Sei sulla pagina 1di 23

Chapter 2

Complex analysis

2.1 Basics
Given z0 ∈ C, r > 0 we define

D(z0 , r) := {z ∈ C : |z − z0 | < r} (open disk),


D(z0 , r) := {z ∈ C : |z − z0 | 6 r} (closed disk).

A subset U ⊆ C is called open if U = ∅ or if for every z0 ∈ C there is δ > 0 with


D(z0 , δ) ⊂ U . A subset U ⊆ C is called closed if U c := C \ U is open. It is known
that a subset of C is compact (i.e., every open cover of the set has a finite subcover)
if and only if it is closed and bounded (see the Prerequisites).
In what follows, U is a non-empty open subset of C and f : U → C a function.
We say that f is holomorphic or analytic in z0 ∈ U if
f (z) − f (z0 )
lim exists.
z→z0 z − z0
In that case, the limit is denoted by f 0 (z0 ). We say that f is analytic on U if f is
analytic in every z ∈ U ; in that case, the derivative f 0 (z) is defined for every z ∈ U .
We say that f is analytic around z0 if it is analytic on some open disk D(z0 , δ) for
some δ > 0. Finally, given a not necessarily open subset A of C and a function
f : A → C, we say that f is analytic on A if there is an open set U ⊇ A such that
f is defined on U and analytic on U . An everywhere analytic function f : C → C
is called entire.

35
For any two analytic functions f, g on some open set U ⊆ C,we have the usual
rules for differentiation (f ±g)0 = f 0 ±g 0 , (f g)0 = f 0 g+f g 0 and (f /g)0 = (gf 0 −f g 0 )/g 2
(the latter is defined for any z with g(z) 6= 0). Further, given a non-empty set U ⊆ C,
and analytic functions f : U → C, g : f (U ) → C, the composition g ◦ f is analytic
on U and (g ◦ f )0 = (g 0 ◦ f ) · f 0 .
Recall that a power series around z0 ∈ C is an infinite sum

X
f (z) = an (z − z0 )n
n=0

with an ∈ C for all n ∈ Z>0 . The radius of convergence of this series is given by
 −1
p
n
R = Rf = lim sup |an | .
n→∞

We state without proof the following fact.


P∞ n
Theorem 2.1. Let z0 ∈ C and f (z) = n=0 an (z − z0 ) a power series around
z0 ∈ C with radius of convergence R > 0. Then f defines a function on D(z0 , R)
which is analytic infinitely often. For k > 0 the k-th derivative f (k) of f has a power
series expansion with radius of convergence R given by
X∞
(k)
f (z) = n(n − 1) · · · (n − k + 1)an (z − z0 )n−k .
n=k

In each of the examples below, R denotes the radius of convergence of the given
power series.

z
X zn
e = , R = ∞, (ez )0 = ez .
n=0
n!

X z 2n
cos z = (eiz + e−iz )/2 = (−1)n , R = ∞, cos 0 z = − sin z.
n=0
(2n)!

iz −iz
X z 2n+1
sin z = (e − e )/2i = (−1)n , R = ∞, sin 0 z = cos z.
n=0
(2n + 1)!
∞  
X α
(1 + z)α = zn, R = 1, ((1 + z)α )0 = α(1 + z)α−1
n
n=0  
α α(α − 1) · · · (α − n + 1)
where α ∈ C, = .
n n!

X (−1)n−1 n
log(1 + z) = ·z , R = 1, log 0 (1 + z) = (1 + z)−1 .
n=1
n

36
2.2 Cauchy’s Theorem and some applications

For the necessary definitions concerning paths, line integrals and topology we re-
fer to the Prerequisites. In the remainder of this course, a path will always be a
piecewise continuously differentiable path. Recall that for a piecewise continuously
differentiable path γ, say γ = γ1 + · · · + γr where γ1 , . . . , γr are paths with continu-
ously differentiable parametrizations gi : [ai , bi ] → C, and for a continuous function
P Rb
f : γ → C we have γ f (z)dz = ri=1 aii f (gi (t))gi0 (t)dt.
R

Theorem 2.2 (Cauchy). Let U ⊆ C be a non-empty open set and f : U → C an


analytic function. Further, let γ1 , γ2 be two paths in U with the same start point and
end point that are homotopic in U . Then
Z Z
f (z)dz = f (z)dz.
γ1 γ2

Proof. Any textbook on complex analysis.

Corollary 2.3. Let U ⊆ C be a non-empty, open, simply connected set, and f :


U → C an analytic function. Then for any closed path γ in U ,
I
f (z)dz = 0.
γ

Proof. The path γ is homotopic in U to a point, and a line integral along a point is
0.

Corollary 2.4. Let γ1 , γ2 be two contours (closed paths without self-intersections


traversed counterclockwise), such that γ2 is contained in the interior of γ1 . Let
U ⊂ C be an open set which contains γ1 , γ2 and the region between γ1 and γ2 .
Further, let f : U → C be an analytic function. Then
I I
f (z)dz = f (z)dz.
γ1 γ2

Proof.

