we define fˆ(0) using
this value for the integral, then we have
π
fˆ(0) = − i,
2
which is exactly half way between limt→0+ fˆ(t) and limt→0− fˆ(t).
The next example involves a different technique – no residues – but the Cauchy
theorems are still involved.
Example 5.3.6. Find the Fourier transform of the normal distribution function
1
h(x) = √ e−x /2 .
2
2π
Solution: We have
∞
1
e−(x
2
+2itx)/2
hˆ(t) = dx.
2π −∞
5.3. Fourier Transforms 171
it
−a b
= e dx.
2π −∞
We will use Cauchy’s Integral Theorem to show that the last integral above is
independent of t.
Fix a, b > 0 and consider the path γ that traces once in the positive direction
around the boundary of the rectangle with vertices at −a, b, b + ti, −a + ti (see
Figure 5.3.2).
The integral of e−z /2 over this path is 0 by Cauchy’s Theorem. Hence,
2
b t
e−x /2 dx + e−(b+iy) /2 i dy
2 2
0=
−a 0
(5.3.4) b t
e−(x+it) e−(−a+iy)
2 2
− /2
dx − /2
i dy.
−a 0
which means that the integral on the left is independent of t and can be evaluated
by evaluating the integral on √ the right. The integral on the right is a standard
calculus problem. Its value is 2π, and it is obtained by expressing the square of
the integral as a double integral over the plane and then evaluating this using polar
coordinates. We leave it as an exercise (Exercise 5.3.1).
In view of (5.3.3) and the above calculation, we have
e−t /2 √ e−t /2
2 2
hˆ(t) = 2π = √ = h(t).
2π 2π
172 5. Residue Theory
In other words, the normal distribution function is its own Fourier transform.
∞
1
(5.3.5) fˇ(x) = fˆ(−x) = √ eitx f (t) dt.
2π −∞
∞
1
(5.3.6) f (x) = fˆˇ(x) = √ eitx fˆ(t) dt.
2π −∞
There are several versions of this result, depending on the hypotheses that
are assumed satisfied by f and fˆ. They all have rather technical proofs that are
complicated by the fact that we are working with improper integrals. One fairly
standard version assumes that both f and fˆ are absolutely integrable. A proof of
this version can be found in [8]. We will not attempt to prove any of the versions of
this result here. Instead, we will just illustrate the formula in a couple of examples.
Example
5.3.7. By Example 5.3.3 the function f (x) = 1/(1 + x2 ) has fˆ(t) =
−|t|
π/2 e as Fourier transform. Compute the inverse Fourier transform of fˆ and
verify that the Fourier Inversion Formula holds in this case.
Solution: We have
∞
1
fˆˇ(x) = √ eixt π/2 e−|t| dt
2π −∞
0 ∞
1
= e(ix+1)t dt + e(ix−1)t dt
2 −∞ 0
1 1 1 1
= − = 2 = f (x).
2 ix + 1 ix − 1 x +1
Proof. Since f has a simple pole at 0, its Laurent series expansion in DR (0) has
the form
a−1
f (x) = + a0 + a1 z + · · · + an z n + · · · ,
z
5.3. Fourier Transforms 173
where a−1 = Res(f, 0). The integral of the first term of this expansion over γ is
θ2
ia−1 dθ = i(θ1 − θ2 ) Res(f, 0),
θ1
while the integral of the nth term for n ≥ 0 is
θ2
ian r n+1 ei(n+1)θ dθ,
θ1
which has absolute value less than or equal to |an |r n+1 (θ2 − θ1 ). Since this has
limit 0 as r → 0, the proof is complete.
Example 5.3.9. Given real numbers a < b, find the Fourier transform of the
function f which is equal to 1 for a ≤ x ≤ b and is equal to 0 otherwise. Verify
that the Fourier Inversion Formula holds in this case.
Solution: By the definition of f ,
b
1 i e−ibt − e−iat
fˆ(t) = √ e−itx dx = √ .
2π a 2π t
Then fˆˇ is given by the improper integral
∞ it(x−b)
i e − eit(x−a)
fˆˇ(x) = dt.
2π −∞ t
We may write this first as
−r it(x−b) ∞ it(x−b)
i e − eit(x−a) e − eit(x−a)
lim dt + dt ,
r→0 2π −∞ t r t
and then as
−r ∞
i eit(x−b) eit(x−b)
lim dt + dt ,
r→0 2π −∞ t r t
(5.3.7) −r ∞ it(x−a)
i eit(x−a) e
− lim dt + dt ,
r→0 2π −∞ t r t
provided these limits exist.
We evaluate the two limits in (5.3.7) using an indented contour. We use the
contour described in Figure 5.3.3 in the case where x − b > 0 in the first limit and
where x − a > 0 in the second limit. Since the integrands are analytic except at 0,
the integral around this contour is 0. As in the proof of Theorem 5.3.2, the integrals
along the vertical lines and the horizontal line at z = ic have limit 0 as L, R, and c
approach infinity. By the previous theorem, the integral around the semicircle has
limit iπ as r → 0. It follows that the first of the two limits in (5.3.7) is 1/2 if x > b,
while the second limit is 1/2 if x > a.
For the cases x < b and x < a we use the reflection of the contour in Figure
5.3.3 through the x-axis. This has the result of changing the sign of the integral
around the semicircle. Thus, the limits are −1/2 for the case x < b and the case
x < a.
Of course, when x = b, the first limit is limr→0 (ln r − ln r) = 0. The same result
holds for the second limit when x = a.
174 5. Residue Theory
ic
−L −r r R
We conclude that fˆˇ(x) = 1 if a < x < b and is 0 otherwise. Note that this is
equal to f (x) except at the points a and b where f is not continuous.
1. Prove that
∞ √
e−x
2
/2
dx = 2π
−∞
on the set of complex numbers λ for which this integral exists. If u is a real number
such that e−xu f (x) is an absolutely integrable function of x, then the function
e−xλ f (x) will also be integrable for all λ with Re(λ) ≥ u and the integral F (λ) will
be an analytic function of λ on the open half-plane {λ ∈ C : Re(λ) > u} with limit
0 as λ → 0 in this half-plane. In fact, since
∞
F (s + it) = e−ixt e−xs f (x) dx,
0
√
the Laplace transform of f at s + it is just 2π times the Fourier transform of
the function g(x) which is e−xs f (x) for x ≥ 0 and is 0 for x < 0. If the Fourier
Inversion Formula applies to g and gˆ, then it results in a formula for recovering f
from its Laplace transform F . Specifically, with λ = s + it,
s+i∞
exs ∞ 1
(5.4.2) f (x) = F (s + it) eixt dt = F (λ) exλ dλ.
2π −∞ 2πi s−i∞
We will call this the inverse Laplace transform of F .
On the open half-plane where F is defined and analytic, the integral in (5.4.2)
is independent of s and, if x < 0, it has limit 0 as s → ∞. If F can be extended
to a function which is analytic on C except at finitely many singularities, all to the
left of Re(λ) = u, then we may calculate the integral in (5.4.2) using residues. That
is, we compute
1
F (λ) eixλ dλ,
2πi γ
where γ traverses the boundary of a rectangle with interior containing the singu-
larities of F and with right edge along the vertical line Re(λ) = u. As in the proof
of Theorem 5.3.2, this integral is independent of the rectangle chosen and, on the
one hand, is equal to the integral on the right in (5.4.2) and, on the other hand, is
equal to the sum of the residues of F (λ) exλ . Thus, if f is continuous on (0, ∞) and
has a Laplace transform F which is analytic except at finitely many singularities,
then
f (x) = Res(F (λ) exλ , λi ),
where {λi } is the set of singularities of F .
λ
Example 5.4.1. Find the inverse Laplace transform of F (λ) = .
(λ − 2)(λ + 1)2
176 5. Residue Theory
Solution: The function F (λ) exλ has a simple pole at λ = 2 and a pole of
order 2 at λ = −1. By Theorem 5.1.2
2 e2x 2
Res(F (λ) exλ , 2) = = e2x ,
(2 + 1)2 9
while
d λ eλx 3x − 2 −x
Res(F (λ) e , −1) =
xλ
= e .
dλ λ − 2 λ=−1 9
Thus, f (x) = 19 [(3x − 2) e−x +2 e2x ] for x ≥ 0.
where g(z) = f (z)z t−1 on C \ [0, +∞) and z t−1 is defined as above.
Proof. The path γ for our path integral is the path which goes along the upper
edge of the cut from > 0 to R, then along the circle γR of radius R centered at
the origin, then along the lower edge of the cut from R to and finally around the
origin along the circle γ of radius (see Figure 5.4.1).
5.4. The Laplace and Mellin Transforms 177
γR
γε
It remains to show that the conditions on f (z)z t−1 ensure the integrals around
the two circles in (5.4.5) have limits 0 as R → ∞ and → 0. If M (R) is the
maximum value of |f (z)| on the circle of radius R, then
f (z)z t−1
dz ≤ 2πM (R)Rt .
γR
1
The calculation of the Mellin transform for the function is a key step in
1+x
the developement of the properties of the gamma function in Chapter 9.
1 π
Example 5.4.3. Show that the Mellin transform of f (x) = is .
1+x sin(πt)
Solution: According to the preceding theorem, the Mellin transform of this
function f will exist if 0 < t < 1, since it is for these values of t that
|z|t |z|t
lim = 0 = lim .
z→0 |1 + z| z→∞ |1 + z|
The only singularity of f (z) occurs when z = −1, and the corresponding residue,
Res(g, −1) for g(z) = f (z)z t−1 , is
(−1)t−1 = eπ(t−1)i = − eπti .
Thus, by Theorem 5.4.2, if 0 < t < 1, then
∞ t−1
x π
dx = .
0 1+x sin(πt)
1. Find the Laplace transforms of the functions f (x) = 1 and g(x) = x. On what
subset of C are these defined?
2. Find the Laplace transform of the function f of Example 5.4.1. Does it coincide
with the function F of this example?
λ
3. Find the inverse Laplace transform of F (x) = 2 .
(λ + 4)
1
4. Find the inverse Laplace transform of F (x) = .
1 + λ2
5. If f is continuous on [0, ∞), differentiable on (0, ∞), f (0) = a, and the Laplace
transform F of f exists for Re(λ) > u, what is the Laplace transform of f in
terms of F ?
6. If f satisfies the differential equation f + af + bf = g on (0, ∞) and f and
f are continuous on [0, ∞), describe the Laplace transform F of f in terms of
the Laplace transform G of g and the initial values (f (0) and f (0)) of f and
f . Hint: Use the result of the previous exercise.
7. Can a non-zero constant function be the Laplace transform of a function f on
[0, ∞)?
5.5. Summing Infinite Series 179
1
8. Find the Mellin transform of f (x) = .
1 + x2
1
9. Find the Mellin transform of f (x) = .
∞ 2/3 1 + x3
x
10. Find dx.
0 1 +x
11. Show
that if h(y) = f (ey ), then the Mellin transform of f can be expressed as
∞
h(y) ety dy. Note the similarity to the inverse Fourier transform.
−∞
12. Verify the second equality in (5.4.5).
Some Properties of the Cotangent. Since sin(πz) has a zero of order one at
every integer and no other zeroes on C, the meromorphic function
cos(πz)
cot(πz) =
sin(πz)
has a simple pole at each integer and no other poles. The residue of this function
is 1/π at each n (Exercise 5.1.5). Thus,
Lemma 5.5.1. The function π cot(πz) has a simple pole with residue 1 at each
integer, and no other poles.
The function π cot(πz) is obviously not bounded near ∞, since it has a pole
at each integer. However, it turns out that there is a sequence of closed paths,
converging to ∞, on which this function is bounded. The paths in question are the
180 5. Residue Theory
(N+1/2)(−1+i) (N+1/2)(1+i)
(N+1/2)(−1−i) (N+1/2)(1−i)
paths γN , where N is a positive integer and γN traces once in the positive direction
around the square with vertices at
(N + 1/2)(1 + i), (N + 1/2)(−1 + i), (N + 1/2)(−1 − i), and (N + 1/2)(1 − i)
(see Figure 5.5.1).
Lemma 5.5.2. There is a positive number R such that | cot(πz)| ≤ 2 on γN (I) for
each N ≥ R.
π cot(πz)
Proof. The function h(z) = is analytic inside γN except at integer
z
points, where it has poles. By the Residue Theorem
N
π cot(πz)
(5.5.3) = Res(h, n).
γN z
n=−N
Proof. The singularities of πf (z) cot(πz) occur at the integers and at the points
z1 , · · · , zn . At an integer n not in E, this function has residue f (n), since cot(πz)
has a simple pole with residue 1 at n and f has no singularity at n. If N is large
enough so that the zj are all inside γN , then the Residue Theorem implies
m
1
(5.5.5) πf (z) cot(πz) dz = f (n) + Res(πf (z) cot(πz), zj ).
2πi γN N j=1
n∈Z0
Thus, to prove the theorem, we simply need to show that the integral on the left
converges to 0 as N → ∞.
Choose R and M large enough that (5.5.4) holds, the inequality in Lemma
5.5.2 holds, and |zj | ≤ R for each j. Then f is analytic in the annulus
A = {z : R < |z| < ∞}
and vanishes at infinity. It therefore has a Laurent expansion, in this annulus,
involving only negative powers of z. Say,
c−1 c−2 c−n
f (z) = + 2 + ··· + n + ··· .
z z z
By Lemma 5.5.3,
π cot(πz)
dz = 0.
γN z
182 5. Residue Theory
and so
1 1 c−1
(5.5.6) πf (z) cot(πz) dz = π f (z) − cot(πz) dz.
2πi γN 2πi γN z
However, the function
c−1 c−2 c−n
f (z) − = 2 + ···+ n + ···
z z z
q(z)
has the form in A, where q is analytic in A and has limit c−2 at infinity.
z2
This implies that, for some R1 > R, the function |q| is bounded by some positive
constant M1 on {z : |z| ≥ R1 }. Then, by Lemma 5.5.2, the integrand of the integral
on the right in (5.5.6) is bounded above by 2M M1 /|z|2 . Since |z| > N on γN and
the length of the path γN is 8N + 4, we have that
1 2M M1 (8N + 4)
πf (z) cot(πz) dz ≤ ,
2πi N2
γN
cot z
where Z0 is the set of non-zero integers. The residue of at 0 is −1/3 by
z2
Example 5.1.9. Since the coefficient of z in the Laurent expansion of π cot(πz) is
π 2 times the coefficient of z in the Laurent expansion of cot z, we conclude that
π cot(πz)
has residue π 2 /3 at 0. Then,
z2
∞
1 1 1 π2
= = .
n=1
n2 2 n2 6
n∈Z0
∞
1
4. Find .
n=1
n4
N
1
5. Prove that if w ∈ C \ Z, then π cot πw = lim .
N →∞ n+w
n=−N
6. Assuming the result of the previous exercise, prove that if w ∈ C \ Z, then
∞
1 1
π cot πw = + .
w n=1 w2 − n2
π
7. Prove that the function has a pole at each integer n with residue (−1)n
sin(πz)
and no other poles.
∞
8. Derive a method for summing a series of the form (−1)n f (n), where f is a
−∞
meromorphic function with a finite number of poles. Hint: Use the preceding
exercise and the method of Theorem 5.5.4.
Chapter 6
Conformal Mappings
If U and V are open subsets of the complex plane, then just how different are U
and V from the standpoint of the study of analytic functions? If there is a analytic
map h : U → V with an analytic inverse h−1 : V → U , then the answer is, “they
are not different at all.” In this case, we shall say that h is a conformal equivalence
between U and V , and that U and V are conformally equivalent.
If h : U → V is a conformal equivalence, then each analytic function g on
V gives rise to an analytic function g ◦ h on U , and each analytic function f on
U gives rise to an analytic function f ◦ h−1 on V . This establishes a one-to-
one correspondence between the analytic functions on U and those on V . This
correspondence preserves sums and products and exponentials and logarithms and
roots. The same formulas define one-to-one correspondences between meromorphic
functions on U and on V , and between functions with isolated singularities on U and
on V . Similarly, paths, closed paths, and homotopies in U all have corresponding
objects in V . Indices of paths are preserved by this correspondence.
Clearly then, it is important to know when two open sets U and V are con-
formally equivalent. That is the subject of this chapter. The main theorem of the
chapter is the Riemann Mapping Theorem, which says that every connected, simply
connected, proper open subset of C is conformally equivalent to the open unit disc.
185
186 6. Conformal Mappings
be the matrix which rotates every vector through an angle θ. Then A−1 · dh(z0 ) is
an angle preserving matrix which sends (1, 0) to a positive multiple (r, 0) of itself.
