Very Short Review About Complex Analysis
Very Short Review About Complex Analysis
Rudi Weikard
2 2
1 1
0
-2 0
-1
0 -1
1
-2
2
Preface iii
Chapter 1. The complex numbers: algebra, geometry, and topology 1
1.1. The algebra of complex numbers 1
1.2. The topology and geometry of complex numbers 2
1.3. Sequences and series 3
Glossary 33
Complex Analysis, also called the Theory of Functions, is one of the most important
and certainly one of the most beautiful branches of mathematics. This is due to the fact
that, in the case of complex variables, differentiability in open sets has consequences which
are much more significant than in the case of real variables. It is the goal of this course to
study these consequences and some of their far reaching applications.
As a prerequisite of the course familiarity with Advanced Calculus1 is necessary and,
generally, sufficient. Some prior exposure to topology may be useful but the necessary
concepts and theorems are provided in the appendix if they are not covered in the course.
In order to help you navigate the notes an index, a list of symbols, and a glossary of
important terms (not explicitly defined in the text) are appended at the end.
There are many textbooks on our subject; several of the following have been consulted
in the preparation of these notes.
• Lars V. Ahlfors, Complex analysis, McGraw-Hill, New York, 1953.
• John B. Conway, Functions of one complex variable, Springer, New York, 1978.
• Walter Rudin, Real and complex analysis, McGraw-Hill, New York, 1987.
• Elias M. Stein and Rami Shakarchi, Complex analysis, Princeton University Press,
Princeton, 2003.
These lecture notes are not intended to be encyclopedic; I tried, rather, to be peda-
gogic. As a consequence there are many occasions where assumptions could be weakened
or conclusions strengthened. Also, very many interesting results living close by have been
skipped and other interesting and fruitful subjects have not even been touched. This is just
the very beginning of the most exciting and beautiful Theory of Functions.
Thanks to my classes of Fall 2013, 2016, and 2019 who caused many improvements to
the notes and found a large number of mistakes (but probably not all).
1
Among many sources my Advanced Calculus notes would serve in this respect. They may be found
at http://people.cas.uab.edu/~weikard/teaching/ac.pdf.
iii
CHAPTER 1
A subset S of C is called convex if it contains the line segment joining x and y whenever
x, y ∈ S. Every disk is convex.
1.2.6 Connectedness. A subset S of C is called connected if S ∩ A ∩ B 6= ∅ whenever
A, B ⊂ C are open sets neither of which covers S but whose union does.
The empty set, all singletons, and all convex sets are connected. To prove the last claim
assume that the convex set S is not connected and pick points x ∈ S ∩ A and y ∈ S ∩ B.
The fact that {t ∈ [0, 1] : (1 − t)x + ty ∈ A} has a supremum leads to a contradiction.
1.2.7 Connected components. If S is a subset of the complex plane and x ∈ S, let C(x)
subsets of S which contain x. Any union of sets in C(x)
be the collection of all connected S
is connected. In particular, T = C∈C(x) C is a connected subset of S. Any larger subset
of S is not connected. T is called a component of S. The components of S are pairwise
disjoint and their union equals S.
The components of open sets are open. There are at most countably many of them.
2.3. Integration
2.3.1 Integrals of complex-valued functions over intervals. Let [a, b] be a closed
interval in R and f a complex-valued function on [a, b]. Then Re f and Im f are real-valued
functions on [a, b]. We say that f is Riemann integrable over [a, b] if Re f and Im f are. The
integral is defined to be
Z b Z b Z b
f= Re f + i Im f.
a a a
The integral is linear and satisfies
Z b Z b
f ≤ |f |.
a a
Rb
To prove this inequality let z = a
f and, when this is different from 0, set α = |z|/z. Then
Rb
consider a αf which is real.
2.3.2 Paths. Let Ω ⊂ C be a non-empty open set and [a, b] a non-trivial bounded interval
in R. A continuous function γ : [a, b] → Ω is then called a path in Ω. If the derivative γ 0
exists and is continuous on [a, b], γ is called a smooth path.
2.4. CONTOURS 7
The points γ(a) and γ(b) are called initial point and end point of γ, respectively. A
path is called closed if its initial and end points coincide. The range of γ, i.e., the set
{γ(t) : t ∈ [a, b]} is denoted by γ ∗ . Note that γ ∗ is compact and connected. The number
Rb 0
a
|γ (t)|dt is called the length of the smooth path γ.