37
Let z0 , z1 be points on γ1 , γ2 respectively,
and let α be a path from z0 to z1 lying in-
side the region between γ1 and γ2 without
self-intersections.

Then γ1 is homotopic in U to the path α + γ2 − α, which consists of first traversing


α, then γ2 , and then α in the opposite direction. Hence
I Z I Z  I
f (z)dz = + − f (z)dz = f (z)dz.
γ1 α γ2 α γ2

Corollary 2.5 (Cauchy’s Integral Formula). Let γ be a contour in C, U ⊂ C an


open set containing γ and its interior, z0 a point in the interior of γ, and f : U → C
an analytic function. Then
I
1 f (z)
· dz = f (z0 ).
2πi γ z − z0

Proof.

Let γz0 ,δ be the circle with center z0


and radius δ, traversed counterclockwise.
Then by Corollary 2.4 we have for any suf-
ficiently small δ > 0,

I I
1 f (z) 1 f (z)
· dz = · dz.
2πi γ z − z0 2πi γz0 ,δ z − z0

38
Now, since f (z) is continuous, hence uniformly continuous on any sufficiently small
compact set containing z0 ,
I I
1 f (z) 1 f (z)
· dz − f (z ) = · dz − f (z )

0 0

2πi z − z 2πi z − z

γ 0 γz0 ,δ 0

Z 1 2πit

f (z 0 + δe ) 2πit

= 2πit
· δe dt − f (z0 )
0 δe
Z 1

f (z0 + δe ) − f (z0 ) dt 6 sup |f (z0 + δe2πit ) − f (z0 )|
2πit

=
0 06t61

→ 0 as δ ↓ 0.

This completes our proof.

We now show that every analytic function f on a simply connected set has an
anti-derivative. We first prove a simple lemma.

Lemma 2.6. Let U ⊆ C be a non-empty, open, connected set, and let f : U → C


be an analytic function such that f 0 = 0 on U . Then f is constant on U .

Proof. Fix a point z0 ∈ U and let z ∈ U be arbitrary. Take a path γz in U from z0


to z which exists since U is (pathwise) connected. Then
Z
f (z) − f (z0 ) = f 0 (w)dw = 0.
γz

Corollary 2.7. Let U ⊂ C be a non-empty, open, simply connected set, and f :


U → C an analytic function. Then there exists an analytic function F : U → C
with F 0 = f . Further, F is determined uniquely up to addition with a constant.

Proof (sketch). If F1 , F2 are any two analytic functions on U with F10 = F20 = f , then
F10 − F20 is constant on U since U is connected. This shows that an anti-derivative
of f is determined uniquely up to addition with a constant. It thus suffices to prove
the existence of an analytic function F on U with F 0 = f .

39
Fix z0 ∈ U . Given z ∈ U , we define F (z)
by Z
F (z) := f (w)dw,
γz

where γz is any path in U from z0 to z.


This does not depend on the choice of γz .
For let γ1 , γ2 be any two paths in U from
z0 to z. Then γ1 − γ2 (the path consisting

of first traversing γ1 and then γ2 in the opposite direction) is homotopic to z0 since


U is simply connected, hence
Z Z I
f (z)dz − f (z)dz = f (z)dz = 0.
γ1 γ2 γ1 −γ2

F (z+h)−F (z)
To prove that limh→0 h = f (z), take a path γz from z0 to z and then the
line segment [z, z + h] from z to z + h. Then since f is uniformly continuous on any
sufficiently small compact set around z,
Z Z  Z
F (z + h) − F (z) = − f (w)dw = f (w)dw
γz +[z,z+h] γz [z,z+h]
Z 1  Z 1 
= f (z + th)hdt = h f (z) + (f (z + th) − f (z))dt .
0 0

So
Z 1
F (z + h) − F (z)
− f (z) = (f (z + th) − f (z))dt
h
0

6 sup |f (z + th) − f (z)| → 0 as h → 0.
06t61

This completes our proof.

Example. Let U ⊂ C be a non-empty, open, simply connected subset of C with


0 6∈ U . Then 1/z has an anti-derivative on U .
For instance, if U = C \ {z ∈ C : Re z 6 0} we may take as anti-derivative of
1/z,
Log z := log |z| + iArg z,

40
where Arg z is the argument of z in the interval (−π, π) (this is called the principal
value of the logarithm).
On {z ∈ C : |z − 1| < 1} we may take as anti-derivative of 1/z the power series

X (z − 1)n
(−1)n−1 .
n=1
n

2.3 Taylor series


Theorem 2.8. Let U ⊆ C be a non-empty, open set and f : U → C an analytic
function. Further, let z0 ∈ U and R > 0 be such that D(z0 , R) ⊆ U . Then f has a
Taylor series expansion

X
f (z) = an (z − z0 )n converging for z ∈ D(z0 , R).
n=0

Further, we have for n ∈ Z>0 ,


I
1 f (z)
(2.1) an = · dz for any r with 0 < r < R.
2πi γz0 ,r (z − z0 )n+1

Proof. We fix z ∈ D(z0 , R) and use w to indicate a complex variable. Choose r with
|z − z0 | < r < R. By Cauchy’s integral formula,
I
1 f (w)
f (z) = · dw.
2πi γz0 ,r w − z