Since it preserves angles between vectors, it must also send every other vector to a
multiple of itself. We claim that, in fact, it multiplies each vector by r. Suppose
(0, 1) is sent to (0, t). Then the vector (1, 1) is sent to (r, t) and, if this is to be a
multiple of (1, 1), we must have r = t. This implies that A−1 · dh(z0 ) is r times the
identity matrix. Hence, dh(z0 ) has the form (6.1.2). This implies that h satisfies
the Cauchy-Riemann equations at z0 and, hence, has a complex derivative at z0 .
This derivative is non-zero because dh(z0 ) is non-singular.
Note that a conformal map h from U onto V need not be one-to-one. In fact,
the exponential function ez is a conformal map of C onto C which is not one-to-one.
However, the following is true, based on the Inverse Mapping Theorem (Theorem
4.5.5).
z 1+z
i
1−z
z
√ i 1+z
1−z
z 2 1+z
π log 1−z
Example 6.1.6. Find a conformal equivalence from the unit disc to the first quad-
rant of the plane.
Solution: The upper half-plane consists of all points which, in polar coor-
dinates, have the form z = r eiθ with r > 0 and 0 < θ < π, while the open
first quadrant consists of points of this form with 0 < θ < π/2. It follows that
√
the principal branch of the square root function (which sends r eiθ to r eiθ/2 for
−π < θ < π) is a conformal equivalence from the upper half-plane to the first quad-
rant with inverse function w → w2 . It follows from this and the previous example
that
1+z
h(z) = i
1−z
is a conformal equivalence from the unit disc to the first quadrant, provided we use
the principal branch of the square root function.
Example 6.1.7. Find a conformal equivalence from the unit disc onto the strip
S = {z : −1 < Im(z) < 1}.
Solution: The principal branch of the log function is a conformal equivalence
of the open right half-plane onto the strip {z : −π/2 < Im(z) < π/2}. Thus, we
190 6. Conformal Mappings
z sin(z)
−1 1
−π/2 π/2
can construct a conformal equivalence from the unit disc to S by composing the
following series of conformal maps: First, we map the unit disc to the right half-
1+z
plane using z → , as in Example 6.1.4; we follow this by the map log from
1−z
the right half-plane to the strip {z : −π/2 < Im(z) < π/2}; and finally, we divide
by π/2. The composition of these maps is
2 1+z
h(z) = log ,
π 1−z
which is the required conformal equivalence.
Example 6.1.8. Find a conformal equivalence from the set
A = {z : −π/2 < Re(z) < π/2, Im(z) > 0}
to the upper half-plane.
Solution: Recall from Exercise 1.3.14 that the function sin z can be written
as
(6.1.3) sin(x + iy) = sin x cosh y + i cos x sinh y,
while its derivative cos z can be written as
(6.1.4) cos(x + iy) = cos x cosh y − i sin x sinh y.
The first of these formulas can be used to prove that sin is one-to-one on the set
A (Exercise 6.1.9), while the second can be used to prove that its derivative, cos z,
is non-vanishing on A (Exercise 6.1.10). Therefore, sin is a one-to-one conformal
map of A onto some open set B.
To identify B we note that sin takes the interval [−π/2, π/2] on the real line
to the interval [−1, 1]. It takes the half-lines {z : Re(z) = ±π/2, Im(z) > 0} to the
half-lines [1, ∞) and [−∞, −1]. Hence, it must take A to either the upper or lower
half-plane. Since it takes i to i sinh 1 and sinh 1 > 0, it must take A to the upper
half-plane. Hence, sin is the required conformal equivalence.
6.2. The Riemann Sphere 191
Analytic Functions on the Riemann Sphere. In what follows, the phrase an-
alytic at z0 , applied to a function on a subset of S 2 , will mean defined and analytic
in a neighborhood of z0 .
We know what it means for a function to be analytic at a finite point of S 2 .
What does it mean for a function f (z) to be analytic at ∞? We make sense of this
by, again, using the coordinate function w = 1/z, which is 0 at z = ∞.
Definition 6.2.1. We will say that a complex-valued function f (z) is analytic
at ∞ if f (1/w) is analytic at w = 0 (meaning defined and analytic in a deleted
neighborhood of 0 with a removable singularity at 0).
1−z
Example 6.2.2. Show that f (z) = is analytic at ∞ if we set f (∞) = −1.
1+z
Solution: According to the above definition, we must show that f (1/w) is
analytic at w = 0. However,
1 − 1/w w−1
f (1/w) = =
1 + 1/w w+1
and this is, indeed, analytic at w = 0.
The function 1/z that has played a major role in the discussion to this point,
can not only be thought of as a new coordinate function on S 2 , it can also be
thought of as a mapping of S 2 to itself which interchanges 0 and ∞. We will call
this mapping s. Thus,
s(z) = 1/z,
for all z ∈ S 2 (including 0 and ∞, where, of course, 1/0 = ∞ and 1/∞ = 0).
According to our definitions, a function f is analytic at infinity if and only if
f ◦ s is analytic at 0, while f is analytic at a point z0 where it has the value ∞ if
and only if s ◦ f is analytic at z0 .
Theorem 6.2.4. The mapping s is a conformal equivalence of S 2 onto itself, that
is, it is a one-to-one analytic map of S 2 onto itself with an analytic inverse map.
Proof. Clearly s is one-to-one and onto and is its own inverse function, since
s ◦ s(z) = z. It is also analytic at every point, with the possible exceptions of 0
and ∞. However, s(1/w) = w is analytic at w = 0 and so s is analytic at z = ∞.
Similarly, 1/s(z) = z is analytic at z = 0 and so s is analytic at 0. This completes
the proof.
194 6. Conformal Mappings
map
w
ψ : P 1 (C) \ {q} → C, where ψ(z, w) = .
z
Both φ and ψ are one-to-one and onto. The inverse map for φ sends the complex
number z to the equivalence class containing (z, 1), while the inverse map for ψ
sends w to the equivalence class containing (1, w). Then, on C\{0}, the composition
ψ ◦ φ−1 is defined and it is, in fact,
1
ψ ◦ φ−1 (z) = .
z
This means that P 1 (C) may also be thought of as two copies of C glued together
along C \ {0} by identifying the point z in one copy of C \ {0} with the point 1/z in
the other copy. Definitions of continuity and complex differentiability for functions
on P 1 (C) may then be defined at points on each copy of C. These definitions will
agree on the overlap because 1/z is an analytic function of z on C \ {0}. The
space P 1 (C), with topology and notion of analytic function as defined above, is
one-dimensional complex projective space.
It is apparent that our description of P 1 (C) is just another description of the
Riemann Sphere S 2 .
Proof. We show that each linear fractional transformation h has an inverse func-
tion which is also a linear fractional transformation. In fact, if
az + b
h(z) = ,
cz + d
and we solve the equation w = h(z) for z, the result is
−dw + b
z= = g(w).
cw − a
This suggests that g is an inverse function for h. In fact, if we calculate h ◦ g, then
for all w ∈ S 2 except w = ∞ and w = −a/c we get
(ab − cd)w
h ◦ g(w) = .
ab − cd
Since ab − cd = 0, we conclude that h ◦ g(w) = w. The cases w = ∞ and w = −a/c
can be checked separately (Exercise 6.3.1), and so h ◦ g(w) = w for all w ∈ S 2 . A
similar calculation shows that g ◦ h(z) = z for all z ∈ S 2 . Hence, h has an inverse
function h−1 = g : S 2 → S 2 , and it is also a linear fractional transformation.
Since h has an analytic inverse function, defined on all of S 2 , it is a conformal
automorphism of S 2 . Thus, each linear fractional transformation is a conformal
automorphism of S 2 .
To prove the converse, we let f be a conformal automorphism of S 2 and proceed
to show that it has the form claimed in the theorem. We first consider the case
where f (∞) = ∞. By Theorem 6.2.4, s(z) = 1/z is a conformal automorphism of
S 2 . Thus, the composition
1
q(z) = s ◦ f ◦ s(z) =
f (1/z)
is also a conformal automorphism of S 2 , and it takes 0 to 0. Thus, q is a one-to-one
analytic function with a zero at 0. This zero must be of order one since a conformal
map has non-vanishing derivative at each point. This implies that f (1/z) has a
pole of order 1 at z = 0 and is analytic and finite at every other point of S 2 . Then
f (z) has a pole of order 1 at ∞ and is analytic and finite everywhere on C. It
follows that z −1 (f (z) − f (0)) is analytic and finite everywhere on S 2 . Hence, it is
constant by Liouville’s Theorem. If this constant is a and b = f (0), then f is the
affine transformation f (z) = az + b. Such a transformation is of the form (6.3.1),
with c = 1 and d = 0, and, hence, is a linear fractional transformation.
We next consider the case where f (∞) = k = ∞. If we let
1
p(z) = ,
z−k
6.3. Linear Fractional Transformations 199
The preceding result can be very useful in making a quick determination of the
image of a set under a linear fractional transformation.
Example 6.3.3. Find the image of the unit disc under the transformation
2z
h(z) = .
z−i
200 6. Conformal Mappings
Proof. We first show that there is a linear fractional tranformation that takes a
given ordered triple of complex numbers {w1 , w2 , w3 } to {0, 1, ∞}. In fact,
(w2 − w3 )(z − w1 )
h(z) =
(w2 − w1 )(z − w3 )
does the job.
Now suppose we have a triple {w1 , w2 , w3 } of distinct points of S 2 , where one
of the points is ∞. Then we can still find an h taking this triple to {0, 1, ∞}. For
example, if w3 = ∞, then
z − w1
h(z) =
w2 − w1
will suffice.
Now to get a linear fractional transformation h which takes {w1 , w2 , w3 } to
{z1 , z2 , z3 }, we choose h1 taking {w1 , w2 , w3 } to {0, 1, ∞} and h2 taking {z1 , z2 , z3 }
to {0, 1, ∞} and then set h = h−1 2 ◦ h1 .
It remains to prove that h is unique. Suppose g is another linear fractional
transformation that takes {w1 , w2 , w3 } to {z1 , z2 , z3 }. With h = h−1 2 ◦ h1 , as in the
previous paragraph, we have that f = h2 ◦g◦h−1 1 is a linear fractional transformation
that takes {0, 1, ∞} to {0, 1, ∞}. If
az + b
f (z) = ,
cz + d
then f (0) = 0 implies that b = 0, f (∞) = ∞ implies that c = 0, and f (1) = 1
implies that a/d = 1. We conclude that f is the identity transformation – that is,
f (z) = z. This, in turn, implies that
g = h−1 −1
2 ◦ f ◦ h1 = h2 ◦ h1 = h.
maps w to 0, and 0 to −w. It also maps the unit circle to itself since, if |u| = 1,
then u−1 = u, and so
u − w u − w u − w
= =
1 − wu u−1 − w u − w = 1.
Since hw maps the unit circle to itself and a point, w, inside the unit circle to
another point, 0, inside the unit circle, it must map the unit disc D onto itself (it
has to map it to some open set in S 2 which has the unit circle as boundary and
there are only two choices: D and the complement of its closure). It follows that
hw is a conformal automorphism of D.
Suppose that h is some other conformal automorphism of D that takes w to 0.
Then h ◦ h−1w is a conformal automorphism of D that takes 0 to 0. By Theorem
3.5.6 there is a complex number u of modulus one such that h ◦ h−1w (z) = uz. On
replacing z by hw (z), this implies that h(z) = uhw (z). Thus, we have proved the
following theorem.
Theorem 6.3.5. The conformal automorphisms of the unit disc are the linear
fractional transformation of the form
z−w
h(z) = u = uhw (z),
1 − wz
where |u| = 1 and |w| < 1.
1+z
6. The linear fractional transformation h(z) = takes the open unit disc D
1−z
to the right half-plane. Show that h sends the set consisting of those points in
D that lie outside another circle which intersects the unit circle at right angles
and passes through the point z = 1 to a set of the form
{z : Re(z) > 0, Im(z) < a} or {z : Re(z) > 0, Im(z) > −a}
for some positive a. What is a?
7. A certain linear fractional transformation h takes the ordered triple of points
{a1 , b1 , c1 } on circle number one to the ordered triple {a2 , b2 , c2 } on circle
number two. Just by looking at these points, how can you tell whether h maps
the inside of circle one to the inside or to the outside of circle two?
8. Prove Theorem 6.3.6.
9. Prove that, for each w ∈ D, the mapping hw of (6.3.3) has h−w as its inverse
transformation.
10. Find a conformal automorphism of D which takes 1/2 to 0 and has derivative
3i/4 at z = 0.
11. Find a conformal automorphism of D which takes 1/2 to i/3 and has positive
derivative at 1/2.
12. If w ∈ D, prove that among all analytic functions that map D into D, map w
to 0, and have a positive derivative at w, the one with the largest derivative
is the function hw of (6.3.3). Hint: Use Schwarz’s Lemma (Lemma 3.5.5).
13. Each non-singular 2 × 2 complex matrix
a b
A=
c d
determines a linear fractional transformation
az + b
φA (z) = .
cz + d
Show that this correspondence A → φA is a group homomorphism – that is,
show that it takes the product AB of two matrices A and B to the composition
φA ◦ φB of the corresponding linear fractional transformations.
14. What is the kernel of the map A → φA of the preceding exercise – that is, for
which matrices A is φA the identity transformation z → z?
The last step in this program – showing that there is an h ∈ F with derivative of
maximal modulus at z0 – involves showing that any sequence in F has a subsequence
which converges to a function which is also in F. This means we must prove that
F is a normal family. We take up this task first.
Normal Families.
Definition 6.4.1. Let U be an open subset of C. A collection F of analytic func-
tions defined on U is called a normal family if each sequence in F either converges
uniformly to infinity on each compact subset of U or has a subsequence which
converges uniformly on each compact subset of U to an analytic function.
Normal families containing sequences which converge to infinity will not occur
in this chapter, but they will play a role in the next chapter in the proof of the Big
Picard Theorem.
Here we encounter another one of the amazing properties of analytic functions:
It takes very little for a family of analytic functions to be a normal family. In fact,
it is enough that the family be uniformly bounded, where F is said to be uniformly
bounded on U if there is an M such that |f (z)| ≤ M for every z ∈ U and every
f ∈ F.
Theorem 6.4.2 (Montel’s Theorem). Given an open set U ⊂ C, each uniformly
bounded sequence of analytic functions on U has subsequence which converges uni-
formly on each compact subset of U . Hence, each uniformly bounded family of
functions on U is a normal family.
Let gn = fnn . We will show that {gn } converges uniformly on every compact
subset of U – not just pointwise at each rational point zj .
If w ∈ U , we choose r > 0 such that the closed disc D2r (w) is contained in U .
Then for each z ∈ Dr (w) we have
Dr (z) ⊂ D2r (w) ⊂ U.
By Cauchy’s estimates, each f ∈ F satisfies
M
|f (z)| ≤ .
r
Then, for any two points z, z ∈ Dr (w) we have
z
M
(6.4.2) |f (z) − f (z )| = f (λ) dλ ≤ |z − z |.
z r
Note that the bound on the right in this estimate is independent of the function
f ∈ F. In particular, it holds for each function in the sequence {gn }.
r
Given > 0, we choose δ = . If z ∈ Dr (w), there are points with rational
3M
coefficients arbitrarily close to z. Let zj be one such point with |z − zj | < δ. By
(6.4.2),
M r
|gn (z) − gn (zj )| < = .
r 3M 3
We next choose N such that
|gn (zj ) − gm (zj )| < whenever n, m ≥ N.
3
We can do this because {gn (zj )} is a convergent, hence Cauchy, sequence for each
j. Then
|gn (z) − gm (z)| ≤ |gn (z) − gn (zj )| + |gn (zj ) − gm (zj )| + |gm (zj ) − gm (z)| < ,
if n, m ≥ N . This proves that gn is uniformly Cauchy on the disc Dr (w) and,
hence, converges uniformly on this disc.
If K is a compact subset of U , then the discs of the form Dr (w), where w ∈ K
and r is chosen so that D2r (w) ⊂ U , form an open cover of K. Since K is compact,
it is covered by finitely many of these discs. Since {gn } converges uniformly on
each of these discs, it also converges uniformly on K.
w
f
0
z0
Since g is analytic, its image g(U ) is open by the Open Mapping Theorem
(Theorem 4.5.8). It follows that we can find a closed disc Dr (w0 ) ⊂ g(U ) with
0 < r < ∞. Since g 2 = f and f is a one-to-one function, the image of g cannot
contain a non-zero point and its negative, since then these two points would be
taken to the same point by g 2 . Thus no point of the reflection through the origin,
Dr (−w0 ), of Dr (w0 ) is contained in g(U ). It follows that
|g(z) + w0 | > r
for all z ∈ U , and this means that the function
r
p(z) =
g(z) + w0
is a one-to-one conformal map of U into the unit disc. If p(z0 ) = w, then we can
compose p with the conformal automorphism hw of the unit disc (6.3.3) to obtain
a one-to-one conformal map hw ◦ p of U into D which takes z0 to 0. Thus, F is not
empty.