Let γ : [a, b] → Ω be a smooth path. The coordinate transform ϕ(t) = (1 − t)a + tb,
t ∈ [0, 1], gives rise to a new smooth path γ̃ = γ ◦ ϕ defined on [0, 1] with the same initial
and end points and the same range as γ. Similarly, η(t) = γ(a + b − t), t ∈ [a, b], called the
opposite of γ, also has the same range as γ but switches initial and end points. We use the
notation η = γ.
2.3.3 Integrals along smooth paths. Suppose γ : [a, b] → C is a smooth path in C and
f : γ ∗ → C is continuous. Then the number
Z b
(f ◦ γ)γ 0
a
R
is well-defined. It is called the integral of f along γ and denoted by γ
f . We have the
estimate Z
f ≤ sup(|f |(γ ∗ ))L(γ)
γ
2.4. Contours
2.4.1 Contours. A contour in Ω is a finite ordered list of smooth paths in the non-empty
open set Ω. Among the contours we introduce an associative binary relation ⊕ which assigns
to a and b, lists of length m and n, respectively, the concatenation of these lists, denoted
by a ⊕ b, i.e., a listLof length m + n. We will write γ1 γ2 for γ1 ⊕ ( γ2 ).
n Sn
Suppose Γ = k=1 γk is a contour in Ω. Then we denote the compact set k=1 γk∗ by
Γ∗ . Moreover, if f is continuous on Γ∗ we define
Z Xn Z
f= f.
Γ k=1 γk
2.4.2 Closed contours. A contour γ1 ⊕...⊕γn is called closed when there is a permutation
π of {1, ..., n} such that the end point of γk coincides with the initial point of γπ(k) for
k = 1, ..., n.
2.4.3 Connected contours. A particularly important instance of a contour Γ = γ1 ⊕...⊕γn
is when for k = 1, ..., n − 1, the end point of γk coincides with the initial point of γk+1
(perhaps after reordering the indices). A contour of this type is called a connected contour .
The initial point of γ1 is called the initial point of Γ while the end point of γn is called the
end point of Γ. We say a contour connects x to y, if it is a connected contour with initial
point x and end point y.
If Γ is a connected contour, then Γ∗ is connected.
8 2. COMPLEX-VALUED FUNCTIONS OF A COMPLEX VARIABLE
2.4.4 Polygonal contours. Let z1 , ..., zn+1 ∈ C. We define γk (t) = (1 − t)zk + tzk+1 for
t ∈ [0, 1] and k = 1, ..., n. Then γk is a smooth path with initial point zk and end point
zk+1 . Note that γk∗ is the line segment joining zk and zk+1 . The length of the line segment
equals the length of γk . Moreover, γ1 ⊕ ... ⊕ γn is a connected contour which we denote by
hz1 , ..., zn+1 i. For instance, when n = 3 and z4 = z1 , Γ = hz1 , z2 , z3 , z1 i is a closed contour
tracing the circumference of a (possibly degenerate) triangle.
2.4.5 Primitives. Suppose Ω is a non-empty open set and F 0 = f on Ω. Then F is called
a primitive of f . If F is a primitive of f and c is a constant, then F + c is also a primitive
of f .
IfRγ : [a, b] → Ω is a smooth path and the continuous function f has a primitive F in Ω,
then γ f = F (γ(b)) − F (γ(a)).
The primitives of the zero function in a non-empty connected open set are precisely the
constant functions.
P∞
such that |gn (z)| ≤ Mn for all z ∈ S and that the series n=1 Mn converges. Then the
following statements hold:
P∞
(1) The series Pn=1 gn (z) converges absolutely for every z ∈ S.
∞
(2) The series n=1 gn converges uniformly in S.
2.5.4 Uniform convergence and continuity. Let S ⊂ C and n 7→ fn a sequence of
continuous complex-valued functions on S which converges uniformly to a function f : S →
C. Then f is continuous on S.
2.5.5 Uniform convergence and integration. Let [a, b] be a bounded interval in R.
Suppose that n 7→ fn : [a, b] → C is a sequence of functions which are Riemann integrable
over [a, b] and that this sequence converges uniformly to a function f : [a, b] → C. Then f
is Riemann integrable and
Z b Z b
f = lim fn .
a n→∞ a
Of course, the analogous result holds for series: Suppose that n 7→ gn : [a, b] →
P∞C is a
sequence of functions which are Riemann integrable over [a, b] and that the series n=0 gn
converges uniformly to a function g : [a, b] → C. Then g is Riemann integrable and
Z b ∞ Z b
X
g= gn .
a n=0 a
Sketch of proof. Let Γ = γ1 ⊕ ... ⊕ γn where the γk are smooth paths defined on
[ak , bk ], respectively. Using 2.6.6 we see that IndΓ is analytic. Now define Fk for each γk as
in 3.1.2 so that Fk /(γk − z) is constant. In particular, Fk (bk ) = (γk (bk ) − z)/(γk (ak ) − z).