We rewrite the integrand. We have


 −1
f (w) f (w) f (w) z − z0
= = · 1−
w−z (w − z0 ) − (z − z0 ) w − z0 w − z0
∞   n ∞
f (w) X z − z0 X f (w)
= · = n+1
· (z − z0 )n .
w − z0 n=0 w − z0 n=0
(w − z0 )

The latter series converges uniformly on γz0 ,r . For let M := supw∈γz0 ,r |f (w)|. Then

M |z − z0 | n
 
f (w) n
sup
n+1
· (z − z0 ) 6 =: Mn
w∈γz0 ,r (w − z0 ) r r

41
P∞
and n=0 Mn converges since |z − z0 | < r. Consequently,
I
1 f (w)
f (z) = · dw
2πi γz0 ,r w − z
I ∞  
1 X f (w) n
= · (z − z0 ) dw
2πi γz0 ,r n=0 (w − z0 )n+1

( I )
X 1 f (w)
= (z − z0 )n · dw .
n=0
2πi γz0 ,r (w − z0 )n+1

Now Theorem 2.8 follows since by Corollary 2.4 the integral in (2.1) is independent
of r.

Corollary 2.9. Let U ⊆ C be a non-empty, open set, and f : U → C an analytic


function. Then f is analytic on U infinitely often, that is, for every k > 0 the k-the
derivative f (k) exists, and is analytic on U .

Proof. Pick z ∈ U . Choose δ > 0 such that D(z, δ) ⊂ U . Then for w ∈ D(z, δ) we
have
∞ I
X
n 1 f (w)
f (w) = an (w − z) with an = n+1
· dw for 0 < r < δ.
n=0
2πi γz,r
(w − z)

Now for every k > 0, the k-th derivative f (k) (z) exists and is equal to k!ak .

Corollary 2.10. Let γ be a contour in C, and U an open subset of C containing γ


and its interior. Further, let f : U → C be an analytic function. Then for every z
in the interior of γ and every k > 0 we have
I
(k) k! f (w)
f (z) = · dw.
2πi γ (w − z)k+1

Proof. Choose δ > 0 such that γz,δ lies in the interior of γ. By Corollary 2.4,
I I
1 f (w) 1 f (w)
k+1
· dw = · dw.
2πi γ (w − z) 2πi γz,δ (w − z)k+1

By the argument in Corollary 2.9, this is equal to f (k) (z)/k!.

42
We prove a generalization of Cauchy’s integral formula.
Corollary 2.11. Let γ1 , γ2 be two contours such that γ1 is lying in the interior of
γ2 . Let U ⊂ C be an open set which contains γ1 , γ2 and the region between γ1 , γ2 .
Further, let f : U → C be an analytic function. Then for any z0 in the region
between γ1 and γ2 we have
I I
1 f (z) 1 f (z)
f (z0 ) = dz − dz.
2πi γ2 z − z0 2πi γ1 z − z0

Proof. We have seen that around z0 the function f has a Taylor expansion f (z) =
P∞ n
n=0 an (z − z0 ) . Define the function on U ,

f (z) − a0
g(z) := (z 6= z0 ); g(z0 ) := a1 .
z − z0
The function g is clearly analytic on U \ {z0 }. Further,

g(z) − g(z0 ) X
= an (z − z0 )n−2 → a2 as z → z0 .
z − z0 n=2

Hence g is also analytic at z = z0 . In particular, g is analytic in the region between


γ1 and γ2 . So by Corollary 2.4,
I I
g(z)dz = g(z)dz.
γ1 γ2

Together with Corollaries 2.5, 2.4 this implies


I I
1 a0 1 a0
f (z0 ) = a0 = · dz − · dz
2πi γ2 z − z0 2πi γ1 z − z0
I I
1 f (z) 1 f (z)
= · dz − · dz.
2πi γ2 z − z0 2πi γ1 z − z0

2.4 Isolated singularities, Laurent series, mero-


morphic functions
We define the punctured disk with center z0 ∈ C and radius r > 0 by
D0 (z0 , r) = {z ∈ C : 0 < |z − z0 | < r}.

43
If f is an analytic function defined on D0 (z0 , r) for some r > 0, we call z0 an
isolated singularity of f . In case that there exists an analytic function g on the
non-punctured disk D(z0 , r) such that g(z) = f (z) for z ∈ D0 (z0 , r), we call z0 a
removable singularity of f .

