Está en la página 1de 31

The ring of entire functions

Jonas Bjermo
U.U.D.M. Project Report 2004:15
Examensarbete i matematik, 20 pong
Handledare och examinator: Karl-Heinz Fieseler
Juni 2004
Department of Mathematics
Uppsala University
The ring of entire functions.
Examensarbete
Matematiska Institutionen
Uppsala Universitet
Jonas Bjermo
August 19, 2004
1
Contents
1 Introduction 2
2 Holomorphic and meromorphic functions 2
2.1 Holomorphic functions 2
2.2 Meromorphic functions 5
3 Preparation for the ideal theory in O(G) 8
3.1 Ideals 8
3.2 Convergent series and sequences of complex functions 9
3.3 Innite products of complex functions 11
3.4 Mittag-Leer series 13
3.5 Greatest common divisors 15
3.6 Wedderburns lemma 16
4 Ideal structure in O(G) 17
4.1 Principal ideal theorem 17
4.2 Fixed and free ideals 19
4.3 Maximal ideals 21
5 Prime ideals 26
6 References 29
2
Acknowledgment
I am most grateful to my supervisor Karl-Heinz Fieseler for presenting the
idea of this thesis and for answering all my questions about this thesis and
related subjects.
1 Introduction
In 1940, O. Helmer [5] proved that, in the ring of all entire functions, every
nitely generated ideal is a principal ideal. In this text we show, based on [6,
chapter 6.3], that this is also valid in rings of functions holomorphic in arbi-
trary open connected subsets G C denoted O(G). To prove this, several
tools from complex function theory are used, such as the concept of greatest
common divisors which depends on Weierstra products, and Mittag-Leer
series which help us to prove the important lemma of Wedderburn.
Then, as in [3] and [4], we consider the ideals in the ring O(G) with
the dierence that we use the concept of divisors to describe them. We
introduce so called lters and deal with them instead of the distributions
of zeros for the functions in the ideal. Both maximal and prime ideals are
considered. We will see that the factor ring O(G)/m, where m is maximal
is isomorphic to C both when the functions in m have at least one common
zero and when they have no common zeros.
2 Holomorphic and meromorphic functions
2.1 Holomorphic functions
We will start to recall some results from complex analysis. A domain is a
non-empty, open subset of C. From now on it will always be denoted D. A
connected domain is called a region and will be denoted G in the rest of the
text.
Denition A function f: D C is called holomorphic in the domain D
if f is complex-dierentiable at every point of D; f is called holomorphic
at c D if there is an open neighborhood U of c lying in D such that the
restriction f[
U
of f to U is complex dierentiable in U.
Remark The set of all points at which a function is holomorphic, is al-
ways open in C. A function which is holomorphic at c is always complex-
dierentiable at c, but the converse do not need to be true.
3
The set of all function holomorphic in D is denoted O(D). Let us recall
from complex analysis that for a power series

n=0
a
n
(z c)
n
there exists
a unique number R 0 (possibly R = +), such that if [z c[ < R, the
series converges, and if [z c[ > R, the series diverges. The convergence
is uniform and absolute on every closed disc in A = z C: [z c[ < R.
R is called its circle of convergence. We also know that a power series

n=0
a
n
(z c)
n
is holomorphic inside its radius of convergence. On the
other hand we recall that a function holomorphic on a domain D equals the
Taylor series

n=0
f
(n)
(c)
n!
(z c)
n
on every open disc U centered at c and
contained in D. That is, f is holomorphic on a domain D if and only if f
equals a convergent power series around c on some open disc B
r
(c) D.
Denition Let f O(D), then the set Z(f) = z C: f(z) = 0 is called
its zero-set.
Denition If M is any subset of O(G), let Z(M) = Z(f)[f M.
The zero-set Z(f) of a function f ,= 0 holomorphic in a region G is a
discrete and relatively closed (and hence nite or countably innite) subset
of G [7 p. 232].
Now we dene the order of a zero of a holomorphic function.
Denition Let f O(D) and let c D, then f has a zero of order m at c
if f(c) = 0, ..., f
m1
(c) = 0, f
m
(c) ,= 0.
The function f being holomorphic means that we can expand it in a
Taylor series f(z) =

n=0
f
(n)
(c)
n!
(z c)
n
. Then we see that f has a zero
of order m if and only if in a neighborhood of c, we can write f(z) =
(z c)
m
h(z) where h is holomorphic at c, h(z) =

m
f
()
(c)
!
(z c)
m
and h(c) = f
(m)
(c)/m! ,= 0.
The order is denoted o
c
(f) and we set o
c
(f) := if f equals zero near
c.
We have a rule for computation of the order.
Theorem 2.1 (Product rule) If the functions f and g are holomorphic near
c, then o
c
(fg) = o
c
(f) + o
c
(g).
Proof Let o
c
(f) = n and o
c
(g) = m. Using the fact that n + = , the
denition of the order and Taylors theorem, we get that o
c
(fg) = m + n =
o
c
(f) + o
c
(g).
It is easy to see that the set O(D) becomes a commutative ring with
unity with the pointwise addition and multiplication. The units in O(D)
are exactly the nowhere vanishing functions. We will now see that if D is
connected, which means that D is a region in C then the ring O(D) is an
integral domain.
4
Theorem 2.2 (Identity Theorem [7 p. 228]) The following statements about
a pair f, g of holomorphic functions in a region G C are equivalent:
1. f = g.
2. The coincidence set w G[f(w) = g(w) has a cluster point in G.
3. There is a point c G such that f
(n)
(c) = g
(n)
(c) for all n N.
Proof 1) 2) is trivial.
2) 3) Set h := f g O(G). Then by the hypothesis the zero-
set M := w G[h(w) = 0 has a cluster point c G. If there is an
m N with h
(m)
(c) ,= 0 then we consider the smallest such m. From
above we see that we have the factorization h(z) = (z c)
m
h
m
(z) with
h
m
(z) =

m
h
()
(c)
!
(z c)
m
O(B) for all open balls B centered at
c in G and h
m
(c) ,= 0. Because of its continuity h
m
is zero free in some
neighborhood U B of c. Then M (U c) = , that is c is not a
cluster point of M. This contradiction shows that there is no such m, that
is h
(n)
(c) = 0 for all n N, hence f
(n)
= g
(n)
for all n N.
3) 1) Set h := f g, and let S
k
= Z(h
(k)
. Each S
k
is relatively closed
in G because h
(k)
O(G) is continuous. Then S :=

0
S
k
is relatively
closed in G. However, if z
1
S then the Taylor series of h in any open
B C centered at z
1
equals the zero series. Hence h
(k)
[
B
= 0 for every
k N, that is B S. This implies that S is open in G. Since G is a region
(hence connected), G is the only non-empty subset of G which is both open
and closed. Hence S = G, and therefore f = g.
A consequence of the Identity Theorem is
Theorem 2.3 The domain D C is connected if and only if the ring O(D)
is an integral domain.
Proof ) Suppose f, g O(D), f is not the zero function but fg is, that
is f(z)g(z) = 0 for all z D. There is some c D where f(c) ,= 0 and
a neighborhood U D of c in which f is zero-free. Then g(U) = 0 and
since D is a region (hence connected) this means that g(D) = 0 by the
Identity Theorem. That is, g = 0, the zero element of the ring O(D).
) If D is not a region then it would not be connected. It would be
possible to express D as the disjoint union of the non-empty open sets D
1
and D
2
. Let us dene f, g in D by
f(z) := 0 for z D
1
and f(z) := 1 for z D
2
,
g(z) := 1 for z D
1
and g(z) := 0 for z D
2
.
These functions are holomorphic in D, and neither is the zero function
in O(D) but fg is. This contradicts the hypothesis that O(D) is an integral
domain.
5
2.2 Meromorphic functions
The Taylor expansion does not apply to functions that fail to be holomorphic
at some points. For such functions we have another expansion called the
Laurent expansion. Let r
1
0, r
2
> r
1
, and c C, and consider the region
A = z C: r
1
< [z c[ < r
2
. Let us recall from complex analysis that if a
function f is holomorphic in A, it can be written f(z) =

n=0
a
n
(z c)
n
+

n=1
b
n
(zc)
n
where both series on the right side of the equation converge
absolutely on A and uniformly on any set of the form B

