You are on page 1of 199

Lee A.

Rubel
with James E. Colliander

ENTIRE AND
MEROMORPHIC
FUNCTIONS

Springer
Universitext
Editors (North America): S. Axler, F.W. Gehring, and P.R. Halmos

Aksoy/Khamsi: Nonstandard Methods in Fixed Point Theory


Aupetit: A Primer on Spectral Theory
BoossBleecker: Topology and Analysis
Borkar: Probability Theory; An Advanced Course
Carleson/Gamelin: Complex Dynamics
Cecil: Lie Sphere Geometry: With Applications to Submanifolds
Chae: Lebesgue Integration (2nd ed.)
Charlap: Bieberbach Groups and Flat Manifolds
Chern: Complex Manifolds Without Potential Theory
Cohn: A Classical Invitation to Algebraic Numbers and Class Fields
Curtis: Abstract Linear Algebra
Curtis: Matrix Groups
DiBenedetto: Degenerate Parabolic Equations
Dimca: Singularities and Topology of Hypersurfaces
Edwards: A Formal Background to Mathematics I alb
Edwards: A Formal Background to Mathematics II alb
Foulds: Graph Theory Applications
Gardiner: A First Course in Group Theory
Giirding/Tambour: Algebra for Computer Science
Goldblatt: Orthogonality and Spacetime Geometry
Hahn: Quadratic Algebras, Clifford Algebras, and Arithmetic Witt Groups
Holmgren: A First Course in Discrete Dynamical Systems
Howe/Tan: Non-Abelian Harmonic Analysis: Applications of SL(2, R)
Howes: Modem Analysis and Topology
Humi/Mlller: Second Course in Ordinary Differential Equations
Hurwitz/Kritikos: Lectures on Number Theory
Jennings: Modem Geometry with Applications
Jones/Morris/Pearson: Abstract Algebra and Famous Impossibilities
Kannan/Krueger: Advanced Real Analysis
Kelly/Matthews: The Non-Euclidean Hyperbolic Plane
Kostrikin: Introduction to Algebra
Luecking/Rubel: Complex Analysis: A Functional Analysis Approach
MacLane/MoerdJjk: Sheaves in Geometry and Logic
Marcus: Number Fields
McCarthy: Introduction to Arithmetical Functions
Meyer: Essential Mathematics for Applied Fields
Mines/Richman/Rultenburg: A Course in Constructive Algebra
Moise: Introductory Problems Course in Analysis and Topology
Morris: Introduction to Game Theory
Porter/Woods: Extensions and Absolutes of Hausdorff Spaces
RamsayfRlichtmyer: Introduction to Hyperbolic Geometry
Relsel: Elementary Theory of Metric Spaces
Rickart: Natural Function Algebras
Rotman: Galois Theory
RubelColliander: Entire and Meromorphic Functions

(continued after index)


Lee A. Rubel
With assistance from James E. Colliander

Entire and Meromorphic


Functions

Springer
Lee A. Rubel James E. Colliander
Department of Mathematics Department of Mathematics
University of Illinois, Urbana-Champaign University of Illinois, Urbana-Champaign
Urbana,IL 61801-2917 Urbana, IL 61801-2917
USA USA
(deceased)

Editorial Board
S. Axler F.W. Gehring P.R. Halmos
Department of Mathematics Department of Mathematics Department of Mathematics
Michigan State University University of Michigan Santa Clara University
East Lansing, MI 48824 Ann Arbor, MI 48109 Santa Clara, CA 95053
USA USA USA

Mathematics Subject Classification (1991): 30Dxx, 30D35

Library of Congress Cataloging-in-Publication Data


Rubel, Lee A.
Entire and meromorphic functions / Lee A. Rubel with assistance
from James E. Colliander.
p. cm. - (Universitext)
Includes bibliographical references and index.
ISBN 0-387-94510-5 (softcover : alk. paper)
1. Functions, Entire. 2. Functions, Meromorphic. 3. Nevanlinna
theory. 1. Colliander, James E. II. Title.
QA353.E5R83 1995
515'.98-dc20 95-44887

Printed on acid-free paper.

With 2 illustrations.

®1996 Springer-Verlag New York, Inc.


All rights reserved. This work may not be translated or copied in whole or in part without the written
permission of the publisher (Springer-Verlag New York, Inc., 175 Fifth Avenue, New York, NY 10010,
USA), except for brief excerpts in connection with reviews or scholarly analysis. Use in connection
with any form of information storage and retrieval, electronic adaptation, computer software, or by sim-
ilar or dissimilar methodology now known or hereafter developed is forbidden.
The use of general descriptive names, trade names, trademarks, etc., in this publication, even if the for-
mer arc not especially identified, is not to be taken as a sign that such names, as understood by the Trade
Marks and Merchandise Marks Act, may accordingly be used freely by anyone.

Production managed by Laura Carlson; manufacturing supervised by Jacqui Ashri.


Camera-ready copy prepared using the authors' AMS-TeX files.
Printed and bound by R.R. Donnelley & Sons. Harrisonburg, VA
Printed in the United States of America.

987654321
ISBN 0-3 87-945 1 0-5 Springer-Verlag New York Berlin Heidelberg SPIN10424824
Dedicated to the Memory of Steven B. Bank
Student, Colleague, Teacher, Friend
Contents

1. Introduction ................................... 1

2. The Riemann-Stieltjes Integral ..................... 3

3. Jensen's Theorem and Applications ................. 6

4. The First Fundamental Theorem of Nevanlinna Theory . 9

5. ....................
Elementary Properties of T(r, f) 12

6. The Cartan Formulation of the Characteristic ......... 16

7. The Poisson-Jensen Formula ...................... 20

8. Applications of T(r) ............................. 23

9. A Lemma of Borel and Some Applications ........... 26

10. The Maximum Term of an Entire Function ........... 30


11. Relation Between the Growth of an Entire Function
...............
and the Size of Its Taylor Coefficients 40
12. Carleman's Theorem ............................. 45
13. A Fourier Series Method ......................... 49
14. The Miles-Rubel-Taylor Theorem on Quotient
Representations of Meromorphic Functions .......... 78
15. Canonical Products .............................. 87
Viii Entire and Meromorphic Functions

16. Formal Power Series ............................. 93

17. Picard's Theorem and the Second Fundamental


Theorem ...................................... 99

18. A Proof of the Second Fundamental Theorem ......... 113

19. "Two Constant" Theorems and the Phragm6n-Lindelof


Theorems..................................... 121

20. The Pblya Representation Theorem ................. 124

21. Integer-Valued Entire Functions .................... 139

22. On Small Entire Functions of Exponential-Type with


Given Zeros ................................... 146

23. The First-Order Theory of the Ring of All Entire


Functions ..................................... 158

24. Identities of Exponential Functions ................. 175

References ......................................... 182

Index ............................................. 185


1
Introduction

Mathematics is a beautiful subject, and entire functions is its most beautiful


branch. Every aspect of mathematics enters into it, from analysis, algebra,
and geometry all the way to differential equations and logic.
For example, my favorite theorem in all of mathematics is a theorem
of R. Nevanlinna that two functions, meromorphic in the whole complex
plane, that share five values must be identical. For real functions, there is
nothing that even remotely corresponds to this.
This book is an introduction to the theory of entire and meromorphic
functions, with a heavy emphasis on Nevanlinna theory, otherwise known as
value-distribution theory. Things included here that occur in no other book
(that we are aware of) are the Fourier series method for entire and mero-
morphic functions, a study of integer valued entire functions, the Malliavin-
Rubel extension of Carlson's Theorem (the "sampling theorem"), and the
first-order theory of the ring of all entire functions, and a final chapter on
Tarski's "High School Algebra Problem," a topic from mathematical logic
that connects with entire functions.
This book grew out of a set of classroom notes for a course given at the
University of Illinois in 1963, but they have been much changed, corrected,
expanded, and updated, partially for a similar course at the same place in
1993. My thanks to the many students who prepared notes and have given
corrections and comments.
In order to discover and prove interesting, deep, or powerful theorems
in this area, what we most need is more examples of interesting meromor-
phic functions-I would guess that the number of fundamentally different
examples known is about 20 or 30. One promising source of such examples
is the Painleve transcendents (see [14], pp. 438-444, and [29]).
However, in spite of a growing literature on these functions, the unfor-
2 Entire and Meromorphic Functions

tunate fact is that the "proofs" are incomplete and not rigorous-indeed,
there still is not a satisfactory proof that the Painleve transcendents of
even the first kind (i.e., solutions of w" = 6w2 + z) are meromorphic in
the full complex plane. Basic notions like "fixed singularity" and "movable
singularity," however intuitively appealing, have never been given rigorous
definitions.
It is hard to see how this lamentable situation will improve since, the
world being as it is, there is little "glory" attached to proving theorems
that have already been "proved."
The subject of entire and meromorphic functions has been growing for
many decades, and will continue to grow forever. It is hoped that this book
will give the novice reader a good introduction to the subject, or the expert
some new insights. This book could "easily" have been four or five times
it length, since the subject is so extensive, but to use my favorite saying,
"enough is too much."
LEE A. RUBEL

Lee Rubel died on March 25, 1995. As my teacher, the way his per-
sonality merged into his mathematics always inspired me. I sincerely hope
that readers of this book find similar inspiration.
JAMES E. COLLIANDER
October 18, 1995
2
The Riemann-Stieltjes Integral

We give here a brief summary of some of the basic facts about the Riemann
Stieltjes integral. Those unfamiliar with the subject are urged to read
Chapter 9 of Mathematical Analysis by Apostol [1].
Throughout this section, a and b are real numbers, usually a < b, and f
and a are real-valued functions defined on the closed interval [a, b]. When
f and a are suitably restricted, we will define f b f da as the Riemaun-
Stieltjes integral of f with respect to a. When a(x) = x for all x in [a, b],
fae f da is the ordinary Riemann integral of f, and many of the familiar
properties of the Riemann integral extend to the Riemann-Stieltjes integral.
Definition. A partition of [a, b] is an ordered (n + 1)-tuple

P = {z0,z1,...,xn} with x0 = a,

xn = b, and x j _ 1 < x j for j = 1,... , n.


Definition. A selection or from a partition P is an ordered n-tuple a =
{t1i ... , tn} such that

xj_1 <tj <xj for j=1,...,n.


Definition. A Riemann-Stieltjes sum (of f with respect to a) is a sum of
the form
S(P, or : J, a') = > f (tk){a(xk) - a(xk_01.
k=1
4 Entire and Meromorphic Functions

Definition. P' is a refinement of P; written P C P', means that each x


that occurs in P also occurs in P'.
Definition. f is integrable with respect to a; written f E R(a), means that
there exists a number A such that, for each c > 0, there exists a partition
P, such that P. C P; and if a is a selection from P, then

JS(P,a: f,a) - Al < e.

It is easily seen that A is uniquely determined if f E R(a), so that we


may write
b
fda = / b f (x) da(x) = A.
Ja a

Theorem. If f is continuous on [a, b] and a is monotone on [a, b], then


f E R(a) and a E R(f).
Theorem (Integration by parts). If f E R(a), then a E R(f) and

rb rb
/ f da = f(b)a(b) - f(a)a(a) - J a df.
fa a

Under suitable hypotheses, the usual linearity properties and formulas


for "change of variables" hold, e.g.,

r
J(Af +B g)da=A J fda+B r gda

J
a
b fd(Aa + BO) = A. f b fda + B
a
f
n
b fdf3
- 6 +
fa ' fa The

f
Jn
b f (x)da(x) =/ b f (a)a'(x)dx
b
fda
jd
=

h=fog, (3=aog,
a = g(c), b = g(d),

where (f o g)(x) = f (g(x))-


2. The Riemann-Stieltjes Integral 5

Theorem. If a is a step function that jumps ak at points xk, k = 1,


2, ... , n in [a, b] and f is continuous on [a, b], then
b

J f(x)da(x) = E f(xk)ak
a k=1

Remark. Given a finite sum E akbk, we may write

E akbk =

where B is a step function that jumps bk at xk and a is a continuous


function such that a(xk) = ak. Integration by parts now becomes "partial
summation" and we have the formula
aobo + albs + ... aNbN = Ao(bo - b1) + ... + AN-1(bN-1 - bN) + ANbN,
where Ak
First Mean Value Theorem. If a is nondecreasing on [a, b], f E R(a),
and

M = sup{f (x) : x E [a, b]}


m = inf{ if (x) : x E [a, b]},
then there is a number c, m < c < M, such that

b f (x)da(x) = c f b da(x) = c[a(b) - a(a)].


Ja n

If f is continuous, then c = f(xo) for some xo E [a, b]. In particular,

m J b da(x) < I b f (x)da(x) < M b da(x).


a a Ja
Second Mean Value Theorem. Suppose that a is continuous and f is
nondecreasing on [a, b]. Then there is an xo E [a, b] such that
b rxo 6

f(x)da(x) = f(a) / da(x) + f(b) da(x).


f /a 1.0

Corollary (Bonnet's theorem). If g is continuous and f is nonnegative


and nondecreasing on [a, b], then
b

Ja f(x)9(x)dx = f(b) 1.0b 9(x)dx


for a suitable choice of xo E [a, b].
3
Jensen's Theorem and Applications

One of our most useful tools is Jensen's Theorem, which can be used to
relate the distribution of zeros of an entire function to its growth. We prove
Jensen's Theorem using the Gauss Mean Value Theorem.
Gauss Mean Value Theorem. Suppose u is a harmonic function in D.
Then the value of u at the center is equal to the average of the boundary
values of u. That is,

U(O)
= 2w
j u(e1t)
dt

Proof. Form the analytic function f (z) whose real part is u(x, y). Apply
Cauchy's integral formula to evaluate f at zero, then take the real part and
the theorem is proved.
Jensen's Theorem. If f is meromorphic in IzI < R, if r < R, and if
//
f(z) = akzk + ak+izk+1 +--- (ak O 0)
is the Laurent expansion of f around zero, then
rx T
(3.2) I log f (rei8) I dO =1og Iakl+ log n - log +k log r,
21r
r <r P <r Pn

where the zeros of f are z5 = r3ei8' and the poles off are wi = pies#J, not
counting zeros or poles at the origin.
Proof. With no loss in generality, assume f (0) = 1 and R = 1:
f(z) T7rz-wn
(3.3) F(z) = z - zn 11
n
Fln ( 1-2nZ
3. Jensen's Theorem and Applications 7

F is an analytic function with no zeros and no poles. Since log IFS is a


harmonic function in the disk, the Gauss Mean Value (3.1) implies

(3.4) log I F(reie) I dO = log If (0)1 + E log pn - E log r,,,.


Zx J
But 2x f% log I F(re'B)I dO =
2-
f" r log If (re'B)I since I 1_ I = 1 on
oz
IzI =1ifIzoI<1.
Let n(r, f) be the number of poles of f in the closed disk I zI < r, counted
according to multiplicity. Thus n (r, f 'a) counts the number of a-points
[an a-point is a point z satisfying f(z) = a] of f. Let n(r) = it (r, j),
which counts the zeros of f. Let k+ = max(k, 0) and k- min(k, 0), so
that k+ - k- = k. Then,

(3.5) E log
r,. <r
! + k+ log r =
fr
+
log t d{n(t) - n(0)} + k+ log r

with u = log t, du = - t and v = n(t) - n(0). An integration by parts


givest

r+ n(t)
[n(t) - n(0)] log t o+ + n(O) dt + k+ log r
t
Ir n(t) - n(0)
dt + k+ log r.
+ t
We define:

n(t,3) -n(0, f) rr + k+ log r


N(r,1I=_k+logr+Jr dt = log
0+t rn<r rn

N(r,f)-k- log r +
jr n (t, f)-n(0,t) dt = log
r + k-log r.
+ t Pn
Pn<r

Remarks. n(r, f) counts the number of poles of f in the disk Izl < r.
N(r, f) is a useful average of the counting function. We usually normalize
f so that f (O) = 1, and in this case k+ = k- = n(0, f) = it ( 0,.!)
log IakI=0.
As a typical application of Jensen's Theorem, we prove the next result.
tThe 0+ indicates that we integrate from, say, a to r, where e is smaller than the
smallest positive modulus of a zero.
8 Entire and Meromorphic Functions

Theorem. Given z1, z2.... with 0 < Iz1I = rj < 1, there exists a bounded
holomorphic function f in the unit disk whose zeros are precisely the zj if
and only if
E(1 - rj) < oo.