Theorem 2.12. Let U ⊆ C be a non-empty, open set and f : U → C an analytic


function. Further, let z0 ∈ U , and let R > 0 be such that D0 (z0 , R) ⊆ U . Then f
has a Laurent series expansion

X
f (z) = an (z − z0 )n converging for z ∈ D0 (z0 , R).
n=−∞

Further, we have for n ∈ Z,


I
1 f (z)
(2.2) an = · dz for any r with 0 < r < R.
2πi γz0 ,r (z − z0 )n+1

Proof. We fix z ∈ D0 (z0 , R) and use w to denote a complex variable. Choose r1 , r2


with 0 < r1 < |z − z0 | < r2 < R.
By Corollary 2.11 we have
I I
1 f (w) 1 f (w)
(2.3) f (z) = · dw − · dw =: I1 − I2 ,
2πi γz0 ,r2 w − z 2πi γz0 ,r1 w − z

say. Completely similarly to Theorem 2.8, one shows that


∞ I
X
n 1 f (w)
I1 = an (z − z0 ) with an = · dw.
n=0
2πi γz0 ,r2 (w − z0 )n+1

Notice that for w on the inner circle γz0 ,r1 we have


 −1
f (w) f (w) f (w) w − z0
= =− · 1−
w−z (w − z0 ) − (z − z0 ) z − z0 z − z0
X∞
= − f (w)(z − z0 )−m−1 (w − z0 )m .
m=0

Further, one easily shows that the latter series converges uniformly to f (w)/(w − z)

44
on γz0 ,r1 . After a substitution n = −m − 1, it follows that


!
−1
I X
I2 = f (w)(w − w0 )m (z − z0 )−m−1 · dw
2πi γz0 ,r2 m=0
−1 I
X
n 1 f (w)
= − an (z − z0 ) , with an = · dw.
n=−∞
2πi γz0 ,r1 (w − z0 )n+1

By substituting the expressions for I1 , I2 obtained above into (2.3), we obtain



X
f (z) = I1 − I2 = an (z − z0 )n .
n=−∞

Further, by Corollary 2.4 we have


I
1 f (w)
an = · dw
2πi γz0 ,r (w − z0 )n+1

for any n ∈ Z and any r with 0 < r < R. This completes our proof.

Let U ⊆ C be an open set, z0 ∈ U and f : U \ {z0 } → C an analytic function.


Then z0 is an isolated singularity of f , and there is R > 0 such that f has a Laurent
series expansion
X∞
f (z) = an (z − z0 )n
n=−∞

converging for z ∈ D0 (z0 , R). Notice that z0 is a removable singularity of f if an = 0


for all n < 0.
The point z0 is called

• an essential singularity of f if there are infinitely many n < 0 with an 6= 0;

• a pole of order k of f for some k > 0 if a−k 6= 0 and an = 0 for n < −k; a pole
of order 1 is called simple;

• a zero of order k of f for some k > 0 if ak 6= 0 and an = 0 for n < k; a zero


of order 1 is called simple.

45
Notice that if f has a zero of order k at z0 then in particular, z0 is a removable
singularity of f and so we may assume that f is defined and analytic at z0 . Moreover,
z0 is a zero of order k of f if and only if f (j) (z0 ) = 0 for j = 0, . . . , k − 1, and
f (k) (z0 ) 6= 0.
For f, z0 as above, we define

ordz0 (f ) := smallest k ∈ Z such that ak 6= 0.

Thus,
z0 essential singularity of f ⇐⇒ ordz0 (f ) = −∞;
z0 pole of order k of f ⇐⇒ ordz0 (f ) = −k;
z0 zero of order k of f ⇐⇒ ordz0 (f ) = k.
Further, ordz0 (f ) = k if and only if there is a function g that is analytic around z0
such that f (z) = (z − z0 )k g(z) for z 6= z0 and g(z0 ) 6= 0.

Lemma 2.13. Let R > 0 and let f, g : D0 (z0 , R) → C be two analytic functions.
Assume that g 6= 0 on D0 (z0 , R), and that z0 is not an essential singularity of f or
g. Then

ordz0 (f + g) > min ordz0 (f ), ordz0 (g) ;
ordz0 (f g) = ordz0 (f ) + ordz0 (g);
ordz0 (f /g) = ordz0 (f ) − ordz0 (g).

Proof. Exercise.
You may compare this with the function ordp for prime numbers p defined in Chapter
1. Both ordz0 and ordp are examples of discrete valuations. A discrete valuation on
a field K is a surjective map v : K → Z ∪ {∞} such that v(0) = ∞; v(x) ∈ Z for
x ∈ K, x 6= 0; v(xy) = v(x) + v(y) for x, y ∈ K; and v(x + y) > min(v(x), v(y)) for
x, y ∈ K.

Let U ⊂ C be an open set. A set S called discrete in U in U if S ⊂ U and


for every compact set K with K ⊂ U , the intersection S ∩ K is finite. By the
Bolzano-Weierstrass Theorem, S is discrete in U if and only if S ⊂ U and S has
no limit point in U , that is, there are no z0 ∈ U and an infinite sequence {zn } of
distinct elements of S such that limn→∞ zn = z0 .

46
A meromorphic function on U is a complex function f with the following prop-
erties:
(i) there is a set S discrete in U such that f is defined and analytic on U \ S;
(ii) all elements of S are poles of f .
We say that a complex function f is meromorphic around z0 if f is analytic on
D0 (z0 , r) for some r > 0, and z0 is a pole of f .
It is easy to verify that if f, g are meromorphic functions on U then so are f + g
and f · g. Later we will see that if U is connected, and g is a meromorphic, non-
zero function on U , then the set of zeros of f is discrete in U . The zeros of g
are poles of 1/g, and the poles of g are zeros of 1/g. Hence 1/g is meromorphic
on U . Consequently, if U is an open, connected subset of C, then the functions
meromorphic on U form a field.