1
,
2
= z C:
1

[z c[
2
where r
1
<
1
<
2
< r
2
. The series is called the Laurent
series or the Laurent expansion around c in A. The series

n=0
a
n
(z c)
n
,
respectively,

n=1
b
n
(zc)
n
is called its regular, respectively, principal part.
When r
1
= 0 in the Laurent expansion we have a special case. In this
case, f is holomorphic in z : 0 < [z c[ < r
2
, and we say that c is an
isolated singularity of f.
Denition Let f be holomorphic in a domain D except for one point c D,
i.e., f is holomorphic in D c, then c is called an isolated singularity of f.
As seen above, if f is holomorphic in z : 0 < [z c[ < r
2
, then we can
expand f in a Laurent series: f(z) = ... +
b
n
(zc)
n
+... +
b
1
(zc)
+a
o
+a
1
(z c)+
a
2
(z c)
2
+ ... valid for 0 < [z c[ < r
2
. There are three dierent types of
isolated singularities depending on the number of coecients b
n
,= 0. If all
the coecients b
n
,= 0 is zero then we say that c is a removable singularity,
and if the number of b
n
are innite c is called an essential singularity. We
are interested in the case when all but a nite number of the coecients b
n
is zero.
Denition If c is an isolated singularity of f and if all but a nite number
of the b
n
in the Laurent expansion are zero, then c is called a pole of f. The
highest integer k such that b
k
,= 0, is called the order of the pole.
Denition A meromorphic function on D is a pair (f, P), where P D is
discrete and f: D P C is holomorphic with a pole at each point of P.
The set P is called the pole-set of f. Usually we write f instead of (f, P)
and P = P(f).
The set of all function meromorphic in a domain D is denoted /(D).
Remark Because the pole-set is allowed to be empty, the functions holo-
morphic in D are also meromorphic in D.
The pole-set of each function f ,= 0 meromorphic in D is, just as for the
zero-set of a holomorphic function, a discrete and relatively closed subset of
6
D. Then it follows that the pole-set of each function meromorphic in D is
either empty, nite, or countably innite [7 p. 316].
From complex analysis it is known that poles (locally) arise via the
formation of reciprocals of holomorphic functions.
Proposition 2.4 Let f O(Dc) and let c D be an isolated singularity
of f, then c is a pole of order m if and only if there exists a function g
O(D), g(c) ,= 0, such that f(z) =
g(z)
(zc)
m
for z D c.
Proof ) By denition c is a pole i f(z) =
b
k
(zc)
k
+
b
k1
(zc)
k1
+ ... +
b
1
(zc)
+

n=0
a
n
(z c)
n
=
1
(zc)
k
(b
k
+ b
k1
(z c) + ... + b
1
(z c)
k1
+

n=0
a
n
(z c)
n+k
), where b
k
,= 0. This is valid in D c. Let g(z) =
b
k
+ b
k1
(z c) + ... + b
1
(z c)
k1
+

n=0
a
n
(z c)
n+k
, this implies that
g O(D) and g(c) = b
k
,= 0.
) Let g O(D) where g(z) =

n=0
a
n
(z c)
n
and set f =
g(z)
(zc)
n
.
Then f =

n=0
a
n
(z c)
nm
, and by denition, it has a pole at c of order
m 1.
Theorem 2.5 Let f O(D c), then f has a zero of order m at c if and
only if
1
f(z)
has a pole of order m at c. If

f O(D) and

f(c) ,= 0 then

f(z)
f(z)
also has a pole of order m at c.
Proof The function f O(D c) has a zero of order m i f(z) = (z
c)
m
h(z), where h(z) is holomorphic at c with h(c) ,= 0. Set g(z) :=
1
h(z)
,
g O(D), g ,= 0. Then f(z) = (z c)
m 1
g(z)
, by prop 2.4 this is equivalent
to that
1
f(z)
has a pole of order m at c.
As we already know, a function f with a pole of order m at c can be
written f(z) =

1
n=m
b
n
(zc)
n
+

n=0
a
n
(zc)
n
. We can then see that to
ask when f has a pole of order m at c is the same as asking when (z c)
m
f
is bounded.
Theorem 2.6 Let m N, m 1 and let f O(D c). Then f has
a pole of order m at c if and only if there is a neighborhood U of c lying
in D and positive nite constants M

, M

such that for all z U c


M

[z c[
m
[f(z)[ M

[z c[
m
.
Proof ) The function f has a pole of order m at c i h :=
1
f
has a
zero of order m at c, which can be written in the form (z c)
m

h for an

h O(U), U D small enough. Let M

= inf
zU
[

h(z)[
1
> 0 and
M

= sup
zU
[

h(z)[
1
< . The claim now follows from [f(z)[ = [z
c[
m
[

h(z)[
1
.
) [(z c)
m
f(z)[ M

for z U c shows that (z c)


m
f is bounded
near c and [(z c)
m1
f(z)[ M

[z c[
1
shows that (z c)
m1
f is not
bounded near c. That is, c is a pole of order m.
7
Let f O(D). We say that f increases uniformly to around c, written
lim
zc
f(z) = , if for every nite M there is a neighborhood U of c in D
such that inf
zU\{c}
[f(z)[ M
Then a consequence of theorem 2.6 is the following
Corollary 2.7 The function f O(D c) has a pole at c if and only if
lim
zc
f(z) = .
In view of this corollary, at each pole, we choose f(z) := for z P(f).
So meromorphic functions are special mappings from D C to C .
If f, g /(D) with pole-sets P(f), P(g) are given, then P(f) P(g) is
also discrete and relatively closed in G. In GP(f) P(g) both f and g are
holomorphic, and hence f +g, fg are holomorphic. For each c P(f)P(g),
f and g can be written as Laurent expansions with nite principal part in an
open neighborhood U of c, with U (P(f) P(g)) = c. Then both f +g
and fg can be written as Laurent expansions with nite principal part. The
point c is then either a pole or a removable singularity of f +g and fg. Thus
the pole-set of these functions are subsets of P(f) P(g), which are discrete
and relatively closed in G. This implies that f + g, fg /(D). From
the rules of calculating with holomorphic functions it follows that /(D)
is a commutative ring with unity with respect to pointwise addition and
multiplication. The ring O(D) is a subring of /(D).
In the ring O(D), division of an element f is possible only when the
value of f is zero-free in D. But in the ring /(D), division of functions
which have zeros in D is possible.
Now we dene the zero-set of a meromorphic function.
Denition Let f /(D), then Z(f) is called the zero-set of f if Z(f) is
the zero-set of the holomorphic function f[(D P(f)) O(D P(f)).
Remark Z(f) is relatively closed in G and Z(f) P(f) = .
Now we dene the order function for a meromorphic function. If a func-
tion f ,= 0 is meromorphic, then it can be developed uniquely into a Laurent
series