Hence
n
Y γk (bk ) − z
exp(2πi IndΓ (z)) = = 1.
γk (ak ) − z
k=1
By Theorem 2.6.5 IndΓ (z) is an integer. Continuity shows that the value of IndΓ is constant
on sufficiently small disks within Ω. Hence the sets {z ∈ Ω : IndΓ (z) = k} are open for every
k ∈ Z.
3.1.4 Example. Consider the smooth paths γ1 (t) = exp(it), γ2 (t) = exp(it)/2, and γ3 (t) =
exp(−it)/2 all defined on [0, 2π] and the contours Γ1 = γ1 , Γ2 = γ1 ⊕ γ2 , and Γ3 = γ1 ⊕ γ3 .
The contours are closed so that IndΓk (z) is defined for z = 0, z = 3i/4, and z = −2.
The number IndΓ (z) is also called the winding number of the contour Γ around z. The
above examples hint at the origin of the name.
Sketch of proof. Assume first that z0 6∈ ∆. T∞Let ∆0 = ∆. Given ∆n let ∆n+1 be one
of the four triangles constructed in 3.2.3. Then n=1 ∆n consists of one point only, say w.
3.3. CONSEQUENCES OF CAUCHY’S THEOREM 13
Sketch of proof. Using 2.6.4 we may express f (k) (z0 ) in terms of a Taylor coefficient
and then compute that coefficient employing a combination of 3.3.1 and 2.6.6.
P3.3.8
∞
Zeros and their order. Let z0 be a zero of the holomorphic function f and
n
n=0 n (z − z0 ) the Taylor series of f about z0 . Then z0 is an isolated point in the
a
set of all zeros of f unless an = 0 for all n ∈ N0 . If z0 is an isolated zero of f , the number
m = min({k ∈ N : ak 6= 0}) is called the order or the multiplicity of z0 as a zero of f .
In this case we have f (z) = (z − z0 )m g(z) where g is holomorphic in the domain of f and
g(z0 ) 6= 0.
3.3.9 The set of zeros of a holomorphic function. Let Ω be a non-empty open
connected subset of C and f a holomorphic function on Ω. Let Z(f ) be the set of zeros of
f , i.e., Z(f ) = {a ∈ Ω : f (a) = 0}, and A the set of limit points of Z(f ) in Ω. Then A and
Ω \ A are open and, in consequence, either Z(f ) = Ω or else Z(f ) is a set of isolated points.
3.3.10 Analytic continuation. Let f and g be holomorphic functions on Ω, a non-empty
open connected subset of C. Then the following statements hold:
(1) If f (z) = g(z) for all z in a set which contains a limit point of itself, then f = g.
(2) If f (n) (z0 ) = g (n) (z0 ) for some z0 and all n ∈ N0 , then f = g.
Suppose f is a holomorphic function on Ω and that Ω0 is another open connected set
which intersects Ω. Then there is at most one holomorphic function on Ω∪Ω0 which coincides
with f on Ω. This function (if it exists) is called the analytic continuation of f to Ω ∪ Ω0 .
Sketch of proof. By A.3.1 Γ∗ ×B(z0 , r) is compact and A.2.3 gives that ϕ|Γ∗ ×B(z0 ,r)
is uniformly continuous so that, for all t, |ϕ(γ(t), z) − ϕ(γ(t), z0 )| < ε if |z − z0 | is sufficiently
small.
3.4.2 Differentiation under the integral. Let Ω, Γ, ϕ, and g be defined as in 3.4.1. Fur-
thermore suppose that ϕ(u, ·) is holomorphic in Ω for each u ∈ Γ∗ and denote its derivative
by ψ(u, ·), so that ϕ(u, ·) is a primitive of ψ(u, ·) for every fixed u ∈ Γ∗ . If ψ : Γ∗ × Ω → C
is continuous, then g is holomorphic in Ω.
R
Sketch of proof. Note that we have ϕ(u, z)−ϕ(u, z0 ) = hz0 ,zi ψ(u, ·) and let h(z0 ) =
R
Γ
ψ(·, z0 ). Then
n Z bk Z 1
g(z) − g(z0 ) X
− h(z0 ) ≤ |ψ(γk (t), u(s)) − ψ(γk (t), z0 )|ds|γk0 (t)|dt
z − z0 ak 0
k=1
i.e., convergence of a Laurent series requires convergence of both series of positive and
negative powers.