2.5 Residues, logarithmic derivatives


Let z0 ∈ C, R > 0 and let f : D0 (z0 , R) → C be an analytic function. Then f has
a Laurent series expansion converging on D0 (z0 , R):

X
f (z) = an (z − z0 )n .
n=−∞

We define the residue of f at z0 by


res(z0 , f ) := a−1 .
In particular, if f is analytic at z0 then res(z0 , f ) = 0. By Theorem 2.12 we have
I
1
res(z0 , f ) = f (z)dz
2πi γz0 ,r
for any r with 0 < r < R.
Theorem 2.14 (Residue Theorem). let γ be a contour in C. let z1 , . . . , zq be in the
interior of γ. Let f be a complex function that is analytic on an open set containing
γ and the interior of γ minus {z1 , . . . , zq }. Then
I q
1 X
f (z)dz = res(zi , f ).
2πi γ i=1

47
Proof. We proceed by induction on q. First let q = 1. Choose r > 0 such that γz1 ,r
lies in the interior of γ. Then by Corollary 2.4,
I I
1 1
f (z)dz = f (z)dz = res(z1 , f ).
2πi γ 2πi γz1 ,r

Now let q > 1 and assume the Residue


Theorem is true for fewer than q points.
We cut γ into two pieces, the piece γ1 from
a point w0 to w1 and the piece γ2 from w1
to w0 so that γ = γ1 + γ2 . Then we take a
path γ3 from w1 to w0 inside the interior
of γ without self-intersections; this gives
two contours γ1 + γ3 and −γ3 + γ2 .

We choose γ3 in such a way that it does not hit any of the points z1 , . . . , zq and both
the interiors of these contours contain points from z1 , . . . , zq . Without loss of gener-
ality, we assume that the interior of γ1 +γ3 contains z1 , . . . , zm with 0 < m < q, while
the interior of −γ3 + γ2 contains zm+1 , . . . , zq . Then by the induction hypothesis,
I I I
1 1 1
f (z)dz = f (z)dz + f (z)dz
2πi γ 2πi γ1 2πi γ2
I I
1 1
= f (z)dz + f (z)dz
2πi γ1 +γ3 2πi −γ3 +γ2
m q q
X X X
= res(zi , f ) + res(zi , f ) = res(zi , f ),
i=1 i=m+1 i=1

completing our proof.

We have collected some useful facts about residues. Both f, g are analytic func-
tions on D0 (z0 , r) for some r > 0.
Lemma 2.15. (i) f has a simple pole or removable singularity at z0 with residue α
α
⇐⇒ lim (z − z0 )f (z) = α ⇐⇒ f (z) − is analytic around z0 .
z→z0 z − z0
(ii) Suppose f has a pole of order 1 at z0 and g is analytic and non-zero at z0 . Then

res(z0 , f g) = g(z0 )res(z0 , f ).

48
(iii) Suppose that f is analytic and non-zero at z0 and g has a zero of order 1 at z0 .
Then f /g has a pole of order 1 at z0 , and

res(z0 , f /g) = f (z0 )/g 0 (z0 ).

Proof. Exercise.

Let U be a non-empty, open subset of C and f a meromorphic function on U


which is not identically zero. We define the logarithmic derivative of f by

f 0 /f.

Suppose that U is simply connected and f is analytic and has no zeros on U . Then
f 0 /f has an anti-derivative h : U → C. One easily verifies that (eh /f )0 = 0. Hence
eh /f is constant on U . By adding a suitable constant to h we can achieve that
eh = f . That is, we may view h as the logarithm of f , and f 0 /f as the derivative of
this logarithm. But we will work with f 0 /f also if U is not simply connected and/or
f has zeros or poles on U . In that case, we call f 0 /f also the logarithmic derivative
of f , although it need not be the derivative of some function.
The following facts are easy to prove: if f, g are two meromorphic functions on
U that are not identically zero, then
(f g)0 f0 g0 (f /g)0 f0 g0
= + , = − .
fg f g f /g f g
Further, if U is connected, then
f0 g0
= ⇐⇒ f = cg for some constant c.
f g
Lemma 2.16. Let z0 ∈ C, r > 0 and let f : D0 (z0 , r) → C be analytic. Assume
that z0 is either a removable singularity or a pole of f . Then z0 is a simple pole or
(if z0 is neither a zero nor a pole of f ) a removable singularity of f 0 /f , and

res(z0 , f 0 /f ) = ordz0 (f ).

Proof. Let ordz0 (f ) = k. This means that f (z) = (z − z0 )k g(z) with g analytic
around z0 and g(z0 ) 6= 0. Consequently,
f0 (z − z0 )0 g 0 k g0
=k + = + .
f z − z0 g z − z0 g

49
The function g 0 /g is analytic around z0 since g(z0 ) 6= 0. So by Lemma 2.15,
res(z0 , f 0 /f ) = k.

Corollary 2.17. Let γ be a contour in C, U an open subset of C containing γ and


its interior, and f a meromorphic function on U . Assume that f has no zeros or
poles on γ and let z1 , . . . , zq be the zeros and poles of f inside γ. Then
q
f 0 (z)
I
1 X
· dz = ordzi (f ) = Z − P,
2πi γ f (z) i=1

where Z, P denote the number of zeros and poles of f inside γ, counted with their
multiplicities.