m
a

(z c)

with a

C, m Z and a
m
,= 0.
Denition If f ,= 0, f /(D) the number m in the Laurent series above
is called the order of f at c and is denoted o
c
(f).
From the denition it follows directly that for an f meromorphic at c:
1. f holomorphic at c o
c
(f) 0
2. If m = o
c
(f) < 0, then c is a pole of f of order m.
As for holomorphic functions the product rule 2.1 is also valid for mero-
morphic functions.
Now we prove that the ring of all functions meromorphic in a region is
a eld.
8
Theorem 2.8 Let u /(D), then u is invertible, that is uv = 1 for some
v /(D), if and only if the zero-set Z(u) is discrete in D.
Proof ) For c D, uv = 1 implies that u(c) = 0 i v(c) = and
u(c) = i v(c) = 0. This means that Z(u) = P(v) and P(u) = Z(v), then
by the denition of a meromorphic function the zero-set Z(u) is discrete.
) The set Z(u) P(u) is discrete and relatively closed in G. Choose
v := 1/u, then v is holomorphic in G Z(u) P(u) and have a pole at every
point of Z(u). Every point c P(u) is a removable singularity of v because
lim
zc
1/u(z) = 0. Hence v /(D)
From this theorem we see that the quotient of two elements f, g /(D)
exists in the ring /(D) if Z(g) is discrete in D.
A consequence of theorem 2.8 is
Corollary 2.9 Let G C be a region, then /(G) is a eld.
Proof If f /(G), f ,= 0 and G is a region, then G P(f) is a region
and f[(G P(f)) is a holomorphic function which is not the zero element
of O(G). Therefore Z(f) is discrete in G and then f is invertible. Hence,
every element of /(G) 0 is invertible and therefore /(G) is a eld.
Because /(G) is a eld it contains no proper ideals ,= 0. The ring
O(G) on the other hand, has an interesting ideal structure.
3 Preparation for the ideal theory in O(G)
To prove some important result about the ideal structure in the integral
domain O(G) we make use of some important tools. These are the lemma
of Wedderburn and the concept of greatest common divisors. In the proof
of Wedderburns lemma we make use of Mittag-Leer series which will also
be considered. We start with ideals.
3.1 Ideals
Since we are going to consider the ideal structure in the ring of functions
holomorphic in open connected subsets of C we give some basic denitions.
In this subsection R is a commutative ring with unity.
Denition Let R be a ring. A subset I R, I ,= is called an ideal in R
if I is an additive subgroup and ax I for all a I, x R.
Now, let M ,= be any subset of R and L be the set of all nite linear
combinations i.e. L = u =

n
i=1
r
i
f
i
: r
i
R, f
i
M. Then L is an ideal
in R and M is said to generate L.
9
Denition An ideal I in a ring R is called nitely generated if there is a
nite set that generates I. A ring R is called Noetherian if every ideal in R
is nitely generated.
If we can choose M with only one element, we have a special case.
Denition An ideal I in a ring R is called principal if it is generated by an
element f, i.e. if I = rf : r R for some f R. A integral domain R is
called a principal ideal domain if every ideal is a principal ideal.
Remark A nitely generated ideal generated by M = f
1
, ..., f
r
is usually
denoted I = Rf
1
+... +Rf
n
, and a principal ideal generated by f is usually
denoted (f) or Rf.
Denition An ideal I in a ring R is called maximal provided that, I ,= R
and there does not exist an ideal J ,= R such that I J.
Denition An ideal I ,= R is called a prime ideal if ab I implies that
either a I or b I for a, b R.
3.2 Convergent series and sequences of complex functions
Let X be a metric space.
Denition A sequence of functions f
n
: X C is called uniformly conver-
gent in A X to f: A C if for every > 0 there exists an n

N such
that [f
n
(x) f(x)[ < for all n n

and all x A.
A series

of functions converges uniformly in A if the sequence s


n
=

n
f

of partial sums converges uniformly in A.


Remark The limit functions f and

f

are uniquely determined.


We introduce the supremum semi-norm [f[
A
:= sup[f(x)[: x A
for subsets A X and functions f: X C. The set V := f: X
C: [f[
A
< is a C-vector space and the mapping f [f[
A
fullls:
1. [f[
A
= 0 f[
A
= 0.
2. [cf[
A
= [c[[f[
A
.
3. [f + g[
A
[f[
A
+[g[
A
. for all f, g V , c C.
We see that the sequence f
n
converges uniformly in A to f exactly when
lim[f
n
f[
A
= 0.
Since convergence does not need to be uniform on the whole X, we
introduce the following
10
Denition A sequence of functions f
n
: X C is called locally uniformly
convergent in X if every point x X lies in a neighborhood U
x
in which
the sequence f
n
converges uniformly.
A series

f

is called locally uniformly convergent in C when its asso-


ciated sequence of partial sums is locally uniformly convergent in X.
Uniform convergence implies locally uniform convergence. Let A
1
, ...A
m
be subsets of X. If the sequence f
n
: X C converges uniformly in each of
these subsets, then it also converges uniformly in the union A
1
... A
m
.
A consequence of this is
Theorem 3.1 If the sequence f
n
converges locally uniformly in X, then it
converges uniformly on each compact subset K of X.
Proof Every point x K has an open neighborhood U
x
in which f
n
is
uniformly convergent. The open cover U
x
: x K of the compact set K
admits a nite subcover, say U
x
1
, ..., U
x
m
. Then f
n
converges uniformly in
U
x
1
... U
x
m
, and therefore so in the subset K of this union.
Denition A sequence or series converges compactly in X if it converges
uniformly on every compact subset of X.
So we may reformulate 3.1:
Corollary 3.2 Local uniform convergence implies compact convergence.
Let X be a metric space. We call X locally compact if each of its points
has at least one compact neighborhood. Then it is easy to see the following
Theorem 3.3 If X is locally compact, then every compactly convergent se-
quence or series in X is locally uniformly convergent in X.
In locally compact spaces local uniform convergence and compact con-
vergence are equivalent. This is also valid in domains in C since they are
locally compact.
To guarantee that every re-arrangement of a series will converge locally
uniformly we introduce the following concept.
Denition A series

of functions f

: X C is called normally conver-


gent in X if each point of x has a neighborhood U which satises

[f

[
U
<
.
By the Weierstra majorant criterion [7 p.103], we see that every series
which is normally convergent in X is locally uniformly convergent in X.
Since locally uniform convergence implies compact convergence we get that:
If

is normally convergent in X, then

[f

[
K
< for every compact
subset K X.
11
If the metric space is locally compact then locally uniform convergence
and compact convergence is equivalent, hence the converse is true:
If X is locally compact and

[f

[
K
< for every compact subset
K X, then

f

is normally convergent in X.
Since domains D in C are locally compact then

f

is normally con-
vergent in D if and only if

[f

[
K
< for every compact set K D.
We end this section with the rearrangement theorem.
Theorem 3.4 If

0
f

converges normally in X to f, then for every bi-


jection : N N the rearranged series

0
f
()
also converges normally in
X to f.
Proof For every point x X there is a neighborhood U for which

[f

[
U
<
. Since for every bijection of N every rearrangement of a absolutely
convergent series of complex numbers converges to the same limit as the
original one (see [7 p.28]), we have

[f
()
[
U
< and

0
f
()
(x) = f(x).
This is valid for every x X, hence

f
()
converges normally in X to f.
3.3 Innite products of complex functions
Denition Let (a

)
k
be a sequence of complex numbers. The sequence
(

n
=k
)
nk
a

of partial products is called an innite product with the factors


a

. It is written

=k
a

or

a

. In general, k = 0 or k = 1.
We introduce the partial products p
m,n
:= a
m
a
m+1
...a
n
=

n
=m
a

for
k m n.
Denition The product

a

is called convergent if there exists an index


m such that the sequence (p
m,n
)
nm
has a limit a
m
,= 0.
We call a := a
k
a
k+1
...a
m1
a
m
the value of the product and we write

:= a
k
a
k+1
...a
m1
a
m
= a. The following result is easy to see, namely:
That a product

being convergent is equivalent to the fact that at most


nitely many factors are zero and the sequence of partial products consisting
of the nonzero elements has a limit ,= 0. We also see that a convergent
product

a

is zero if and only if at least one factor is zero.