Lemma 6.4.4. Let U , z0 , and F be as above. If f ∈ F and f does not map U onto
D, then there is a g ∈ F with
|g (z0 )| > |f (z0 )|.
Proof. Let w ∈ D be a point which is not in the image of f (see Figure 6.4.1).
Recall from the previous section that the map hw of (6.3.3) is a conformal automor-
phism of the unit disc which sends w to 0. Thus, hw ◦ f (z) = 0 for all z ∈ U . As
in the proof of the previous lemma, this means that hw ◦ f has an analytic square
root q. If q(z0 ) = λ, then λ2 = w, q 2 = hw ◦ f , and
hw (0) (1 − |w|2 ) (1 − |λ|4 )
q (z0 ) = f (z0 ) = f (z0 ) = f (z0 ).
2q(z0 ) 2λ 2λ
The function q is a one-to-one conformal map of U into the unit disc, but q(z0 ) is
not 0 – it is λ. To get an element of F we compose q with hλ . The resulting map
g = hλ ◦ q belongs to F and satisfies
(1 − |λ|4 ) (1 + |λ|2 )
(6.4.3) g (z0 ) = hλ (λ)q (z0 ) = f (z0 ) = f (z0 ).
2λ(1 − |λ|2 ) 2λ
206 6. Conformal Mappings
Since 0 < (1 − |λ|)2 = 1 − 2|λ| + |λ|2 , we have 2|λ| < 1 + |λ|2 . Since a conformal
map has non-vanishing derivative, |f (z0 )| > 0. It then follows from (6.4.3) that
|g (z0 )| > |f (z0 )|.
Theorem 6.4.5 (Riemann Mapping Theorem). Let U be a proper, connected open
subset of C with the property that every non-vanishing analytic function on U has
an analytic square root. Then there is a conformal equivalence of U onto the unit
disc D.
1. Is the set of all conformal automorphisms of the unit disc a normal family?
If so, then every sequence of such maps has a convergent subsequence. Is the
limit necessarily also a conformal automorphism of the unit disc?
6.5. The Poisson Integral 207
If U is simply connected and has a simple closed curve as its boundary, then
the Dirichlet problem always has a solution. One proof that this is so, uses the
Riemann Mapping Theorem. This proof proceeds as follows: We exhibit an explicit
solution to the Dirichlet problem for the unit disc D and any continuous function
208 6. Conformal Mappings
Assuming this theorem, we will proceed with the proof that the Dirichlet prob-
lem always has a solution for a set U as above. In this section we deal with the
first step – the case where U is the unit disc D. The next section is devoted to
showing how to use the Riemann Mapping Theorem to attack the problem on more
complicated sets.
The Poisson Kernel. Recall the linear fractional transformation from Example
6.1.4:
1+z
(6.5.1) .
1−z
This is a one-to-one conformal map of the unit disc onto the right half-plane.
The Poisson kernel is constructed from the real part of this function. This was
calculated in Example 6.1.4 by multiplying numerator and denominator in (6.5.1)
by the conjugate of the denominator. The result is
1+z 1 − |z|2
(6.5.2) Re = .
1−z |1 − z|2
If z = r eiθ is the polar form of z, then
|1 − z|2 = |1 − r eiθ |2 = 1 − 2r cos θ + r 2
and so the expression on the right in (6.5.2) becomes
1 − r2
(6.5.3) Pr (θ) = .
1 − 2r cos θ + r 2
This may also be written as
∞
Pr (θ) = r |n| einθ
−∞
by Exercise 6.5.1.
The Poisson kernel is the function
1 − r2
Pr (θ − t) =
1 − 2r cos(θ − t) + r 2
(6.5.4) it
1 + e−it z e +z
= Re = Re .
1 − e−it z eit −z
It produces a harmonic function on D when it is integrated with respect to t against
a continuous function on the boundary of D.
6.5. The Poisson Integral 209
The integral defining the harmonic function u in the above theorem is called
the Poisson integral.
Boundary Values. We will show that the function u, defined in the above
theorem, has the property that
lim u(z) = g(w)
z→w
for each w on the unit circle. This implies that we get a continuous extension of
u to the closed unit disc D if we define u(w) to be g(w) at each w ∈ ∂D. Before
proving this, we need to establish some properties of the function Pr (θ).
Lemma 6.5.4. The Poisson kernel Pr (θ − t) of (6.5.4) satisfies
π
1
Pr (θ − t) dt = 1
2π −π
for every r < 1 and every θ.
Proof. We have
π π
1 1 eit +z
(6.5.6) Pr (θ − t) dt = Re dt .
2π −π 2π −π eit −z
We can easily evaluate this using residue theory. In fact,
π it π
1 e +z 1 1 + e−it z it
dt = i e dt
2π −π e −z
it 2πi −π eit −z
1 1 + z/w
= dw.
2πi |w|=1 w − z
The last expression can be evaluated by applying the Residue Theorem to the
function
1 + z/w w+z
= .
w−z w(w − z)
This has residue −1 at w = 0 and residue 2 at w = z. Hence,
π it
1 e +z
dθ = 2 − 1 = 1.
2π −π eit −z
Thus, its real part is also 1. In view of (6.5.6), the proof is complete.
210 6. Conformal Mappings
Proof. The idea of the proof is this: When r is near 1, Pr (θ − t) is nearly 0 for all
t ∈ [−π, π] except near θ, where the function has a large spike. This means that
only values of g(eit ) for t near θ contribute significantly to the integral (6.5.7). The
proof below makes this claim precise.
The function Pr (θ) is a positive, even function of θ. Since its denominator is
increasing on the interval [0, π], this function is decreasing as θ moves away from 0
in either direction. It follows that, if η is a fixed number between 0 and π, then
(6.5.8) Pr (θ − t) ≤ Pr (η) when η ≤ |θ − t| ≤ π.
Putting (6.5.10) together with (6.5.12) and the analogous estimate for the in-
tegral on [−π, θ − η], we conclude that
π
1
iθ
g(e )Pr (θ) dθ − g(e ) <
iθ
2π
−π
for 1 − δ < r < 1 and for every θ ∈ [−π, π]. This proves the Theorem.
The next theorem tells us that the Poisson integral yields a solution to the
Dirichlet problem for the unit disc and an arbitrary continuous boundary function
g.
Theorem 6.5.6. If g is a real-valued continuous function on the unit circle and u
is defined on the closed unit disc D by
2π
1
(6.5.13) iθ
u(r e ) = g(eit )Pr (θ − t) dt
2π 0
if r < 1 and by u(eiθ ) = g(eiθ ) on the unit circle, then u is continuous on D and
harmonic on D.
Proof. In view of Theorem 6.5.3, the only thing left to prove is that
(6.5.14) lim u(z) = g(eiθ0 ) for all θ0 .
z→eiθ0
The previous lemma almost says this, but not quite. The limit there is taken only
along radial lines, and this is not exactly what is meant by (6.5.14). We need to do
a bit more work.
Because the convergence in Lemma 6.5.5 is uniform in θ, given > 0 we may
choose a δ1 > 0 such that
|u(r eiθ ) − g(eiθ )| ≤
2
for all θ and all r with 1 − δ1 < r < 1. We may also choose δ2 such that
|g(eiθ ) − g(eiθ0 )| <
2
whenever |eiθ − eiθ0 | < δ2 . If δ = min{δ1 , δ2 }, then |r eiθ − eiθ0 | < δ implies both
|r eiθ − eiθ | < δ1 and | eiθ − eiθ0 | < δ2 , from which it follows that
|u(r eiθ ) − g(eiθ0 )| ≤ |u(r eiθ ) − g(eiθ )| + |g(eiθ ) − g(eiθ0 )| < .
This proves (6.5.14).
212 6. Conformal Mappings
2. Prove the uniqueness theorem (Theorem 6.5.1) for solutions of the Dirichlet
problem.
3. By inspection, find a solution to the Dirichlet problem on D if the boundary
function g is g(eit ) = cos t. Use your answer to evaluate
2π
1 (1 − r 2 ) cos t
dt.
2π 0 1 − 2r cos(θ − t) + r 2
4. Show that the Poisson kernel Pr (θ − t) is a harmonic function of z = r eiθ on
the open unit disc.
5. Equation (6.5.5) expresses the harmonic function u, given by the Poisson in-
tegral, as the real part of an analytic function on D. Use this equation to find
an integral formula for a harmonic conjugate to u.
6.6. The Dirichlet Problem 213
6. Show that if g is a positive, continuous function on the unit circle, then the
analytic function f defined by equation (6.5.5) has positive real part and its
imaginary part vanishes at z = 0.
7. Let f be analytic on the open unit disc D. If the real part of f is positive and
extends to be continuous on D, then prove that f is a function of the form
given in the previous exercise plus an imaginary constant.
8. Supply the details of the proof that (6.5.15) gives a solution to the Dirichlet
problem for DR (z0 ) with boundary function g.
9. Prove Theorem 6.5.7.
10. Show that if u is a real-valued, non-negative harmonic function on an open set
containing DR (z0 ), then
R−r R+r
u(z0 ) ≤ u(z) ≤ u(z0 )
R+r R−r
for 0 ≤ r = |z − z0 | < R. This is Harnack’s Inequality. Hint: Use Theorem
6.5.7.
11. If u is harmonic and non-negative on D and u(0) = 1, give upper and lower
bounds for u(1/2). Hint: Use the result of the previous exercise.
12. Prove that if un is a sequence of positive harmonic functions on a connected
open set U , and if un (z0 ) → 0 at one point z0 ∈ u, then un (z) → 0 at every
point z of U . Hint: Use the result of Exercise 10.
2
2
h
0
0
Example 6.6.4. Find a function on the closed unit disc which is continuous except
at ±1, is harmonic on the open unit disc, and is 0 on the lower half of the unit
circle and 2 on the upper half of the unit circle.
Solution: In Example 6.1.7 we showed that the function
2 1+z
h(z) = log
π 1−z
is a conformal equivalence from the unit disc to the strip {z : −1 < Im(z) < 1}.
This map takes the upper semicircle to the line Im(z) = 1 and the lower semicircle
to the line Im(z) = −1. It follows that a solution to the problem may be obtained
by composing the solution to the previous problem with h (see Figure 6.6.1). This
yields
2 1+z
u(z) = 1 + Im(h(z)) = 1 + arg ,
π 1−z
so that
2 2y
u(x, y) = 1 + tan−1
π 1 − x2 − y 2
is our solution. Here, for points on the unit circle x2 + y 2 = 1, we interpret this
expression to mean its limit as the point (x, y) is approached from within the unit
disc. Note that the resulting function u is, indeed, continuous on the closed disc,
except at the points ±1 where the boundary function fails to be continuous.
√
Example 6.6.5. Let g be the function on R which is 1 − x2 for −1 < x < 1 and
is 0 for all other values of x. Find a harmonic function on the upper half-plane
which has g as its boundary function and which has limit 0 as z → ∞ in the upper
half-plane.
Solution: We use the results of Example 6.1.8. There we showed that sin z is
a conformal equivalence of the strip A = {z : −π/2 < Re(z) < π/2, Im(z) > 0}
onto the upper half-plane. If we compose g with sin z, we get the function q on ∂A
which is cos x on the real line between −π/2 and π/2 and is 0 on the vertical lines
Re(z) = ±π/2. The real part e−y cos x of eiz is a harmonic function on A which
has q as boundary function and has limit 0 as z → ∞ in A. Hence, we obtain a
solution u to our problem by setting
−1
f (z) = ei sin (z)
and u(z) = Re(f (z))
for z in the upper half-plane. We can simplify this.
If w = sin−1 (z), then sin w = z and cos2 w = 1 − z 2 . Since, for every z in the
open upper half-plane, 1 − z 2 is in the domain of the principal branch of the square
6.6. The Dirichlet Problem 217
i z z+1/z
-1 1 -2 0 2
√ √
root function (denoted · ), we have cos w = ± 1 − z 2 . Since cos w has positive
real part if w ∈ A (see (6.1.4)), and the principal branch of the √
square root function
takes values in the right half-plane, we conclude that cos w = 1 − z 2 . Then
eiw = cos w + i sin w = 1 − z 2 + iz,
and so
f (z) = 1 − z 2 + iz.
Then our solution u is
u(z) = Re 1 − z 2 + iz
(6.6.2)
1
= −y + 1 − x2 + y 2 + (1 − x2 + y 2 )2 + 4x2 y 2 .
2
This last equality is left as an exercise (Exercise 6.6.13).
Example 6.6.6. Let B be the open upper half-plane with the upper half of the
closed unit disc removed. That is,
B = {z : |z| > 1, Im(z) > 0}.
Solve the Neumann problem for B: Find a function u on B which is harmonic on
B and has normal derivative 0 at every boundary point of B.
Solution: The same problem on the upper half-plane has an easy solution: the
function Re(z) = x is harmonic on the upper half-plane and its gradient (1, 0) is
parallel to the real axis at each point. Thus, its normal derivative on the real axis
is 0. To get a solution to our original problem, we need only compose this solution
with a conformal equivalence of B to the upper half-plane. In fact
1
z→z+
z
is such a map (Exercise 6.6.9). Thus, our solution is
1
u(z) = Re z +
z
x
=x+ 2 .
x + y2
Any constant multiple of this solution is also a solution. If this solution represents
the potential function for a fluid flow problem, then the corresponding velocity
218 6. Conformal Mappings
1. Solve the Dirichlet problem for the upper half-plane and the boundary function
which is 1 on the positive real axis and 0 on the negative real axis. Hint: Think
about the function arg(z) on the upper half-plane.
2. Use the result of the preceding exercise to find a different way of solving the
problem in Example 6.6.4 .
3. Solve the Dirichlet problem on the first quadrant with boundary function g
which is 1 on the real axis and 0 on the imaginary axis.
4. Solve the Dirichlet problem on the half-disc {z : |z| < 1, Im(z) > 0} with
boundary function which is 0 on the line [−1, 1] and 1 on the semicircle where
|z| = 1. Hint: Use the result of Exercise 6.1.5.
5. Solve the Dirichlet problem for the open set consisting of the points inside the
unit circle and outside a circle C which intersects the unit circle at right angles
at the points 1 and i. The boundary function is the function which is 1 on the
part of the boundary which lies on the unit circle and is 0 on the part which
lies on C. Hint: see Exercise 6.3.6.
1 1
6. Show that , for x ∈ R, is the real part of . Use this to solve the
1 + x2 1 − ix
1
Dirichlet problem on the upper half-plane with boundary function on
1 + x2
the real line and limit 0 as z → ∞.
7. Modify the approach used in the preceding problem to find a solution to the
1
Dirichlet problem on the lower half-plane for the boundary function .
1 + x2
8. For each w ∈ C show that the equation w = z + 1/z has two solutions for z,
one is the inverse of the other, and they are distinct if and only if w = ±2.
Show that the solution with |z| ≥ 1 has imaginary part with the same sign as
Im(w), while if |z| < 1, then z and w have imaginary parts with opposite sign.
9. Show that the transformation z → z + 1/z is a conformal equivalence of the
exterior of the unit circle onto the complex plane with the interval [−2, 2]
removed. Also show that the restriction of this transformation to
is a conformal equivalence from this set to the upper half-plane. Hint: Use the
result of the previous exercise.
10. Find a solution to the Neumann problem for the set consisting of the exterior
of the unit disc. Hint: Use the result of Exercise 9.
11. Let U and V be open sets in the plane and let h : U → V be analytic. Prove
that u ◦ h is harmonic on U for every harmonic function u on V .
12. Suppose U is a bounded open set and g is a continous function on ∂U . Prove
that if u is a solution to the Dirichlet problem on U with boundary function
6.6. The Dirichlet Problem 219
converges on the open unit disc D to a function analytic on D. In fact, this function
is
1
f (z) = ,
1−z
on D, and it can obviously be extended to a function, given by the same algebraic
expression, which is analytic on the much larger set C \ {1}.