Proof. By Theorem 2.14 and Lemma 2.16 we have


q q
f 0 (z)
I
1 X
0
X
· dz = res(zi , f /f ) = ordzi (f ) = Z − P.
2πi γ f (z) i=1 i=1

2.6 Unicity of analytic functions


In this section we show that two analytic functions f, g defined on a connected open
set U are equal on the whole set U , if they are equal on a sufficiently large subset
of U .
We start with the following result.

Theorem 2.18. Let U be a non-empty, open, connected subset of C, and f : U → C


an analytic function. Assume that f = 0 on an infinite subset of U having a limit
point in U . Then f = 0 on U .

Proof. Our assumption that U is connected means, that any non-empty subset S of
U that is both open and closed in U , must be equal to U .
Let Z be the set of z ∈ U with f (z) = 0. Let S be the set of z ∈ U such that z is
a limit point of Z. By assumption, S is non-empty. Since f is continuous, we have
S ⊆ Z. Any limit point in U of S is therefore a limit point of Z and so it belongs to

50
S. Hence S is closed in U . We show that S is also open; then it follows that S = U
and we are done.
Pick z0 ∈ S. We have to show that there is δ > 0 such that D(z0 , δ) ⊂ S. There
is δ > 0 such that f has a Taylor expansion

X
f (z) = an (z − z0 )n
n=0

converging on D(z0 , δ). Assume that f is not identically 0 on D(z0 , δ). Then not
all coefficients an are 0. Assume that am 6= 0 and an = 0 for n < m, say. Then
f (z) = (z − z0 )m h(z) with h(z) = ∞ n−m
P
n=m an (z − z0 ) . Since h(z0 ) = am 6= 0 and
h is continuous, there is δ1 > 0 such that h(z) 6= 0 for all z ∈ D(z0 , δ1 ). But then
f (z) 6= 0 for all z with 0 < |z − z0 | < δ1 , contradicting that z0 ∈ S.
Hence f is identically 0 on D(z0 , δ). Clearly, every point of D(z0 , δ) is a limit
point of D(z0 , δ), hence of Z. So D(z0 , δ) ⊂ S. This shows that indeed, S is
open.

Corollary 2.19. Let U be a non-empty, open, connected subset of C, and let f :


U → C be an analytic function that is not identically 0 on U . Then the set of zeros
of f in U is discrete in U , i.e., every compact subset of U contains only finitely
many zeros of f .

Proof. Suppose that some compact subset of U contains infinitely many zeros of f .
Then by the Bolzano-Weierstrass Theorem, the set of these zeros would have a limit
point in this compact set, implying that f = 0 on U .

Corollary 2.20. Let U be a non-empty, open, connected subset of C, and f, g :


U → C two analytic functions. Assume that f = g on an infinite subset of U having
a limit point in U . Then f = g on U .

Proof. Apply Theorem 2.18 to f − g.

Let U, U 0 be open subsets of C with U ⊂ U 0 . Let f : U → C be an analytic


function. An analytic continuation of f to U 0 is an analytic function g : U 0 → C
such that g(z) = f (z) for z ∈ U . It is often a difficult problem to figure out whether
such an analytic continuation exists. The next Corollary shows that if it exists, it
must be unique.

51
Corollary 2.21. Let U, U 0 be non-empty, open subsets of C, such that U ⊂ U 0 and
U 0 is connected. Let f : U → C be an analytic function. Then f has at most one
analytic continuation to U 0 .

Proof. . Let g1 , g2 be two analytic continuations of f to U 0 . Then g1 (z) = g2 (z) =


f (z) for z ∈ U . Since U has a limit point in U 0 , it follows that g1 (z) = g2 (z) for
z ∈ U 0.

Another consequence of Theorem 2.18 is the so-called Schwarz’ reflection priciple,


which states that analytic functions assuming real values on the real line have nice
symmetric properties.
Corollary 2.22 (Schwarz’ reflection principle).
Let U be an open, connected subset of C,
such that U ∩ R 6= ∅ and such that U is
symmetric about R, i.e., z ∈ U for every
z ∈ U . Further, let f : U → C be a non-
identically zero analytic function with the
property that

{z ∈ U ∩ R : f (z) ∈ R}

has a limit point in U .


Then f has the following properties:
(i) f (z) ∈ R for z ∈ U ∩ R;
(ii) f (z) = f (z) for z ∈ U ;
(iii) If z0 and r > 0 are such that D0 (z0 , r) ⊂ U , then ordz0 (f ) = ordz0 (f ).