Proof If such an f exists, we may suppose f (0) = 1 and apply Jensen's


Theorem:
x
N (r, Flog * = i! f x log If (reie) I do.
rj<r

Since f is bounded, N (r, 1) is bounded. N (r, j) is nondecreasing,


so s log ( ) < oo. But (exercise) s log (-) < oo if and only if
log(1-r5)<oo.
In the other direction, suppose E log (s) < oo and let

x-z9
Pn(z)=
1-zjz

The P,, form a normal family since I P,, (z) I < 1 for all n and all z E
D. Passing to a subsequence if necessary, the Pn converge uniformly on
compact subsets of D to a bounded holomorphic function f. Can f be
identically zero? Since IP,,(0)l = I H (-zi)I = rI r we have If (0),
rji° r1. Hence If (0) 1 > 0 since log rj' (J) _ E 0° log < 00.
4
The First Fundamental Theorem of
Nevanlinna Theory

Rewriting Jensen's Theorem, we get

(4.1) 27rf log I f (re'B)I dO = log IakI + N (r,


)- N(r, f),

where N is a kind of average number of poles of f.


f
For positive numbers x, let us write

log+ x = max(O, log x) = log[max(1, x)]

so that
1
log x =log+ x - 1og+
x
We now list some simple properties of log+:
(a) log+(x1 x2 ... - xn) < 1og+ x1 + log+ x2 + ... + log+ xn.
(b) log+(xl + x2 + + xn) < 1og+ XI + 1og+ x2 + + 1og+ xn + log n.
In particular,
(c) log+(x1 + x2) < 1og+ xl + 1og+ x2 + log 2.
From (c), we get

log+lx - al < log+Ixl + log+Ial + log 2


log+lxl < log+Ix - al + log+Ial + log 2,
8o that
(d) Ilog+lx - al - log+IxI I < log+Ial + log 2.
10 Entire and Meromorphic Functions

We may write

log If (Te`')I d9 log+If(reie)I dB


"
(4.2) - 2 Jr" log I
f (reie)
dB.

Let m(r, f) = i" f ""log+I f (re") I dB. Then we may rewrite Jensen's The-
orem as

m(r,f)-m(r'f) loglakI+N(r'f) - N(r, f)


or

(4.3) m(r,f)+N(r,f)=1ogIakI+m(r'f)+N\r'fllI.
Notice that N(r, f) counts the poles of f (with a certain kind of aver-
aging) that is the averaged number of times f takes the value oo, while
m(r, f) measures the tendency of f to take the value oo. Hence, the quan-
tity m(r, f) + N(r, f) measures, in some sense, the total affinity of f for
the value oo. Similarly, m (r, -1) + N (r, t) measures the total affinity
of f for the value zero. So the above version of Jensen's Theorem asserts
that the total affinity of f for oc is the same as the total affinity of f for
the value zero, modulo a bounded function of r. The first fundamental
theorem is based on the observation that, for any constant a, the affinity
of f - a for oo is essentially the same as that for f, while the affinity of
f - a for zero is, of course, the affinity of f for a. The theorem states that
m (r, 1 Q) + N (r, 1) is independent of a, modulo a bounded function
of r. Here we use the convention that if a = oo, then f 1. means f.
Fix a E C. Then N(r, f) = N(r, f - a) since z is a pole off if and only
if z is a pole of f - a. From property (d) of log+ we obtain
I m(r, f - a) - m(r, f )l < log+Ial + log 2.
We define
T(r, f) = m(r, f) + N(r, f ).
T is called the (Nevanlinna) characteristic of f.
From Jensen's Theorem, we have

T (r, f) = T (rl f) +.O(r)

T(r, f - a) = 7' r, f -a) + c(r; a),


where O(r) = ¢(r; 0) = log IakI and O(r; a) = log Iak(a)I. Here ak(a) is
the first nonvanishing coefficient in the Laurent expansion of f - a at the
origin. For each a E C, S(r; a) is a bounded function of r.
4. The First Fundamental Theorem of Nevanlinna Theory 11

First Fundamental Theorem of Nevanlinna Theory. If f is mero-


morphic in IzI < R, where 0 < R < oo, then

(4.4) T (r, f a ) = T (r, f) + 4(r; a)


1

where 1¢(r;a)l < log+'jal + Ilog+Iak(z)lI + log2 for all r with 0 < r < R.
Proof. This is simply a rephrasing of Jensen's Theorem using the new no-
tation we have introduced.
Since it is customary to work modulo bounded functions of r, we may
sometimes abuse the notation and write things like T(r, f) = T(r, f - a)
when we mean only T(r, f) = T(r, f - a) + 0(1). The characteristic plays
a central role in the theory of meromorphic (and entire) functions.
5
Elementary Properties of T (r, f)

In this chapter, we present basic properties of the characteristic function.

(5.1) T(r, hIh2) < T(r, hl) +T(r, h2)


(5.2) T(r,hi+h2) <T(r,hl)+T(r,h2)+log2
(5.3) T(r, h2) T(r, hl) + T(r, h2) + 0(1)
(5.4) T(r,±f')=nT(r,f); n,=1,2,....
The proofs of (5.1), (5.2), and (5.4) are simple consequences of the prop-
erties of log+ and of the fact that N(r, f) "counts poles," while (5.3) follows
from (5.1) and the first fundamental theorem, which implies that

+
T (rhJh1) = nT(r, h) E T(r, hj) + n log 2.
0 0

The proof of (5.5) is by induction based on the identity:

n n
> hjhj = ho + h> hjh'-1
0 1

to which (5.1) and (5.2) are applied in the obvious way.


5. Elementary Properties of T(r, f) 13

Definition. To each function A (r) that is positive, continuous, and non-


decreasing on 0 < r < oo, we associate the class A of functions f that are
meromorphic in Izi < oo and that satisfy

T(r, f) < AA(Br)

for positive constants A and B as r --+ R. If f is entire and satisfies this


condition, we say that f is an entire function of finite A-type.
It is easy to verify that A is a field, and we call any such field a A-field.
Remarks. Some A-fields have been studied heavily. The case A(r) =
max(l,rP) is especially important and we denote the corresponding A-field
by A,,. In case f E A,,, we say that f is of order at most p and of expo-
nential type. In case p = 1, we say simply that f is of exponential type.
The intersection, SIP = APB, consists of all functions of order at most
p. We shall discuss order and type in much greater detail later on. Notice
that the next theorem implies that the fields of all meromorphic functions
(i) of order at most p exponential type and (ii) of order at most p are
algebraically closed in the field of all meromorphic functions on C.
Theorem. Each A -field is algebraically closed in the field of all meromor-
phic functions on C.
By this we mean the following. Suppose that f, fo, fl,..., fn are mero-
morphic in C, that fn is not identically zero, that fi E A for j = 0, ... , n,
and that
(5.6) fo + f1f + f2f2 +. _ _ + fnfn = 0.
Then the theorem asserts that f E A. Notice that we do not prove the
existence of an f that satisfies (5.6).
Proof. 1 om (5.6) we may write

T(r,fn)=T (r_1fifi).
.
i=0

Then by (5.3), (5.4), and (5.5) we get

n
T(r, f) < ET(r, fi)+O(1) as r--> oc
i-o
and thus T(r, f) < AA(Br) of appropriate constants A and B.
Definition. Let B denote the ring of all of those functions that belong to
A and are holomorphic in C.
14 Entire and Meromorphic Functions

Theorem. Each P-ring is algebraically closed in the ring of all functions


holomorphic in DR.
This follows directly from the preceding theorm.
We give a detailed proof of the theorem only outlined in [12, p. 54],
reversing the notation for f and g.
Clunie's Theorem. Let f (z) be a transcendental entire function, let g(z)
be a nonconstant entire function, and let <p(z) = f (g(z)). Then

7'(r, 'P) _, 00
T(r, g)
asr -+ o0.
Proof. We may and do assume that f (w) has infinitely many distinct zeros
at w1i w2, --+ 00- (Otherwise, we could replace f by f - A for a suitable
constant A. We are implicity using the fact that if f is an entire function
that takes each complex number a as a value only finitely many times, then
f must be a polynomial. Take this as an exercise. [Hint: Hurwitz's The-
orem, the Casorati-Weierstrass Theorem, and Liouville's Theorem.] This
fact is also a consequence of several of the later results in this book, like
Picard's theorem.) Then, for any integer P,
P
N r, 1 > N r, 1
'P V-1 g(w) - W.
because the averaged counting function N is a monotone increasing func-
tion of the pole set. We also want
P /
(5.8) im.(r,!) >Emf r, 1 ) -0(1).
v=1 \ g(w) - W.
Fix P and let

(5.9) 0<6< 10 min{Iwi-w1I:iAj,


Write

(5.10) f(w) = (w - WO" ... (w - wp)mP (w),


where is nonzero at each wi, and where we also choose b so small that
(w)96 Ofor 0<Iw-w;I<bfor all i-1,2,...,P. Sayl4(w)I>e>0
for all w within 5 of wi, i = 1, 2, ... , P.
Define
n
(5.11) E = U{z : Ig(z) - wig < 5}.
5. Elementary Properties of T(r, f) 15

For z E E, we have
P
(5.12) log+ > E log+ I I- M
1
If (9(x))I _, 1
g(x) - w

for a suitable constant M depending on P, 6, and e. But

1 " 1
d9
27r _x log I g(rei8) - w;

is asymptotic as r -i oo, to

eie) I dB,
2a f Elog+ 9(

where
E; = 10: I9(reie) - w;I < S}
because the integral over the remaining part is less than log+
We conclude, using (5.12), that

T (r,!) > PT(r,g)+0(1)

for any integer P, and the result follows.


6
The Cartan Formulation of the
Characteristic

We begin with some remarks on convex functions.


Lemma. If /3 is a nondecreasing function, then B(t) = f6 /3(s) ds is a
convex function of t.
Corollary. If a is a nondecreasing function, then A(r) = for a(x) is a
convex function of log r, that is, A(et) is a convex function of t.
The corollary follows directly from the lemma since
r
A(et) = J e, a(x) dx = 1 t /3(s) ds, where /3(s) = a(e°).
o x o0

Of course, we assume that a is small enough near 0 so that the integral


exists: In most of our applications, we will have a(x) = 0 for 0 < x < xo
for some x0 > 0. Formally, a condition that B is convex is that B'(t) shall
be nondecreasing, and here B'(t) =#(t). Similarly, rA'(r) = a(r).
Proof. After a simple normalization, it is seen that we must prove that

fox ,0(s) ds < x1 1,D(s) ds for 0 < x < 1.


o

This will be the case if I fo /3(s) ds is nondecreasing. But fo /3(s) ds =


z
fo l3(xt) dt, which is obviously nondecreasing since 0 is. Suppose y > x.
Then we see
foI[fl(Yt)

f0
0
/3(yt) dt - Jl O(xt) dt =
p
- /3(xt)] dt > 0.
6. The Cartan Formulation of the Characteristic 17

Roughly speaking, B is a convex function if and only if B can be repre-


sented as in the lemma. Similarly, A is logarithmically convex if and only
if A can be represented as in the corollary.
We now reformulate the first fundamental theorem following a procedure
due to Cartan.
Theorem. For a certain constant C,
X
leap
T(r, f) = C + 2 N r, f dcc
x
=C+ dW dt
lr {
2-.T
Th l t, f le"W

In particular, T(r, f) is a nondecreasing convex function of log r.


Proof. Apply Jensen's Theorem to the function f - e" P for some real con-
stant Sp:
(6.1)
x
2a 1-"r log
f(e:e) - e"P dO = log Iak((P)I + N r, f 1 4, J - N(r, f).

Thinking of the Laurent series for f - e'w and the definitions of N and
ak(,p), it is obvious that
(i)Ifk<0,ak(cp)=ak and k(W) = k for all WER.
(ii)If k=0,ak(cc)=ak-e'and k(W)=k=0 f o r all W E
(iii) If k > 0, ak(<p) = -e"v and k(cp) = 0. Therefore, since k+(V) = 0
foripE]R,
dcp = 0.

With at most one exception [namely if f (0) = e=' for some cP E R] we have

n (0, ) = 0.
f e"°

Now, by Jensen's Theorem for any constant b, we get


x
(6.2) I log Ib - e"°I d7r =log+ dbl.
x
(Check the cases where IbI > 1 and IbI < 1.)
Let us integrate (6.1) with respect to cp:
(6.3)

2w f * log jf(reie) - e'IPjd8 } dip

_, dV+ "
j.- f., N (rf_C1) d,-N(r,f)
18 Entire and Meromorphic Functions

Applying Fubini's Theorem to the left-hand side (LHS) and using (6.2)
yields

1 " 1

2zr,
"
logIf(re'e)-e"°I d8 dip=
--
1
f m(r,f) dO=m(r,f)

Hence we obtain the result:


(6.4)

m(r, f) + N(r, f) = 27r


,I log f ak(co)I dW +
27r
/ N l r, f I ess)
x x

The first integral is a constant and we are done.


Remark. We could define T°(r) = 2 f' N (r, T - dcp as the Car-
tan characteristic of f, but by a minor abuse of notation we shall write
T(r) = -2rf "xN
- (r, f ,,,) d<p since we customarily work modulo bounded
functions of r anyway.
Interpretation. Roughly, we have

r dT = L(r),

where
L(r) = --
r?
/ n r,
1 eI\ d<p.
JJJ ,r f- ;

What does the function L measure? The function f (rese) may be con-
sidered as a mapping of the circumference OD, = {z : IzI = r} into the
Riemann sphere. n (r, f counts the number of times the point e""
is covered by this map. Hence, f ,,n (r, fir) dcp measures the total
arc length of the unit circumference (counting multiplicity) covered by the
mapping f. In other words, the more heavily the mapping f covers the
unit circumference, the faster T grows. Thus, T measures the covering
properties of f.
We outline here another characteristic, the Ahlfors-Shimizu character-
istic TA (r, f), which behaves much like the Nevanlinna characteristic but
which has as an enlightening geometric interpretation. For full details, see,
from which our presentation is abstracted.
We define
TA(r, a) = MA(r, a) + N(r, a),
where
N(r, a) =
fr nadt,
t
6. The Cartan Formulation of the Characteristic 19

as before, but
1
za 1
MA (r, a) log [w, a] dB,
27r fo
where

[w, a] - I

1+ a x 1+I w
I
aI
Iz and [x, a] =
1+Iaz
1

Now [w, a] is the distance on the Riemann sphere between the points on
the sphere to which w and a correspond via stereographic projection.
It is easy to see that

IT(r,oo) -TA(r,oc) -log+If(a)II < 2 log 2.

By Green's Theorem, one can show that


j2ir j21r I f ;e)p dp
log 1 + If(re)I2 dO+n(r, f) _ X (1+ If(Peie)Ix]z
Denote the right side by A(r), divide by r, and integrate the resulting
identity from 0 to r to get

A(9 dt = N(r, f) log 1 + if (reie)I z dO -log 1 + If (0) 12.


Ir + 2-j
If we make a rotation of the Riemann sphere, which corresponds to the
transformation
1 + aw
w= ,
w-a
where w = f (z), and call the resulting function to = F(z), we can derive
the first fundamental theorem for the Ahlfors-Shimizu characteristic.
Theorem. If f is meromorphic in IzI < R, where 0 < R < oo, then for
every finite or infinite a and r with 0 < r < R we have,

TA (r, 00) = dt = N(r, a) + MA(r, a) - MA(0, a).


f'.