According to the Inverse Function Theorem, a function f which is analytic on
an open set U and which has a non-zero derivative at a point z0 ∈ U has an analytic
inverse function defined in some neighborhood W of f (z0 ). Although the theorem
assures us of an inverse function just in some possibly very small neighborhood
of f (z0 ), we would naturally like to know on just how large a set can the inverse
function be defined.
There are many problems of this nature, problems where a solution in the form
of an analytic function is known or is known to exist locally, in a neighborhood of
a point, and the question arises as to just how much we can enlarge the domain
on which this function is defined, is analytic, and is still a solution to the original
problem. Questions of this type are questions of analytic continuation.
One method of analytic continuation, using the Schwarz Reflection Principle,
will at the end of the chapter lead to the existence of a special analytic function (an
221
222 7. Analytic Continuation and the Picard Theorems
analytic covering map) from the unit disc onto C \ {0, 1}. Using this function, one
can turn an entire function with two distinct points of the plane not in its range
into an entire function with range contained in the unit disc. Such a function must
be constant. This leads directly to the proofs of the Big and Little Picard theorems.
Schwarz Reflection. Figure 7.1.1 illustrates the situation described in the next
theorem. In what follows, if V is a subset of C, then V − will denote {z : z ∈ V }.
Theorem 7.1.1 (Schwarz Reflection Principle). Let U be an open set which is
symmetric about the x-axis R. Let A = U ∩ R and let V be the part of U in the
open upper half-plane. If f is a function which is continuous on V ∪ A, analytic on
V , and real-valued on A, then f can be continued to a function analytic on all of
U by setting its continuation equal to f (z) on V − .
Now A is an open subset of the real line and, hence, is a union of open line
segments. Let L be one of these line segments. Then the set W = V ∪ L ∪ V − is
an open subset of U and f is continuous on W and analytic on W except possibly
at points of L. However, it is a simple consequence of Morera’s Theorem that a
function which is continuous on an open set W and analytic on W except possibly
at the points of a line segment in W is actually analytic on all of W (Exercise
3.2.14). Since f is analytic on each of the open sets W = V ∪ L ∪ V − for L one of
the line segments making up A, it is anaytic on the union of these sets, and U is
this union. This completes the proof.
Note that (7.1.1) implies that ρ is its own inverse function on U . If κ(z) = z,
then κ ◦ ρ and ρ ◦ κ are both analytic functions and are inverse functions of one
another. This implies that κ◦ρ is a one-to-one conformal map. Thus, each reflection
defined on U is the conjugate of a conformal equivalence from U to U − .
Theorem 7.1.6. If C is any simple analytic curve in C, then there is a reflection
ρ through C defined on some domain V containing C. This reflection is unique
in the sense that another reflection through C, defined on a neighborhood V1 of C,
must be equal to ρ on the connected component of V ∩ V1 which contains C.
Proof. The proof is the same as the proof of the Schwarz Reflection Principle. We
set
g(z) = f (ρ(z))
Example 7.1.8. What is the reflection through an arc on the unit circle and what
does the previous theorem say about continuing an analytic function across such
an arc?
Solution: The map
1 1 iθ
ρ(z) = or ρ(r eiθ ) = e
z r
is defined and conjugate analytic on C \ {0}. It satisfies ρ ◦ ρ(z) = z. It also fixes
each point on the unit circle since, for z on the unit circle, |z|2 = zz = 1, which
implies z = 1/z. It follows that, for any domain U which meets the unit circle in
an arc C and is taken to itself by ρ, the unique reflection through C, defined on U ,
is the map ρ.
For such a U and C, the previous theorem implies that any function analytic
on the part V of U lying on one side of C, continuous on V ∪ C, and real-valued
on C can be analytically continued to an analytic function on all of U .
226 7. Analytic Continuation and the Picard Theorems
r
z
.
0 D
U U∪D
Although the function element (fn , Dn ) of the above definition, seems to de-
pend on the choice of partition 0 = t0 < t1 < · · · < tn+1 = 1 of [0, 1] and sequence
(f1 , D1 ), (f2 , D2 ), · · · , (fn , Dn ) of function elements, it turns out that, up to equiv-
alence, it is actually independent of these choices.
Theorem 7.2.2. Given a curve beginning at z0 and ending at w and an analytic
function element (f0 , D0 ) at z0 , any two analytic continuations of (f0 , D0 ) along γ
are equivalent as analytic function elements at w.
The next theorem is almost obvious, but there is some work to be done to show
the required chain of overlapping discs can be chosen as in Definition 7.2.1. We
leave the details to the exercises (Exercise 7.2.1).
Theorem 7.2.3. Suppose (f0 , D0 ) is a function element at z0 and γ is a curve
joining z0 to w. If there is an open set U , containing D0 and γ(I), and an analytic
function f on U such that f = f0 on D0 , then (f0 , D0 ) can be analytically continued
along γ.
The converse of the above theorem is not true. It may seem that it should be
true, since the union of the overlapping discs in Definition 7.2.1 is an open set U
containing γ(I), and it appears that the functions fj fit together to define a single
function, on this union, that agrees with f0 on D0 . However, this is not the case,
due to the fact that the curve γ may cross itself, as in Figure 7.2.2, and the crossing
point may be contained in two of the discs Dj and Dk . The corresponding functions
fj and fk may not be equivalent as function elements at this point. In fact, the
curve might be a closed curve, so that z0 = γ(0) and w = γ(1) are the same point.
In this case the analytic continuation of (f0 , D0 ) along γ leads to another function
element (fn , Dn ) at z0 . This function element need not be equivalent to (f0 , D0 ).
Example 7.2.4. Give an example of the phenomenon referred to in the previous
paragraph.
Solution: Let log denote the principal branch of the log function and consider
the analytic function element (log, D1 (1)). If γ is the unit circle traversed once
in the counterclockwise direction beginning at γ(0) = γ(1) = 1, then (log, D1 (1))
can be analytically continued along γ (Exercise 7.2.2), but the resulting function
element on the final disc will differ from that on the first disc by 2πi.
There is no domain containing the unit circle to which (log, D1 (1)) can be
analytically continued. If there were such a domain U and a continuation f of
(log, D1 (1)) to U , then f would have to agree with the principal branch of the log
function on all of U \ [−∞, 0], but this function has a 2πi jump discontinuity across
the half-line [−∞, 0] and certainly cannot be analytically continued across it.
(f0 , D0 ) can be analytically continued along γ1 and along γ2 , then are the resulting
function elements at w equivalent? Not necessarily, as the following example shows:
Example 7.2.5. Can analytic continuations of a function element at z0 along
different paths from z0 to w lead to non-equivalent function elements at w?
Solution: Let the initial function element be (log, D1 (1)), where log is the
principal branch of the log function, z0 = 1, and w = −1. Clearly this can be
continued along the curve γ1 consisting of the upper half of the unit circle traversed
from 1 to −1 and along the curve γ2 consisting of the lower half of the unit circle
traversed from 1 to −1 . However, the resulting function elements at −1 = γ1 (1) =
γ2 (1) differ by 2πi.
The Monodromy Theorem. The next theorem gives conditions under which
we can be sure that the analytic continuations of a function along two curves from
z0 to w are necessarily equivalent function elements, in contrast to the preceding
example.
Theorem 7.2.6 (Monodromy Theorem). Let U be a connected open set in C, z0
and w points of U , and (f0 , D0 ) an analytic function element at z0 , with D0 ⊂ U .
Suppose
(a) (f0 , D0 ) can be analytically continued along every curve in U ; and
(b) γ0 and γ1 are homotopic curves in U joining z0 to w.
Then the continuations of f along γ0 and γ1 are equivalent function elements at w.
Example 7.2.7. Consider the function element φ consisting of the principal branch
of the log function and the disc D1 (1). Set
Show that the analytic continuations along any two curves from 1 to −1 in A have
equivalent terminal function elements at −1.
Solution: The branch of the log function defined by restricting arg(z) to lie
in the inteval (−π/2, 3π/2) agrees with the principle branch of the log function
on D1 (1). By Theorem 7.2.3, φ can be analytically continued along any curve in
A. The set A is simply connected and so any two curves in A from 1 to −1 are
homotopic. In view of the Monodromy Theorem, continuations of φ along two such
curves will have equivalent terminal function elements at −1.
The Monodromy Theorem allows us, in the next theorem, to conclude that if
a function element can be analytically continued along every curve in an open set
U and if the open set is simply connected, then there is an analytic continuation of
f to all of U .
g h
f
U W
There is another class of maps h for which a kind of lifting result holds. This
is the class of analytic covering maps.
Definition 7.3.1. An analytic map h : V → W is called an analytic covering map
if, for each w0 ∈ W , there is a neighborhood A of w0 , contained in W , such that
h−1 (A) is the disjoint union of a collection {Bj } of open subsets of V with the
property that, for each j, h is a conformal equivalence of Bj onto A.
Example 7.3.2. Show that z → ez is an analytic covering map from C to C \ {0}.
Solution: The function exp, defined by exp(z) = ez , is certainly an analytic
map of C onto C \ {0}. We claim that each disc D in C \ {0} has inverse image
exp−1 (D) consisting of a disjoint union of open sets on each of which exp is a
conformal equivalence onto D.
Given a disc D ⊂ C \ {0}, there is a ray from 0 to ∞ disjoint from D (e.g.,
take the ray from 0 to ∞ which passes through the center of D and rotate it by the
angle π). This means that there is an angle θ0 such that every z for which ez ∈ D
satisfies
θ0 + 2n π < Im(z) < θ0 + 2(n + 1)π
for some integer n. Thus, exp−1 (D) is a disjoint union of open sets
Bn = exp−1 (D) ∩ {z : θ0 + 2n π < Im(z) < θ0 + 2(n + 1)π}.
On each Bn , exp is a conformal equivalence from Bn to D, with inverse function
equal to the branch of the log function for which arg takes values between θ0 + 2n π
and θ0 +2(n+1)π (see Figure 7.3.2). Thus, by definition, exp is an analytic covering
map.
(θ0+2π)i
B0
D
θ0 i exp
B−1 θ0
(θ0−2π)i
B−2
(θ0−3π)i
B−3
2πi
πi exp D
−πi
in the strip π < Im(z) < 2π. The map exp is a conformal equivalence of the first of
these onto D, but only maps the second one onto the lower half of D (see Figure
7.3.3). Since this problem persists no matter how small a disc we choose centered
at 1, the map exp is not an analytic covering map from V to C \ {0}.
This function serves to lift f through h, but only on the set D0 . The main work of
the proof is to show that the analytic function element (g0 , D0 ) can be analytically
continued along any curve in U beginning at z0 . If we can do this, then the Mon-
odromy Theorem implies that (g0 , D0 ) can be analytically continued to an analytic
7.3. Analytic Covering Maps 235
The preceding theorem is, on the one hand, a powerful application of the Mon-
odromy Theorem and, on the other hand, is the key ingredient in our proofs of the
Picard theorems in the next section.
236 7. Analytic Continuation and the Picard Theorems
C2
C1
by reflection through a circle (Example 7.1.8 – the Schwarz Reflection Principle for
circles). We begin by proving a few simple facts about reflection through a circle.
Reflection Through a Circle. Suppose C1 and C2 are two circles in the plane
that intersect one another in two points. If the tangents to the two circles are
perpendicular at each of these two points, then we will say that the circles meet at
right angles (Figure 7.4.1). Actually, it turns out that if the tangents are perpen-
dicular at one of the points of intersection, then they are also perpendicular at the
other point of intersection (Exercise 7.4.1).
Theorem 7.4.1. If circles C1 and C2 meet at right angles, and A is the arc of C1
that lies inside C2 , then reflection through A maps C2 onto itself.
Theorem 7.4.2. With C1 , C2 as above, the reflection through C1 takes any other
circle C, which meets C2 at right angles, to another circle which meets C2 at right
angles or to a line through the center of C2 .
238 7. Analytic Continuation and the Picard Theorems
√ √
An Analytic Covering Map. The three points −1, 1/2 + i 3/2, 1/2 − i 3/2
are equidistant points on the unit circle C. According to Exercise 7.42, we can join
each pair of these points with an arc of a circle that meets C at right angles. These
three arcs then bound an open curvilinear triangle V0 (see Figure 7.4.2).
The open set V0 is clearly simply connected. By the Riemann Mapping Theo-
rem, there is a conformal equivalence h : V0 → H, where H is the upper half-plane.
By Theorem 6.5.2, the map h extends to a continous map of the closure of V0 to
the closure of H in S 2 , which takes the boundary of V0 to the boundary of H in S 2 .
By composing with a linear fractional transformation, if necessary,
√ we √may choose
this map in such a way that it takes the points −1, 1/2 − i 3/2, 1/2 + i 3/2 to the
points ∞, 0, 1.
The function h is real-valued on each of the three “edges” of V0 . By the circle
version of the Schwarz Reflection Principle, h can be continued by reflection across
each of these edges. This results in h being defined and analytic in the set V1
described in Figure 7.4.3. Note that each of the three new cells on which h is
defined in this way is contained in the unit disc D (Theorem 7.4.1) and is bounded
by three circles that meet the unit circle C in right angles (Theorem 7.4.2).
7.4. The Picard Theorems 239
Note also that, while h maps V0 to the upper half-plane, it maps each of the
new cells of V1 to the lower half-plane, since each is a reflection of V0 through one
of its edges.
We can now analytically continue h to a still larger domain V2 by reflecting
across each of the circular arcs that bound V1 . This results in h being defined in
the set V2 represented in Figure 7.4.3. Note that h now maps each of the new cells
in V2 into the upper half-plane. We can clearly continue this process by induction
to create an increasing sequence {Vn } of open sets to which h may be analytically
continued. The union of these open sets is D (Exercise 7.4.5), and so the result is
an analytic function h which maps the open unit disc D onto a subset of the plane
which contains the upper and lower open half-planes and the intervals (−∞, 0),
(0, 1), and (1, ∞) on the real line. The points ∞, 0, and 1 are not in the image
because every triangular cell that occurs in the above construction has all of its
vertices on the unit circle. Thus, h is an analytic map of D onto C \ {0, 1}.
Theorem 7.4.3. The analytic map h : D → C \ {0, 1}, described above, is an
analytic covering map.
Proof. Think of the disc D as being partitioned into light cells and dark cells,
separated by arcs of circles, as in Figure 7.4.3. The dark cells are mapped into the
upper half-plane by h and the light cells are mapped into the lower half-plane. Each
arc separating two of these cells is mapped to one of the intervals I1 = (−∞, 0),
I2 = (0, 1), I3 = (1, ∞).
For j = 1, 2, 3, let Pj be the union of the open upper and lower half-planes and
the open interval Ij . Then each Pj is an open set consisting of S 2 with a closed arc
removed.
Let Δ be any open disc in C \ {0, 1}. Then Δ lies entirely inside Pj for at
least one j. We fix a j for which this is true. The inverse image of Pj under h
consists of the union of all the light and dark open cells together with some of the
arcs separating them – those arcs which h maps to Ij . The collection of these arcs
is pairwise disjoint, and each one of them separates exactly one light cell from one
dark cell. The union of the arc and the two cells it separates is an open set on
which h is a conformal equivalence onto Pj . No two distinct open sets of this form
can overlap since, if they did, the overlap would have to be a cell with boundary
containing two arcs mapping to Ij under h. Let {Uk }∞ k=1 be the collection of open
240 7. Analytic Continuation and the Picard Theorems
sets of this form. This is a pairwise disjoint collection of open sets with union equal
to h−1 (Pj ). Also,
h−1 (Δ) ⊂ Uk ,
k
and so h−1 (Δ) is the disjoint union of the open sets Bk = h−1 (Δ) ∩ Uk . The map h
is a conformal equivalence of Bk onto Δ for each k. It follows that h : D → C\{0, 1}
is an analytic covering map.
Proof. If f does not take on the values z0 and z1 , then the function
f (z) − z0
z1 − z0
does not take on the values 0 and 1. Thus, we may as well assume that f itself has
this property. Then f is an analytic function from the simply connected set C to
C \ {0, 1}. Since the map h : D → C \ {0, 1} of the previous theorem is an analytic
covering map, Theorem 7.3.4 implies that f may be lifted through h to an analytic
function g : C → D such that
f = h ◦ g.
However, this means that g is a bounded entire function and, hence, is a constant,
by Liouville’s Theorem. This, of course, forces f to also be a constant.
The Big Picard Theorem. The proof of the big Picard Theorem depends on
the following theorem. Recall the definition of a normal family (Definition 6.4.1).
Theorem 7.4.5. Let U be a connected subset of the plane. Then the set of analytic
functions on U with values in the set C \ {0, 1} is a normal family.