Isolated singularities
We call the point z0 a removable singularity, if an = 0 for all n < 0. It is called a pole of
order m or of multiplicity m (where m > 0), if an = 0 for all n < −m and a−m 6= 0. The
point z0 is called an essential singularity, if it is neither a pole nor a removable singularity,
i.e., if the set {n ∈ Z : an 6= 0} is not bounded below.
4.1.3 Removable singularities. If z0 is a removable singularity of a function f with a
Laurent expansion on the punctured disk B(z0 , r) \ {z0 }, then there is an analytic contin-
uation of f to B(z0 , r). In particular, in this case f has a limit at z0 . Conversely, if z0
is an isolated singularity of f and limz→z0 (z − z0 )f (z) = 0, then z0 is in fact a removable
singularity of f . To see this define h(z0 ) = 0 and h(z) = (z − z0 )f (z) for z 6= z0 . Then one
may show that h is holomorphic near z0 .
4.1.4 Poles. The following statements about poles hold.
(1) The point z0 is a pole of order m of a function f if and only if 1/f has an analytic
continuation for which z0 is a zero of order m.
(2) If z0 is a pole of f and M is any positive real number, then there is a positive δ
such that |f (z)| ≥ M for all z ∈ B(z0 , δ) \ {z0 }.
(3) If z0 is an isolated singularity of f and there is a natural number m such that
limz→z0 (z − z0 )m+1 f (z) = 0, then z0 is either a removable singularity of f or else
a pole of order at most m.
(4) If f has a pole of order m at z0 , then there are complex numbers c1 , ..., cm so that
m
X ck
z 7→ f (z) −
(z − z0 )k
k=1
17
18 4. ISOLATED SINGULARITIES
Pm k
has a removable singularity at z0 . The sum k=1 ck /(z−z0 ) is called the principal
part or singular part of f at z0 .
4.1.5 Meromorphic functions. Let P be a set of isolated points in Ω without a limit
point in Ω and f a holomorphic function on Ω \ P . If no point of P is an essential singularity
of f , then f is called meromorphic on Ω.
4.1.6 The Casorati1-Weierstrass theorem. Suppose f is a holomorphic function defined
on B 0 = B(z0 , r) \ {z0 }. If z0 is an essential singularity of f , then f (B 0 ) is dense in C.
Sketch of proof. Assume on the contrary that there is a a ∈ C and δ > 0 such that
B(a, δ) ∩ f (B 0 ) = ∅. Then consider the function g(z) = 1/(f (z) − a) on B 0 which has a
removable singularity at z0 . If z0 is a zero of g of order m ∈ N then f has a pole of order
m. If m = 0, then f has a removable singularity.
4.2.4 Counting zeros and poles. Suppose f is a non-zero meromorphic function on the
open set Ω, Z is the set of its zeros, and P is the set of its poles. For any point z0 ∈ Ω there is
a unique integer M (z0 ) denoting the smallest index for which the coefficient in the Laurent
expansion of f about z0 is non-zero. In fact, if z0 is a zero or a pole of f its multiplicity is
M (z0 ) or −M (z0 ), respectively.
Let Γ be a closed contour in Ω\(Z ∪P ) such that IndΓ (w) = 0 whenever w ∈ C\Ω. There
are at most finitely many zeros and poles in those components of Ω which have a non-zero
index. To see this note that A ∪ Γ∗ is a compact subset of Ω if A = {z ∈ C : IndΓ (z) 6= 0}.
It follows that Z 0
1 f X
= M (z) IndΓ (z).
2πi Γ f
z∈Z∪P
A zoo of functions
In this chapter we investigate briefly the most elementary functions of analysis. Polyno-
mials and exponential functions have been introduced earlier since they are too important
to postpone their use.
is called a polynomial (of a single variable). The integer n is called the degree of p if an is
different from 0. It is then called the leading coefficient of p. The zero function is also a
polynomial but no degree is assigned to it. Any polynomial is an entire function.
If p and q are polynomials of degree n and k, respectively, then p + q and pq are also
polynomials. The degree of p + q is the larger of the numbers n and k unless n = k in which
case the degree is at most n. The degree of pq equals n + k.
5.1.2 The fundamental theorem of algebra. Suppose p is a polynomial of degree n ≥ 1.
Then there exist numbers a and z1 , ..., zn (not necessarily distinct) such that
n
Y
p(z) = a (z − zk ).
k=1
To prove this use either Rouchés theorem or the fact that 1/p would be entire if p had no
zeros.