Theorem 3.5 If

=0
a

converges, then a
n
:=

=n
a

exists for all n


N. Moreover, lim a
n
= 1 and lima
n
= 1
Proof We assume that a :=

a

,= 0. Then a
n
= a/p
0,n1
. Since
limp
0,n1
= a, it follows that lim a
n
= 1. And lima
n
= 1 holds because
a
n
,= 0 and a
n
= a
n
/ a
n+1
.
Now we let X denote a locally compact metric space.
12
Denition Let f

((X) be a sequence of continuous functions on X


with values in C. The innite product

f

is called compactly convergent


in X if, for every compact set K X there exists an index m such that
the sequence p
m,n
:= f
m
f
m+1
...f
n
, n m, converges uniformly on K to a
non-vanishing function

f
m
.
For each point x X f(x) :=

(x) C exists and we call the function


f: X C the limit of the product and write f =

.
The continuity theorem [7 p. 95], says that if the sequence f

((X)
converges locally uniformly in X, then the limit function f = limf
n
is
likewise continuous on X. Because locally uniform convergence and compact
convergence coincide in the locally compact metric space X, it follows from
the continuity theorem and theorem 3.5 that
if

f

converges compactly to f in X, then f is continuous in X and


the sequence f

converges compactly in X to 1.
In view of this we write the factors of a product

f

in the form f

=
1 + g

and the sequence g

converges compactly to zero if



f

converges
compactly.
Denition A product

f

with f

= 1 + g

((X) is called normally


convergent in X if the series

g

converges normally in X.
It is quite easy to see that [6 p. 7] if

0
f

converges normally in X,
then for every bijection : N N, the product

0
f
()
converges nor-
mally in X and the product converges compactly in X. It actually converges
compactly to a function f: X C, see the rearrangement theorem [6 p. 8].
We will need the next theorem (3.7) in section 4.2. First we recall from
complex analysis that if a sequence f
n
of functions holomorphic in D con-
verges compactly there to f, then f is holomorphic in D. This is known as
Weierstra convergence theorem. This theorem can be sharpened as follows:
Let G be a bounded region and f
n
a sequence of functions which are
continuous on G and holomorphic in G. If the sequence f
n
[G converges
uniformly on G, then the sequence f
n
converges uniformly in G to a limit
function which is continuous on G and holomorphic in G (see [7 p. 269] for
proof).
From this we get the following
Lemma 3.6 If A is a discrete subset of G and f
n
O(G) is a sequence
which converges compactly in GA, then the sequence f
n
converges compactly
in the whole of G.
13
Theorem 3.7 If f =

f

, f

O(G), is normally convergent in G, then


the sequence

f
n
=

n
f

O(G) converges compactly in G to 1.


Proof Let

f
m
,= 0. Then A := Z(

f
m
) is discrete and relatively closed in G.
All the partial products p
m,n1
O(G), n > m, are non-vanishing in G A
and

f
n
(z) =

f
m
(z)(
1
p
m,n1
(z)) for all z G A.
The sequence 1/p
m,n1
converges compactly in GA to 1/

f
m
. Therefore
by lemma 3.6, this sequence also converges compactly in G to 1.
3.4 Mittag-Leer series
We know that in every Laurent expansion,

1
b

(z c)

was called the


principal part. It is called nite if almost all b

vanish. A distribution of
principal parts on D is dened to be a function that maps every point
d D C to a principal part q
d
, such that the set T = z D[(z) ,= 0,
is discrete. This set is called the support of .
Every function h holomorphic in D except for some isolated singularities
can be written as a Laurent series with a principal part around its singu-
larities, every such function determines thus a principal part distribution
PD(h) where the support is the set of nonremovable singularities.
If all except a nite number of the b
n
vanish in all the principal parts of
the Laurent expansions for some function h holomorphic in D d
1
, d
2
, ...,
the points d
1
, d
2
, ... are poles. Then h is a meromorphic function in D.
Thus h is meromorphic in D if and only if PD(h) is a distribution of nite
principal parts with the pole-set P(h) as support. We now pose the following
problem:
For every principal part distribution with support T, construct a func-
tion h O(D T) with PD(h) = .
We are able to construct such functions with help of certain series called
Mittag-Leer series for a principal part distribution , and show that every
such series represents a meromorphic function with principal part distribu-
tion equal to . Let us dene Mittag-Leer series.
We let T = d
1
, d
2
, ... be the support of the principal part distribution
and arrange the points d
1
, d
2
, ... in a sequence where every point occurs
once. If the origin belongs to T we set d
1
= 0. The principal part is uniquely
described by the sequence (d

, q

) where q

:= (d

).
14
Denition A series h =

=1
(q

) is called a Mittag-Leer series for


the principal part distribution (d

, q

) on D if the following holds


1. g

is holomorphic in D;.
2. the series h converges normally in D d
1
, d
2
, ....
Proposition 3.8 If h is a Mittag-Leer series for a principal part distri-
bution (d

, q

), then h O(D d
1
, d
2
, ...) and PD(h) = (d

, q

).
Proof Because the series converges normally in D d
1
, d
2
, ..., h is holo-
morphic in D d
1
, d
2
, .... Since all the summands q

, ,= n, are
holomorphic in a neighborhood U D of d
n
, the series

=n
(q

) con-
verges uniformly on every compact disc in U to a function f
n
O(U) in U,
by Weierstra convergence theorem [2 p. 213]. Since h = f
n
+ (q
n
g
n
)
we have h q
n
= f
n
g
n
in U d
n
where f and g
n
are holomorphic
at d
n
, it follows that q
n
is the principal part of h at d
n
, n 1. Hence
PD(h) = (d

, q

).
Remark If the principal parts are nite, then the points d

are poles. That


means that in the Mittag-Leer series for the distribution of that nite
principal parts, every summand is meromorphic in D. Which implies that
the series is a normally convergent series of meromorphic functions in D.
The terms g

in

=1
(q

) which force the series to converge are called


the convergence producing summand of the Mittag-Leer series.
We now consider the case D = C. Every function q

O(C d

)
can be expanded to a Taylor series around 0, which converges in the disc
of radius [d

[, 2. We denote by p
k
the kth Taylor polynomial for q

around 0. We show that these polynomials force the series to converge.