Proof. We will prove the theorem in the case where U is an open disc Δ. The
proof that the theorem for general U follows from this special case will be left to
the exercises.
Let Δ be any open disc and let z0 be its center. Let F denote the set of analytic
functions on U with values in C \ {0, 1}. Let {fn } be a sequence in F. We will show
that this sequence converges uniformly to ∞ on compact subsets of Δ or it has
a subsequence which converges uniformly on compact subsets of Δ to a function
analytic on Δ.
Suppose {fn } does not converge uniformly to ∞ on compact subsets of Δ.
Then there is a disc Δ1 with the same center and with Δ1 ⊂ Δ and an R > 0 such
that fn (Δ1 ) has elements with modulus less than R for infinitely many n. This
means that we can choose a subsequence of {fn } which itself has no subsequence
converging uniformly to ∞ on Δ1 . Without loss of generality, we may replace
{fn } by this subsequence. That is, we may assume that {fn } has no subsequence
converging uniformly to ∞ on Δ1 .
Let h : D → C \ {0, 1} be the analytic covering map of Theorem 7.4.3. Then,
since Δ is simply connected, Theorem 7.3.4 implies each fn can be lifted through
7.4. The Picard Theorems 241
Proof. We may as well assume that z0 = 0. If f does not have the property stated
in the conclusion of the theorem, then there is a disc Dr (0) on which f fails to take
on at least two complex values. We may assume these values are 0 and 1, since
otherwise we may compose f with a linear fractional transformation which takes
the two values missed by f to 0 and 1.
Thus, we may assume f is an analytic function from Dr (0) \ {0} to C \ {0, 1}
with an essential singularity at 0. We define a sequence of functions with the same
properties by setting
fn (z) = f (z/n) for n = 1, 2, · · · .
By the previous theorem, this sequence converges uniformly on compact subsets of
Dr (0) to ∞ or it has a subsequence converging uniformly on compact subsets of
Dr (0) to an analytic function. In the first case, 1/f is bounded and, hence, has
a removable singularity at 0. This means it extends to an analytic function with
a zero of some finite order at 0. This is impossible, since it implies that f has
a pole at 0 rather than an essential singularity. The second case implies that f
itself is bounded and, hence, has a removable singularity at 0. This also violates
the hypothesis that f has an essential singularity at 0. Thus, our assumption that
f misses two values in some disc centered at 0 has led to a contradiction. This
completes the proof.
242 7. Analytic Continuation and the Picard Theorems
1. Prove that if two circles intersect in two points and their tangents are perpen-
dicular at one point of intersection, then the tangents are also perpendicular
at the other point of intersection.
2. Prove that, given two distinct points on a circle C1 , there is a unique circle C2
which meets C1 at right angles at these two points.
3. Prove that if Cr and CR are circles of radius r and R, respectively, which are
tangent at a point, then the reflection of CR through Cr is a circle of radius
rR
2R + r
if neither circle is inside the other and is
rR
2R − r
if Cr is inside CR .
4. The set Vn in the construction preceding Theorem 7.4.3 has boundary consist-
ing of a set of circular arcs of different sizes with endpoints on the boundary
of the unit disc. Prove by induction that the largest of these circular arcs has
radius (2n + 1)−1 R0 , where R0 is the radius of each of the three circular arcs
comprising the boundary of V0 . Hint: Use the calculation of the preceding
exercise.
5. Prove that the union of the open sets Vn , described in the paragraph preceding
Theorem 7.4.3, is the unit disc D. Hint: Use the result of the preceding
exercise.
6. Prove that if f and 1/f are both entire functions, then the image of f is exactly
C \ {0}.
7. Prove that if f is a non-constant entire function and b2 = 4ac, then the function
g(z) = af 2 (z) + bf (z) + c
must have a zero.
8. Is there a non-constant analytic function from the plane with one point re-
moved to the plane with two points removed? Justify your answer.
9. If U is a simply connected open set, show that the set of all analytic functions
on U with values in C \ {0, 1} is a normal family (see Section 6.4).
10. Suppose r < 1, R < 1 and h is a conformal equivalence from the annulus
AR = {z : R < |z| < 1} to the annulus Ar = {z : r < |z| < 1}. Suppose
also that h extends to a continuous map from AR to Ar which takes the unit
circle to itself and the disc of radius R to the disc of radius r. Prove that
h can be continued to a conformal equivalence of the unit disc onto the unit
disc that takes 0 to 0. Conclude from this that R = r. Hint: Do the analytic
continuation through a series of steps involving reflection through a circle, as
in Exercise 7.1.11; the first of these steps involves reflecting the annulus AR
through the circle of radius R.
11. Prove that if F is a family of analytic functions on a connected open set U
and if for each disc Δ in U the family of restrictions of elements of F to Δ is
a normal family, then F itself is a normal family.
7.4. The Picard Theorems 243
12. Prove that an entire function which is not a polynomial takes on every complex
value but one infinitely often.
Chapter 8
Infinite Products
where {z1 , z2 , · · · , zn } are the zeroes of p. It turns out that product expansions
of a similar type (but with infinitely many factors) are possible for other analytic
functions.
Since the exponential function converts sums to products, we can expect that
the theory of infinite products will be closely related to the theory of infinite sums.
245
246 8. Infinite Products
Theorem 8.1.2. If {uk } is a sequence of complex numbers, then the infinite product
(8.1.2) converges to a non-zero number p if and only if the infinite sum
∞
(8.1.3) log uk
k=1
Proof. We have to be careful here, because the log function converts products to
sums only up to ±2πi. However, it is true that log uv = log u + log v if u and v
have positive real parts since, in this case, log u and log v have imaginary parts in
(−π/2, π/2) and log uv has imaginary part in (−π, π). Thus, log uv and log u+log v
cannot differ by a non-zero multiple of 2πi.
We define the partial products pn as in (8.1.1). If p = limn→∞ pn exists and is
non-zero, then
lim log(pn /p) = 0,
n→∞
It follows that pn /pm = (pn /p)(pm /p)−1 is in the right half-plane for n, m ≥ N . In
particular, un+1 = pn+1 /pn is in the right half-plane for n ≥ N . Thus,
for all n > N . Since the left side of this equality converges as n → ∞, so does the
right side. This implies the convergence of the series (8.1.3).
Conversely, if this series converges and we let
n
λn = log un
k=1
be its nth partial sum, then the sequence {λn } converges to a number λ. Since
p n = eλn ,
Proof. Let λn (z) be the nth partial sum of the infinite sum and pn (z) the nth
partial product of the infinite product. Then the uniform convergence of the series
on S implies that λn (z)−λ(z) converges uniformly to 0 on S. Since the exponential
function is continuous at 0, this implies that
pn (z)
= eλn (z)−λ(z)
p(z)
converges uniformly to 1.
The fact that each pn is bounded on S implies that Re(λn ) is bounded above
on S. This and the uniform convergence imply that Re(λ) is bounded above on
S, which implies that p(z) = eλ(z) is bounded on S. It follows that pn = (pn /p)p
converges uniformly to p on S.
If the series (8.1.4) converges uniformly on S, then there is a K such that |ak (z)| ≤
1/2 for k ≥ K and for all z ∈ S. If we use (8.1.6) with w = ak , it follows that one
of the two series
∞ ∞
| log(1 + ak (z))| and |ak (z)|
k=K k=K
converges uniformly on S if and only if the other one does also. Hence, if (8.1.4)
converges uniformly, then
∞
log(1 + ak (z))
k=K
converges uniformly and absolutely. By the previous two theorems, this implies the
uniform convergence of
∞
(1 + ak (z))
k=K
to a function on S with no zeroes. It follows that (8.1.5) converges uniformly on S,
the limit is unaffected by rearrangements of the factors, and each of its zeroes is a
zero, with the same order, of the product of the factors 1 + ak (z) for k < K.
Example 8.1.5. Prove that the infinite product
∞
(8.1.7) (1 − z 2 /k2 )
k=1
converges uniformly on DR (0). Hence, by the previous theorem, the infinite product
(8.1.7) also converges uniformly on DR (0) for each R and, hence, on each bounded
subset of C.
When applied to infinite products, this immediately implies the following corol-
lary.
Corollary 8.1.8. Let {uk } be a sequence of analytic functions on a connected open
set U . If the product
∞
f (z) = uk (z)
k=1
converges uniformly on compact subsets of U to a function f which is not identically
0, then the infinite sum
∞
uk (z)
uk (z)
k=1
converges uniformly to f /f on compact subsets of U \ S, where S is the set of
zeroes of f .
Example 8.1.9. Show that the function
∞
z2
f (z) = πz 1− 2
k
k=1
Solution: Note that the infinite product in the expression for f converges
uniformly on each compact disc in the plane by Example 8.1.5.
By the previous theorem,
∞ ∞
f (z) 1 −2z/k2 1 2z
= + = + .
f (z) z 1 − z /k
2 2 z z 2 − k2
k=1 k=1
the nth partial sum of the series (8.1.8) can be rewritten as the sum that appears
in (8.1.9).
10. Prove that if f , g and h are the functions of the previous exercise, then the
logarithmic derivative of g is h (z) = π cot πz − f (z)/f (z).
11. With h as above, prove that h is bounded on the strip 0 ≤ Re(z) ≤ 1 (use
(8.1.8)). Show that this implies it is bounded on the entire plane and, hence,
is constant.
12. With h as above, prove that h (0) = 0 and, hence, that h is identically 0 and
h is a constant. Then use the fact that limz→0 z −1 sin z = 1 to show that this
constant is 0. Conclude that
∞
z2
sin(πz) = πz 1− 2 .
k
k=1
of − log(1 − z) about z = 0 and so, although Ep (1) = 0, the sequence Ep (z) will
converge uniformly to (1−z)(1−z)−1 = 1 on each disc of radius less than 1 centered
at 0. More precisely:
Theorem 8.2.1. Each Ep (z) is an entire function with the following properties:
(a) the only zero of Ep (z) occurs at z = 1;
(b) if |z| ≤ 1, then |Ep (z) − 1| < |z|p+1 .
Proof. Part (a) is obvious. To prove Part (b), we note that the derivative of
1 − Ep (z) is (Exercise 8.2.1)
(1 − Ep (z)) = −Ep (z) = z p ez+z
2
/2+···+z p /p
(8.2.1) .
Since this has a zero of order p at z = 0, the function 1 − Ep (z) has a zero of order
p + 1 at z = 0.
The function (8.2.1) has a power series expansion about 0 with all of its coeffi-
cients non-negative real numbers, since this is true of the exponential function and
the function z + z 2 /2 + · · · + z p /p. It follows that the function
1 − Ep (z)
h(z) =
z p+1
also has non-negative real numbers as coefficients for its power series expansion
about 0. This implies that the maximum value achieved by |h(z)| for |z| ≤ 1 is
h(1) = 1. Part (b) follows from this.
252 8. Infinite Products
By Theorem 8.2.2 this infinite product converges to an entire function with the
required zeroes.
Weierstrass Factorization. The Weierstrass Theorem for the plane leads im-
mediately to the Weierstrass Factorization Theorem for entire functions:
Theorem 8.2.5. Let f be an entire function which is not identically zero. Let m
be the order of the zero of f at 0, and let {zk } be a list of the non-zero zeroes of
f counting multiplicity. Then there exist non-negative integers p1 , p2 , · · · and an
entire function h such that
∞
f (z) = eh(z) z m Epk (z/zk ).
k=1
The sequence {pk } may be chosen in any way which satisfies ( 8.2.2).
where the product is over all non-zero integers k. Note that if the factors for k and
−k in this product are paired, the result is
(1 − z/k) ez/k (1 + z/k) e−z/k = 1 − z 2 /k2 .
We conclude from Exercise 8.1.12 that eh(z) = π, that
sin(πz) = πz (1 − z/k) ez/k
k=0
Proof. Either U or its image under some linear fractional transformation will con-
tain ∞. Thus, we may as well assume ∞ ∈ U . Then the complement of U in S 2 is
a compact subset K of the plane.
Since {zk } has no limit point in U , the distance between zk and K must ap-
proach 0 as k → ∞. It follows that we may choose a sequence {wk } of points of K
such that lim |zk − wk | = 0.
We set
∞
zk − wk
f (z) = Ek .
n=1
z − wk
8.2. Weierstrass Products 255
The function f is analytic in U and has {zk } as a list of its zeroes counting multi-
plicity.
changing its poles and principal parts, in such a way as to end up with an infinite
series which does converge.
For each n > 1, the function gn is analytic on an open set containing the closed
disc Drn−1 (0). Hence, it is the uniform limit on this closed disc of its power series
at 0. It follows that there is a polynomial pn such that
|gn (z) − pn (z)| < 2−n for |z| ≤ rn−1 .
If we set f1 = g1 and fn = gn − pn for n > 1, then, for each m > 1, the series
∞
fn (z)
n=m+1
is defined as a meromorphic function on Drm (0) and has the required poles and
principal parts at those points of S which lie in this disc. Since this is true for each
m, and lim rm = R, f is meromorphic on all of DR (0) and has the required poles
and principal parts.
for all z with |z| sufficiently large. The infimum of all such numbers t is called the
order of f .
p
For each non-negative integer p, the function ez is an entire function of finite
order p. More generally:
Example 8.3.2. Show that eh(z) is an entire function of finite order p if h is a
polynomial of degree p.
Solution: If t > p, then limz→∞ |z|−t |h(z)| = 0. This implies that there is an
R > 0 such that
|h(z)| < |z|t for |z| > R.
Then
| eh(z) | ≤ e|z|
t
(8.3.1) for |z| > R.
Since such a statement is true for all t > p, by definition eh(z) has finite order at
most p.
On the other hand, if t < p, then limz→∞ |z|−t |h(z)| = +∞. Hence, there is no
R for which (8.3.1) holds. We conclude that the order of f is at least p and, hence,
is equal to p.
Non-Vanishing Entire Functions of Finite Order. It turns out that the func-
tions eh(z) of the preceding example are the only entire functions of finite order
which are non-vanishing.
To prove this, we will need the following theorem of Borel-Carathéodory relating
the growth of the real part of an analytic function to the growth of the the absolute
value of the function.
Theorem 8.3.3. Suppose 0 < r < R and let g be a function analytic on an open
set containing DR (0). Then
2r R+r
|g(z)| ≤ sup{Re(g(w)) : |w| = R} + |g(0)| if |z| ≤ r.
R−r R−r
This concludes the proof in the case where g(0) = 0. This general case follows
from applying this result to the function g0 (z) = g(z) − g(0). The details are left
to the exercises.
Theorem 8.3.4. An entire function f with no zeroes has finite order p if and only
if p is a non-negative integer and f has the form
f (z) = eh(z) ,
Proof. In view of Example 8.3.2, we need only show that every non-vanishing
entire function f of finite order p has the above form.
Since f has no zeroes and the plane is simply connected, there is an entire
function h such that
f (z) = eh(z) for all z ∈ C.
Since f has finite order p, for each t > p there is an M > 0 such that
This implies
Re(h(z)) ≤ |z|t for |z| ≥ M.
We apply the previous theorem with r > M and R = 2r to conclude
Since this is true for all r > M , Exercise 3.3.9 implies that h must be a polynomial
of degree at most t. Since t was an arbitrary number greater than p, we conclude
that h is a polynomial of degree at most p. If it were a polynomial of degree less
than p, then f would have order less than p. Hence, the degree of the polynomial
h is exactly p. This, of course, implies that p is a non-negative integer.
260 8. Infinite Products
Canonical Products. Given a sequence {zk }, we let μ be the inf of the numbers
t such that
∞
1
(8.3.2) < ∞.
|zk |t
k=1
If there is no such t, then we set μ = ∞. The number μ is called the exponent of
convergence for the sequence {zk }.
If {zk } has finite exponent of convergence μ, then we can write down a con-
vergent Weierstrass product (8.2.3), using {zk }, in which the sequence {pk } is a
constant p. We choose p to be the smallest integer such that μ < p + 1. Then the
condition
∞
1
(8.3.3) <∞
|zk |p+1
k=1
is satisfied. Hence, by Theorem 8.2.2, the Weierstrass product
∞
(8.3.4) f (z) = Ep (z/zk )
k=1
converges. This is called the canonical product for the sequence {zk }.
The significance of the choice of p made for the canonical product is that, with
this choice, the resulting product is an entire function with order λ equal to the
exponent of convergence μ of the sequence {zk }. The next theorem yields part of
what is needed to prove this. The remainder of the proof will come in the next
section.
Theorem 8.3.5. The canonical product for a sequence {zk }, with finite exponent
of convergence μ, is an entire function of finite order λ ≤ μ.