5.1.3 Zeros of polynomials. The coefficients ak of a monic polynomial are given as
symmetric polynomials in terms of the zeros. As a matter of fact,
X
an−k = (−1)k zj1 ...zjk .
1≤j1 <...<jk ≤n
Pn Qn
In particular, an−1 = − k=1 zk and a0 = (−1)n k=1 zk . They vary continuously with the
zeros.
That the converse is also true can be proved Pwith the aid of Rouché’s
Pn theorem. The
n
precise statement is as follows: Suppose f (z) = k=0 ak z k and g(z) = k=0 bk z k are two
monic polynomials of degree at most n. If z0 is the only zero of f in B(z0 , r) and if the
multiplicity of z0 is m then the following holds: For every ε ∈ (0, r) there is a δ > 0 such
that, if |ak − bk | < δ for k = 0, ..., n, then g has precisely m zeros in B(z0 , ε) (counting
multiplicities).
21
22 5. A ZOO OF FUNCTIONS
5.1.4 Among the entire functions only polynomials grow like powers. Suppose f
is an entire function and k 7→ rk is an increasing unbounded sequence of positive numbers.
Furthermore, suppose there are positive real numbers N and C such that |f (z)| ≤ C|z|N
for all z lying on any of the circles |z| = rk . Then f is a polynomial whose degree is less
then or equal to N .
It is in fact enough to require Re(f (z)) ≤ C|z|N to arrive at the same conclusion. To see
R 2π
this use the Laurent expansion 3.4.6 of f to obtain 2πrkn an = 0 f (γk (t)) exp(−int)dt for
R 2π
n ∈ Z. Therefore, using that an = 0 if n < 0, we get πrkn an = 0 Re f (γk (t)) exp(−int)dt
R 2π R 2π
for n ∈ N and 2π Re a0 = 0 Re f (γk (t))dt. We also have 0 CrkN exp(−int)dt = 0 so that
2π
πrkn |an | ≤ 0 (CrkN − Re f (γk (t)))dt = 2πCrkN − 2π Re a0 . Hence an = 0 when n > N .
R
Thus the real (or the imaginary part) of a polynomial grows roughly at the same pace as
the modulus. This is, in fact, a special case of a more general result, the Borel-Carathéodory1
theorem, which allows to estimate the modulus of an entire function by its real (or imaginary)
part.
Note that f˜ is continuous at every point of Ω, in the sense of A.1.2, even the poles.
1Constantin Carathéodory (1873 – 1950)
5.3. EXPONENTIAL AND TRIGONOMETRIC FUNCTIONS 23
is called the exponential function. It is an entire function and exp0 (z) = exp(z).
If x, y ∈ C, then exp(x + y) = exp(x) exp(y). In particular, exp(z) is never zero and
exp(−z) = 1/ exp(z).
The exponential function has period 2πi, i.e., exp(z + 2πi) = exp(z). If p is any period
of exp, then there is an integer m such that p = 2mπi. Let a be a fixed real number and
Sa = {z ∈ C : a < Im(z) ≤ a + 2π}. Then exp |Sa maps Sa bijectively to C \ {0}.
5.3.2 Trigonometric functions. The trigonometric functions are defined by
exp(iz) + exp(−iz)
cos(z) = ,
2
and
exp(iz) − exp(−iz)
sin(z) = .
2i
They are entire functions, called the cosine and sine function, respectively.
The following properties, familiar from the real case, extend to the complex case:
(1) Derivatives: sin0 (z) = cos(z) and cos0 (z) = − sin(z).
(2) The Pythagorean theorem: (sin z)2 + (cos z)2 = 1 for all z ∈ C.
Thus there are (at most) countably many different functions LΓ which are defined this way,
even though there are many more contours Γ connecting 1 to a point in Ω. In fact, for
each m ∈ Z there is a Γ0 such that LΓ (z) − LΓ0 (z) = 2mπi. This shows that defining the
logarithm as an antiderivative comes with certain difficulties (to say the least). At the same
time this behavior gives rise to a lot of interesting mathematics.
LΓ is a holomorphic function on Ω with derivative 1/z. Moreover, exp(LΓ (z)) = z for
all z ∈ Ω but LΓ (exp(z)) may well differ from z by an integer multiple of 2πi.