Proof. We first show that the function z 7→ f (z) is analytic on U . Indeed, for
z0 ∈ U , the limit
 
f (z) − f (z0 ) f (z) − f (z0 )
lim = lim = f 0 (z0 )
z→z0 z − z0 z→z0 z − z0
exists.
Notice that for every z ∈ U ∩ R with f (z) ∈ R, we have f (z) = f (z). So by our
assumption on f , the set of z ∈ U with f (z) = f (z) has a limit point in U . Now
Corollary 2.20 implies that f (z) = f (z) for z ∈ U . This implies (i) and (ii).

52
We finish with proving (iii). Our assumption implies that f has a Laurent series
expansion
X∞
f (z) = an (z − z0 )n
n=−∞

converging on D0 (z0 , r). Then for z ∈ D0 (z0 , r) we have z ∈ D0 (z0 , r) and


! ∞
X X
f (z) = f (z) = an (z − z0 )n = an (z − z0 )n ,
n=−∞ n=−∞

which clearly implies (iii).

2.7 Analytic functions defined by integrals


In analytic number theory, quite often one has to deal with complex functions that
are defined by infinite series, infinite products, infinite integrals, or even worse,
infinite integrals of infinite series. In this section we have collected some useful
results that allow us to verify in a not too difficult manner that such complicated
functions are analytic.
We start with a general theorem on analytic functions defined by an integral.
Since we couldn’t find a good reference for a result of this type, we have included a
proof.

Theorem 2.23. Let D be a measurable subset of Rm , U an open subset of C and


f : D × U → C a function with the following properties:
(i) f is measurable on D × U (with U viewed as subset of R2 );
(ii) for every fixed x ∈ D, the function z 7→ f (x, z) is analytic on U ;
(iii) for every compact subset K of U there is a measurable function MK : D → R
such that
Z
|f (x, z)| 6 MK (x) for x ∈ D, z ∈ K, MK (x)dx < ∞.
D

Then the function F given by


Z
F (z) := f (x, z)dx
D

53
is analytic on U , and for every k > 1,
Z
(k)
F (z) = f (k) (x, z)dx,
D

where f (k) (x, z) denotes the k-th derivative with respect to z of the analytic function
z 7→ f (x, z).

Proof. Fix z ∈ U . Choose r > 0 such that D(z, r) ⊂ U , and let 0 < δ < 21 r. We
show that for w ∈ D(z, δ), F (w) can be expanded into a Taylor series around z; then
it follows that F is analytic on D(z, δ) and so in particular in z. Let M (x) : D → R
be a measurable function such that |f (x, w)| 6 M (x) for x ∈ D, w ∈ D(z, r) and
R
D
M (x)dx < ∞.
For w ∈ D(z, δ) we have by Corollary 2.5,
Z Z ( I )
1 f (x, ζ)
F (w) = f (x, w)dx = · dζ dx.
D D 2πi γz,2δ ζ − w
By inserting
 −1
f (x, ζ) f (x, ζ) f (x, ζ) w−z
= = 1−
ζ −w (ζ − z) − (w − z) ζ −z ζ −z

X f (x, ζ)
= · (w − z)n
n=0
(ζ − z)n+1
we obtain
Z ( I ∞
! )
1 X f (x, ζ)
F (w) = n+1
(w − z)n dζ dx
D 2πi γz,2δ n=0
(ζ − z)
Z (Z 1 ∞ 2πit
! )
X f (x, z + 2δe )
= 2πit n
(w − z)n dt dx.
D 0 n=0
(2δe )
We apply the Fubini-Tonelli Theorem (see the Prerequisites). Note that since
|w − z| < δ,
Z (Z 1 X ∞ ! )
f (x, z + 2δe2πit )

2πit n
(w − z)n dt dx
D 0 n=0
(2δe )
Z (Z 1 X ∞
! ) Z
−n
6 M (x)2 dt dx 6 2M (x)dx < ∞.
D 0 n=0 D

54
So the conditions of the Fubini-Tonelli Theorem are satisfied, and in the expression
for F (w) derived above we can interchange the integrations and the summation.
Performing this interchange and then applying Corollary 2.10, we obtain
∞ 1
f (x, z + 2δe2πit )
X Z Z  
n
F (w) = (w − z) dt dx
n=0 D 0 (2δe2πit )n
∞ Z ( I ) !
X 1 f (x, ζ)
= (w − z)n · dζ dx
n=0 D 2πi γz,2δ (ζ − z)n+1

f (n) (x, z)
X Z 
n
= (w − z) · dx .
n=0 D n!

This shows that indeed, F (w) has a Taylor expansion around z converging on
D(z, δ). So in particular, F is analytic in z. Further, F (k) (z) is equal to k! times
R
the coefficient of (w − z)k , that is, D f (k) (x, z)dx. This proves our Theorem.

We deduce a result, which states that under certain conditions, the pointwise
limit of a sequence of analytic functions is again analytic.

Theorem 2.24. Let U ⊂ C be a non-empty open set, and {fn : U → C}∞ n=0 a
sequence of analytic functions, converging pointwise to a function f on U . Assume
that for every compact subset K of U there is a constant CK < ∞ such that

|fn (z)| 6 CK for all z ∈ K, n > 0.


(k)
Then f is analytic on U , and fn → f (k) pointwise on U for all k > 1.