Proof. We choose p to be the smallest integer such that μ < p + 1, and let t be
any number in the range μ < t < p + 1.
We claim that there is a positive constant A such that
t
(8.3.5) |Ep (z)| ≤ eA|z|
for all z.
If |z| ≤ 1/2, this follows from (8.1.6) with w = Ep (z) − 1 and Theorem 8.2.1.
These combine to show that
| log Ep (z)| ≤ 2|z|p+1 ≤ 2|z|t ,
and this implies (8.3.5) holds with A = 2.
If |z| > 1/2, then |z|k ≤ 2t−k |z|t , and so
p
Re(z k )
log |Ep (z)| = log |1 − z| +
k
k=1
p
≤ |z| + |z|k ≤ (p + 1)2t |z|t .
k=1
t
Thus, (8.3.5) holds with A = (p + 1)2 in this case.
8.3. Entire Functions of Finite Order 261
where
∞
B=A 1/|zk |t .
k=1
The series in this expression converges because t is larger than the exponent of
convergence μ.
Since for any s > t, we have B|z|t ≤ |z|s for |z| sufficiently large, it follows that
f has finite order at most t. Since t was an arbitrary number strictly between μ
and p + 1, we conclude that f has order at most μ.
One might guess, based on Theorem 8.3.4, that the order of an entire function
of finite order must be a non-negative integer. This is not the case, as is shown by
the following example.
Example 8.3.6. Find an entire function with finite order 1/2.
Solution: The function
∞
sin πz z2
= 1− 2
πz k
k=1
for every t > 1/2 and for no smaller values of t. Hence, the sequence {1/k2 } has
exponent of convergence 1/2. Since 0 is the smallest integer p such that 1/2 < p+1,
the preceding theorem implies that the canonical product
∞
z
f (z) = 1− 2
k
k=1
1. Finish the proof of Theorem 8.3.3 by showing that, if it is true in the case
where g(0) = 0, then it is true in general.
2. Show that a polynomial has finite order 0.
3. Show that sin z, z −1 sin z, and cos z all have finite order 1.
4. If f is an entire function of order λ(f ), k is a non-negative integer, and g(z) =
f (z k ), then prove that λ(g) = kλ(f ), where λ(g) is the order of g. √
5. Prove that if g(z) is an even entire function of finite order λ and f (z) = g( √ z),
then f is an entire function of finite order λ/2. In particular, show that cos z
has order 1/2.
6. Prove that the order of the sum or product of two entire functions is less than
or equal to the maximum of the orders of the two functions.
7. What is the order of the entire function esin z ?
8. Suppose f is an entire function which satisfies the inequality |f (z)| ≤ |z||z| for
|z| sufficiently large. Prove that f has finite order at most 1.
9. Find the exponent of convergence of the following sequences: {2k }, {kr } (r >
0), {log k}.
10. Given an arbitrary non-negative real number μ, show that there is a sequence
of complex numbers {zk } with exponent of convergence μ.
11. Does the order of an entire function necessarily have to be the same as the
exponent of convergence of its sequence of zeroes? Justify your answer.
Jensen’s Formula.
Theorem 8.4.1. If f is analytic in an open set containing the disc Dr (0), f has
no zeroes on the boundary of this disc, f (0) = 0, and z1 , z2 , · · · , zn are the zeroes,
counting multiplicity, of f in Dr (0), then
2π
|f (0)|r n 1
log = log(|f (r ei θ)|) dθ.
|z1 | · |z2 | · · · |zn | 2π 0
This function is analytic and non-vanishing in an open set containing the closed
unit disc D, and has the same modulus on the unit circle as does f . Thus, g has an
analytic logarithm in an open set containing D. Then log |g(x)| is the real part of
an analytic function in this set and, hence, is harmonic. The Mean Value Theorem
for harmonic functions implies that
2π
|f (0)| 1
(8.4.1) log = log |g(0)| = log |f (eiθ )| dθ.
|z1 | · |z2 | · · · |zn | 2π 0
To prove the theorem for general r, it suffices to apply (8.4.1) with f replaced
by the function f (rz). If f has zeroes at z1 , z2 , · · · , zn in the disc Dr (0), then
f (rz) has zeroes z1 /r, z2 /r, · · · , zn /r in the unit disc D. Thus, the equation of the
theorem follows directly from (8.4.1) applied to f (rz).
This leads to the following estimate on the number of zeroes of an entire function
inside a disc Dr (0).
Theorem 8.4.2. If f is an entire function with |f (0)| = 1, n(r) is the number
of zeroes of f inside a disc Dr (0), and M (2r) is the supremum of |f (z)| on the
boundary of D2r (0), then
log M (2r)
n(r) ≤ .
log 2
Zeroes of Functions of Finite Order. The preceding theorem has the following
consequence for entire functions of finite order.
Theorem 8.4.3. Let f be an entire function of finite order λ and with f (0) = 0.
Let {zk } be a list of the zeroes of f , counted according to multiplicity and indexed in
the order of increasing modulus, and let μ be the exponent of convergence of {zk };
then μ ≤ λ.
264 8. Infinite Products
Proof. We claim that, for each t > λ, there are constants N , C > 0 and q > 1
such that
(8.4.2) |zk |t ≥ Ckq for all k ≥ N.
Assuming this, we conclude that the series
∞
1
|zk |t
k=1
which converges for q > 1. This, in turn, implies the exponent of convergence μ is
at most λ.
To complete the proof, we must verify the claim concerning (8.4.2). In doing
this, we may as well assume that |f (0)| = 1, since, if this is not so, we may make it
so by replacing f by f divided by a constant times a power of z. Such a replacement
will have no effect on whether the above claim is true.
Let rk = |zk |. Since the zeroes are indexed in such a way that the modulus is
a non-decreasing function of k, there are at least k zeroes of f with modulus less
than or equal to rk . By Theorem 8.4.2,
log M (2rk )
k≤ ,
log 2
where M (2rk ) is the sup of |f (z)| on the circle |z| = 2rk .
We choose s with λ < s < t. Since f has order λ, there is an R such that
rk ≥ R implies
s
M (2rk ) ≤ e(2rk ) .
Hence, for rk ≥ R,
(2rk )s
k≤ .
log 2
This implies
(log 2)t/s t/s
rkt ≥ k = Ckq ,
2t
where
(log 2t/s ) t
C= and q = > 1.
2t s
This is true provided rk = |zk | > R. However, since lim zk = ∞, there is an N such
that k > N implies |zk | > R. This completes the proof.
The above theorem, when combined with Theorem 8.3.5, yields the following
corollary.
Corollary 8.4.4. The canonical product for a sequence {zk } with exponent of con-
vergence μ has finite order λ = μ.
8.4. Hadamard’s Factorization Theorem 265
Hadamard’s Theorem. In the proof of the next theorem, we will need the
following estimates on the size of the inverse Ep−1 (z) of the function Ep (z).
Lemma 8.4.5. If p is a non-negative integer, p ≤ t ≤ p + 1, and z ∈ C, then there
is a constant A such that
1 t
(8.4.3) ≤ eA|z|
|Ep (z)|
if |z| ≥ 2 or |z| ≤ 1/2.
Thus, (8.4.3) holds with A = 2 in this case. If we choose A = max{2, p}, then
(8.4.3) holds in both cases.
where m is the order of the zero of f at 0, {zk } is a list of the other zeroes of f
counting multiplicity, and h(z) is a polynomial of degree at most p.
Let t be any number with λ < t ≤ p + 1 and let r ≥ 1 be any radius which is
not one of the numbers |zk |. We factor g(z) as g(z) = g1 (z)g2 (z), where
(8.4.5) g1 (z) = f (z)z −m Ep−1 (z/zk )
|zk |≤2r
and
(8.4.6) g2 (z) = Ep−1 (z/zk ).
|zk |>2r
Suppose |z| = 4r = R. Then |z/zk | ≥ 2 for all k with |zk | ≤ 2r. By the
previous lemma, there is a positive constant A1 such that
eA1 |z/zk | .
t
|g1 (z)| ≤ |f (z)|
|zk |≤2r
|f (z)| ≤ e|z|
t
and, hence,
t
(8.4.7) |g1 (z)| ≤ eB1 r ,
where
∞
1
t
B1 = 4 1 + A1 .
|zk |t
k=1
The infinite series in this expression converges by Theorem 8.4.3. Since g1 (z) is
an entire function (once the removable singularities at the zk with |zk | < 2r are
removed), if the inequality (8.4.7) holds for |z| = 4r = R, it must hold for all z in
the disc |z| ≤ R, by the Maximum Modulus Principle. In particular, this inequality
holds for all z with |z| = r.
Also if |z| = r, then |z/zk | < 1/2 if |zk | > 2r, and the previous lemma implies
that there is a constant A2 such that
eA2 |z/zk | ≤ eB2 r ,
t t
(8.4.8) |g2 (z)| ≤
|zk |>2r
where
∞
1
B2 = A 2 .
|zk |
k=1
Since t is an arbitrary number larger than λ and less than or equal to p + 1, g has
order at most λ. This completes the proof.
8.4. Hadamard’s Factorization Theorem 267
This chapter is devoted to developing some of the properties of two special func-
tions of a complex variable – the gamma function and the zeta function. These
functions are of great importance in modern mathematics. The developement of
their properties provides a very instructive practical application of many of the
techniques developed in the preceding chapters.
The zeta function is the subject of one of the most famous unsolved problems
in mathematics – the Riemann Hypothesis. This conjecture arose from Riemann’s
attempt to settle an old conjecture concerning the rate of growth of the number
π(x) of primes less than or equal to x as the positive number x increases. In the
process, Riemann developed (but did not completely prove) a formula for π(x). This
formula involves the zeroes of the zeta function in the strip 0 < Re(z) < 1, and its
study led Riemann to conjecture that all these zeroes lie on the line Re(z) = 1/2.
If true, this would have been helpful in both the proof of Riemann’s Formula and
its use in analyzing the growth of π(x).
The methods introduced by Riemann eventually led to proofs by others of the
result on the growth of π(x) that he was seeking. This result is now known as
the Prime Number Theorem. These proofs use information about the location of
the zeroes of the zeta function, but not, of course, the information proposed in the
Riemann Hypothesis, since it has never been proved.
One of the reasons this chapter is included in the text is so that we may describe
the Riemann Hypothesis and its connection to the Prime Number Theorem. For
completeness, we conclude the chapter with a proof of the Prime Number Theorem.
This proof makes strong use of the results on infinite products presented in the
previous chapter.
We begin the chapter with a discussion of the gamma function.
269
270 9. The Gamma and Zeta Functions
Proof. For 0 < r < s we define a function Γr,s on the right half-plane by
s
Γr,s (z) = e−t tz−1 dt.
r
The function e−t tz−1 = e−t+(z−1) log t is continuous as a function of (t, z) in [r, s]×C
and is analytic in z for each fixed value of t. By Exercise 3.2.16, Γr,s is analytic
on the entire plane. We will show that as s → ∞ and r → 0 the functions Γr,s
converge uniformly on each strip of the form
S = {z : a ≤ Re(z) ≤ b} with 0 < a < b.
The limit function is then necessarily analytic on the right half-plane and is, by
definition, Euler’s function Γ.
If x = Re(z), then
| e−t tz−1 | = e−t tx−1 .
Thus, if z is in the strip S, then
(9.1.2) | e−t tz−1 | ≤ ta−1 for t ≤ 1.
If t ≥ 1, then
| e−t tz−1 | ≤ e−t tb−1 on S.
The function e−t tb+1 is continuous and has limit 0 at infinity. It is, therefore,
bounded on [1, ∞), by a positive number K. Thus,
(9.1.3) | e−t tz−1 | ≤ Kt−2 for t ≥ 1.
−2
Since t is integrable on (0, 1] and Kt is integrable on [1, ∞), inequalities (9.1.2)
a−1
and (9.1.3) imply that the improper integrals of e−t tz−1 on (0, 1] and on [1, ∞) both
exist (see Theorem 5.2.2). Hence, the improper integral defining Γ exists for each
z ∈ S.
To show that Γ(z) is analytic, we will show that Γr,s converges uniformly to Γ
on each strip of the form S as r → 0 and s → ∞. In fact, from (9.1.2) and (9.1.3)
we conclude
r ∞
|Γ(z) − Γr,s (z)| ≤ ta−1 dt + Kt−2 dt ≤ r a /a + K/s.
0 s
Given > 0 the right side of this inequality is less than whenever r < (a/2)1/a
and s > 2K/. It follows from this that Γr,s (z) converges uniformly to Γ(z) for
z ∈ S as r → 0 and s → ∞. This completes the proof.
9.1. Euler’s Gamma Function 271
Proof. We have ∞
Γ(z + 1) = e−t tz dt.
0
Integrating by parts with u = tz and dv = e−t dt yields
∞
∞
Γ(z + 1) = −tz e−t 0 + e−t ztz−1 dt = zΓ(z).
0
Zeroes of Gamma. It turns out that Γ has no zeroes. To prove this requires de-
riving another functional equation. The derivation involves a pair of computational
lemmas.
We define Euler’s beta function, B(z, w), by
1
B(z, w) = (1 − s)z−1 sw−1 ds,
0
Proof. If we set z = x and w = 1 − x for x ∈ (0, 1), then since Γ(1) = 1, Lemma
9.1.5 implies
1
Γ(x)Γ(1 − x)
Γ(x)Γ(1 − x) = = B(x, 1 − x) = (1 − s)x−1 s−x ds.
Γ(1) 0
t
Then the substitution s = leads to
t+1
∞ x−1 ∞ −x
t t−x dt t
Γ(x)Γ(1 − x) = 1− −x 2
= dt.
0 t+1 (t + 1) (t + 1) 0 1+t
We use the previous lemma to evaluate the last integral and conclude that
π
Γ(x)Γ(1 − x) = .
sin πx
Since this identity holds for x ∈ (0, 1), the Identity Theorem implies that it contin-
ues to hold when x is replaced by any complex number z for which the functions
involved are defined.
This theorem has the following corollary, the proof of which is left as an exercise
(Exercise 9.1.4).
Corollary 9.1.8. The gamma function has no zeroes.
Product Formula for Γ. The fact that e−t = limn→∞ (1−t/n)n can be exploited
to express Γ as an infinite product of the sort studied in the previous chapter. The
first step in deriving this formula is to show the following:
Theorem 9.1.9. The identity
n
Γ(x) = lim (1 − t/n)n tx−1 dt
n→∞ 0
holds for all x > 0.
for n > a. The first term on the right converges to 0 as n → ∞ by (9.1.4) and the
second term can be made less than any given by choosing a large enough, because
the improper integral defining Γ converges.
274 9. The Gamma and Zeta Functions
Theorem 9.1.10. The entire function 1/Γ can be represented as the infinite prod-
uct
∞
1 1 + z/k
=z ,
Γ(z) (1 + 1/k)z
k=1
where this product converges uniformly on each disc of finite radius.
Proof. The integral in the previous theorem may be evaluated using a repeated
application of integration by parts (Exercise 9.1.9). The result is
nx n!
Γ(x) = lim .
n→∞ x(x + 1) · · · (x + n)
for x > 0.
If we invert this, divide both numerater and denomenator by n!, and note that
nx = n−1 x
k=1 (1 + 1/k) , we obtain
∞
1 + x/k
1
=x
Γ(x) (1 + 1/k)x
k=1
for x > 0.
If we can show that this infinite product converges, not just for x > 0, but
uniformly on each disc of finite radius in the complex plane, then the result will be
an entire function which agrees with the entire function 1/Γ(z) on the positive real
axis. This implies the two entire functions agree on all of C. Thus, the proof will
be complete if we can show that
∞
1 + z/k
(9.1.5) z
(1 + 1/k)z
k=1
converges uniformly on each compact disc. This product is very nearly a Weierstrass
product, as studied in the previous chapter. This fact can be used to prove the
uniform convergence on compact discs. The details are left to Exercise 9.1.10.
1. Show that, for z real and positive, Γ(z) is the Mellin transform of a certain
function. What function? (see Section 5.4).
2. Prove that Γ(n) = (n − 1)! if n is a positive integer.
Γ(z + n + 1)
3. Prove that z(z + 1)(z + 2) · · · (z + n) = .
Γ(z)
4. Prove Corollary 9.1.8.
(−1)n
5. Prove that the residue of Γ(z) at −n is for n = 0, 1, 2, · · · .
∞ n!