For the geometrical interpretation of TA, note that if S is the Riemann


sphere (of diameter 1), da is the element of area in the z-plane near the
point z, and dA is the corresponding element of area on S, then
do
dA = (1 + Iz12)2

Hence irA(r) is exactly the area (counting multiplicity) of the image on the
.R.iemann sphere of { IzI < r} by w = f (z). The rotation described above
leaves A(r) invariant and replaces MA(r, oo), N(r, oo) by MA(r, a), N(r, a).
We see that TA(r, oo) can be interpreted as an average of the spherical area
of the image of disks under the mapping w = f (z).
7
The Poisson-Jensen Formula

The material we present in this chapter is a specialization of some general


results of potential theory. Our presentation is in the context of analytic
function theory.
We consider functions f holomorphic in DR = {z E C : IzI < R}. We
denote u = Ref, choose r < R, and write z = re'B, w = Re''P.

( 7.1) u(re'B)
-
The Poisson Formula.

Pu =
1

27r
fir
u(Re"')
R2 - r2
RZ - 2rRcos(B - gyp) r2 dip'

The Poisson Kernel.


R2 - r
P = P(z, w) = P(reie, Re"p) =
R2 - 2rRcos(B - gyp) + r2
(7.2)
_ JwI2- Iz12 = Rew + z
w - zJ2 w-z

Proof. Without loss of generality, assume R = 1:

1 2 __ 1- Iz12 IzI2
wJw-z12
w-z+ l-zw (w-z)(1-zw)
By the Cauchy integral formula,

1f + ] dw = f(z)
2ai w -i f(w) [ w 1- x 1 - xw
7. The Poisson-Jensen Formula 21

After parametrizing the integral with respect to the angle ep and taking
real parts, we get the Poisson formula.
Remark. For z = 0, Poisson's formula reduces to the Gauss Mean Value
Theorem,
u(0)
2a fx
u(e'w) dcp.

The Poisson formula is the "invariant form" of the Gauss Mean Value
Theorem in the following sense. Choose Z E D and define
w+z
T,; : D -+ D by Tzw = for w ED.
j+-,w
Let
F=foTzf U=uoTZ.
If = i Yw then w = 1 w and
dw 1- zw dA
w IA-z12 A
so that
1 U(w) dw = 1 r u(A)1 - Iz12 dA
u(z) = U(O) =
tai w 2a: J IA - z12 A '

which is the Poisson formula in different notation.


This generalization of the Gauss Mean Value Theorem has a natural
application which generalizes Jensen's Theorem.
The Poisson-Jensen Formula.
Suppose that f is meromorphic in the disk DR = {z E C : IzI < R}, r < R.
Then,
(7.3)
x
° 2a log If (ReiB) I R2 - 2rR R2
- r2
log l f (re'B) I d`p
cos(e - gyp) + r2
+ E log I BR(z : z1,)I - log I BR(z : w1,)i - k log R
Iz.I <R <R

where B is the Blaschke factor defined by

BR(z a)
R2-
z az

and the z are the zeros of f, the w are the poles of f, and k is the order
of the zero or pole at the origin.
22 Entire and Meromorphic Functions

Corollary. If f is holomorphic, then

log If(reie)I < J Plog IfI

In other words, if f is holomorphic, then log If I is dominated by its


Poisson integral. Notice that for z = 0 the Poisson-Jensen formula reduces
to Jensen's Theorem. One way to prove the Poisson-Jensen formula is to
show that it is the invariant form of Jensen's Theorem. We choose another
proof.
Proof of the Poisson-Jensen Formula. If f is holomorphic and has no zeros,
then there is a branch F of log f and log If I = ReF so that the formula
follows as a special case of the Poisson formula. Also, if A denotes the left-
hand side and p the right-hand side, notice that A(fg) = \(f) +.\(g) and
p(fg) = p(f) + p(g). So it is enough to prove the formula for holomorphic
f. Now consider g(z) = f(z)/IIBR(z : supposing that f(0) 36 0. Since
IBR(w : z.) I = 1 for IwI = R, the formula follows on applying the Poisson
formula to g. And if f (O) = 0, consideration of f (z)/zk leads to the general
case.
8
Applications of T(r)

Theorem. If f is holomorphic and M(r) = sup[{If (z)[ : IzI < r}, then
for anyR>r
T(r) < log+ M(r) < R + TT(R).

Proof. Since f is holomorphic, T(r) = m(r) and


x
m(r) = Z f log+ If (Te`8)I d8 < log+ M(r).

Also, by the corollary to the Poisson-Jensen formula,

2 f.I=R P log if I.
l
log If (reie) <

It is easy to verify that


0<P< R+r
R-r
so that

and the theorem is proved.


We now can prove the following extension of the Liouville Theorem,
Which is left as an exercise.
Exercise. Suppose that f is meromorphic in IzI < oc and that T(r) is
bounded. Then f is a constant.
24 Entire and Meromorphic Functions

Definition. Suppose that A(r) is positive, continuous, and nondecreasing


for r > 1 and furthermore is slowly increasing in the sense that a rT is
bounded. Let A' be the class of all entire functions f such that

log' M(r) = O(a(r)).

It is easy to see that A' is a ring. One of our exercises is to show that
f E A* if and only if f E t, where I consists of those entire functions f for
which T(r, f) = O(a(r)). In the case where A(r) = max(1, rp), we see that
the notions of f being of finite order, of order at most p, and of order at
most p exponential type are the same whether defined by the characteristic
or the logarithm of the maximum modulus. It also follows that each A*
ring is algebraically closed in the ring of all entire functions.
We say that a meromorphic function f in the unit disc is of bounded
characteristic to mean that T(r, f) is bounded. Next we characterize the
functions of a bounded characteristic in the unit disk D.
Theorem. A function f meromorphic in D is of bounded characteristic if
and only if there exist bounded functions A and B, holomorphic in D, such
that f = A/B.
Proof. It is easy to see that if f = A/B, then f is of bounded characteristic.
In the other direction, suppose that f is of bounded characteristic and,
without loss of generality, suppose f (0) = 1. Since T(r, f) is bounded, it
follows that N(r, f) and N(r,) are bounded, so that

log 1
rn
< oo and s log 1 < 00.
pn

Here, as before, {rnes°" } and { pneiw^ } are the zeros and poles of f.
By the theorem of Chapter 4, there exist bounded functions cp and
holomorphic in D, such that the zeros of cp are the zeros of f and the zeros
of ib are the poles of f. Thus, g = is of bounded characteristic and has
no zeros or poles.
It is enough to show that g has a representation g = A/B. Note that
even though g is holomorphic with no zeros and T(r,g) = 0(1), it does not
follow that g is bounded; witness g(z) = 11 Z'
Proceeding with the proof, there exists a function h, holomorphic in D,
such that g = eh. Writing h = u + iv, we have

1 A lu(re`9)1 dO < M < oo for r < 1,


27r J
since m(r, g) and m(r, 9) are bounded, and IgI = exp u. Now for R < 1,
writing
hR(z) = h(Rz) = UR(z) + ivR(Z),
8. Applications of T(r) 25

we have ('
1

27r
J ,,,.
IuR(reie)I dO < M for R < 1.

Now we write
uR = 71R - U-; 4 = maX(0,UR).
Let

pR(z)
IF tip
- dco
27r, _x eiv z
x
aR(z) 27r
f u-R (etP) e "P

e.
+z
- z d/p.
x
It is easy to verify that aR and 6R are holomorphic in D. Let

9R = eRP ` iAR
/ ABR

R
where
AR = exp(-aR),
BR = exp(-/3R),
and AR is an appropriate real constant. Now
ReaR>0 and ReIR>0
so that
IARI<1 and IBRI<1.
Since the families {AR} and {BR} are uniformly bounded, they are
normal families; and since the unit circumference is compact, we may write,
for a suitable sequence of R approaching 1,
limAR = A, limBR = B, IAI < 1, IBS < 1, limAR = A.
It is easy to see that
lim gR = g.
We therefore get the required representation

B
Provided only that B is not identically zero. But
IBR(O)I = exP(-I3R(O))
and
q
R(0) = t dO < M.
The proof is complete.
9
A Lemma of Borel and Some
Applications

Definition. A set E of real numbers has length < t, written IEJ < e,
means that there is a countable union of intervals [an, bn], an < bn, that
contains E and such that

E(bn - an) < e.

Definition. tEJ = inf{t : JEJ < e}.


Lemma (trivial). If JE1J < el and JE21 < e2i then

IE1UE2l<e1+e2.

Bore! Lemma. Suppose that µ(r) is defined for all r > ro, that u is
nondecreasing, and that p(ro) > 1. Then for each a > 1

µ (r + aA(r)

except in a set Ea such that AEQI < a61


Remark. The inequality of the Borel Lemma estimates p at a point greater
than r by using the value that µ takes at r. Intuitively, if the inequality
fails in "too big" a set, the function u will become infinite "too soon" and
will not be defined for all r.
9. A Lemma of Borel and Some Applications 27

P r o o f . Let E = E,, = {r > ro : µ (r + ) > ap(r)}. Let c > 0 be


given. We proceed to define a sequence {rn} and an allied sequence {r,}
by induction.
Let r1 = inf{r : r E E}. Suppose rl,...,rn_1 have been constructed
together with numbers Ek > 0, k = 1, ... , n- I so that El +C2 +- .. En- 1 < E
and rk + Ek E E. Let r' = rk + Ek + µ(Tk+Ck
Now define rn = inf{r : r E E and r > rn_1}. Choose c,, > 0 so that
El + + En < E and so that rn + En E E and proceed.
This procedure will terminate after n steps if and only if there does not
exist an r E E satisfying r > rn. We have

(9.2) rn < rn + En < rn < ra+l

Now (r,, rn+1) fl E is empty by construction.


Claim. There exist only finitely many rn or else rn - oo. For otherwise,
there would be a finite r such that rn -> r and, by (10.2), r, -+ r. However,
by the construction we have for all k

1 1
rk - rk = Ek + 0,
A(rk + Ek) - p(r) >

which is a contradiction.
Claim. E C Un=1[rn, r ,1.
Pick an arbitrary x E E. Let rno = max[rn : rn < x]. This makes sense
by the previous claim. Now rno < x. Suppose x > rno. Then rno+1 < x,
an immediate contradiction. Therefore rno < x < rno, which proves the
claim.
So we have constructed a countable collection of intervals whose union
contains the set E. We estimate E(rn - rn). Notice that

IA(rn) = it rn + En + 1
A(rn+En)) -> ap(rn + En) -> al,(rn)
80 that p(rn+1) > ap(rn). Therefore, p(rn+1) > anp(r1) > an. Hence

1 i
<
(rn - rn)
- En +
U(rn + En) /
E +
N'(rn + En)
But
1 1 -E 00
1

µ(rn + En) {u(rn) n=1


en-1 a-
It now follows that IEJ < as l + e, and the lemma is proved.
28 Entire and Meromorphic Functions

// hypotheses \on µ and a,


Corollary. Under the same

(9.3) µ (r l l + µr)) J < aµ(r)

except for r in a set En, where EQ has logarithmic length < -aas
(By this we mean that Ea = exp Ea, where IEaI < t. We write IEalbg =

Proof. Let µ1(y) = µ(exp y). Then µ (exp (y + v ) ) < ap(exp y) for
y f Ea, where IEa1 < aal by the Borel Lemma. But
1 1
exp >1+
µ(exp y) u(exp y)
so that

µI 1+ I exp, y) < aµ(exp y) for y V Ea,


u(exp y)
and the result follows on writing r = exp V.

Application of the Borel Lemma to Nevanlinna Theory


We already have proved that if f is entire, then

T(r) < log M(r) < R +TT(R) if R > r.

Choose
R=r 1+ T(r) ) 1

to get

log M(r) < (2T(r) + 1)T (,r(i+ T(r)) )


Unless f is a constant, T(r) --+ oo so that 2T(r) + 1 < ZT(r) for large r.
Applying the Borel Lemma we find for entire functions f,
(9.4) logM(r) < 3(T(r))2
except for a set of finite logarithmic length.
Definition. We say that A(r) -' L means that there is a set E of finite
eff
logarithmic length such that Jim A(r) = L as r oo.
r¢E
We attach a similar meaning to expressions like A(r)
= B(r) , etc.
a B(r), A(r)
eff
We have, in effect, proved the next result.
9. A Lemma of Borel and Some Applications 29

proposition. If f is a nonconstant entire function, then

log log log M(r) - log log T (r).


eff

In a certain sense, this says that T(r) and log M(r) have the same size
for most of the r. The log log takes a lot of punishment.
propostion. Suppose that a real function f has a continuous, increasing
derivative on [1, oo], and that lim f (x) = oo. Then
s-00

log log(f(x))
- log log(x f'(x)).
P oof. Write v(t) = tf'(t), and suppose, without loss of generality, that
f (l) = 0. Then
f(x) =f f'(t) dt j u(t).
Hence f (x) < v(x) log x. But for some e > 0, we must have v(x) > ex for
x large, so that

(9.5) f(x) < (v(x))2 for large x.

In the other direction, if y < x, then


fZ
f (x) ? v(t) d > v(y) log (x )
y

Choose x = y + ftv) to get

v(y) < f (11 )


f (y) log (1 + !(v))

Since log(1 + t) > for t near 0 and positive, we have, if y is large,


2t
P(Y) <_ 2f (y)f (y + f v) . Applying the corollary of the Borel Lemma, we
get

(9.6)
v(y) < e(f(x))2.
eff

Equations (9.5) and (9.6) together imply the result.


10
The Maximum Term of an Entire
Function

We will give in this chapter a proof that a suitable entire function can grow
as fast as we please.
Let f (z) = E a"z" be an entire function; ao 0 0

An=lanl
Izl = r.

For each r, the sequence A0, Ajr, A2r2, ... converges to zero. Therefore
we can define

(10.1) B(r) = max(Ao, A, r, A2r2, ... ).

B(r) is called the maximum term for r. A term Akrk is a maximum term
if Akrk = B(r).
Since each Akrk is a nondecreasing function of r (increasing if k 0,
Ak # 0), B(r) is nondecreasing. B(r) is also continuous and unbounded.
We define the rank of the maximum term as

(10.2) u(r) = sup(n : A"r" = B(r)).

It follows immediately that if n < µ(r), then A"r" < AM(r)rµ(r), and if
n > µ(r), then A"r" < A,(r)ra(r). Therefore we also can write the rank of
the maximum term as
p(r) = sup(n : A"r" > B(r)).
10. The Maximum Term of an Entire Function 31

Clearly, ja(r) is a nondecreasing, integer-valued function of r. Let


( g = - log A, if A,, 0
(10.3)
gn=oo if

Since f is entire, we have

(10.4) lim A ° = oo so that lim gn = oo.


n_00 n-00 n
Let cn be the point (n, g(n)) on the plane. From (10.4) it follows that below
any straight line of finite slope there is only a finite number of points cn.

91

C4

GI

C3

1 2 3 4 5 6

The Newton Polygon

This property of the c, enables us to construct the Newton polygon ir(f)


of the entire function f.
We construct the polygon as follows:
Among the segments oc, consider those of minimal slope. From these
segments of minimal slope choose the one that is the longest; denote this
Segment by cc-k,- Repeat this selection procedure starting with the point
ck, to obtain the point ck21 and so on.
The vertices of the polygon are 'yo, -yi, 7i, . , where ryi = ck; _
(ki, g(ki)) for i = 1, 2.... and -yo = cka = co. The x-coordinates of the
vertices -to, -rl,... are called the principal indices.
Let Gn be the y-coordinate of the point on r(f) whose x-coordinate is
n. Let
(10.5) An = exp(-Gn).
32 Entire and Meromorphic Functions

An follows
is called the logarithmic convezifcation of An. Since An = exp(-gn),
it immediately that

Gn = gn if n is a principal index
Gn < gn for all n.

Given r > 0, we have

log April < log B(r) = log Anr", where n = µ(r).


Since
log Ap = -gp
log An = -9n,
we have
p log r - gp < n log r - 9n
or

(10.6) gp?9n+(p-n)logr.
But

(10.7) y=gn+(x-n)logr
is the equation of the straight line through cn with slope log r.
Equation (10.6) says that all points of 7r(f) lie above the line (10.7); this
line is "tangent" to rr(f ). Call this line D,. Thus, µ(r) is the rightmost
point of contact of D,. with rr(f) [because µ(r) is the largest value of n at
which we can have equality in (10.6)]. Hence the values of p(r) are the
principal indices.
Since, for x = 0, (10.7) yields y = gn-n log r = - log Anrn = - log B(r),
it follows that D, cuts the y axis at - log B(r). Two immediate conse-
quences of this are:
(a) Given fl, f2 entire such that rr(fl) = 7r(f2), then

R(r : fl) = p(r : f2)


1 B(r : fl) = B(r : f2)
(b) Among all entire functions, h(z) _ Anzn is the largest function
that has the same µ and B as f.
Let
-An_1
(10.8) Rn
An

We call the Rn the corrected ratios.