Theorem 3.9 (Mittag-Leer) For every principal part distribution (d

, q

)
1
in C, there exists a Mittag-Leer series in C of the form q
1
+

=2
(q


p
k

), where p
k

: = k

th Taylor polynomial of q

around 0.
Proof Since the sequence (p
k
)
k1
converges compactly in B
|d

|
to q

, for
every 2 we can choose a k

N such that [q

(z) p
k

(z)[ 2

for all
z with [z[
1
2
[d

[. Since lim d

= , there is for every compact set K C


a n such that K lies in all the discs B1
2
|d

|
with n. Therefore

n
[q

p
k

[
K

n
2

< ,
which means that the series is normally convergent in C d
1
, d
2
, ....
The polynomials p
k

are holomorphic in C, hence the series is a Mittag-


Leer series in C for (d

, q

)
1
.
15
Now the following corollary is clear.
Corollary 3.10 Every principal part distribution on C with support T is
the principal part distribution of a function holomorphic in C T .
Corollary 3.11 Every function h that is meromorphic in C can be repre-
sented by a series

h

that converges normally in C, where each summand


h

is rational and has at most one pole in C.


Proof By Mittag-Leers theorem, corresponding to the principal part dis-
tribution PD(h) there is a Mittag-Leer series

h C with polynomials as
convergence producing summands. Because the function h is meromorphic,
all principal parts are nite. This implies that all the summands of this
series are rational functions with exactly one pole in C. The dierence h

h
is holomorphic in all of C and can therefore be expanded to a Taylor series.
Hence it is normally convergent.
3.5 Greatest common divisors
In our way to prove the principal ideal theorem in section 4 we are going to
use the concept of greatest common divisors in the ring O(G), where G C
is a region. The most important result shown below is that there exists
a greatest common divisor for every non-empty subset S in the integral
domain O(G).
Denition A map D: G Z is called a divisor on G if its support [D[ :=
z G[D(z) ,= 0 is discrete in G.
Let h be meromorphic in a region G with both zero-set Z(h) and pole-set
P(h) discrete in G. Then (h): G Z determines, by z o
z
(h), a divisor
with support Z(h) P(h). Such divisor is called a principal divisor. A
divisor D is called positive if D(z) 0 for all z G. Positive divisors are
also called distributions of zeros. We denote the set of all positive divisors
T
+
(G) Holomorphic functions f have positive divisors (f).
The basic arithmetic concepts are dened as usual.
Denition A function f O(G) is called a divisor of g O(G) if g = f h
with h O(G).
Remark Note that divisor means two dierent things here.
There is a connection between divisibility for elements f, g ,= 0 and their
principal divisors (f), (g).
Theorem 3.12 Let f, g O(G) 0. Then f divides g if and only if
(f) (g).
16
Proof The function f divides g if and only if h := g/f O(G). This occurs
if and only if o
z
(h) = o
z
(g) o
z
(f) 0 for all z G, i.e. if and only if
(f) (g).
Denition Let S O(G), S ,= , then f O(G) is called a common
divisor of S if f divides every element g of S. Let f be a common divisor of
S, then f is called a greatest common divisor of S if every common divisor
of S is a divisor of f. It is denoted by f = gcd(S).
Remark Greatest common divisors are unique up to unit factors. Despite
this we speak about the greatest common divisor f of a set S and denote it
f =gcd(S).
Denition A set S O(G), S ,= is called relatively prime if gcd(S)= 1.
Theorem 3.13 Every function f O(G) with (f) = min(g) : g S, g ,=
0 is a gcd of S ,= 0.
Proof Let f O(G). Then (f) = min(g) : g S, g ,= 0 implies that
(f) (g), which means that f divides g for all g S. That is f is a common
divisor to S. Choose any other common divisor h of S, then (h) (g) which
implies that (h) (f) and h divides f. Hence f is a gcd(S).
Theorem 3.14 Every non-empty subset S in O(G) has a gcd.
Proof Given S ,= , choose f O(G) such that (f) = min(g)[g S, g ,=
0. Such a function exists by the Weierstra product theorem [6 p.92].
Theorem 3.15 The set S O(G), S ,= is relatively prime if and only if
the functions in S have no common zeros in G.
Proof ) Suppose gcd(S)= 1 and assume that
gS
Z(g) ,= . Then a
function vanishing on a point of
gS
Z(g) ,= with order equal to one
would have order less or equal to every g S. Hence it would be a common
divisor of S and thus divide 1, contradiction.
) Let
gS
Z(g) = = Z(1). Then 1 has order less or equal to every
g S. Hence 1 =gcd(S).
3.6 Wedderburns lemma
This lemma will help us later on.
Lemma 3.16 Let u, v O(G) be relatively prime. Then there are functions
a, b O(G) such that au + bv = 1.
17
Proof Since u, v are relatively prime, we know from theorem 3.15 that
Z(u) Z(v) = . Then the pole set of the function 1/uv is equal to
the disjoint union of the pole sets of 1/u and 1/v, that is P(1/uv) =
P(1/u) P(1/v). Because 1/uv is a meromorphic function, by corollary
3.11 of Mittag Leers theorem, it can be written as a normally convergent
series

h

, where each summand has at most one pole. This series can be
rearranged to a sum of two series a
1
=

h

1
and b
1
=

h

2
, where the
poles of the summands

h

1
belongs to P(1/u) = Z(u) and the poles of
the summands in h

2
belongs to P(1/v) = Z(v). The sums represents mero-
morphic functions, therefore we have: 1/uv = a
1
+b
1
, where a
1
, b
1
/(G).
Now we dene a: = vb
1
and b: = ua
1
both holomorphic in G, then it follows
that au + bv = 1.
4 Ideal structure in O(G)
4.1 Principal ideal theorem
We recall from algebra that for example in the ring Z or F[x] when F is
a eld, every ideal is principal and in the ring F[x
1
, ..., x
n
] every ideal is
nitely generated. The situation in the ring O(G) is a bit dierent. We
know from the previous section that a greatest common divisor exists for
every subset of O(G).
Proposition 4.1 If f O(G) is a gcd of the nitely many functions f
1
, ..., f
n

O(G), then f can be written as a sum f = a
1
f
1
+ a
2
f
2
+ ... + a
n
f
n
, where
a
1
, ..., a
n
O(G).
Proof By induction on n. Let f ,= 0. For n = 1 it is true. Let n > 1
and choose a function

f: = gcdf
2
, ..., f
n
. By the induction hypothesis we
know that

f = a
2
f
2
+ ... + a
n
f
n
where a
2
, ..., a
n
O(G). Now, because
f = gcdf
1
,

f we have 1 = gcdf
1
/f,

f/f. Then by Wedderburns lemma
f
1
/f and

f/f satisfy the equation 1 = (f
1
/f)a + (

f/f)b with functions
a, b O(G), which implies f = f
1
a +

fb = f
1
a + (f
2
a
2
+ ... + f
n
a
n
)b. Let
a
1
: = a and a

: = a

b, 2. Hence f = a
1
f
1
+ ... + a
n
f
n
.
We are now able to prove our main theorem of this section.
Theorem 4.2 (Principal ideal theorem) An ideal I O(G) is nitely gen-
erated if and only if it is a principal ideal.
Proof ) We reformulate the claim as follows: If I is generated by f
1
, ..., f
n
,
then I = O(G)f, where f = gcdf
1
, ..., f
n
. Every non-zero subset in O(G)
18
has a gcd, then by proposition 4.1, O(G)f I. Since f divides all of
the f
1
, ..., f
n
, we see that f
1
, ...f
n
O(G)f; hence I O(G)f therefore
I = O(G)f, which means that I is a principal ideal.
) This direction is trivial.
Since f divides every f
1
, ..., f
n
the functions f
1
, ..., f
n
vanish where f
does, and since f = a
1
f
1
+ ... + a
n
f
n
the function f vanish at the common
points where f
1
, ..., f
n
vanish, we have Z(f) =
i=1,...,n
Z(f
i
).
However, the next example shows that there exists ideals in O(G) that
are not nitely generated.
Example Let G be a region and let A G be innite, discrete and closed in
G. The set a := f O(G) : f(a) = 0 for all except a nite number of a
A is an ideal in O(G). Assume that a is nitely generated, i.e. a =
O(G)f
1
+... +O(G)f
n
. Then there exists a k
j
N such that for all k k
j
,
f
j
(a
k
) = 0, j = 1, .., n. Choose k
0
:=maxk
1
, ..., k
n
, then for h a, h =