−rt z−1
6. Prove that, for r > 0 and Re(z) > 0, 0 e t dt = r −z Γ(z).
−π
7. Prove that Γ(z)Γ(−z) = .
z sin πz
9.2. The Riemann Zeta Function 275
1
8. Prove that e−t −(1 − t/n)n ≤ for all t ∈ [0, n]. Hint: Show that the
ne
−t
maximum of the function h(t) = e −(1 − t/n)n on [0, n] occurs at a point t0
where h(t0 ) = e−t0 t0 /n. Then show that this number is less than or equal to
1
.
ne
9. Using integration by parts, prove that if x > 0, then
n n
t nx n!
1− tx−1 dt = .
0 n x(x + 1) · · · (x + n)
10. Prove that the infiinite product
∞
1 + z/k
1
=z
Γ(z) (1 + 1/k)z
k=1
A Product Formula for the Zeta Function. Let {p1 , p2 , p3 , · · · } be the set of
prime numbers written in increasing order. Then we have
∞
Theorem 9.2.1. For Re(z) > 1, ζ(z) = n=1 (1 − p−z
n )
−1
.
Proof. The fact that the infinite product converges for Re(z) > 1 follows from
∞ Re(z)
Theorem 8.1.4 and the fact that n=1 pn converges if Re(z) > 1. Then
∞
∞
ζ(z)(1 − 2−z ) = n−z − (2n)−z = n−z ,
1 1 n∈S1
where S1 is the set of odd natural numbers. An induction argument using the same
technique then shows that for each natural number k
k
ζ(z) (1 − p−z
n ) = n−z ,
n=1 n∈Sk
276 9. The Gamma and Zeta Functions
where Sk is the set of natural numbers not divisible by any of the first k primes.
Since the right side of this equation has limit 1 as k → ∞, the theorem follows.
This theorem has the following two corollaries. We leave the proofs to the
exercises (Exercises 9.2.2 and 9.2.3).
Corollary 9.2.2. There are infinitely many primes.
Corollary 9.2.3. The zeta function has no zeroes in the region Re(z) > 1.
The Function ξ. Our next goal is to extend the zeta function to be a meromor-
phic function on the entire plane. We will do this by expressing the zeta function
in terms of the gamma function and a certain entire function ξ.
We begin the development of ξ by making the substitution t = n2 s2 π in the
formula (9.1.1) defining Γ. The result is
∞
ds
e−n s π s2z .
2 2
Γ(z) = 2n2z π z
0 s
If we divide by n2z π z and sum over n = 1, 2, 3, · · · , we obtain
∞ ∞
−z ds
e−n s π s2z
2 2
(9.2.2) ζ(2z)Γ(z)π = 2 if Re(z) > 1.
n=1 0
s
We will use the result of Exercise 9.2.7 to prove that it is legitimate to move the
summation inside the integral in the expression on the right. We estimate the size
of each integrand in this series as follows:
−n2 s2 π 2z−1
≤ e−ns s2 Re(z)−1 .
2
e s
The functions on the right are positive and their sum is
∞
s2 Re(z)−1
e−ns s2 Re(z)−1 =
2
(9.2.3) .
n=1
es2 −1
For each z, this series converges uniformly on each closed subinterval of (0, ∞).
2
Furthermore, since es −1 ≥ s2 , if Re(z) > 2, the function on the right in (9.2.3) is
less than or equal to s2 Re(z)−3 and, hence, has finite integral over [0, 1] if Re(z) > 2.
2 2
Since es −1 ≥ es /2 if s ≥ 1, the function on the right in (9.2.3) is less than or
equal to 2 e−s sRe(z)−1 on [1, ∞) and, hence, has finite integral on [1, ∞). It follows
2
that this function has finite integral on [0, ∞). Thus, by the result of Exercise 9.2.7,
the sum can be taken inside the integral in (9.2.2). Doing so yields
∞ ∞
−z ds
e−n s π s2z
2 2
(9.2.4) ζ(2z)Γ(z)π = 2 for Re(z) > 2.
0 n=1
s
The function ξ is obtained by multiplying this expression by z(z − 1)/2. Thus, for
Re(z) > 2,
∞
z(z − 1) −z/2 ds
(9.2.7) ξ(z) = ζ(z)Γ(z/2)π = z(z − 1) H(s)sz .
2 0 s
π
∞
= f (θ + 2πn)Pr (θ) dθ
−∞ −π
∞ (2n+1)π
(9.2.8) = f (θ)Pr (θ) dθ
−∞ (2n−1)π
∞
= f (θ)Pr (θ) dθ
−∞
∞
√
= 2π r |n| fˆ(n).
−∞
Note that the third step in this calculation uses the hypothesis that the series
defining g converges uniformly absolutely .
As r → 1 the integral on the left in (9.2.8) converges to
∞
2πg(0) = 2π f (2πn),
n=1
by Lemma 6.5.5, while the sum on the right converges to
∞
√
2π fˆ(n)
−∞
278 9. The Gamma and Zeta Functions
∞
since, by hypothesis, the series −∞ fˆ(n) converges absolutely. We conclude that
∞
∞
1
f (2πn) = √ fˆ(n),
n=−∞
2π n=−∞
as required.
1
If g(x) = √ e−x /2 is the normal distribution function from Example 5.3.6, then
2
2π
∞
√ √
G(s) = 2π g(ns 2π).
−∞
Proof. If we break the integral on the right side of (9.2.6) into an integral over
[1, ∞) and an integral over [0, 1] and make the substitution s → s−1 in the latter
integral, the result is
∞ ∞
ds ds
H(s)sz + H(s−1 )s−z .
1 s 1 s
9.2. The Riemann Zeta Function 279
It is useful to note that the above formula can be put in a slightly different
form by using Theorem 9.1.2. This theorem, with z replaced by z/2, implies that
(z/2)Γ(z/2) = Γ(z/2 + 1).
Then (9.2.11) becomes
π z/2 ξ(z)
(9.2.12) ζ(z) = .
(z − 1)Γ(z/2 + 1)
if Re(z) ≥ 2.
∞
6. Let u(t) = n=1 un (t) be the sum of a series of positive continuous func-
tions on (0, ∞) and suppose this series converges uniformly on closed bounded
intervals of (0, ∞). Prove that
∞ ∞ ∞
un (t) dt = u(t) dt.
n=1 0 0
Hint: Either
both sides are infinite or one of them is finite.
∞
7. Let h(t) = n=1 hn (t) be the sum of an infinite series of continuous functions
on (0, ∞) and suppose
|hn (t)| ≤ un (t) for all n, t,
∞
where n=1 un is a positive termed series which satisfies the conditions of the
previous exercise. Prove that the improper integral of h on R converges and
∞ ∞ ∞
h(t) dt = hn (t) dt.
0 n=1 0
8. Show that the Poisson Summation Formula can, under appropriate hypotheses
on f and fˆ, be reformulated as
∞
∞
√
f (n) = 2π fˆ(2πn).
n=−∞ n=−∞
9. Use the form of the Poisson Summation Formula derived in the previous ex-
ercise to show that
∞
1 e2π +1
= π .
−∞
1 + n2 e2π −1
1
Hint: The Fourier transform of is calculated in Example 5.3.3.
1 + x2
The next five exercises outline an alternative to the approach used in Theorem 9.2.7
to prove that ζ extends to be meromorphic in the plane.
10. Formula (9.2.4) was developed using the substitution t = n2 sπ in the integral
defining Γ. Use the substitution t = ns in a similar way to derive the formula
∞
1 z−1
ζ(z)Γ(z) = s ds for Re(z) > 1.
0 es −1
11. For complex numbers w and z, define (−w)z−1 to be e(z−1) log(−w) , where log
is the principal branch of the log function. Show that this function is analytic
except for a cut on the positive real line, that its limit as w approaches the
positive real number s from above is e(z−1)(log s−πi) , and that its limit as w
appproaches s from below is e(z−1)(log s+πi) .
9.3. Properties of ζ 281
γr
12. Using this definition for (−w)z−1 , consider the contour integral
1
(9.2.13) η(z) = w −1
(−w)z−1 dw,
γr e
where γr is the contour indicated in Figure 9.2.1, r < 2π is the radius of
the indicated circle, and the two horizontal lines are a distance ≤ r above
and below the positive real axis. Prove that this integral exists for all z, is
independent of r and , and defines an entire function η(z).
13. By passing to the limit as → 0 and r → 0 in the integral 9.2.13, prove that
η(z) = −2πi sin(πz)Γ(z)ζ(z) if Re(z) > 1.
14. Use the previous exercise and an identify involving Γ to prove that
1
ζ(z) = − Γ(1 − z)η(z) for Re(z) > 1.
2πi
Conclude that ζ has a meromorphic extension to the plane with a single simple
pole at z = 1.
9.3. Properties of ζ
The expressions (9.2.11) and (9.2.12) for ζ and the properties of Γ and ξ lead to a
wealth of information about ζ. Ultimately, this information will lead to a proof of
the Prime Number Theorem in the last section of this chapter.
Proof. The function Γ(z/2 + 1) has no zeroes, and the function ξ is entire. Thus,
(9.2.12) implies that the only pole of ζ is at z = 1 and it is a simple pole. The √fact
that the residue is 1 follows from the fact, proved in the exercises, that Γ(1/2) = π,
and from (9.2.10), which implies ξ(1) = 1/2.
282 9. The Gamma and Zeta Functions
Theorem 9.3.2. The zeta function has a zero of order 1 at each negative even
integer.
Proof. Except for the zeroes of ζ that occur at negative even integers (due to the
poles of Γ), the functions ζ and ξ have the same zeroes. Since ζ has no zeroes in
the region Re(z) > 1 by Corollary 9.2.3, and since ξ is symmetric about 1/2, it
follows that ζ has no zeroes outside the strip 0 ≤ Re(z) ≤ 1 except the negative
even integers.
Here, in both equations, the summation on the left is over all primes p.
Proof. If Re(z) > 1, then |p−z | < 1/2 for all primes p and so log(1 − p−z ) is
defined and analytic if log is the principal branch of the log function. Furthermore,
by Theorem 9.2.1 we have for Re(z) > 1
−z
(9.3.3) exp − log(1 − p ) = (1 − p−z )−1 = ζ(z).
p
−z
Hence, − p log(1 − p ) is an analytic logarithm for ζ(z) on Re(z) > 1. If we
expand this function in a power series in p−z , the result is (9.3.1). On differentiating
(9.3.1), we obtain (9.3.2).
Theorem 9.3.5. The zeta function has no zeroes outside the strip
0 < Re(z) < 1
except those which occur at negative even integers.
Since
(sz−1 + s−z ) = s−1/2 (sz−1/2 + (s−z+1/2 )
= 2s−1/2 cosh ((z − 1/2) log(s)),
(9.3.7) can be rewrittten as
∞
(9.3.8) ξ(z) = (s2 H (s)) s−1/2 cosh ((z − 1/2) log(s)) ds.
1
The terms of this series are clearly positive for s ≥ 1 and so the function itself is
positive. Also, the power series coefficients in the expansion of cosh w about 0 are
real and non-negative. Since log(s) ≥ 0 for s ≥ 1, the theorem follows (see Exercise
9.37).
Theorem 9.3.7. There is a constant R such that |ξ(1/2 + z)| ≤ r r for all z ∈ C
with |z| = r > R.
Proof. Since ξ(z + 1/2) has a power series expansion in z with real non-negative
coefficients, its maximum absolute value on any disc Dr (0) is achieved at z = r.
However, if n is an integer such that 1/2 + r ≤ 2n ≤ 5/2 + r, then by (9.2.7) and
the fact that ξ is increasing on the positive real line (Exercise 9.3.2),
ξ(1/2 + r) ≤ ξ(2n) = n(2n − 1)ζ(2n)Γ(n)π −n.
Now ζ is decreasing on (1, ∞). Thus,
ζ(2n) ≤ ζ(2) if n ≥ 1,
and
Γ(n) = (n − 1)!
if n is a positive integer. Thus,
ξ(1/2 + r) ≤ 2n n! ζ(2) ≤ 2ζ(2) nn+1 ≤ r r
if r is sufficiently large (since n ≤ 5/4 + r/2) (Exercise 9.3.9).
The following corollary is a direct consequence of the above theorem and the
results of sections 8.3 and 8.4. We leave the details to the exercises.
Corollary 9.3.8. The function ξ is an entire function of finite order at most
1. Consequently, the zeroes of ξ (and, hence, the zeroes of ζ that lie in the strip
0 < Im(z) < 1) form a sequence with exponent of convergence at most 1.
9.3. Properties of ζ 285
where q is a polynomial of degree at most 1, and the product is over all zeroes σ of
ξ(1/2 + w). However, the zeroes of this function are symmetric about 0 and so, if a
zero σ appears in this product, then so does its negative. The exponential factors
ew/σ and e−w/σ cancel and so
ξ(1/2 + w) = eq(w) (1 − w/σ)
σ
as long as it is understood that the factors involving a given σ and its negative −σ
are to be grouped together. If this is done, then the product expansion becomes
ξ(1/2 + w) = eq(w) (1 − w2 /σ 2 ).
Im(σ)>0
It is this product that actually converges (see the discussion of the product expan-
sion of sin(πz) in Example 8.2.6).
Now ξ(1/2 + w) is an even function of w (symmetric about 0) and so is the
above product. It follows that the polynomial q must also be an even function.
Since the only even polynomials of degree at most 1 are constants, we conclude
that
(9.3.9) ξ(1/2 + w) = c (1 − w/σ)
σ
converges as long as the factors involving ρ − 1/2 and its negative are grouped
together (Exercise 9.3.11), we obtain an infinite product expansion
z
ξ(z) = c1 1− .
ρ
ρ
where the numbers ρ are the zeroes of ξ(z) and the factors are arranged so that ρ
and 1 − ρ are grouped together. The product converges uniformly on each disc of
finite radius.
This, in turn, implies that the zeta function has the following infinite product
expansion:
Theorem 9.3.10. The zeta function satisfies
1 π z/2 z
ζ(z) = 1− ,
z(z − 1) Γ(z/2) ρ ρ
where the product is over the zeroes ρ of ζ in the strip 0 < Re(z) < 1, and the
factors involving ρ and 1 − ρ are grouped together in the infinite product.
8. Show that ∞
dt
e−t tz
1 t
is an entire function of z with all non-negative coefficients in its power series
expansion about 0.
9. Verify the claim made in the last sentence of the proof of Theorem 9.3.7 – that
is, show that 2ζ(2)nn+1 ≤ r r if r is sufficiently large and n ≤ 5/4 + r/2.
10. Prove the identity (9.3.10).
11. Show why the product in (9.3.11) converges if the terms are grouped as indi-
cated.
where the “other” terms are of lower order in x, and ρ ranges over all zeroes of ζ
in the strip 0 < Re(z) < 1. Note that the sums over n are actually finite sums for
each fixed x and ρ, since x1/n and xρ/n are less than 2 if n is large enough.
288 9. The Gamma and Zeta Functions
An integration by parts argument shows that the first (and presumably domi-
nant) term in the expansion (9.4.2) may be rewritten as
x
x dt
Li(x) = − 2
+ c.
log x 2 (log t)
x
The second term in this expression, when divided by , has limit 0 at infinity
log x
(Exercise 9.4.1). Thus, if the remaining terms in Riemann’s Formula for π(x), when
x
divided by , also have limit 0 at infinity, and if it is legitimate to take the limit
log x
inside the sum in the first term, then (9.4.1) follows.
In dealing with the terms involving the zeroes ρ of ζ in Riemann’s Formula, it
would be useful if the zeroes of ζ in the strip 0 < Re(z) < 1 all satisfied Re(z) < r
for some r < 1. Riemann suspected that this was true and, in fact, he conjectured
that all such zeroes actually lie on the line Re(z) = 1/2. This is the Riemann
Hypothesis.
Actually, Riemann did not give a complete proof of his formula (9.4.2) for
π(x) and, in fact, he did not even prove that the infinite series in this formula
converges. Both facts were eventually proved, but the difficulties involved in these
proofs and in determining the contribution of the terms involving the zeroes of ζ to
the asymptotic behavior of π(x) led to the introduction of another function ψ(x)
which also measures the density of primes and which satisfies a simpler and more
natural formula analogous to (9.4.2).
Eventually, Hadamard and de la Vallée-Poussin in 1896 proved (9.4.1). It is
now known as the Prime Number Theorem. The proofs of Hadamard and de la
Vallée-Poussin as well as other classical proofs of this result are based on the fact
that there are no zeroes of the zeta function on the line Re(z) = 1. We will present
one such proof in the next section.