10. The Maximum Term of an Entire Function 33

Geometrically, log Rn is the slope of the side of 7r(f) joining the points
whose x-coordinates are n - 1 and n. Rn is nondecreasing and Rn -0 oc
as n --* oo.
Without loss of generality, assume that Ao = ao = 1 [note that p(0) = 0].
From the definition of Rn, it follows that eGn = R1 R2 Rn.
Since p(r) runs through the principal indices, gu(r) = G,u(r), it follows
that
rn(r)
B(r) = Ri.R2...Rn
Taking logarithms,

log B(r) = p(r) log rFlog


p(r)

F
11 (r)
r
Rk = log Rk .

But
Ar)
1og =
j log t dp(t), where p(t) _
Rk <t
1.

Integrating by parts we get:


jlogd,L(t)1or,(t)l+ Ir t) dt
t t o+ t

or

(10.9) log B(r) = f pt) dt,


0

and by the lemma in Chapter 6 it follows that B(r) is a convex function of


log r.
Relation between B(r) and M(r)
From Cauchy's inequality, lanl < r-"M(r), we get
fanIrn < M(r) or
(10.10)
B(r) < M(r).
Remark. M(r) < F(r) A' rn.
Note: Choose p > p(r). Then R,, > r because log R. > log r for p > p(r),
as we see on interpreting log Rp as a slope. Then for q > p we can write
r9-P+1
A r9 = e-Ggrq = eG,-irP-1

Since e-Go-, rP-i < B(r), we have


rq-P+1 r 9-P+1
`44r4 = B(r)1VrP+1 < B(r) /
P RP
34 Entire and Meromorphic Functions

because the slopes of the edges of the Newton polygon are increasing. We
have:
P-1 00
F(r) _ e_c"r"
+ : e-c°r"
0 p
00 / 4-Ptl
p B(r) + B(r)
4 =P
(k)
= pB(r) + B(r)
Rp r
r.
Consequently,

F(r) < B(r) [p + Rpr rl provided p > µ(r).

As a heuristic guide, let us try the choice p = µ(x) for x > r, supposing
for the moment that µ(x) > µ(r). Then

F(r) < B(r) [µ(x) +


r r] < B(r) I µ(x) + 1

Rµ(
_ r1
I
The last inequality is justified as follows: R,,(...) is the slope between (p -
1,Gp_1) and (p,Gp). This slope grows without bound, so there exists an
x so that for p > x we have the inequality. Write x = r + y; y > 0. Then

F(r) < B(r) [ILr+Y)+].


Now write y = so that

F(r) < B(r) [µ (r + t) + t] .


We try to make

t=µ(r+tr\ I r I

µ r+ µ rr+f )
by choosing t.
As a first approximation we choose t = p(r). For that choice z =
t
r + u . Since we want to guarantee that p > µ(r), our actual choice is
p = µ (r + n ; ) + 1. We then get:

F(r) < B(r) I p f r+ pr) + $L(r) + 1


J J
10. The Maximum Term of an Entire Function 35

or
F(r) < B(r) L2µ 1 \\r + A('.) f + 11 .

Using the Borel Corollary (Chapter 9) we find / 1

F(r) < 3B(r)p(r)


eff

so that

(10.11) B(r) < M(r) 3B(r)p(r).


eff

For functions of finite order, say order p, we have

log M(r) = o(r") as r - oo if p' > p.

In this case,

log B(r) = o(rP) as r -i oo if p' > p.


We have
logB(r)=ptt) dt=o(nl )
Jr
andwe know that
< f2, p(t) dt
p(r) log 2 < f2r M(t) dt = o(r°)

Hence,

(10.12) p(r) = o(r°).

Together with (10.11), this implies that if either log M or log B is o(r' ),
0:5 p, then both are. Now, from (10.11) we have

logM(r) log3+logB(r)+logp(r)
eff
=log3+o(r')+p1ogr.
From B(r) < M(r) we get

(10.13) logB(r)
limsup < 1.
r-+oo logM(r)
Also,

log M(r) < log 3 + log B(r) + log p(r)


< 1 + log 3 + log p(r)
log B(r) ;ff log B(r) log B(r)
36 Entire and Meromorphic Functions

Writing F(r) as
n
F(r) Anrn = E AnRn (R)
and using the definition of B(r), we get for R > r:

F(r) < B(R)


(R)n B(r)RR r
so that

(10.14) B(R) < M(r) < B(R)RR-R


r
Let us choose R = r + Brf and again apply the Borel Corollary (9.3):

Br r sWMr
B(r) < M(r) < B r+ r
(r)
B(rr))
B(r) < M(r) < B [ r+ (B(r) + 1).

Hence,
B(r) < M(r) < 2B(r)2.
eff

Taking logarithms twice we obtain

(10.15) log log M(r) log log B(r).


eff

From (10.11) we get

B(r) <M(r) 1.

If p(r) = o(rP), p < oo, then p (r + µ (,) < µ(2r) = o(rP) so that

log B(r) < log M(r) < log B(r) + p log r + constant.

Then
logB(r) < 1 < logB(r) + plogr
log M(r) - - log M(r) log M(r)'
We shall prove that = o(1), which implies:
10. The Maximum Term of an Entire unction 37

Theorem 10.1. If f is of finite order, then


log B(r) - log M(r).

Assertion: iog1M r) = o(l) (unless f is a polynomial).


proof. Otherwise, we can find a sequence {rn } such that rn oo and such
that
logM(rn) < clogrn or M(rn) < rn. But Cauchy's inequality says that

M(rn < rn
rn .
lakl < rn
It follows that ak = 0 for k > c, and hence f is a polynomial.
In general, we have:
Theorem 10.2. If liminfr. ogr = c < oo, then f is a rational func-
tion.
Proof. We again find an increasing sequence {rn}, rn -+ oo such that

T(rn) :5 clog rn

so that
N(rn, f) < clog r.-
We may suppose without loss of generality that f (O) 0, oo. Then:

Irn(t'f) dtc1ogrt

and for s > 0 : f e " EiLL dt < c log r,,. Then


t
r
n(s, f) logs < c log rn or
clog rn
n(s' f) <
log 8

Let rn -- oo, and we see that n(s, f) < c.


Therefore, f has at most c poles al, ... , ak with k < c. Now, multiplying
f by
(x - al) (z - ak) and applying the previous result to the holomorphic
function so obtained, we complete the proof of the theorem.
Suppose we are given M'(r) such that log M'(r) is logarithmically con-
'0". We shall assume that M'(r) can be written in the form

M'(r) = 1 r 2(t) dt; 7(t) increasing.


0
38 Entire and Meromorphic Functions

We already have proved (in Chapter 7) that fo 4 dt, where 'y(t) is in-
creasing, is logarithmically convex. Suppose further that log M'(r) = o(rP)
for some p < oc. Then there exists an entire function f such that
log M(r, f) - log M'(r). Thus, for any such M'(r) there is an entire func-
tion f whose maximum modulus grows essentially like M'(r).
The idea of the proof is the following: Take 1A(t) = ry(t). Define B(r) _
for L(tl dt and draw a Newton polygon associated with p and B. This gives
the An. Let f be the associated function f = E An z". Then

log M(r) - log B(r) = log M'(r).

Since µ(t) must be integer-valued, for the actual proof we take

µ(t) = [ry(t)] where


[x] = greatest integer not exceeding x.

We have then
7(t) - 1 < Fi(t) 5 7(t)
Since log M'(r) = o(rP), it follows that -1 = o(rP) so that µ(r) - log B(r).
It remains to be shown that

log B(r) = log M'(r) + O(log r).

But
-log-
ro
-
< ro p(t) -t 7(t) dt < 0

so that
IlogB(r) - log M'(r) l = O(log r).
Hence log M(r) '- log M'(r), and the proof is complete.
Dropping the hypothesis of finite order, we can still get the following
result: Given that log M'(r) is logarithmically convex, it is possible to find
an entire function f such that M(r) > M'(r).
Proof

log B(r) < log M(r)


log B(r) > log M'(r) + O(log r).

The term O(log r) is easily disposed of by multiplying by a suitable poly-


nomial.
Another result along this line is the following:
10. The Maximum Term of an Entire Function 39

Theorem 10.3. Given any continuous function M'(r), there exists an


entire function f such that M(r) > M'(r).
proof. It is easy to see that any such function M'(r) has an increas-
ing majorant such that log M'(r) is logarithmically convex, and, indeed,
log M'(r) = fo ry=f dt. Now proceed as in the previous theorem and the
proof is complete. Thus, there exist entire functions that grow as rapidly
as we please.
We conclude this chapter with some more estimates on B(r). We know
that:
r
(10.16) B(r) < F(r) < B(r) I r + + 11
A(r)
and that for R > r
F(r) < B(r) R R r.
We can refine our results using the fact that
(t)
log B(r) = r dt.
J t

Since f," 11 dt < fo dt, we get


r r
logB(R)=1RFU(t)) dt>1 RAM dt>µ(r)logR
so that
log B(R)
(10.17) AN <
log R
Choose R = r + Ioe r R Then we have

log B r+ log B r )
k(r) < i
log (1 + og r )
<2logB r+ log rB(r) logB(r).

From the Borel Corollary it follows that


(10.18) µ(r) < 3log2 B(r).
eff
But then (10.11) gives

F(r) < 3B(r) loge B(r).


Hence,
log F(r)
- log B(r).
eff
11
Relation Between the Growth of an
Entire Function and the Size of Its
Taylor Coefficients

Let F be an entire function and M(r) be its maximum modulus for lzi = r.
Suppose A is a positive continuous increasing function for r >_ 1 such that
W r is bounded.
Definition. If log M(r) = O(a(r)), we say f is of finite A-type and write
fEA
Proposition 11.1. Eanzn is of finite A-type if and only if there exists a
constant K such that lanl < `Tn(*) for each n.
Proof. Suppose Eanzn is of finite A-type. Then, since log M(r) = O(a(r)),
there exists a constant K such that M(r) _< eKA(r). Now the Cauchy
inequality gives Ianl < M , which gives the result.
xA(r) xa(r)
Conversely, suppose IanI <_ 2r n = so that
IanIrn < 2-ne"J1(r).
Thus
Elanlr' S eKX(" ,
so that M(r) < eKa(r) and log M(r) = O(a(r)).
Definition. An entire function f is of order< p if for each p' > p there
exist constants A = A(p) and K = K(p) such that
I f(z)I < AeKIa1' for all Z.
11. The Growth of an Entire Function and Its Taylor Coefficients 41

Definition. An entire function f is of order p if it is of order < p but not


of order < po for any po < p.
We have the following facts immediately from the definitions:
log M(r) < log+ A + Kr"
log+ log+ M(r) < log+ log+ A + log+ K + p' 1og+ r + log+ 2.

Thus we have 1o +Iogrm r < p+ o(1) as r -+ oc.


Now let A = lim sup log+ Og+ M(r). Then by the above observation we
r-00
have A < p' for all p' > p and hence A < p. In the other direction, we have
r,K(r)
+
log
< A + o(1) or

Therefore,
ee(X+o(1))1oB*
M(r) < = erxro(').

Thus, for every A' > A and r large (r > ro) we have M(r) < erg' or, more
generally, for all r, M(r) < Aerx'. Hence f is of order < A' for all A' > A
and thus f is of order A. We therefore have proved:
Proposition 11.2. For any entire function f,
log+ log+ M(r)
P = lim sup
r-oo log r

Definition. Suppose f is an entire tunction of order p and that If (z)I <


AeKkI° for all z. Then we say f is of order p, type at most K. The type 7-
of f is the infimum of those numbers K such that f is of type at most K.
Definition. We say f is of finite-type if r is finite; we say f is of minimal-
type if r = 0; and we say f is of mean-type if it is of finite type but not of
minimal-type.
Definition. We say f is of growth (p, r) if either it is of order < p or it is
of order p and type < r. A function of growth (1, r) is called a function of
exponential-type.
Proposition 11.3. For any entire function f of order p,
log+ M(r)
r = lim sup rP
r-oo
The proof is straightforward.
proving the following propositions we shall use the elementary fact
that
RR/e
m, Ry
y=(RJR/e-e R/e
e
42 Entire and Meromorphic Functions

Proposition 11.4. Given an entire function f, let


n log n
a = limsup I .
n-.oo log T;;---

Then a = p.
Proof. Take o > p. Then Janlrn < M(r) < er' for large r. Thus IanI <
r-n er' or log j an j < r°- n log r. Now choose r = (o) ° which is large for
n large. Therefore we have
nn 1 > n log n - n .
log jan l < n - n log or log
0' or v JanI o, a o,

Hence
n log n< n log n
log TG.j
-alogn-a-
o
< a+ 0(1) as n --> oo.

Therefore a < o and, since o can be chosen arbitrarily close to p, a < p.


Now take ,0 > a so that nlogI n < Q for large n and, without loss of
e
generality, for all n. Thus n log n < Q log 1 or nn/ft < or Ian I < ;;;70
Hence IanIrn < it and therefore B(r) < sup. x , where B(r) is the
maximum term of the series Elanlrn (see Chapter 10). If we let y =
= sup v" v = sups v = eR/e =
=/9 v v
and R = L, we have B(r1iP) < supx ;UY
er/Pe. Therefore B(r) < exp(rP/efe), log B(r) < " and log log B(r) <
Q tog r - tog Qa Hence lim supr_ao 1-g+loglog'
y
a(r) <
Q But log M(r) -
log B(r), and so we have p < Q and hence p < a. We therefore must have
P = a.
Proposition 11.5. If f = Eanzn is of order < p, then

r= 1ep lim sup njanIoln.


n-oo

Proof. If f is of order p and type r, take r' > r. Using the Cauchy
inequality we have
Ianl < m( < SL'' for large r. We now minimize the expression on the
right. Its logarithm is r'rP - n log r, and setting the derivative of this equal
1/p
to zero, r'prP-1- r = 0, we choose r = (?P) . Thus Ian < _

() . Hence nJanIP/n < er'p, so that we have limsupnjan1P/n


and therefore lim sup nI an IP/" < er p.
n-.oo
< er'p

n-.oo
11. The Growth of an Entire Function and Its Taylor Coefficients 43

Now take,3 > a limsupn. nlanl1/n. Then for large n we have Ian) <

`/ n/p
so that Ianlrn < (n
*+/P
rn. Hence B(r) <
B(r) < max.., "Op " rs. If we let y = v and R = eflr, we have
n/P
rn or

(e#p)vyrv v
B(rl/P) < max = max I eyr) = max RY = eR/e = epr.

Hence B(r) < efirv and thus lira supr_,,. log B r < 6. But log B(r)
logM(r) so that r < Q and thus r < eP nlanIP/n, which com-
pletes the proof.

Proposition 11.6. The orders and types of an entire function f and of


its derivative f are the same.
This follows easily from the formulas for order and type in terms of the
power series coefficients.
Corollary to Proposition 11.5. Write f (z) = E nz". Then f is of
exponential-type if and only if Eanl;" has a finite radius of convergence.
Proof. We use Stirling's formula n! - n"e " 2rrn.
Il/n N nl Il/n
Now we have nl n a-" N elanll/n Furthermore, the ra-
n I1/n
dius of convergence of EanSn is 19uieu
P1 a" 11/ ". Hence, since n n
elanll/", f is of exponential-type by Proposition 11.5 if and only if
limuP,,-. Ianll/" < oo, as was to be proved.
Definition. To the function f given by f (z) = En z" we associate the
function 4'(w) = Ean za-Fr . Then f is of exponential-type if and only if 0
is holomorphic at oo and 4) is called the Borel transform of f.
Indeed, it is easily seen that we have a one-one linear correspondence
between entire functions of exponential-type and functions fi that are holo-
morphic near oo with 4)(oo) = 0.
Proposition 11.7. The Borel transform off is an analytic continuation
f
of the Laplace transform of f. More specifically, 4'(w) = +°° f (t)e-'w dt
in some right half-plane.
Proof. First, fo a-tw dt = ZT-FT if Rew > 0. Now we estimate
n!
f+°°[f(t) - sn(t)]e'tw dt, where sa(t) is the nth partial sum of Eat'.