Proof. The set U can be covered by disks D(z0 , δ) with z0 ∈ U , δ > 0, such that
D(z0 , 2δ) ⊂ U . We fix such a disk D(z0 , δ) and prove that f is analytic on D(z0 , δ)
(k)
and fn → f (k) pointwise on D(z0 , δ) for k > 1. This clearly suffices.
Let z ∈ D(z0 , δ), k > 0. Then by Corollary 2.10, we have
I
(k) k! fn (ζ)
fn (z) = · dζ
2πi γz0 ,2δ (ζ − z)k+1
Z 1 Z 1
fn (z0 + 2δe2πit )2δe2πit
= k! · dt = k! gn,k (t, z)dt,
0 (z0 + 2δe2πit − z)k+1 0

55
say. By assumption, there is C < ∞ such that |fn (w)| 6 C for w ∈ D(z0 , 2δ), n > 0.
Further, for t ∈ [0, 1] we have |z0 + 2δe2πit − z| > δ. Hence

(2.4) |gn,k (t, z)| 6 C · 2δ/δ k+1 = 2Cδ −k for n, k > 0.

Notice that for k > 0, t ∈ [0, 1], z ∈ D(z0 , δ) we have


f (z0 + 2δe2πit )2δe2πit
gn,k (t, z) → gk (t, z) := .
(z0 + 2δe2πit − z)k+1
Thanks to (2.4) we can apply the dominated convergence theorem, and obtain for
every fixed z ∈ D(z0 , δ), k > 0,
Z 1
(k)
fn (z) → k! gk (t, z)dt.
0

Applying this with k = 0 and using fn → f pointwise, we obtain


Z 1
f (z) = g0 (t, z)dt for z ∈ D(z0 , δ).
0

It follows from Theorem 2.23 that the right-hand side, and hence f , is analytic on
D(z0 , δ), and moreover,
Z 1 Z 1
(k) (k)
f (z) = g0 (t, z)dt = k! gk (t, z)dt = lim fn(k) (z) for z ∈ D(z0 , δ), k > 1,
0 0 n→∞

(k)
where g0 is the k-th derivative of g0 with respect to z. Indeed, g0 (t, z) is measurable
on [0, 1] × D(z0 , δ) and for every fixed t, the function z 7→ g0 (t, z) is analytic on
D(z0 , δ). Further, by (2.4) and since gn,0 (t, z) → g0 (t, z), we have |g0 (t, z)| 6 2C for
t ∈ [0, 1], z ∈ D(z0 , δ). So all conditions of Theorem 2.23 are satisfied.
This completes our proof.
Corollary 2.25. Let U ⊂ C be a non-empty open set, and {fn : U → C}∞ n=0
a sequence of analytic functions, converging to a function f pointwise on U , and
uniformly on every compact subset of U .
(k)
Then f is analytic on U and fn → f (k) pointwise on U for every k > 1.

Proof. Take a compact subset K of U . Let ε > 0. Then there is N such that
|fn (z) − fm (z)| < ε for all z ∈ K, m, n > N . Choose m > N . Then there is C > 0
such that |fm (z)| 6 C for z ∈ K since fm is continuous. Hence |fn (z)| 6 C + ε for
z ∈ K, n > N . Now our Corollary follows at once from Theorem 2.24.

56
Corollary 2.26. let U ⊂ C be a non-empty open set, and {fn : U → C}∞ n=0
a sequence of analytic functions, converging to a function f pointwise on U and
uniformly on every compact subset of U . Then

fn0 (z) f 0 (z)


lim =
n→∞ fn (z) f (z)

for all z ∈ U with f (z) 6= 0, where the limit is taken over those n for which fn (z) 6= 0.

Proof. Obvious.

Corollary 2.27. Let U ⊂ C be a non-empty open set and {fn : U → C}∞ n=0 a
sequence of analytic functions. Assume that for every compact subset K of U there
are reals Mn,K such that |fn (z)| 6 Mn,K for z ∈ K and ∞
P
n=0 Mn,K converges. Then
  (k)
(i) ∞
P∞
= ∞
P P (k)
n=0 fn is analytic on U , and n=0 fn n=0 fn for k > 0,
Q∞
(ii) n=0 (1 + fn ) is analytic on U .

Proof. Our assumption on the functions fn implies that both the series ∞
P
Q∞ n=0 fn and
the infinite product n=0 (1 + fn ) converge uniformly on every compact subset of U
(see the Prerequisites). Now apply Corollary 2.25.

Corollary 2.28. Let U , {fn }∞


n=0 be as in Corollary 2.27 and assume in addition
Q∞
that fn 6= −1 on U for every n > 0. Then for the function F = n=0 (1 + fn ) we
have ∞
F 0 X fn0
= .
F n=0
1 + fn

Proof. Let Fm := m
Q
n=0 (1 + fn ). Then Fm → F uniformly on every compact subset
of U . Hence by Corollary 2.26,
m
F0 Fm0 X fn0
= lim = lim
F m→∞ Fm m→∞
n=0
1 + fn

which clearly implies Corollary 2.28.

57

Potrebbero piacerti anche