(9.4.3) ψ(x) = log p = mp log p,
pm ≤x p≤x
where the first sum is over all prime powers pm ≤ x. For a given prime p, the term
log p appears in this sum as many times as there are positive powers of p less than
or equal to x. This number is the mp which appears in the second sum. It can also
be described as the largest number m such that pm ≤ x.
We will show that if the function ψ satisfies limx→∞ ψ(x)/x = 1, then the
Prime Number Theorem follows.
Lemma 9.4.1. Let x and y be real numbers greater than 1. Then
ψ(x) ψ(x)
≤ π(x) ≤ y + .
log x log y
9.4. The Riemann Hypothesis and Prime Numbers 289
Proof. We have
log p
π(x) = π(y) + 1≤y+ if y < x,
(9.4.4) log y
y<p≤x y<p≤x
π(x) ≤ y if y ≥ x,
where the sums are over primes p in the indicated range.
By definition,
(9.4.5) ψ(x) = mp log p = log pmp .
p≤x p≤x
The above result says that the Prime Number Theorem will follow if we can
prove that limx→∞ ψ(x)/x = 1. There are direct proofs of this; however, they
involve serious difficulties with improper integrals and conditionally convergent se-
ries. It turns out that these difficulties can be made to disappear if we follow a
similar approach, but use the integral of ψ rather than ψ itself as the main focus
of attention. Thus, we set
x
(9.4.6) φ(x) = ψ(u) du.
1
This is another function which measures the density of primes. Furthermore, the
Prime Number Theorem follows from an appropriate estimate on its asymptotic
behavior.
ψ(x)
Properties of φ. It turns out that, lim = 1, from which the Prime
x→∞ x
Number Theorem follows, provided
φ(x) 1
(9.4.7) lim 2
= .
x→∞ x 2
This is proved using a kind of reverse L’Hôpital’s Rule. Specifically:
Lemma 9.4.3. Let f be a positive, increasing function on [1, ∞) and suppose r > 0.
Then
f (x) r+1 x
lim = lim f (u) du,
x→∞ xr x→∞ xr+1 1
Proof. If we were to assume that the limit on the left exists, then the equality
would follow from applying L’Hôpital’s rule to the limit on the right. However,
we are assuming only that the limit on the right exists, and so we must proceed
differently (see Exercise 9.4.6).
Using the fact that f is increasing and positive, we conclude that, for α < 1
and β > 1,
x βx
1 1
f (u) du ≤ f (x) ≤ f (u) du.
(1 − α)x αx (β − 1)x x
On dividing by xr , this becomes
x βx
1 f (x) 1
f (u) du ≤ ≤ f (u) du.
(1 − α)xr+1 αx xr (β − 1)xr+1 x
x
If we set F (x) = 1 f (u) du, then this can be rewritten as
F (x) − F (αx) f (x) F (βx) − F (x)
≤ r ≤ ,
(1 − α)xr+1 x (β − 1)xr+1
9.5. A Proof of the Prime Number Theorem 291
or as
1 F (x) F (αx) f (x) 1 F (βx) F (x)
− αr+1 ≤ ≤ β r+1 − .
1−α x r+1 (αx)r+1 x r β−1 (βx)r+1 xr+1
If we set
F (x) F (αx) F (βx)
L = lim = lim = lim ,
xr+1
x→∞ x→∞ (αx)r+1 x→∞ (βx)r+1
then the above inequality implies that
1 − αr+1 f (x) f (x) β r+1 − 1
L ≤ lim inf r ≤ lim sup r ≤ L.
1−α x→∞ x x→∞ x β−1
The lemma then follows from this on taking the limit as α and β approach 1, since
1 − αr+1 β r+1 − 1
both and have limit r + 1.
1−α β−1
This leads directly to the following theorem. The details are left to the exercises.
φ(x) π(x)
Theorem 9.4.4. If lim = 1/2, then lim log x = 1.
x→∞ x2 x→∞ x
Proof. In Chapter 5 we showed how to use residue theory to calculate the Fourier
transforms of certain functions (Theorem 5.3.2). The integral that appears in (9.5.1)
is actually the Fourier transform of a function to which Theorem 5.3.2 applies. To
see this, we write
yz ez log y yb
= = eit log y
p(z) p(z) p(b + it)
for z = b + it. Then
b+i∞ z ∞
1 y 1 1
dz = f (t) eit log y dt = √ fˆ(− log y),
2πi b−i∞ p(z) 2π −∞ 2π
where f is the restriction to the real line of the meromorphic function
yb
f (z) = .
p(b + iz)
The function f has limit 0 at infinity since p is a non-constant polynomial. Thus,
by Theorem 5.3.2, if y > 1, then
b+i∞ z
1 y
dz = i Res(f (z) eiz log y , w),
2πi b−i∞ p(z)
w∈B
Solution: Since b > 0, by the previous lemma, if y > 1, the integral is the
yz
sum of the residues of at 0 and −1, which is 1 − 1/y. If y < 1, the lemma
z(z + 1)
implies that the integral is 0, since z(z + 1) has no zeroes to the right of Re(z) = 0.
9.5. A Proof of the Prime Number Theorem 293
xz+1
Proof. We multiply equation (9.3.2) by and integrate. The result is
z(z + 1)
1 b+i∞
xz+1 ζ (z)
− dz
2πi b−i∞ z(z + 1) ζ(z)
(9.5.2) b+i∞ z
1 x x log p
= dz ,
pm
2πi b−i∞ pm z(z + 1)
provided the integrals exist and the integral can be moved inside the summation
on the right. Assuming these things for the moment, we have, by the previous
example,
b+i∞
1 xz+1 ζ (z)
− dz = (x − pm ) log p.
2πi b−i∞ z(z + 1) ζ(z)
pm ≤x
x
The expression on the right is φ(x) = 1 ψ(u) du (Exercise 9.5.1), and so the proof
will be complete if we can verify that the integrals in (9.5.2) exist and the integral
can be brought inside the summation on the right.
The integrand corresponding to n = pm on the right in (9.5.2) is less than or
equal in modulus to
log n xb+1
nb b2 + t2
on the vertical line z = b + it. This has the form cn f (t), where f is a positive
n
integrable function of t on (−∞, ∞) and 1 cn is a convergent series of positive
numbers (see Exercise 9.5.2). By Exercise 9.5.4 this implies that the series of
integrals on the right in (9.5.2) converges and it converges to the integral on the
left.
where ρ ranges over the zeroes of ζ in the strip 0 < Re(z) < 1.
Proof. The integral that appears in Theorem 9.5.3 can also be evaluated by using
the infinite product expansion of Theorem 9.3.10. This theorem implies that the
logarithmic derivative of ζ can be written as
ζ (z) 1 1 log π Γ (z/2) 1
= − + − + ,
ζ(z) 1−z z 2 Γ(z/2) ρ
z−ρ
294 9. The Gamma and Zeta Functions
where, in the last sum, terms involving ρ and 1 − ρ must be grouped together for
the series to converge. The product formula for 1/Γ given in Theorem 9.1.10 leads
to
∞
Γ (z/2) 1 1 1
− = + − log(1 + 1/k) .
Γ(z/2) z 2k + z 2
k=1
Thus,
∞
ζ (z) 1 log π 1 1 1
=− + + + log(1 + 1/k) + .
ζ(z) z−1 2 z + 2k 2 ρ
z − ρ
k=1
or
∞
ζ (z) z z z ζ (0)
(9.5.3) =− − + + .
ζ(z) z−1 2k(z + 2k) ρ
ρ(z − ρ) ζ(0)
k=1
1 xz+1
We next multiply equation (9.5.3) by and integrate along the line
2πi z(z + 1)
Re(z) = b > 1, obtaining
(9.5.4)
b+i∞
1 xz+1 ζ (z)
φ(x) = − dz
2πi b−i∞ z(z + 1) ζ(z)
b+i∞ b+i∞ ∞
1 xz+1 1 xz+1
= dz + dz
2πi b−i∞ (z − 1)(z + 1) 2πi b−i∞ 2k(z + 2k)(z + 1)
k=1
b+i∞ b+i∞
1 xz+1 1 ζ (0) xz+1
− dz − dz.
2πi b−i∞ ρ ρ(z − ρ)(z + 1) 2πi b−i∞ ζ(0) z(z + 1)
Assuming for the moment that the integral can be taken inside each of the
infinite sums, the result is
b+i∞
1 xz+1 ζ (z)
φ(x) = − dz
2πi b−i∞ z(z + 1) ζ(z)
b+i∞ ∞ b+i∞
1 xz+1 1 xz+1
= dz + dz
2πi b−i∞ (z − 1)(z + 1) 2πi b−i∞ 2k(z + 2k)(z + 1)
k=1
1 b+i∞ xz+1 1
b+i∞
ζ (0) xz+1
− dz − dz.
ρ
2πi b−i∞ ρ(z − ρ)(z + 1) 2πi b−i∞ ζ(0) z(z + 1)
Each of these integrals can be evaluated using Lemma 9.5.1. This leads to
∞
x2 x1−2k − 1 xρ+1 − 1 ζ (0)
φ(x) = − − − x,
2 2k(2k − 1) ρ
ρ(ρ + 1) ζ(0)
k=1
9.5. A Proof of the Prime Number Theorem 295
or
∞
x2 x1−2k xρ+1
φ(x) = − − − Ax + B,
2 2k(2k − 1) ρ
ρ(ρ + 1)
k=1
∞
ζ (0) 1 1
where A = and B = + .
ζ(0) 2k(2k − 1) ρ
ρ(ρ + 1)
k=1
It remains to prove that the integral can be taken inside the infinite sums in
(9.5.4). The kth term of the first sum is
xz+1
(9.5.5) .
2k(z + 2k)(z + 1)
The numerator of this fraction is bounded on the vertical line Re(z) = b. With
z = b + it, we estimate the middle factor of the denominator as follows:
|z + 2k|2 = t2 + (b + 2k)2 = t2 + b2 + 4bk + 4k2
≥ t2 + b2 + 4k2 = |z|2 + (2k)2 ≥ 4|z|k.
Thus,
|z + 2k| ≥ 4|z|1/2 k1/2 .
The right factor of the denominator satisfies |z +1| ≥ |z| since z = b+it has positive
real part. Hence, the fraction(9.5.5) has modulus less than or equal to a constant
times |z|−3/2 k−3/2 . It follows that, in the first infinite sum the integral of each term
over Re(z) = b exists and the integral of the sum is the sum of the integrals and
the latter sum is absolutely convergent (see Exercise 9.5.4).
The term involving ρ of the second infinite sum in (9.5.4) is
xz+1
(9.5.6) .
ρ(z − ρ)(z + 1)
The numerator is bounded by xb+1 on Re(z) = b. With z = b + iy, ρ = β + iγ, and
c = b − 1, we estimate the denominator as follows: Since |z − ρ| ≥ (|y − γ| + c)/2
and |z + 1| ≥ (|y| + b + 1)/2 ≥ (|y| + c)/2, we have
|ρ(z − ρ)(z + 1)| ≥ |γ|(|y − γ| + c)(|y| + c)/4.
Thus, the modulus of (9.5.6) is less than or equal to
1
g(y) = 4xb+1 |γ|−1 h(y) where h(y) = .
(|y − γ| + c)(|y| + c)
If we divide the real line into three subintervals by cutting at y = 0 and y = γ,
then, on each of these subintervals, the absolute values in h can be eliminated, and
the integral of h with respect to y can be evaluated using the method of partial
fractions. The result (Exercise 9.5.5) is that the integral of h over each of the
unbounded subintervals is
|γ|−1 log(|γ|/c + 1),
while the integral of h over the bounded subinterval is
2(|γ| + 2c)−1 log(|γ|/c + 1) ≤ 2|γ|−1 log(|γ|/c + 1).
Thus, the integral of g over (−∞, ∞) is less than or equal to
16xb+1 |γ|−2 log(|γ|/c + 1)
296 9. The Gamma and Zeta Functions
5. Verify the claims made near the end of the proof of Theorem 9.5.4 regarding
the integral of
1
h(y) =
(|y − γ| + c)(|y| + c)
over each of the subintervals of (−∞, ∞) created by cutting at y = 0 and
y = γ.
9.5. A Proof of the Prime Number Theorem 297
6. Prove that log(s + 1) ≤ s1/2 for s ∈ (0, ∞). This was used near the end of the
proof of Theorem 9.5.4.
7. Verify the statement about taking the limit inside the integral in the proof
of Theorem 9.5.5. That is, prove that if a series ∞ n=1 un (x) of functions on
[1, ∞) converges uniformly absolutely on [1, ∞), then
∞ ∞
lim un (x) = lim un (x),
x→∞ x→∞
n=1 n=1
provided each limit on the right converges.
8. Prove that if f is a function analytic in an open set containing b + iw and
g(z) = f (b + iz), then g is analytic in an open set containing w and
Res(g, w) = −i Res(f, b + iw).
Bibliography
299
Index
E ◦ , 30 limz→a f (z), 31
Ep (z), 251
Li(x), 287 abelian group, 110
P 1 (C), 194 free, 110
S 2 , 192 absolute convergence, 11
Z(f ), 96 absolute value, 5
C, 2 absolutely integrable function, 161
affine transformations, 198
C2 , 194
analytic at ∞, 192
Γ(z), 270
analytic continuation, 221
Indγ (z), 64
across an analytic curve, 223
Res(f, z0 ), 131 along a curve, 227
arctan z, 18 analytic covering map, 233
E, 30 analytic curve, 223
cos z, 17 analytic function, 34
ez , 14 from S 2 to S 2 , 193
(γ), 52 higher derivatives, 86
exp(z), 14 local factorization, 95, 98
γ
f (z) dz, 45 power series expansion, 84
log z, 21 zeroes, 96
F, 205 analytic function element, 227
∂E, 30 angle preserving maps, 185
∂, 112 annulus, 125
antiderivative, 55
φ(x), 290
existence of, 61
π(x), 269, 287
arctan function, 18
ψ(x), 288
auxiliary polynomial, 92
sin z, 17
√
z, 23 bi-analytic map, 105
tan z, 17 binomial formula, 15, 18
ξ(z), 277 Bolzano-Weierstrass Theorem, 59
ζ(z), 275 Borel-Carathéodory Theorem, 258
az , 23 boundary, 30
z a , 23 boundary map, 112
301
302 Index
properties, 17 properties, 67
inf, 79
factoring polynomials, 91 infinite products
finite order convergence, 246
of an entire function, 258 uniform convergence, 247
Fourier Inversion Formula, 172 integral
Fourier transform, 167 over a 1-chain, 111
of normal distribution, 170 over a path, 45
using residues, 167 Riemann, 43
function φ, 290 integrating using residues, 159
function ψ, 288 interior, 30
functions of finite order, 258 inverse Fourier transform, 172
zeroes, 263 inverse Laplace transform, 175
Fundamental Theorem of Algebra, 90 inversion in C, 7
inversion transformation, 198
gamma function, 270
isolated point, 31, 96
has no zeroes, 273
isolated singularity, 98
infinite product formula, 274
meromorphic continuation, 271 Jensen’s Formula, 262
geometric series, 11 join of two paths, 50
greatest lower bound, 79 Jordan Curve Theorem, 121
group homomorphism, 112
kernel, 202
Hadamard, 288
Hadamard’s Theorem, 265 Laplace transform, 175
harmonic conjugate, 40 Laplace’s equation, 39
existence, 105 Laurent series expansion, 127
harmonic function, 39, 105 least upper bound, 78
maximum principle, 106 length of a path, 52
mean value property, 106 lifting a map, 232
Harnack’s Inequality, 213 through a covering map, 234
heat equation, 214 liminf, 79
heat flow, 214 limit of a function, 31
Heine-Borel Theorem, 54 limit of a sequence, 9
homeomorphic sets, 152 limsup, 79
homeomorphism, 195 linear fractional transformation, 197
homologous cycles, 115 Liouville’s Theorem, 89
homologous to 0, 115 log function, 21
homotopic curves, 145 branches, 22
homotopy, 145 is analytic, 39
hydrodynamics, 214 principal branch, 22
properties, 22
Identity Theorem, 97 logarithm of a function, 97
imaginary part, 5 logarithmic derivative, 248
improper integral, 160 long division of power series, 155
converge absolutely, 160
principal value, 160 Maximum Modulus Theorem, 103
indented contours, 172 Mellin transform, 176
independence of path, 63 using residues, 176
index, 64 meromorphic function, 101
for curves, 144 factorization, 255
locally constant, 144 inverse of, 101
of a cycle, 114 with pole at ∞, 194
304 Index
Sally
The This series was founded by the highly respected
mathematician and educator, Paul J. Sally, Jr.