5.4. THE LOGARITHMIC FUNCTION AND POWERS 25
5.4.3 The principal branch of the logarithm. Let Ω = C \ (−∞, 0]. Then
Z
du
log(z) = ,
γ u
where γ is contour in Ω connecting 1 and z, is uniquely defined for any z ∈ Ω (i.e., log = LΓ
where Γ : [0, 1] → C : t 7→ 1). It is called the principal branch of the logarithm. If z ∈ Ω has
polar representation z = r exp(it) where t ∈ (−π, π) and r > 0, then
log(z) = ln(r) + it.
The range of log is the strip {z ∈ C : | Im(z)| < π}. In particular, log(exp(z)) = z if and
only if | Im(z)| < π.
The Taylor series of z 7→ log(1 + z) about z = 0 has radius of convergence 1 and is given
by
∞
X (−z)n
log(1 + z) = − .
n=1
n
If Re(a) and Re(b) are positive, then log(ab) = log(a) + log(b). The conclusion may be
wrong when the hypothesis is not satisfied.
5.4.4 Powers. Suppose a is a non-zero complex number and L is a branch of the logarithm
whose domain includes a. If b ∈ Z, then ab = exp(bL(a)). We may therefore define
ab = exp(bL(a))
for any b ∈ C. Be aware, though, that, in general, the value of ab depends on the branch of
the logarithm chosen and is therefore ambiguous when b 6∈ Z. In most cases one chooses, of
course, the principal branch of the logarithm and defines ab = exp(b log(a)) (forsaking the
definition of powers of negative numbers). In this case we get ez = exp(z) for all z ∈ C,
after we define e = exp(1).
5.4.5 Power functions. Suppose Ω is a simply connected non-empty open subset of
C \ {0}. Let p be a complex number and L : Ω → C a branch of the logarithm. The function
Ω → C : z 7→ z p = exp(pL(z)) is called a branch of the power function. Each such branch
is a holomorphic function which never vanishes. Note that 1/z p = z −p .
If p ∈ N0 all branches of the associated power function may be analytically extended to
the entire complex plane and any branch gives then rise to one and the same function. The
same is true if p ∈ Z is negative, except that the function may not be extended to 0 since 0
is then a pole of the function.
If p ∈ Q there are only finitely many different branches of z 7→ z p in any simply
connected open set not containing 0. In fact, if m/n is a representation of p in lowest terms,
there will be n branches of z 7→ z p .
5.4.6 Holomorphic functions without zeros. Suppose f is a holomorphic function
defined on a simply connected open set Ω. If f has no zeros, then f 0 /f has an antiderivative
g̃ and f exp(−g̃) is constant. It follows that there is an holomorphic function g on Ω such
that f (z) = exp(g(z)).
5.4.7 A definite integral. We will later use the following result (which is also an instruc-
tive example):
Z 2π
log |1 − exp(it)|dt = 0.
0
26 5. A ZOO OF FUNCTIONS
To see this let f (z) = z −1 log(1 − z) R and note that 0 is a removable singularity of f .
Hence we get for small positive δ that γ1 γ2 f = 0 when γ1 (t) = exp(it), t ∈ [δ, 2π − δ] and
γ2 (t) = (1−t) exp(iδ)+t exp(−iδ), t ∈ [0, 1]. Since Re(log(1−z)) = ln |1−z|, |1−eit | ≥ |t|/2
for sufficiently small t, |1 − γ2 (t)| ≥ |1 − cos δ| ≥ δ 2 /4, and t ln t − t is an antiderivative of
ln t, we obtain the claim upon taking the limit δ → 0.
CHAPTER 6
Entire functions
Moreover,
Y
Epn (z/an )
n
converges to an entire function f . The zeros of f are precisely the numbers an and the order
of z0 as a zero of f equals #{n : z0 = an }.
6.2.3 The Weierstrass factorization theorem. Let f be an entire function and suppose
that an , n ∈ N, are the non-zero zeros of f repeated according to their multiplicity. Then
there is an integer m, an entire function g, and a sequence n 7→ pn ∈ N0 such that
Y
f (z) = z m exp(g(z)) Epn (z/an ).
n
Sketch of proof. We may assume that f (0) = 1 and that the non-zero zeros of f
are labeled by an which are repeated according to their multiplicity and ordered by absolute
value. Thus, if s > ρ and 0 < ε < s − ρ, we obtain from 6.3.5
|an |−s ≤ es nf (|an |)−s/(ρ+ε) ≤ es n−s/(ρ+ε)
when |an | = r > R.
6.4.4 The very little Picard theorem. An entire function of finite order that misses
two values is constant.