We have, for ak =k
00

I/(z) - sn(z)I = 00 Iaklrk = E IaklRk GO k < M(R) E (-r)


n+1 n+1 k
n+1
(r n+l 1
(r n+l Rr, R
= lRl M(R)1- R - \RJ a R-r'
44 Entire and Meromorphic Rinctions

where T' > T. Choose R = 2r so that we get I f(z) - sn(z)I (2e2r'',


Ie-tw I
so that if (z) - s71(z)I < e21 . Now If (t) l < er't and
IZI
= e-t", where
u = Rew; thus we have convergence of the integral and we can interchange
the summation and integration if we take u > 2T'. Thus we have
+00
{ (t)e-tw dt = r +oo (E i t') a-tw dt
1
0 l ` J J
+00 n
Ea:n
J0 nl a-tw an = fi(w)
dt = EWn+l
in some right half-plane, it > 2r', as was to be proved.
As our final result in this chapter we shall give a direct proof that the
order and type of f and f' are the same (Proposition 11.6).
Proof. Let M1(r) = suPe I f'(re'o)I. Then we have from the Cauchy in-
tegral formula, f'(z) = tai if R > r = IzI, that Mi(r) <
M(r) R R Now take R = Ar, where A > 1. Then we get Ml (r) <
.

M(ar) r 1} < M(ar) (A a1 for r > 1. Hence, for r > 1,


a-\

log+ log+ M1 (r) < log+ log+ M(Ar) log AT + log+ log+ la 113 + log+ 2
log r -
log .1r log r log r log r

and thus p1 < p.


In the other direction, supposing without loss of generality that f (0) = 0,
we have f (z) = fo f'(w) dw, where we shall integrate along a ray passing
through the origin. It follows that M(r) < rM1(r). Thus,

log+ log+ M(r) < log+ log+ r + log+ 1og+ M1 (r) + log+ 2
log r - log r

and hence p < p1. Therefore p = p1, as desired.


Similarly for type, we have

log+ ,'l11 (r) < log+ M(ar) AP


rP (Ar)P
+ log+ rP
1A -1i2

Hence 71 < APT for any A > 1 and thus T1 < T. Also, M(r) rMl (r), so
that
to + M(r) < log+ r to + M, r and hence T < T1. Therefore r = T1.
rp - r, +
12
Carleman's Theorem

Let f be holomorphic in Re z > 0 and suppose f has no zeros on z = iy.


Choose p > 0 so that p < (modulus of the smallest zero off in Re z > 0).
Let {zn = rneie^ } be the zeros of f in Re z > 0. Define the following:

E(R) = E(R : f) = E (r, - R2 cos Bn

(proper multiplicity of the zeros taken into account);


n
I (R) = I (R : f) = 1 r log If (it)f (-it)I dt,
2rr (t 2
R12 }

J(R) J(R : f)
;R
1

f "/2
x/2 log If (ReiB)I cos 0 d9,

where the integral is taken along the semicircle of radius R centered at 0.


Then
E(R) = I(R) + J(R) + 0(1).

proof. Let r be the boundary of the "horseshoe" bounded by the semicircle


of radius R, the semicircle of radius p, and the two vertical lines connecting
them:
log f (z) I T2 + R2 dz,
S 27ri Jr L J
46 Entire and Meromorphic Functions

where we assume that f has no zeros on IzI = R and that log f (z) denotes
a branch of log f on F, i.e., log f (z) is some continuous function on F
satisfying exp(log f (z)) = f (z). The proof proceeds by evaluating the
contour integral
r
(12.1) J = 0(1).
Izj=P
Re z>0

Along the negative imaginary axis z = -iy, y > 0, dz = -idy, so


(12.2)
R rR [
27r
logf(-iy) [R2 - y2] dy = 2
J
logf(-iy) - R2] dy.

On z=iy, y>O, dz=i dy, and we have


(12.3)
R
[)2+]
1 !P
dy dy.
tar J R log A W R2 2a<I logf(iy) [y2 - R2
P

On z = Re's (the large semicircle), dz = iReie dO, so we have


a/2
1 1
log f (Reie) (e-2'8 + 1) e'BR dO
tar _w/2 R2
(12.4)
2 x/2
= 2 IrR 2 f x/2 log f (Re'B) cos 0 dO.
The sum of the real parts of (12.2) and (12.3) is
R

tar
log I f (iy)f (-iy) j (2 - R2 dy.

Thus
Re S =1(R) + J(R) + 0(1).
Now integrate S by parts:

u = log f (x) dv =z2(+ R2 I dz

du= f(z) dx v= R2 - z
AZ)
Hence,

S
_, {r 1og fz()rLR2 z 1
finish
1
2ari J (R2
z _ 1
-Z
f'(z)
dz
2ari 2J Istart f(z)
1 x 1 f'(x)
= purely imaginary -
2a- (R2 z f(z) dz.
12. Carleman's Theorem 47

On taking real parts and evaluating the remaining integral by the theory
of residues we find
ReS= E 1

rn
rn
T2--cos0 n .

Remark 1. There is an obvious extension to meromorphic functions.


Remark 2. A formula of Nevanlinna (which stands in the same relation to
Carleman's Theorem as the Poisson-Jensen formula to Jensen's formula)
gives the value of f inside a semicircle from its values on the boundary and
its zeros and poles. (See [5, p. 2].)
Next we present two applications of Carleman's Theorem.
Carlson's Theorem. Suppose f is entire and of exponential-type < it
(i.e., If (z) l < AeBI z1, B < ir) and f (n) = 0 for n = 1, 2,... . Then f - 0.
Remark. Carlson's Theorem also is called the "sampling" theorem. Sup-
pose W(t) is a continuous signal function with support in the interval
[-B, B] with B < 7r. Then the Fourier Transform of .p is
(v(t)e :zt dt = 1
f(z) = O(z) = 1 P(t)e-'zt dt.

vl'2ir R 27 fB
Then, differentiation under the integral and a simple estimate shows that
f is an entire function of exponential-type B < a. If f vanishes on the
positive integers, then f vanishes everywhere by Carlson's Theorem and
therefore W =- 0 as well. By linearity, then, if we know f on the positive
integers, we know it everywhere.
Proof. Suppose f does not vanish identically and satisfies the hypotheses
of Carlson's Theorem. Then we may assume that f has no zeros on the
imaginary axis. Otherwise, translate the plane z H z - e for an appropriate
E > 0. Recalling the notation in Carleman's Theorem, we observe that
E(r) (_j)logR_o(1)
n

_ (R (t2
I(R)
BR
P \
2B
RZ Bt dt <
ifrR B dt < B log R

J(R) 2= = O(1).
aR
Now applying Carleman's Theorem we see

B log R + O(1) < log R.


But a < 1; eventually, this
7r is a contradiction. f must vanish identically.

In Chapter 22 we will present a theorem of Malliavin-Rubel that gives


a very sharp form of Carlson's Theorem.
For the next application, we prove a result that has applications to
1Olynomial approximation theory.
48 Entire and Meromorphic Functions

Theorem. Let a be the difference of two monotone functions on [a, b] and


let f (z) = fQ eZt da(t). Suppose that f (An) = 0 for each An in a sequence
A of numbers in the right half-plane satisfying ERe \an) = oo. Then
f (z) = 0 for all z.
Remark. By use of the Hahn-Banach and Stone-Weierstrass Theorems, it
can be shown that finite sums of the form Ean exp Ant, An E A, are dense
in the space of all continuous functions on [a, b] in the uniform topology
if and only if there is an a (actually, we must allow complex a, but there
is no significant change) for which f exp(Ant) da(t) = 0 for each An E A
implies that f exp(zt) da(t) = 0 for all z. Our theorem above then will
imply that if Er;1 cos Bn = oo, then the sums Ea,, exp(Ant) are dense.
Proof. It is easily verified that f is an entire function. Unless f vanishes
identically, we may suppose that f has no zeros on the imaginary axis,
since we could otherwise translate the y axis. Now f is bounded on the
imaginary axis, say log If I < M there. Hence
I(R)<2M,IR(t2-R2)
dt = O(1).

Now, I f(z)I < Amax(Iell : a < t < b}, where A is a constant. If c >
max(jal, IbI), then If (z)l < Ae" so that J(R) < 0(1) also. Hence E(R) _
O(1).
We have

E(R) > ( - R2 J cos 9n > 4 rE (rn cos On


n
because M1 - - > r if rn < R.
4
Thus
1rn cos On < oo, and on letting R - oo;

we get a contradiction.
13
A Fourier Series Method

The idea presented in this chapter is the following: If f is a meromorphic


function in the complex plane, and if

ck(r,f)= 2x,I (log If(Te:e)I)e-:ke

dO
x

is the kth Fourier coefficient of log If (re") 1, then the behavior of f (z) is
reflected in the behavior of the sequence {ck(r, f)}, and vice versa.
We prove a basic result in Theorem 13.4.5, which characterizes the rate
of growth off in terms of the rate of growth of the ck(r, f) and the density
of the poles of f, generalizing Theorem 1 of [35]. We apply this theorem as
in [35] to obtain estimates for some integrals involving if (z) I and to obtain
information about the distribution of the zeros of an entire function from
information about its rate of growth. Our presentation follows (36].
By these means, we make a study of certain general classes of mero-
morphic and entire functions that include many of the classically studied
classes as special cases. Let A(r) be a positive, continuous, increasing, and
unbounded function defined for all positive r. We say that the meromorphic
function f is of finite A-type to mean that there exist positive constants
A and B with T(r, f) < AA(Br) for r > 0, where T is the Nevanlinna
characteristic. An entire function f will be of finite A-type if and only if
there exist positive constants A and B such that

I f(z)I < exp(AA(BIzI)) for allcomplex z.

If we choose A(r) = rp, then the functions of finite A-type are precisely
the functions of growth not exceeding order p, finite exponential-type. We
50 Entire and Merornorphic Functions

obtain here complete answers to certain basic questions about functions of


finite A-type. For example, in Theorem 13.5.2 we characterize the zero sets
of entire functions of finite A-type. This generalizes the well-known theorem
of Lindelof that corresponds to the classical case A(r) = r". We obtain in
Theorem 13.5.3 a corresponding result for meromorphic functions. Then,
in Theorem 13.5.4, we give necessary and sufficient conditions on A that
each meromorphic function of finite A-type be the quotient of two entire
function of finite A-type. In Chapter 14, we give Miles' proof that these
conditions always hold.
The body of the chapter is divided into five sections, the last two of
which contain the main results. The first three sections are concerned with
various elementary, although sometimes complicated, results on sequences
of complex numbers. The first section discusses the distribution of these
sequences. The "Fourier coefficients" associated with a sequence are defined
in the second section, and several technical propositions involving these
coefficients also are proved there. The third section is concerned with the
property of regularity of the function A, which is closely connected with
the algebraic structure of the field of meromorphic functions of finite A-
type. The fourth section contains the generalizations of the results of [35].
Finally, in the fifth section, the results about the distribution of zeros are
proved.
We urge that, on a first reading, the reader read §4 first and then §5,
referring to §1, §2, §3 for the appropriate definitions and statements of
necessary preliminary results. After this, the complex sequence theory of
the first three sections will seem much more natural.
13.1. An Analysis of Sequences of Complex Numbers
W e study h e r e the distribution o f sequences Z = {zn}, n = 1, 2, 3, ... , with
multiplicity taken into account, of nonzero complex numbers z,a such that
zn --+ oo as n -f oo. Such sequences Z are studied in relation to so-called
growth functions A. We denote by A and B generic positive constants. The
actual constants so represented may vary from one occurrence to the next.
In many of the results, there is an implicit uniformity in the dependence
of the constants in the conclusion on the constants in the hypotheses. For
a more detailed explanation of this uniformity, we refer the reader to the
remark following Proposition 13.1.11.
Let Z = {Zn} be a sequence of nonzero complex numbers such that
limzn=ocasn -- oo.
Definition 13.1.1. The counting function of Z is the function

n(r, Z) = E 1.
IZ,IST
13. A Fourier Series Method 51

Definition 13.1.2. We define


Lr n(t' Z)
N(r, Z) = dt.
t

proposition 13.1.3. We have

N(r,Z) log ICI

proof. Note that

r
E log Ixnl = Jor log (r)t d[n(t, Z)].
Iz+.l<_r

The proposition follows from an integration by parts.


Proposition 13.1.4. We have

n(r, Z) = r N(r, Z).


dr

Proof. Trivial.
Definition 13.1.5. We define, for k = 1, 2,3,... and r > 0,

S(r; k : Z) _
l<r 11 k
klz,.

Definition 13.1.6. We define, for k = 1, 2,3.... and r1, r2 > 0,


S(r1i r2; k : Z) = S(r2; k : Z) - S(r1; k : Z).

When no confusion will result, we will drop the Z from the above nota-
tion and write n(r), S(r; k), etc.
Definition 13.1.7. A growth junction A(r) is a function defined for
O< r < oc that is positive, nondecreasing, continuous, and unbounded.
Throughout this chapter, A will always denote a growth function.
Definition 13.1.8. We say that the sequence Z has finite A-density to
mean that there exist constants A, B such that, for all r > 0,
N(r, Z) < AA(Br).
52 Entire and Meromorphic Functions

Proposition 13.1.9. If Z has finite A-density, then there are constants


A, B such that
n(r, Z) < AA(Br).

Proof. We have
Zr
n(t Z)
n(r, Z) log 2 < dt < N(2r, Z).
1, t

Definition 13.1.10. We say that the sequence Z is A-balanced to mean


that there exist constants A, B such that
AA(Brj) AA(Bkr2)
(13.1.1) IS(r1,r2ik : Z) I _< k +
r1 r2

for all r1, r2 > 0 and k = 1, 2,3,. ... We say that Z is strongly A-balanced
to mean that
AA(Br1) AA(Br2)
(13.1.2) IS(ri, r2i k : Z) I <_ krk + krk
1 2

for all r1,r2>0andk=1,2,3,....


Proposition 13.1.11. If Z has finite A-density and is A-balanced, then
Z is strongly A-balanced.
Remark. Using this result for illustrative purposes, we make explicit here
the uniformity that we leave implicit in the statements of similar results.
The assertion is that if Z has finite A-density with implied constants A,
B, and is A-balanced with implied constants A', B', then Z is strongly
A-balanced with implied constants A", B" that depend only on A, B, A',
B' and not on Z or A.
Proof of Proposition 19.1.11. We observe first that, if r > 0, and if we let
r' = rk1/k, then

(13.1.3) IS(r,r';k)I < 3kr,)


k

To prove this we note that

I S(r, r'; k) I
k
fr tk dn(t),

from which (13.1.3) follows after an integration by parts. Now, for r1,
r2 > 0, let r'1 = rlkl/k and r2 = r2k1/k. Then
IS(ri, r2; k)I _< IS(r', r2; k)I + IS(ri, ri; k)I + I S(r2, rz; k)I.
13. A Fourier Series Method 53

On combining this inequality with (13.1.3), Proposition 13.1.9, and the fact
that k11k < 2, we have

I S(rl, r2; k)I I S(r' , r2i k)I + krk AA(Br1) + krk AA(Br2).
1 2

But, by hypothesis,

IS(ri, r2; k)I < kri AA(Br1) + krz AA(Br2)

for k = 1,2,3,....

Definition 13.1.12. We say that the sequence Z is A-poised to mean


that there exists a sequence a of complex numbers a = {ak }, k = 1, 2, 3....
such that, for some constants A, B, we have, for k = 1, 2,3,. .. and r > 0,
AA(Br)
(13.1.4) jai, i S(r; k : Z) rk

If the following stronger inequality:

AA(Br)
(13.1.5) Iak + S(r; k : Z)I <
krk
holds, we say that Z is strongly A-poised.