n
j=1
g
j
f
j
we have h(a
k
) = 0 for all k k
0
. For any discrete and closed
subset T G we can construct a holomorphic function f with zero set T.
Let T := Aa, then f(a) ,= 0 and f a, but f is not a linear combination
of f
1
, ..., f
n
. Contradiction!
Thus we have proved
Proposition 4.3 No ring O(G) is Noetherian, and therefore never a prin-
cipal ideal domain.
We see that every ideal in O(G) either is principal or not nitely gener-
ated.
Denition An integral domain R is called a unique factorization domain
if the following conditions are valid:
1. Every non-unit element p ,= 0 can be factored into a product of a nite
number of irreducibles.
2. If p
1
, ..., p
n
and q
1
, ..., q
m
are two factorizations of the same element of
R into irreducibles, then n = m and the q
j
can be renumbered so that
for p
i
and q
i
we have p
i
= q
i
u
i
, where u
i
are units in R.
The functions (z c), c G are, up to unit factors, precisely the primes
of O(G). Functions f ,= 0 in O(G) with innitely many zeros in G cannot
be written as the product of nitely many primes. From the fact that such
functions exists, by the general Weierstra product theorem, we get
Proposition 4.4 No ring O(G) is a unique factorization domain.
19
4.2 Fixed and free ideals
An element of O(G) is a unit if and only if it vanishes nowhere, therefore
every function in a proper ideal vanishes somewhere.
We now distinguish between two types of ideals.
Denition An ideal I in O(G) is called xed if
fI
Z(f) ,= , otherwise
it is called free. A point c G is called a zero of an ideal I if f(c) = 0 for
every f O(G).
Remark A xed ideal has at least one zero and a free ideal has no zeros.
From the principal ideal theorem 4.2, we have
Corollary 4.5 Any proper nitely generated ideal is xed.
Denition An ideal is called closed if it contains the limit function of every
sequence f
n
a that converges compactly in G.
We continue with a lemma.
Lemma 4.6 Let I O(G) be an ideal, where c G is not a zero, and let
f, g O(G) where f vanishes only at c. If fg I, then g I
Proof Choose h I with h(c) = 1. Let n := o
c
(f). If n 1 then,
f
zc
g =
f
zc
g(h(z) (h(z) 1)) =
f
zc
gh
h(z)1
zc
fg I because
h(z)1
zc
has a removable singularity at c and gh I and fg I. Applying this n
times gives (
f
(zc)
n
) g I. Since
f
(zc)
n
is invertible, we have g I.
By Weierstra product theorem for arbitrary regions [5 p.93], every
f ,= 0 that is holomorphic in an arbitrary region G C can be written in
the form f = u

1
f

where u is a unit in the ring O(G) and



1
f

converges normally in G and f

has exactly one zero c

in G of order 1.
Proposition 4.7 If I is a closed free ideal in O(G), then I = O(G).
Proof Let f I, f ,= 0 and let f =

be a factorization of f as described
above. Then

f
n
:=

n
f

O(G) converges compactly in G to 1 by


theorem 3.7. We write

f
n
= f
n

1
f

= f
n

f
n+1
. Since

f
0
= f I and f
n
has no zeros in G c

, it follows by induction and lemma 4.6 that



f
n
I
for all n 0. Since I is closed, it follows that 1 I and hence I = O(G).
Theorem 4.8 An ideal I O(G) is a principal ideal if and only if it is
closed.
20
Proof ) Let I = O(G)f, and let g
n
= a
n
f O(G) converge compactly
to g on G. Then a
n
= g
n
/f converges compactly on G Z(f), by lemma
3.6, to a function a O(G). Hence g = af I, so I is closed.
) By 3.14, I has a gcd f in O(G). Then

I := f
1
I is a free ideal in
O(G). Now,

I is closed if I is, therefore by proposition 4.7

I = O(G) which
implies that I = O(G)f.
From this and proposition 4.7 we have
Corollary 4.9 A proper free ideal in O(G) is never nitely generated.
However, a xed ideal is not necessarily nitely generated. Let I be a
free ideal in O(G), then (z c)I with c G would be a xed ideal that is
not nitely generated.
Proposition 4.10 Let I ,= O(G) be a proper free ideal in O(G), then every
function f I has innitely many zeros.
Proof Assume f O(G) has only nitely many zeros. Then there are
f
1
, ..., f
r
I with Z(f, f
1
, ..., f
r
) = , hence 1 = gcd(f, f
1
, ..., f
r
) I. Con-
tradiction!
Because a non-zero polynomial only has a nite number of zeros we get
Corollary 4.11 No non-zero polynomial belongs to a proper free ideal.
We now describe the ideals, especially the free ones, with help of divisors.
As in section 3.5 we denote T
+
(G) the set of positive divisors D T(G),
i.e. such that D(z) 0 for all z G and D(z) > 0 for at least one z G.
Denition A non-empty subset F T
+
(G) is called a lter in T
+
(G) if
it satises
D

D F =D

F and D, D

F =min(D, D

) F
Furthermore let [F[ := [D[; D F.
Proposition 4.12 There is an inclusion preserving bijection between the
set of proper ideals ,= 0 in the ring O(G) and the family of all lters in
T
+
(G), namely O(G) a F := (a) := (f); f a 0
with inverse
F a := I(F) := f O(G); f = 0 (f) F
21
Proof Every function f in an ideal a determines a divisor (f) by the order
function and a contains all other functions that is divided by f. Therefore
(a) fullls, by theorem 3.12 and prop. 4.1, the conditions for a lter.
On the other hand, given a lter F, the set I(F) O(G) is an ideal, since
for f, g I(F)0 we have f+g I(F) because of (f+g) min((f), (g))
F (when we assume f + g ,= 0) while f O(G), g I(F) fg I(F)
is obvious. We have now to show I((a)) = a and (I(F)) = F for an ideal
a O(G) repr. a lter F T
+
(G). The inclusions a I((a)) and
(I(F) F are obvious. Now, if f I((a)) 0, then (f) = (g) with some
g a, hence f = ug with a unit u O(G) resp. f = ug a. On the other
hand, take a divisor D F, write D = (f) with some function f O(G).
But then f I(F) resp. D = (f) (I(F)).
4.3 Maximal ideals
Now we are going to describe maximal ideals of the ring O(G). The maximal
ideals are also described with help of divisors. We will see that the free
maximal ideals have a simpler structure than the xed ones. Further we
consider the factor rings O(G)/m for both free and xed maximal ideals.
We start to recall some results from algebra.
Let N be an ideal of a ring R. Then the subsets a + N = N + a =
a + n[n N of R called the additive cosets of N form a ring R/N with
the binary operations dened by (a + N) + (b + N) = (a + b) + N and
(a + N)(b + N) = ab + N. This ring is called the factor ring of R modulo
N.
The fundamental homomorphism theorem for rings says that if : R
R