To prove this assume that the entire function f of finite order misses the values α and
β, where α 6= β. Then h = (f − α)/(β − α) is entire, of finite growth order, and misses the
values 0 and 1. Hadamard’s theorem 6.4.3 shows that h = exp(g) for some polynomial g.
Since g has no zero it must be constant by the fundamental theorem of algebra, Theorem
5.1.2.
Picard’s3 little theorem states that this so for every entire function. Picard’s great
theorem states that near an isolated singularity a holomorphic function takes every value,
with at most one exception, infinitely often; this is an extension of the Casorati-Weierstrass
theorem.
6.4.5 An entire function of finite non-integral order assumes every value infin-
itely often. If f − α has no zeros it must be of integral order according to Hadamard’s
theorem 6.4.3. The claim follows now since a canonical product with finitely many zeros
has order 0.
A.1. Basics
A.1.1 Topology for a set. Let X be a set. The system τ of subsets of X is called a
topology for X if it has the following properties:
(1) ∅ and X are in τ
(2) Any union of elements of τ is again in τ .
(3) The intersection of two elements of τ is again in τ .
If τ is a topology for X, the pair (X, τ ) is called a topological space. The elements of τ
are called open sets, while their complements are called closed sets.
A set V is called a neighborhood of x ∈ X, if it is open and contains x.
A.1.2 Continuity. Suppose X and Y are topological spaces and f is a function from X
to Y . Then f is called continuous at x ∈ X, if for every neighborhood V of f (x) there is a
neighborhood U of x such that f (U ) ⊂ V . The function f : X → Y is called continuous, if
it is continuous at every point x ∈ X. Equivalently, f is continuous, if the pre-image of any
open subset of Y is an open subset of X.
A.1.3 Metric spaces. Let X be a set. A function d : X × X → [0, ∞) is called a metric
or a distance function in X if it has the following properties:
(1) d(x1 , x2 ) = 0 if and only if x1 = x2 .
(2) d(x1 , x2 ) = d(x2 , x1 ) for all x1 , x2 ∈ X.
(3) d(x1 , x3 ) ≤ d(x1 , x2 ) + d(x2 , x3 ) for all x1 , x2 , x3 ∈ X (triangle inequality).
(X, d) is then called a metric space and the number d(x1 , x2 ) is called the distance
between x1 and x2 .
If X 0 is a subset of X then (X 0 , d0 ) is again a metric space when d0 is the restriction of
d to X 0 × X 0 . Pn
If Rn is equipped with the distance function d defined by d(x, y)2 = k=1 (xk − yk )2 , it
becomes a metric space called the Euclidean space.
A.1.4 Open and closed balls. Let (X, d) be a metric space, x0 a point in X, and r a
non-negative real number. Then the set B(x0 , r) = {x ∈ X : d(x, x0 ) < r} is called the
open ball of radius r centered at x0 while B(x0 , r) = {x ∈ X : d(x, x0 ) ≤ r} is called the
closed ball of radius r centered at x0 . Note that B(x0 , 0) = ∅ so that the empty set is an
open ball. Similarly, the singleton {x0 } = B(x0 , 0) is a closed ball.
A subset of a metric space is called bounded if it is contained in some ball.
A.1.5 Metric spaces are topological spaces. Let (X, d) be a metric space. Then
τ = {U ⊂ X : U is a union of open balls}
31
32 A. TOPOLOGY IN METRIC SPACES
is a topology for X. This follows easily once one proves with the help of the triangle
inequality that the intersection of two open balls is a union of open balls and hence an open
set.
A.1.6 Sequences and their limits. Let (X, d) be a metric space. A function x : N →
X : n 7→ xn is called a sequence in X. We say that the sequence x is convergent if the
following statement holds:
∃L ∈ X : ∀ε > 0 : ∃N > 0 : ∀n > N : d(xn , L) < ε.
If a sequence is not convergent, we say that it diverges or is divergent.
If a sequence x is convergent there is only one element in X to which it converges. This
element is called the limit of the sequence x and is denoted by limn→∞ xn .
A.1.7 Continuity. Let (X, d) and (Y, ρ) be metric spaces and f a function from X to
Y . The definition of continuity given in A.1.2 translates in the context of metric spaces as
follows. The function f is called continuous at a ∈ X if the following statement is true:
∀ε > 0 : ∃δ > 0 : ∀x ∈ X : d(x, a) < δ ⇒ ρ(f (x), f (a)) < ε.