Proposition 13.1.13. If Z has finite A-density and is A-poised, then Z


is strongly A-poised.
Proof. The proof is quite analogous to the proof of Proposition 13.1.11,
based on the substitution r' = rk1/k. We omit the details.
Proposition 13.1.14. A sequence Z is A-balanced if and only if it is
A -Poised, and is strongly A-balanced if and only if it is strongly A-poised.
Proof. We prove only the second assertion, since the proof of the first
sssertion is virtually the same. If it is first supposed that Z is strongly
A-Poised, where {ak} is the relevant sequence, then we have

I S(r1, r2; k)I = I S(r2; k) + ak - ak - S(r1; k)I


Iak+S(r1;k)I +Iak+S(r2;k)I,
so that Z is strongly A-balanced.
Suppose now that Z is strongly A-balanced, with A, B being the relevant
Cftstants. Let

p(A) = inf{p = 1, 2, 3, : lim inf Ar) = 0}.


r-oo rp
54 Entire and Meromorphic Functions

Naturally, we let p(A) = oo in the case liminfA(r)rP > 0 as r --> oo for


each positive integer p. For 1 < k < p(A), we have inf r-'A(Br) > 0 for
r > 0. Thus, there exist positive numbers rk such that
A(Brk) 2 A(Br)
<
k

for r > 0 and 1 < k < p(A). For k in this range, we define

(13.1.6) ak = -S(rk; k).


For those k, if there are any, for which k > p(A), we choose a sequence
0 < P1 < p2 < with pj ---* oo as j --* oc such that

A(Bp')
lim = 0.
j moo PjPW

For values of k, then, such that k > p(A), we define

(13.1.7) ak = jAm -S(pj; k).


-00

To show that the limit exists, we prove that the sequence {S(pj; k)}, j =
1, 2,..., is a Cauchy sequence. Let

Aj,m = S(Pj; k) - S(pm; k) = S(pm, pj; k)

We have )
AA(Bpm)
kpkm
+ AA(Bpj
kp;

Since pk pP(a) for p > 1, it follows from the choice of the pj that 0
as j, m - oo. We now claim that
3Ak (Br)
I ak + S(r; k) I <

For, if 1 < k < p(A), then


AA(Br) + AA(Brk) < 3AA(Br).
I ak + S(r; k)I = IS(rk, r; k)I <
krk krk krk ' -
if k > p(A), then
AA(Br) AA(Bpj) AA(Bp)
I ak+S(r; k)I = j-.oo
lim I S(r, pj; k) I < +limsup
kr j_.oo kpj kp'°
13. A Fourier Series Method 55

Definition 13.1.15. We say that the sequence Z is A-admissible to mean


that Z has finite A-density and is A-balanced.
In view of Propositions 13.1.11 and 13.1.13, the following result is im-
mediate,
proposition 13.1.16. Suppose that Z has finite A-density. Then the
following are equivalent:
(i) Z is A-balanced;
(ii) Z is strongly A-balanced;
(iii) Z is A-poised;
(iv) Z is strongly A-poised;
(v) Z is A-admissible.
In Proposition 13.3.3, we give a simple characterization of A-admissible
sequences in the special case A(r) = r°.

13.2. The Fourier Coefficients Associated with a Sequence


We now present the sequence of so-called Fourier coefficients associated
with a sequence Z of complex numbers, and study its properties. We
will use it in §5 to construct an entire function f whose zero set coin-
cides with Z, and to determine some properties of entire and meromor-
phic functions whose growth is restricted. The reason for calling them
"Fourier coefficients" will become apparent on comparing their definition
with Lemma 13.4.2.
Definition 13.2.1. W e d e f i n e , f o r k = 1, 2, 3, ... ,

k
S'(r;k:Z)=k
1Z-1:5r
lrl
Proposition 13.2.2. We have

S'(r; k : Z)I < -N(er, Z).

PrOof. It is clear that IS'(r;k : Z)j < n(r)/k, and we also have
'Cr n(t)
n(r) < dt < N(er).
J(r
56 Entire and Meromorphic Functions

Definition 13.2.3. Let a = {ak}, k = 1,2,3,,.., be a sequence of


complex numbers. The sequence {ck (r; Z : a)), k = 0, ±1, ±2,..., defined
by
(13.2.1) co(r; Z : a) = co(r; Z) = N(r, Z),

Ck(r;Z:a)= rk {ak+S(r;k:Z)}
(13.2.2) 2

- 2 S'(r; k : Z) for k=1,2,3...,

(13.2.3) c_k(r;Z : a) _ (ck(r;Z;a)) for k=1,2,3 ....


is said to be a sequence of Fourier coefficients associated with Z.

Definition 13.2.4. A sequence {ck(r; Z : a)} of Fourier coefficients


associated with Z is called A-admissible if there exist constants A, B such
that
AA(Br) (k
(13.2.4) Ick(r : x; a) I < = 0, 4:1, ±2.... ).
IkI + 1

Proposition 13.2.5. A sequence Z is A-admissible if and only if there


exists a A-admissible sequence of Fourier coefficients associated with Z.
Proof. Suppose that Z is A-admissible. Then, by Proposition 13.1.16, Z is
strongly A-poised. Let a = (ak), k = 1, 2, 3,..., be the relevant constants,
and form {ck(r; Z : a)} from them by means of (13.2.1)-(13.2.3). Now
Definition 13.2.4 holds for k = 0 and some constants A, B since Z has
finite A-density. For k = ±1, ±2, ±3, ..., we have
rIki 1 ,
ICk(r;Z:a)I < Iak+S(r;k)I+2IS(r;k)I.
2

Then an inequality of the form (13.2.4) holds by Proposition 13.2.2 since


Z has finite A-density, and because Z is strongly A-poised with respect to
the constants {ak}.
On the other hand, suppose that (13.2.4) holds. Then
N(r) = co(r) < AA(Br),
so that Z has finite A-density. Moreover,

2 (ak+S(r;k))I =1ck(r;Z:a)+2S'(r;k)
AA(Br) + N(er) < 2AA(eBr)
IkI + 1 2k k
so that Z is strongly A-poised. By Proposition 13.1.16, it follows that Z is
A-admissible.
13. A Fourier Series Method 57

proposition 13.2.6. Suppose that Z and a = {ak } are such that


Ick(r; Z : a)I < AA(Br). Then {ck(r; Z : a)} is A-admissible. In par-
ticular, there exist constants A', B', depending only on A, B, such that

A'A(B'r)
Ick(r; Z : a)I <
IkI+1I

proof. For k = 1, 2, ... , we have

(13.2.5) Ick(r)I < I ak + S(r; k)I + 2IS'(r; k) I


2

and

(13.2.6) 2 Iak + S(r; k) I < Ick(r)I + 1 IS'(r; k)I.

Since co(r) = N(r) < AA(Br), Z has finite A-density. Then, by Proposition
(13.2.2), IS'(r;k)I < (1/k)O(A(O(r))) uniformly for k = 1,2,3,..., by
which we mean that there are constants A", B" for which IS'(r,k)I <
(1/k)A"A(B"r). From our hypothesis and (13.2.6), it then follows that

r'I ak + S(r; k)I = O(A(O(r))) uniformly for k > 0.

Then, by Proposition 13.1.13, we have that

rklak + S(r; k)I S O(A(O(r))) uniformly for k = 1, 2,3....

Then, using (13.2.5), we have

Ick(r)I < O(A(O(r))) uniformly for k = 1,2,3....

Since c_k(r) = (ck(r)), and since Z has finite upper A-density, the propo-
sition follows immediately.

Definition 13.2.7. The quadratic semi-norm of a sequence {ck(r; Z : a)}


of Fourier coefficients associated with Z is given by
1/Z
( a) 121
Ez(r; Z : a) { I ck(r; Z :
lk__ao
00
58 Entire and Meromorphic Functions

Proposition 13.2.8. The Fourier coefficients {ck(r; Z : a)) are A-


admissible if and only if E2(r; Z : a) < AA(Br) for some constants A,
B.
Proof. First, if
Al (Bir)
Ick(r; Z : a)I <
I+1
then E2 (r; Z : a) < AA(Br), where B = Bl and
00 1 2 1/2

A=A1 > (Ikl+1


fk=-00
On the other hand, suppose there are constants A, B for which E2(r; Z :
a) < AA(Br). Then it is clear that Ick(r; Z : a)I < AA(Br), so that by
Proposition 13.2.6, {ck(r; Z : a)} is A-admissible.

13.3. Sequences That Are A-Balanceable


In this section, we are concerned with the process of enlarging a sequence Z
so that it becomes A-balanced. Growth functions A for which this is always
possible are called regular and give rise to associated fields of meromorphic
functions with special properties; for example, see Theorem 13.5.4. The
principal results of this section are Propositions 13.3.5 and 13.3.6, which
give the simple condition that A be regular. In addition, we give in Propor-
tion 13.3.3 a simple characterization of A-admissible sequences of the case
A(r) = rP.
Definition 13.3.1. The sequence Z is A-balanceable if there exists a
A-admissible supersequence Z' of Z.

Definition 13.3.2. The growth function A is regular if every sequence Z


that has finite A-density is A-balanceable.

Proposition 13.3.3. Suppose that A(r) = rP, where p > 0. Then


(i) the sequence Z is of finite A-density if and only if lim sup r-Pn(r, Z) <
0o as r -+ oc;
(ii) if p is not an integer, then every sequence of finite A-density is A-
admissible;
(iii) if p is an integer, then Z is A-admissible if and only if Z is of finite
A-density and S(r: p : Z) is a bounded function of r;
(iv) the function A(r) = rP is regular.
Proof. To prove (i), we have that n(r) = O(rP) whenever Z has finite A-
density. On the other hand, if limsupr-Pn(r) < oc, then n(r) < Ar" for
some positive constant A, so that

N(r) = fo" t-n(t) dt < Ap lr".


13. A Fourier Series Method 59

To prove (ii), suppose that N(t) < At. Then so long as k 54 p, we have
r2
(A + kl) a(k2)1
ill kt dn(t)
(13.3.1) A
\ Ip - k1 rl r2 1
For, on integrating by parts, we have that the integral is equal to
n(k2) - n(ri) + k J r2 tk+i dt.
2 r1 ri

n(r2) < A . = AA(r2)


T r2 r2
2

n(r1) AA(ri)
k
r1
k
T1

kI f + ),
Ip- \
and the inequality (13.3.1) follows. Hence, so long as p is not an integer,
every sequence Z of finite rP-density is rP-balanced.
To prove (iii), suppose that Z has finite rP-density and that p is an
integer. Then, by (13.3.1), we see that all the conditions that Z be A-
balanced are satisfied except for k = p. For this case, the condition that
S(r1,r2; p) be bounded by ri PAA(Br1) + r2 PAA(Br2) for some A, B is
precisely the condition that S(r; p) be bounded, as is quite easy to see.
To prove (iv), we observe first that if p is not an integer, then A(r) = rP
is trivially regular by (ii). If p is an integer and Z has finite rP-density,
let Z' be the sequence obtained by adding to Z all numbers of the form
w`1Z, where wP = 1, but w 0 1. Then Z' has finite rP-density and
S(r; p : Z') = 0 for all r > 0. Hence, by (iii), Z' is rP-admissible, and it
follows that A(r) = rP is regular.
The next two results give simple conditions, both satisfied in case A(r) _
", that imply that A is regular.
Definition 13.3.4. We say that the growth function A is slowly increasing
to mean that A(2r) < MA(r) for some constant M.
If A is slowly increasing, it is easy to show that for some positive number
A A(r) = O(rP) as r -- oo.
p&oposition 13.3.5. If A is slowly increasing, then A is regular.
imposition 13.3.6. If log A(ex) is convex, then A is regular.
The proofs of these results use the next lemma.
60 Entire and Meromorphic Functions

Lemma 13.3.7. The growth function A is slowly increasing if and only


if there exist an integer po and constants A, B such that
AA(Br) for p > po > 0.
(13.3.2) P(+)+)
dt <

If A(r) = ro, then we may choose p0 = [p] + 1.


Proof. Suppose first that (13.3.2) holds. We may clearly suppose that
B > 1. Then
AA(Br) > r0 ,\(t)
prP Jr tP+1 dt > 2Br tP+1
r- Adt > p(2B)PrP
A(2Br)

whenever p > po. Taking p = po, we have A(2Br) > MA(Br), where
M = A(2B)PO,

so that A(r) is slowly increasing. Suppose next that A(r) is slowly increas-
ing, say A(2r) < MA(r). Then
foo A(t) 00
dt = rkr+lr +i dt < PrP(2k)P < M E (M)k.
k-0 J k=0 k=0
Hence, if po is taken so large that 2PO > M, we have an inequality of the
form (13.3.2). In case A(r) = rP, we have M = 2P, and the final assertion
follows.
Proof of Proposition 13.9.5. Let A he slowly increasing, and let Z be a
sequence of finite A-density. Choose po as in the last lemma so that, for
p>po,
co dt < `4 ( r)
1 A(+1+i
PrP
Define
PO
Z' = U W-kZ,
k=0

where no = po+l, w = exp(2iri/no), and w-kZ = {w-kzn}, n = 1, 2, 3, ...


Then we have S(rl, r2; k : Z') = 0 fork = 1, 2, ... , po, since
,pok =0

so long as wk 0 1, and this is true for k = 1, 2,. .. , p0. Hence, to prove that
Z' is A-balanced, we need consider only k > p0. For such k, with r < r',
we have
k
JS(r,r;k:Z)j= I l < k dn(t,Z).
k r< r' `Zn I k J rr t '
13. A Fourier Series Method 61

On integrating by parts, we have


r, I I r'
k dn(t, Z')I < j) k') +
n(r,

n( ,,k ') + k n(t, Z') dt.


Jr
Since Z is of finite A-density and n(r, Z') = (po + 1)n(r, Z), we have
n(r,Z') < A1A(Blr) for some constants Al, B1 by Proposition 13.1.9.
Since A is slowly increasing, we have A(BBr) < A2A(r) for some positive
constant A2 > 0. To complete the proof of the proposition, we have only
to prove that

ffr r n(t,tk+lZ') dt < A'A(B'r)


krk

for some constants A', B'. However,


1r n(t, Z') f A(t) AA2A(Br)
tk+1 dt
- A2 tk+1 < krk

since k > po.


Proof of Proposition 13.3.6. It is no loss of generality to suppose that
r-PA(r) -> oo as r -+ oc for each p > 0, since otherwise A is slowly in-
creasing by Lemma 13.3.7, and then Proposition 13.3.5 applies. Now for
p = 1, 2,3,..., let R1, be the largest number such that

A(RP) = inf A(r)


19 r>0 rP

and let R0 = 0. Then we have that Ra < R1 < R2 < ... , and that R. - oo
as p -+ oo. Further, by Lemma 13.3.7, r-Pa(r) decreases for r < RP and
increases for r > RP. We also have the inequality

(13.3.3) 2PA(r) < 2A(2r) if r > Rp_1i

since, by the above remark,

A(r) A(2r)
rP-1 - (2r)P-1'

Now let Z be of finite A-density. For convenience of notation, we suppose


that N(r) < A(r) and n(r) < A(r), since we could otherwise replace the
function A(r) by the function AA(Br) for suitable constants A, B. We then
claim that

dn(t) < 4A(r)


1
(13.3.4) /r k
r
62 Entire and Meromorphic Functions

if k > 2P and r < r' < Rp.


To prove (13.3.4), we first integrate by parts, replacing the integral by

nrk) 1- k r tk+l dt.


( r')k /
Now
n(r') < A(r) A(r)
(1J)k - (ri)k<- rk
since r < r' and r-kA(r) is decreasing for r < Rp < Rk. Also,

Jr
n(t) dt
tk+1
rA(1
rrP tk+l-P
dt <
A(r) 1 1

rP (k - p) rk-P'
since t-Pa(t) < r-Pa(r) for r < t < r' < R. Thus,
f 1 A(r) A(r) k A(r)
dn(t)
Fk- < + rk + (k - P) rk

We have k/k(k - p) < 2 since k > 2P, and (13.3.4) follows.