is a ring homomorphism with kernel N, then the map : R/N (R)


given by (x + N) = (x) is an isomorphism. If : R R/N is the
homomorphism given by (x) = x + N, then for each x R, we have
(x) = (x). See [1 p.330].
An ideal m of R is maximal if and only if R/m is a eld [1 p.336].
We also recall that every integral domain D can be enlarged to a eld F
such that every element of F can be expressed as a quotient of two elements
of D. F is called the eld of fractions of D, see [1 section 5.4].
Denition Let F be a eld. A F-algebra A is a ring A such that:
1. A is a vector space under addition.
2. k(ab) = (ka)b = a(kb) for all k F and a, b A.
A F-algebra A is called a division algebra if it as a ring is a division ring.
22
Denition Let A and B be F-algebras. A function f: A B is an F-
algebra homomorphism if it is a ring homomorphism and if f(ka) = kf(a)
for all a A and k F.
From the last denition we see that the elements in C stay xed under
a C-algebra homomorphism.
Theorem 4.13 Let m be an ideal in O(G). Then the following statements
are equivalent:
1. m is a closed maximal ideal in O(G).
2. There exists a point c G such that m = f O(G) : f(c) = 0 =
O(G)(z c).
3. There exists a C-algebra homomorphism : O(G) C with kernel m.
Proof 1) 2) Since m is closed, m = O(G)f and since m is maximal, m
is prime. This implies that f is irreducible, which means that f only has
one zero of order 1.
2) 3) The evaluation map
c
: O(G) C dened by
c
(f) = f(c) is
a homomorphism with kernel m.
3) 1) Let : O(G) C be a C-algebra homomorphism, then z O(G)
implies (z) = c for some c C. Therefore (z c) = 0. Suppose c / G,
then
1
zc
O(G) and we would have 1 = (1) = (
1
zc
)(z c) = 0.
Hence c has to belong to G. Now let f O(G) and write f = f(c) (z
c)
f(z)f(c)
zc
. Since keep complex numbers xed and (z c) = 0, it follows
that (f) = f(c). That is, =
c
where
c
is the evaluation map. By the
fundamental homomorphism theorem, O(G)/m is isomorphic to C, where
m = f O(G) : f(c) = 0 is the kernel of the map : O(G) C . Hence
m is maximal and clearly closed.
Corollary 4.14 If m is a maximal xed ideal, then the residue eld O(G)/m
is isomorphic to C.
Proof Since
c
is onto O(G), O(G)/m is isomorphic to C.
The maximal ideals that are not closed, i.e. free, have a more compli-
cated structure than the closed and hence xed ones. We will now charac-
terize the maximal ideals, both xed and free. From proposition 4.12 we see
that an ideal m O(G) is maximal if and only if M := (m) is a maximal
lter in T
+
(G).
Proposition 4.15 A lter M T
+
(G) is maximal i
min(D

, D) > 0, D M =D

M.
23
Proof ) Suppose M is maximal and the given condition is not valid,
then D

would together with the elements in M generate a lter properly


containing M. Hence M would not be maximal.
) Suppose M is not maximal, that is, there exists a lter N M.
Then for any D

N and D M we have min(D

, D) > 0, hence the given


condition implies that D

M, i.e. N M and N = M.
To a divisor D T
+
(G) we associate its reduced divisor red(D) dened
by
red(D)(z) :=

1 if D(z) > 0 ;
0 otherwise.
A lter F T
+
(G) is called reduced, if D F =red(D) F; we dene
red(F) F to be the unique reduced lter in T
+
(G) with [red(F)[ = [F[.
Proposition 4.16 The lter F T
+
(G) is maximal i red(F) = F and
A G discrete, A B ,= , B [F[ =A [F[.
Proof ) Since red(F) F in any case, the maximality of F implies
red(F) = F. Assume then that A / [F[ would be a discrete subset, such that
A B ,= for all B [F[. Then

F = D : [D[ A B for some B [F[
would be a lter containing F as proper sublter.
) Assume min(D

, D) > 0 for all D F. Then [D

[ [D[ , = for
all D F, hence [D

[ [F[, i.e. [D

[ = [D

[ for some D

F. But then
D

red(D

) = red(D

) red(F) = F, whence D

F.
Note that since the support of a divisor is the zero-set of the associated
function a maximal ideal m is free if [D[ = , D (m).
Theorem 4.17 If m is a maximal free ideal, then O(G)/m contains the
set of all rational functions C(z) as a subeld.
Proof By corollary 4.11, m can not contain a non-zero polynomial. There-
fore the composition C[z] O(G) O(G)/m is an injective map from
C[z] O(G)/m, that is C[z] is isomorphic to a subring of O(G)/m. Be-
cause m is maximal O(G)/m is a eld. Then the subeld C[z] can be
enlarged to the eld of fractions C(z).
Corollary 4.18 The eld C is a subeld of O(G)/m. Hence the residue
eld O(G)/m of a maximal free ideal may be considered as a division algebra
containing C as a proper subeld.
In the remaining part of this section we investigate O(G)/m more closely.
The next two lemmas are going to help us. From now on let G = C. First
we note that:
24
1. If (f) = (g) then f = gh where h is a unit (a nowhere vanishing
function).
2. If z
n

n=1
is any closed, discrete subset of C and if w
n

n=1
is any
sequence of complex numbers then there is an f O(C) such that
f(z
i
) = w
i
, i = 1, 2, ....
3. A eld F is called algebraically closed if every non-constant polynomial
in F[T] has a zero in F.
Let C
N
be the set of all complex-valued sequences.
Lemma 4.19 The eld O(G)/m is algebraically closed.
Proof Let h m be a function with (innitely many) simple zeros, say
a
0
, a
1
, a
2
, .... Consider the following factorization of the quotient map:
O(G) O(G)/(h) K := O(G)/m.
Now the map O(G) C
N
, f (f(a
n
)) has kernel (h) and is, according
to 2), onto, hence O(G)/(h)

= C
N
and K

= C
N
/ m with a maximal ideal
m C
N
. Let p K[T] be a non-constant monic polynomial and lift it to a
monic polynomial q(T) = (q
n
(T))
nN
(C
N
)[T] (C[T])
N
. In particular,
all the polynomials q
n
(T) C[T] are monic of the same degree > 0, so for
any n there is a zero b
n
C of the polynomial q
n
. Then the image in K
of the sequence (b
n
)
nN
is a zero of p(T). So C
N
/ m and hence O(G)/m is
algebraically closed.
Lemma 4.20 The cardinality of O(C)/m is equal to the cardinality of the
continuum.
Proof The dense subset A := Q(i)[z] of O(C) is countable. Consider
the map : A
N
O(C), where A
N
is the set of all sequences (f
n
)
nN
with f
n
A and ((f
n
)
nN
) := 0 if f
n
is not compactly convergent and
((f
n
)
nN
) := lim
n
f
n
else. Since is surjective and [A
N
[ = [R[ = the
cardinality of the continuum, we have [O(C)[ [R[. Hence O(C)/m has as
most [R[ elements. But all complex numbers are incongruent (mod m), so
O(C)/m has at least [R[ elements. Therefore O(C)/m contains precisely
[R[ elements.
From algebra we recall that a eld F is an extension eld of K if K is
a subeld of F. A subset S of F is said to be algebraically independent
over K if for all n > 0, f K[x
1
, ..., x
n
] and pairwise distinct s
1
, ..., s
n
,
f(s
1
, ..., s
n
) ,= 0 implies f = 0.
A transcendence base of F over K is a subset S of F which is alge-
braically independent over K and is maximal in the set of all algebraically
25
independent subsets of F. Moreover, every transcendence base of F over
K has the same cardinality [8. p.315]. The cardinality [S[ of any base S is
called the transcendence degree of F over K and is denoted tr.d.F/K.
In [8. p.317] it is proved that if F
1
and F
2
are algebraically closed eld
extensions of the eld K
1
resp. K
2
with tr.d.F
1
/K
1
=tr.d.F
2
/K
2
, then every
isomorphism of elds K
1