A.2. Compactness
A.2.1 Compact sets. A subset S of a topological space is called compact if every open
cover of SShas a finite subcover . More precisely, whenever V is a family of open
Sn sets such
that S ⊂ V ∈V V then there is a finite subset {V1 , ..., Vn } of V such that S ⊂ k=1 Vk .
A closed subset of a compact set is compact.
A.2.2 Images of compact sets under continuous functions are compact.
A.2.3 Uniform continuity. Let (X, d) and (Y, ρ) be metric spaces and f a function from
X to Y . Then f is called uniformly continuous if the following statement holds:
∀ε > 0 : ∃δ > 0 : ∀x, x0 ∈ X : d(x, x0 ) < δ ⇒ ρ(f (x), f (x0 )) < ε.
If X is compact and f : X → Y is continuous, then it is uniformly continuous.
A.2.4 Sequences in compact metric spaces have convergent subsequences.
A.2.5 The Heine1-Borel2 theorem. The Heine-Borel theorem states that a subset of
Rn is compact if and only if it is closed and bounded. It is a very important result since
it makes it easy to check compactness which in turn is an essential ingredient for several
important results.
A.2.6 The Bolzano3-Weierstrass theorem. Suppose X is a metric space. Every se-
quence in X has a convergent subsequence if and only if X is compact.
annulus: In a metric space an annulus is a set lying between two concentric circles.
binary operation: A binary operation on a set A is a function from A × A to A. It is
customary to express a binary operation as a ? b (or with other symbols in place of ?).
circumference of a triangle: The sum of the edge lengths of the triangle.
congruence of triangles: Two triangles are congruent if their edge lengths respectively
coincide.
dense: A subset A of a metric space (X, d) is called dense in X if every point of X \ A is
a limit point of A.
diameter: In a metric space (X, d) the diameter of a set A ⊂ X is defined to be sup{d(x, y) :
x, y ∈ A}.
group: A set in which an associative binary operation is defined so that an identity and
inverses to any element exist is called a group.
homeomorphic: Two topological spaces X and Y are called homeomorphic, if there is a
bijection f : X → Y such that both f and f −1 are continuous.
identity: An identity (in a set A with a binary operation ?) is an element e ∈ A such that
e ? a = a ? e = a for all a ∈ A.
interior point: A point z is called an interior point of a set S ⊂ C if there is an open disk
about z which is contained in S.
inverse: The inverse of an element a (in a set A with a binary operation ? and identity e)
is an element b ∈ A such that a ? b = b ? a = e.
isolated point: A point x in a metric space (X, d) is called an isolated point of a set
S ⊂ X if there exists an open disk B about x such that B ∩ S = {x}.
limit point: A point x in a metric space (X, d) is called a limit point of a set S ⊂ X, if
B(x, 1/n) ∩ S \ {x} =
6 ∅ for all n ∈ N.
linear: A map F (defined on a vector space) is called linear if F (αu+βv) = αF (u)+βF (v)
for all α, β ∈ C and all u, v in the domain of F .
local maximum: The value u(z0 ) is called a local maximum of a function u : S → R if
u(z0 ) ≥ u(z) for all z in some neighborhood V = B(z0 , r) ∩ S of z0 .
monic: A polynomial is called monic if its leading coefficient is equal to 1.
monomial: A function of the form m : Ck → C : (z1 , ..., zk ) 7→ az1n1 ...zknk when a is
a complex number and n1 , ..., nk are non-negative integers. The number a is called the
coefficient of the monomial while z1 , ..., zk are called the variables.
33
34 GLOSSARY
periodic: A function f defined on C (or R) is called periodic if there exists a complex (or
real) number a 6= 0 such that f (z + a) = f (z) for all z ∈ C (or z ∈ R). The number a is
called a period of f .
permutation: Let A be a finite set. A permutation of A is a bijective map from A to
itself.
polynomial: A finite sum of monomials.
punctured disk: A set is called a punctured disk if it is a disk with the center removed.
P∞ P∞
rearrangement: The series n=1 wn is a rearrangement of the series n=1 zn if there is
bijective sequence k : N → N such that wn = zkn .
similarity of triangles: Two triangles are similar if the ratios of their respective edge
lengths coincide.
singleton: A singleton is a set containing precisely one element.
symmetric polynomial: A polynomial is called symmetric if no permutation of its vari-
ables changes its values.
unit circle: The set of points in C of modulus 1.
unit disk: The set of points in C of modulus less than 1.
zero: x is called a zero of a function f if f (x) = 0.
List of special symbols
nf (r): the number of zeros of an entire function f in the disk B(0, r), 28
35
Index