We now define Z' as follows. For each zn E Z with R.- j < I zn I < RP,
we introduce into Z' the numbers
1 1
Zn, Zn5 ...,wm-1Zni

where m = m(p) = 2P and w = w(m) = exp(27ri/m). We make the


following assertions:

(13.3.5) n(r, Z')-n(Rp_1i Z') = 2P(n(r)-n(Rp_1)) if Rp_1 < r < RP,

(13.3.6) n(r, Z') < 2Pn(r) if r < Rp,

(13.3.7) N(r, Z') < 2PA(r) if r < Rp,

(13.3.8)

jldn(tz')<kjldn(t) if k > 2P and r < r' < Rp.

(13.3.9)
S(r, r'; k : Z') = 0 if r, r' > Rp and k is not a multiple of 2p.
The assertions (13.3.5) and (13.3.6) follow immediately from the defini-
tion of Z', while (13.3.7) follows from (13.3.6) and (13.3.8) follows easily
13. A Fourier Series Method 63

from (13.3.5). To prove (13.3.9), it is enough to prove that S(r, r'; k : Z') =
0 if R;_1 < r < r' < R;, j > p, and k is not a multiple of 2P. But, in this
case, we have
S(r, r'; k : Z') = 7S(r, r'; k : Z),

where
ry= 1+wk+w2k+...+w(m-Uk'

where m = m(j) = 2' and w = w(m) = exp(2iri/m). Since k is not a


multiple of 2P, k is therefore certainly not a multiple of 21, so that wk # 1.
We then have
_ 1 - wkm
ti
1-wk =0,
and our assertion is proved.
We now prove that Z' is A-admissible. To see that Z' has finite A-
density, let r > 0 and let p be such that 4_1 < r < RP. Then, by (13.3.7)
and (13.3.3), we have that N(r, Z') < 2PA(r) < 2A(2r). To see that Z' is
A-balanced, let k be a positive integer and suppose that 0 < r < r'. Write
k in the form 2Pq, where q is odd. Then, by (13.3.9), S(r, r'; k : Z') = 0 if
Rp < r < r'. Suppose that r < R.P. Then S(r, r'; k : Z') = S(r, r"; k : Z'),
where r" = min(r', RP), by (13.3.9). However,

IS(r, r"; k: Z') I< k rI t dn(t, Z').


J
By (13.3.8), this last term does not exceed

tdn(t),
T

and this, in turn, does not exceed 4r-kA(r), by (13.3.4). Consequently, we


always have IS(r, r'; k : Z)j < 4r-kA(r), so that Z' is A-balanced, and the
proof is complete.

13.4. The Fourier Coefficients Associated with a Meromorphic


Function
In this section, we associate a Fourier series with a meromorphic function
and use it to study properties of the function. As we mentioned at the
beguning, the results of this section are generalized versions of the results
of the earlier paper [35], and the proofs are essentially the same. Our
notation follows the notation of [35] and the usual notation from the theory
0(meromorphic and entire functions. Our presentation still follows [36].
We first recall the results from the theory of meromorphic functions that
gill be needed.
64 Entire and Meromorphic Functions

For a nonconstant meromorphic function f, we denote by Z(f) [respec-


tively W(f)] the sequence of zeros (respectively poles) of f, each occurring
the number of times indicated by its multiplicity. We suppose throughout
that f (O) 0 0, oo. It requires only minor modifications to treat the case
where f (0) = 0 or f (0) = oc. By n(r, f) we denote the number of poles of
f in the disc {z : IzI < r}. By N(r, f) we denote the function
r n(t, f) dt,
N(r, f) = t
J
and by m(r, f) the function

m(r, f) = 2x I f (re`B) I d6,


J x log+
where log+ x = max(log z, 0). We have, of course, that n(r, f) = n(r, W (f) )
and N(r, f) = N(r, W(f)). The Nevanlinna characteristic, which measures
the growth of f, is the function
T(r, f) = m(r, f) + N(r, f).
Three fundamental facts about/T(r, f) are that
(13.4.1) T(r,f)=T(r,f/+loglf(0)1,
fg)<\\T(r,f)+T(r,g),
(13.4.2) T(r,
(13.4.3) T(r, f + g) <T(r, f)+T(r,g)+log 2.
An easy consequence of (13.4.1) is that
1 -W
(13.4.4) loglf(re`B)II dO<2T(r,f)+logIf(0)I-
This follows from (13.4.1) by observing that the first term is equal to
m(r, f) + m(r,1/ f ), which is dominated by T(r, f) + T (r, 11f).
For the entire functions f, we use the notation
M(r, f) = sup{I f (z)I : z = r}.
The following inequality relates these two measures of the growth of f in
case f is entire:
(13.4.5) T(r,f) Slog+M(r,f): R+rT(R,f)
for 0 < r < R. We will use (13.4.5) mostly in the form
(13.4.6) T(r, f) < log+ M(r, f) < 3T(2r, f),
which results from setting R = 2r in (13.4.5).
The following lemma, which is fundamental in our method, was proved
in [9] and [351.t We reproduce the proof of [35] here.
tI have a vague memory of seeing this formula in a paper of Frithiof Nevanlinna
published around 1925, but was unable to find it on a recent search.
13. A Fourier Series Method 65

Lemma 13.4.1. If f (z) is meromorphic in I zI < R, with f (0) # 0,


oo, and Z(f) = {zn}, W(f) = {wn}, and if log(f(z)) _ Ek o akzk near
z-0, then for0<r<R we have

(13.4.7) log I f (re'a)I _ E00 ck(r, f)eike,


k=-oo

where the ck(r, f) are given by


r
co (r, f) = log If (0) I + log r - log
I

Ixn Iwk
IEn,_/r
\\l
Iwnl<r
(13.4.8)
=loglf(0)I+NIr, fI -N(r,f).

For k = 1,2,3,...,

k llk
((
ck(r,f) = akr + 2k A;

- l rsl
lEnl<r
[(Zn)
(13.4.9)
1 (r \ - rgr k k

2k lwn/) r
Iwnl<r

Fork=1,2,3,...,
(13.4.10) c_k(r, f) = CA: (r, f)).

There are appropriate modifications if f (0) = 0 or f (0) = oo.


Remark. Observe that in the notation of §1 and §2, formula (13.4.9) be-
comes

ck(r, f) =2akrk + 2 {S(r; k : Z(f)) - S(r; k : W(f))}


(13.4.11)
- 2 {S'(r; k : Z(f)) - S'(r; k : W(f ))}.

PrOof. We may suppose that f is holomorphic, since the result for mero-
morphic functions then will follow by writing f as the quotient of two
lbbmorphic functions. We may suppose further that f has no zeros on
(z : IzI = r}, since the general case follows from the continuity of both
Aides of (13.4.8) and (13.4.9) as functions of r. Formula (13.4.8) is, of
fturse, Jensen's Theorem, and (13.4.10) is trivial since log If I is real. To
Move (13.4.9), write

Ik(r) = ! I [log f (re'®)J cos(k8) dO


66 Entire and Meromorphic Functions

for some determination of the logarithm, and k = 1,2,3,.... Then, by


integrating by parts, we have

This may be rewritten as

4(r)- 2irki Jf (z) xk - rk } dz.


This last integral may be evaluated as a sum of residues, and on taking
real parts we get the kth cosine coefficient of log If 1. Similarly, considering
the integral
Jk(r) = J [log(f (re`e))J sin(kO) d8,
k
where the integration is now between it/2k and 2w + it/2k, we get the kth
sine coefficient. On combining these, we get (13.4.9).
We now define the classes of functions that we shall study.
Definition 13.4.2. Let A be a growth function. We say that f (x) is of
finite A-type and write f E A to mean that f is meromorphic and that
T(r, f) < AA(Br) for some constants A, B and all positive r.

Definition 13.4.3. We denote by AE the class of all entire functions of


finite A-type.

Proposition 13.4.4. Let f be entire. Then f is of finite A-type if and


only if
log M(r, f) < AA(Br) for some constants A, B and all positive r.
Proof. This follows immediately from (13.4.6).
Note in particular that, if A(r) = rP, then f E A if and only if f is of
growth at most order p, exponential-type.
We also note that by inequalities (13.4.2) and (13.4.4), A is a field and
AE is an integral domain under the usual operations.
The main theorem of this chapter is the following.
Theorem 13.4.5. Let f be a meromorphic function. If f is of finite
A-type, then Z(f) and W (f) have finite A-density and there exist constants
A, B such that
AA(Br) (k = 0,±1,±2,... ).
(13.4.12) Ick(r, f)1<
jkI+1
13. A Fourier Series Method 67

In order that f should be of finite A-type, it is sufficient that Z(f) [or W (f )]


have finite A-density and that the weaker inequality

(13.4.13) Ick(r, f)I < AA(Br)


hold for some (possibly different) constants A, B. Thus, in order that f
should be of finite A-type, it is necessary and sufficient that Z(f) have finite
A-density and that (13.4.12) should hold. It is also necessary and sufficient
that Z(f) have finite A-density and that (13.4.13) should hold.
Proof. The order of the steps in the proof will be as follows. We first show
that if f satisfies an inequality of the form (13.4.13), and if either Z(f) or
W(f) has finite A-density, then f must satisfy an inequality of the form
(13.4.12). We then show that if f is of finite A-type, then Z(f) and W(f)
are of finite A-density, and f satisfies an inequality of the form (13.4.13).
Finally, we prove that if Z(f) [or W (f )] has finite A-density and if f satisfies
an inequality of the form (13.4.12), then f must be of finite A-type.
We shall suppose that f (0) = 1. The case f (0) = 0 or f (O) = 00
causes no difficulty since we may multiply f by an appropriate power of z
and the resulting function still will be of finite A-type. This is because if
liminf(A(r)/ log r) = 0 as r - oo, then by the exercise in Chapter 8 the
class A contains only the constants.
Let us suppose that either Z(f) or W (f) is of finite A-density and that
Ick(r, f )I = O(A(O(r))) uniformly for k = 0, ±1, ±2,.... On considering
the case k = 0, we see that both Z(f) and W (f) have finite A-density. It
is enough to prove that f satisfies an inequality of the form (13.4.12) for
k = 1,2,3,..., since c_k is the complex conjugate of ck. We prove this
exactly as Proposition 13.2.6 was proved. From (13.4.11), we have

Ick(r, f)1:5-r"
)I <- I«k + S(r; k : Z(f)) - S(r; k : W(f))I
(13.4.14) 2

+ 1IS'(r;k: Z(f))I + 1IS'(f;k: W(f))I,


2 2
and

IC,k + S(r; k : Z(f)) - S(r; k : W(f))I < Ick(r, f)I


(13.4.15) 2

+ IS'(r;k : Z(f))I + 1IS'(f; k : W(f))I


Then, by Proposition 13.2.2, for Z = Z(f) or Z = W (f ), we have
IS'(r; k : Z)I = k-'O(A(O(r))) uniformly fork > 0.
BY (13.4.14), it is therefore enough to prove that
IQk + S(r; k :Z(f )) - S(r; k : W(f ))I = k-lr-kO(A(O(r)))
uniformly for k = 1, 2,3,....
68 Entire and Meromorphic Functions

But, we already have from (13.4.15) that


Iak+S(r;k : Z(f)) - S(r;k: W(f))I = r-kO(A(O(r)))
uniformly for such k. Replacing r by r' = kl/kr and observing that r' < 2r,
we have that
Ic k + S(r'; k : Z(f)) - S(r'; k : W(f))I = k-1r-kO(A(O(r)))
Thus, the assertion will be proved if we can show that, for Z = Z(f) and
Z = W(f), we have
I S(r, r'; k : Z)I = k-'r-kO(A(O(r))).
This was proved in Proposition 13.1.11 [see (13.1.3)].
Now suppose that f has finite A-type. Then
N(r, W(f)) = N(r, f) <T(r, f),
so that W (f) has finite A-density. By (13.4.1), the function 11f also has
finite A-type. Hence, Z(f) = W(11f) also has finite A-density. To see that
an inequality of the form (13.4.13) holds, note that

Ick(r,f)I=IZxf x{loglf(Te")I}e-:kedel

:5-f f l log If(re16)Il dO < 2T(r, f) + log If(O)I


x
by (13.4.4).
Finally, suppose that W (f) has finite A-density and that (13.4.12) holds.
If Z(f) has finite A-density, we apply the argument below to the func-
tion 1. Then N(r, f) = O(A(O(r))). It remains to prove that m(r, f) _
O(A(O(r))). However,

m(r, f) <_
27r fir lioglf(reied dB,
which, by the Schwarz inequality, does not exceed
(re`B)112

log If dB
27r Tr
By Parseval's Theorem, we have, for suitable constants A, B,
2 00
loglf(re:e)I dB= Ick(r,f)I2
k=-oo

< A2(A(Br))2

Hence, m(r, f) = O(A(O(r))), which completes the proof of the theorem.


k. (IkIl)2
Specializing Theorem 13.4.5 to entire functions, we have the next result.
13. A Fourier Series Method 69

Theorem 13.4.6. Let f be an entire function. If f is of finite A-type,


then there exist constants A, B such that
A(A(Br)) (k
(13.4.16) Ick(r, f)I < = 0,±1,±2,... ).
IkI + 1

It is sufficient, in order that f be of finite A-type, that there exist (possibly


different) A, B such that

(13.4.17) Ick(r, f)I < AA(Br) (k = 0,±1,±2,... ).


Thus, in order that f should be of finite A-type, it is necessary and
sufficient that (13.4.16) should hold, and it is also necessary and sufficient
that (13.4.17) should hold.
Proof. This result is an immediate corollary of Theorem 4.6 since W (f) is
empty in case f is entire.

Definition 13.4.7. For a meromorphic function f, we define


7, Iq

Eq(r, f) _
{j-jlogIf(re19)I d0j Y

x I

Not ice that if f is entire with f (0) = 1, and if a = {ak} is such that
ck(r,f)=ck(r;Z(f):a),k=0,±1,±2,...,then E2(r,f)=E2(r;Z(f)
a), where this last quantity is the one defined in Definition 13.2.7.
Theorem 13.4.8. Let f be an entire function. If f is of finite A-type
and 1 < q < oo, then
(13.4.18) Eq(r, f) < AA(Br)
for suitable constants A, B and all r > 0.
Conversely, if (13.4.18) holds for some q > 1, then f is of finite A-type.
Proof. If f is of finite A-type, then by the Hausdorff-Young Theorem ([511,
p. 190), the L9 norm of log If (reie)I, as a function of 0, is bounded by the
P norm of the sequence {ck }, where (p) + (q) = 1. By Theorem 13.4.6, this
norm is dominated by an expression of the form AA(Br). Conversely,
using Holder's inequality,

Ick(r,f)I <_ 2; JxllogIf(re's)IIdO,

rr l
<2{ I l o g I f (rei8)IIgd0 } < AA(Br)
1
for suitable constants A, B, and it follows from Theorem 13.4.6 that f has
finite A-type.
70 Entire and Meromorphic Functions

Theorem 13.4.9. Let f be a meromorphic function of finite A-type, with


f (0) 36 0, oc. Then for each positive number a there exist positive constants
a, /3 such that, for all r > 0,
x
(()logIf(reb8)I IdO<1+e.
2_ j exp
(13.4.19)
x A

Remark. We have as a consequence that, for all r > 0,

1 1 dO < 1+ e,
27r J xx If (Tese)I011iBr)

which is somewhat surprising, even in case f is entire, since it is by no


means evident that the integral is even finite.
Proof. There is a number /3 > 0 such that

Ick(r,f)I < M (k = 0, ±1, f2, ... ).


A(Qr) (kI -h 1

Let
F(O) = F(O, r) = log If (Te'B)I.
A(Ar)
Then
F(O) = >ykeike, where yk = ck(r,f)
A(Qr)
We may also suppose that the constant M satisfies

IF(0)I dO < M
21r

by Theorem 4.9. By a slight modification of [49] (p. 234, Example 4), we


know that for any such F there exists a constant a > 0, where a depends
only on M and e, such that
x
1 exp(aIF(0)I) dO < 1 + e,
21r J

from which (13.4.19) follows.