= K
2
extends to an isomorphism F
1

= F
2
.
We are now able to state the nal theorem of this section.
Theorem 4.21 Let m be a free maximal ideal, then O(G) is isomorphic as
a ring to C.
Proof Since O(G)/m contains C it has tr.d over Q at least [R[ and by
lemma 4.20 it has tr.d at most [R[ over Q. Hence O(G)/m has tr.d [R[ over
Q. The set of complex numbers also have tr.d [R[ over Q, and both C and
O(G)/m is algebraically closed over Q. Hence by the above O(G)/m

= C.
If we replace the integral domain O(G) with any integral domain with
greatest common divisor in which Wedderburns lemma is true, the principal
ideal theorem is valid.
Moreover, for non-compact Riemann surfaces the theorems of Weierstra
and Mittag-Leer are also valid and hence are Wedderburns lemma and the
existence of greatest common divisors. Therefore, if we replace regions in C
with non-compact Riemann surfaces, all the results in the previous section
remain valid.
26
5 Prime ideals
We are going to investigate the lters of prime ideals.
Denition A lter P T
+
(G) is called prime if D
1
+ D
2
P = D
1

P D
2
P.
Note that a lter P is prime if and only if p := I(P) O(G) is a prime
ideal. We want to investigate that situation more in detail. For a divisor
D T
+
(G) we set
s(D) := supD(z)[z G N .
Lemma 5.1 Let P T
+
(G) be a prime lter. Then P is maximal if and
only if there is a divisor D
0
P with s(D
0
) < .
Proof ) This is clear since for P maximal we have P = red(P).
) Since P is prime, we may assume s(D
0
) = 1. Take now D

T
+
(G)
with min(D

, D) > 0 for all D P, in particular for D = D


0
. Write D
0
=
min(D

, D
0
) + D
1
. Since P is prime, either min(D

, D
0
) P or D
1
P. In
the rst case we obtain that also D

P because of D

min(D

, D
0
) P,
in the second case we would get 0 = min(D
1
, D
0
) P, a contradiction.
Corollary 5.2 If a prime ideal is xed, then it is maximal.
Lemma 5.3 If P T
+
(G) is prime, then so is red(P); in particular M :=
red(P) is maximal.
Proof Assume D
1
+ D
2
red(P). Then [D[ = [D
1
+ D
2
[ with some
D P, and we may even assume D D
1
+ D
2
resp. D = D
3
+ D
4
,
where [D
3
[ = [D
1
[, [D
4
[ = [D
2
[. Since P is prime, we obtain D
3
P or
D
4
P, say D
3
P. But then D
1
red(P).
Let M T
+
(G) be a maximal lter. Denote T := [M[. We shall
investigate the set of lters F M with [F[ = T. Note rst of all that
T T(G) is an ultralter of discrete subsets of G:
1. , T;
2. A B with B G discrete and A T =B T;
3. A, B T =A B T
4. A G discrete, A B ,= , B T =A T.
27
Furthermore note that M can be reconstructed from T, namely M con-
sists of all divisors D, such that there is a subset A T with D(z) > 0 for
all z A.
We dene on T
+
(G) a new relation:
D _ D

: A T with D[
A
D

[
A
.
That relation is obviously reexive and transitive, it becomes even anti-
symmetric if we replace = with the equivalence relation:
D D

: A T with D[
A
= D

[
A
.
But the most interesting fact here is that for any two divisors D, D


T
+
(G) one of the relations D _ D

or D

_ D holds. Assume that D _ D

does not hold, with other words that for A := z G; D(z) < D

(z) one
has A B ,= for all B T. But T being an ultralter, we may conclude
that A T resp. D _ D

. Furthermore note that D

_ D F implies
D

F: Let A T with D

[
A
D[
A
and A = [D
0
[ with some D
0
F.
Then F min(D, D
0
) D

, hence D

F.
The set of divisors is totally ordered with respect to the relation _.
Proposition 5.4 Let F, F

T
+
(G) be lters with [red(F)[ = [red(F

)[ =
T. Then either F F

or F

F.
Proof Suppose F , F

, then for all D F, D , F

, D _ D

does not
hold. Then D

_ D F holds, which implies that D

F. Hence F

F.
So we see that the set of all lters F with [red(F)[ = T is totally ordered
with respect to inclusion. But when is such a lter F prime?
Proposition 5.5 Let F T
+
(G) be a lter with [F[ = T. Then F is
prime if and only if for all n N
+
we have: nD F =D F.
Proof The given condition is obviously necessary, but also sucient: As-
sume D
1
+D
2
F, say D
1
_ D
2
. But then 2D
2
_ D
1
+D
2
F and hence
2D
2
F resp. D
2
F.
From this it follows that an ideal p m for a maximal ideal m O(G)
is prime if and only if f
n
p implies f p. These prime ideals are also
totally ordered in m with respect to inclusion.
To a divisor D M we associate the lter F
D
:= D

T
+
(G); D

_ D.
Furthermore we dene the lter P
D
:=

n=1
nF
D
.
Remark Note here that nF
D
= F
nD
and that P
D
is never empty.
28
Proposition 5.6 The lter P
D
:=

n=1
nF
D
is prime, and it is the largest
non-maximal prime lter contained in the lter F
D
.
Proof Assume mD

P
D
. Then mD

_ nD for all n N, in particular


mD

_ kmD resp. D

_ kD for all k N and thus D

P
D
. The second
part follows from lemma 5.1.
For example, if D = red(D) M is a reduced divisor, then P
D
consists
of all divisors D

, such that for every n N there is a subset A T with


D

(z) n for all z A.


Let D
i
[i I be the set of all divisors D M P. Then, any non-
maximal prime lter P T
+
(G) with [P[ = T is now an intersection
P =

iI
P
i
with prime lters P
i
= P
D
i
of the above type.
29
6 References
1. J. B. Fraleigh, A rst course in abstract algebra, Addison-Wesley Pub-
lishing Company, 1999.
2. J. E. Marsden, M. J. Homan, Basic complex analysis, W. H. Freeman
and Company, New York, 1997.
3. M. Henrikssen, On the ideal structure of the ring of entire functions,
Pac. Journ. Math. 2(1952), 179-184.
4. M. Henrikssen, On the prime ideals of the ring of entire functions,
Pac. Journ. Math. 3(1953), 711-720.
5. O. Helmer, Divisibility properties of integral functions, Duke Math. J.
6(1940), 345-356.
6. R. Remmert, Classical topics in complex function theory, Springer-
Verlag, New York, 1998.
7. R. Remmert, Theory of complex functions, Springer-Verlag, New York,
1998.
8. T. W. Hungerford, Algebra, Springer-Verlag, New York, 1974.
30

También podría gustarte