13.5. Applications to Entire Functions


We present in Theorem 13.5.2 a simple necessary sufficient condition on
a sequence Z of complex numbers that it be the precise sequence of zeros
of some entire function of finite A-type. The condition is that Z should
be A-admissible in the sense of Definition 13.1.15. This generalizes a well-
known theorem of Lindelof (see the remarks following the proof of the
13. A Fourier Series Method 71

Theorem 13.5.1, for constructing an entire function with certain proper-


ties from an appropriate sequence of Fourier coefficients associated with
a sequence of complex numbers). We also prove in Theorem 13.5.4 that
A has the property that each meromorphic function of finite A-type is the
quotient of two entire functions of finite A-type if and only if A is regular in
the sense of Definition 13.3.2. Accordingly, Propositions 13.3.5 and 13.3.6
give a large class of growth functions A for which this is the case, including
the classical case A(r) = re'. Even this case seems to be unknown.
We turn now to our first task, the construction of an entire function
f from a sequence Z and a sequence {ck(r; z : a)} of Fourier coefficients
associated with Z. We recall that we have assumed that Z = {zn} is a
sequence of nonzero complex numbers such that zn --+ 00 as n - oc.
Theorem 13.5.1. Suppose that {ck(r)} = {ck(r; Z : o!)}, k = 0, ±1,
12,..., is a sequence of Fourier coefficients associated with Z such that
for each r > 0, E Ick(r)12 < oo. Then there exists a unique entire function
f with Z(f) = Z, f (0) = 1, and ck(r, f) = ck(r) for k = 0, ±1, ±2,....
Proof. We define
,P(pe"°) = E ck(p)eikw.

Since E Ick(p)12 < oo, this defines as an element of L2[-w, 7r]


for each p > 0 by the Riesz-Fischer Theorem. For p > 0, we define the
following functions:

zn p(zn - z)
(13.5.1) B°(z' Z.) _ Izn1 p2 - znz

(13.5.2)
Pp(z) = H B,, (z; zn),
I z. I

w+z
(13.5.3) K(w; z) _
w-z
dw 1
(13.5.4) Q(z) = exp K(w z)4i(w) w
p ,
21ri J1w1=p

(13.5.5)
fp(z)PP(z)QP(z)
We make the following assertions:
(13.5.6)
The function fp is holomorphic in the disc {z : Iz) < p},
and its zeros there are those zn in Z that lie in this disc.
72 Entire and Meromorphic Functions

(13.5.7) ff(0) = 1.

(13.5.8) ff r < p, then ck(r,f) = ck(r).


Now (13.5.6) is clear from the definition of fP. Also,

Iznl
fP(0) = PP(0)QP(O) = QP(0) R
P
IE^1<P

However,

dw
QP(0) = exp
211ri ,
O(w) = exp{co(p)} = H nl
IW1=P IznISP

and it follows that ff(0) = 1.


To prove (13.5.8), we see by 13.4.1 that it is enough to show that, near
z = 0,
log ff(z) _ E akzk,
k=1
where the ak are such that a = {ak) and ck(r) = ck(r; Z : a). That is,
near z = 0,

(13.5.9) = E kakzk-1.
r (z) k=1

We now make this computation. First we have that


BP' (z; zn) Izn12
- p2 xn - 1

BP(z;za) (zn - z)(p2 - znz) p2 - xnz zn - z


00 (2n 00 k
zk-1 - zk-1
-p2
k=1 k=1 zn

Thus,
PD(z) - Uk,PZk-1
near z = 0,
1'(z) k=1
where
(.)k (i)C
Uk,P =
I ..I_P
>
I .I_P
F o r k = 0,1, 2, ... , we write w = pe"° and ck(p)e'k`° = I kwk. Then by the
definition of ck(p) we see that
Po = N(p, Z)
13, A Fourier Series Method

and

(2n)kl
f2k 2nk +
2k E
Iz, I<P
zn)k -
(k = 1,2,3,...

Then

4(w) = N(p, Z) + >{SZkwk


00 + S1kwk}
k=1
00 (i)k}
N(p, Z) + E {kwk +
1ZkP2k

k=1

so that
2
-'(w)KK(w, z) w 27ri
twz)2 dw,
27ri KIWI=P (w

where

But
1
wk dw = kzk-1 for k = 0,1,2,...,
21W-i
I,,.I=p (w - z)2
and
1 r rl 1 dw = 0 for k=0,1,2,....
2arz IwI=P w (w - z)2
Hence,
00

Q' (z)
QP(x)

= c Vk zk_1,
k=1

where

Vk,P = 20k = (xk + k n) k - \ p2) k


Izn I SP

Hence, near z = 0 we have

f°(z) _ P(z)
00
kakz k 1

fP(z) PP(z) + QP z () k=1

and (13.5.9) is proved

It next follows from (13.5.6)-(13.5.8) that


(13.5.10) if p' > p, then f,,' is an analytic continuation of fP
74 Entire and Meromorphic Functions

For if we define for Izi < p

F(z) = f,(z)
fa(z)'
then
Ck(r, F) = ck (r, fP') - Ck(r, fP) = Ck(r) - ck(r) = 0
for 0 < r < p, and therefore IF(z)l = 1. On the other hand, F(0) = 1, and
it follows that F is the constant function 1.
We now define the function f of Theorem 13.5.1 by setting f (z) = f f(z)
if p > Izj. It is clear that f is entire and, by (13.5.6), that Z(f) = Z. Also,
f(0) = 1, and ck(r, f) = ck(r, fp) for p > r, so that ck(r, f) = ck(r). An
argument analogous to the one used in proving (13.5.10) proves that f is
unique, and the proof of the theorem is complete.
We now characterize the zero sets of entire functions of finite A-type.
Theorem 13.5.2. A necessary and sufficient condition that the sequence
Z be the precise sequence of zeros of an entire function f of finite A-type
is that Z be A-admissible in the sense of Definition 13.1.15, that is, that Z
have finite A-density and be A-balanced.
Proof. If Z = Z(f) for some f E AE, then by Theorem 13.4.6 the se-
quence {ck(r, f)} is a A-admissible sequence of Fourier coefficients associ-
ated with Z and thus Z is A-admissible by Proposition 13.2.5. Conversely,
suppose that Z is A-admissible. Then by Proposition 13.2.5 there exists a
A-admissible sequence {ck(r)} associated with Z. Then by Theorem 13.5.1
there exists an entire function f with Z = Z(f) and {ck(r, f)} = {ck(r)}.
Then by Theorem 14.4.7 and the fact that {ck(r)} is A-admissible, it follows
that f E AE, and the proof is complete.
Remark. This theorem generalizes a well-known result of Lindelof [201,
which may be stated as follows.
Theorem 13.5.3. Let Z be a sequence of compex numbers, and let p > 0
be given. If p is not an integer, then in order that there exist an entire
function of growth at most order p, finite-type, it is necessary and sufficient
that there exist a constant A such that n(r, Z) _< Are. If p is an integer.
it is necessary and sufficient that both this and the following condition be
satisfied for some constant B:

< B.
Iz,.I<r zn

This result follows immediately from Theorem 13.5.2 and the character-
ization of rP-admissible sequences given in Proposition 13.3.3. Our result
shows that, in general, the angular distribution of the sequence of zeros
13. A Fourier Series Method 75

of a function, and not only its density, is involved in an essential way in


determining the rate of growth of the function.
We turn now to the second problem of this section, that of determining
when A is the field of quotients of the ring AE. We first prove the following
result.
Theorem 13.5.4. In order that a sequence Z of complex numbers be the
precise sequence of zeros of a meromorphic function of finite A-type, it is
necessary and sufficient that Z have finite A-density.
Proof. The necessity follows immediately from the fact that if f is a mero-
morphic function, then N(r, f) < T(r, f). For the sufficiency, we remark
first that the method used in proving Theorem 13.5.1 can be used to con-
struct suitable meromorphic functions. Indeed, suppose that we are given
two disjoint sequences Z, W of nonzero complex numbers with no finite
limit point and constants ryk, k = 1,2,3,..., such that the coefficients
defined by
co(r) =N(r, Z) - N(r,W),

ck(r)= 2 {yk+S(r;k:Z)-S(r;k:W)}
- 2 {S'(r; k : Z) - S'(r; k : W) (k = 1, 2,3.... )
c-k(r) =(ck(r)) (k=1,2,3.... )
satisfy E Ick(r) 12 < oc for every r > 0. Then, by defining

Dp(z; w..) = J B,(z; w..)


Iwnl<p

and
PP(z)Qp(z)
fp(z) = Dp(z)
one can show, as in Theorem 13.5.1, that the meromorphic function defined
b3' f(z) = fp(z) for p > Jzi has zero sequence Z, pole sequence W, and
Fourier coefficients {ck(r)}. It is therefore enough to prove that, given a
sequence Z of finite A-density, there exist a disjoint sequence W of finite
A-density and constants ryk, k = 1,2,3,..., such that the ck(r) satisfy
Ick(r)l < AA(Br) for some constants A, B and all r > 0. For then, by
the first part of the proof of Theorem 13.4.5, the ck(r) must satisfy the
stronger inequality
AA(B'r)
Ick(r)I : Jkl + 1 (r > 0)

for some constants A', B', so that the function f synthesized from the ck(r)
Must be of finite A-type by Theorem 13.4.5.
76 Entire and Meromorphic Minctions

Supposing now that Z = {zn } has finite A-density, we define W = {wn }


by wn = zn + En, n = 1, 2, 3, ... , where the En are small complex numbers
so chosen that Iwn I = Izn I, n = 1,2,3,..., all of the numbers wn and zk
are different, and such that

IEnI A(O).
<
Iznl

Then, N(r, W) = N(r, Z) so that W has finite A-density. Hence,

I S'(r; k : Z)I = k-'O(A(O(r)))

and
IS'(r; k : W)I = k-'O(A(O(r))) k=1,2,3,-...
We define
/n`k- (1)k}
rk k>
It remains to prove that

2I'Yk+S(r;k:Z- Sr;k:W
uniformly for k = 1, 2,3,.... Now

2 Ilk+S(r;k:Z)-S(r;k:W)I
rk ()k}
2 k
nl>r
n/ k -
rk

2 Ir [
1 [ (wn)k - (zo)k
IznI>r
(wnzn)k <21
[-` I(wn)k - (zn)_I
InI>r
Izn I2k

However, I(wn)k - (zn)kl C klEnllznik-1, so that we have

2 Iyk+S(r;k:Z)-S(r;k:W)I

IEnl
<1 < ' A(0) < A(r).
rk
2 IE
Iznl j
2 2
1znl>>r Jz,. >r Iznl
13. A Fourier Series Method 77

Theorem 13.5.5. The field A of all meromorphic functions of finite A-


type is the field of quotients of the rings AE of all entire functions of finite
A-type if and only if A is regular in the sense of Definition 18.3.2, that is,
if and only if every sequence of finite A-density is A-balanceable.
proof. First, suppose that A is regular and that f E A. Then Z(f) has finite
A-density by Theorem 13.5.3. There then exists a sequence Z' D Z(f) such
that Z' is A-admissible. [We may suppose, by the remarks preceding the
proof of Theorem 13.4.5, that f (0) 54 0, oo]. Then, by Theorem 13.5.2,
there exists a function g E AE such that Z(g) = Z. Since we have then
that Z(g) C Z(f), the function h = g1 f is entire. However,

T(r,h) <T(r,g)+T (r, )=T(rg)+T(r,f)-logjf(O)j


f
by (13.4.1) and (13.4.2), so that h E AE, and f = g/h is the desired
representation.
Conversely, suppose that A = AE/AE. Let Z have finite A-density.
Then, by Theorem 13.5.3, there exists a function f E A with Z(f) = Z.
We write f = g/h with g, h E AE. Then Z(g) is A-admissible, and Z(g) D
Z(f) = Z, and we have proved that A is regular.
14
The Miles-Rubel-Taylor Theorem
on Quotient Representations of
Meromorphic Functions

Let f be a meromorphic function. In this chapter we describe the work


of Joseph Miles, which completes the work in the last chapter concerning
representations of f as the quotient of entire functions with small Nevan-
linna characteristic. Miles showed that every set Z of finite A-density is
A-balanceable. As a consequence of this and the work of Rubel and Taylor
in the last chapter, there exist absolute constants A and B such that if
f is any meromorphic function in the plane, then f can be expressed as
fl/f2, where f, and f2 are entire functions such that T(r, f;) < AT(Br, f)
for i = 1, 2 and r > 0. It is implicit in the method of proof that for any
B > 1 there is a corresponding A for which the desired representation holds
for all f. Miles' proof is ingenious, intricate, and deep. Miles also showed
that, in general, B cannot be chosen to be 1 by giving an example of a
meromorphic f such that if f = fl/f2, where fl and f2 are entire, then
T(r, f2) i4 O(T(r, f)). We do not give this example here.
In the previous chapter, namely in Propositions 13.3.5 and 13.3.6 using
Theorem 13.5.2, we have obtained the above theorem for special classes of
entire functions. Results in this direction for functions of several complex
variables appear in [16], [17], and [10]. Quotient representations of func-
tions meromorphic in the unit disk are discussed in [2]. The presentation
below follows Mile [24].
We state the theorem.
Theorem. There exists absolute constants A and B such that if f is any
meromorphic function in the plane, then there exist entire functions fl and
14. The Miles-Rubel-Taylor Theorem on Quotient Representations 79

f2 such that f = f, /f2 and such that T(r, f;) < AT(Br, f) for i = 1, 2 and
>0.
Suppose Z = {zn } is a sequence of nonzero complex numbers with jzn I -->
co. We include the possibility that zn = z,n for some n # m. As in the
previous chapter, let

(14.1) n(r, Z)
IZn Kr

and

(14.2) N(r, Z) = f r n(tt Z) dt.


0

It was shown in the previous chapter that the following lemma is suffi f
dent to establish the theorem.
Lemma. Suppose Z = {zn} is a sequence of nonzero complex numbers
with IzzI -- oo. If A(r) = max(1, N(r, Z)), then there exist absolute con-
stants A' and B' and a sequence Z = {-;n} containing Z (with due regard
to multiplicities) such that
(1) N(r, Z) < 5A(4r) r > 0
and,
(ii) for j = 1, 2, 3.... ands > r > 0,

A'A(B'r) A'A(B's)
(;)' 0 8

The argument of the last chapter which shows that this lemma is suffi-
cient to prove the theorem may be summarized as follows. Without loss of
generality we may assume f has infinitely many poles and that f (0) 54 oo.
Let Z be the sequence of poles of f . Recall from the last chapter that
condition (i) of the lemma says that Z has finite density with respect to
the growth function A(r) and condition (ii) says that Z is balanced with re-
spect to the growth function A(r). Let A1(r) denote an arbitrary increasing
unbounded function defined on (0, oo).
In Theorem 13.5.2 we characterized the zero sets of entire functions
such that T(r, ¢) < aa1(lr) for some constants a and Q and all r > 0
as those sets Z* which both have finite density and are balanced with
respect to Al (r). This characterization combined with the above lemma
guarantees the existence of constants Al and B and of an entire function
f2 having zeros precisely on the set Z (counting multiplicities) such that
T(r, f2):5 A1A1(Br) for all r > 0. Hence,
(14.3) T(r, f2):5 A1N(Br, Z) < A1T(Br, f)
80 Entire and Meromorphic Functions

for r > ro(f). Letting fl = f2 f, we see that fl is entire and that T(r, fl) <
(A1 + 1)T(Br, f) for r > ro (f ). For an appropriate complex constant c,
0 < Icl < 1, we have for i = 1, 2 that

(14.4) T(r,cf=)=0 r<ro(f)


and

(14.5) T(r, cf:) < T(r, f;) r > 0.

Letting A = Al+1, we see that f = cfl/cf2 is the desired representation.


It is implicit in the methods of the last chapter and in the proof of the above
lemma that A and B are absolute constants and that to each B > 1 there
corresponds an A for which the representation holds for all f .
We now prove the lemma. For each integer N we let

ZN=Zl{z:2N <IzI<2N+1}.
If ZN # 0, we relabel the elements of ZN as simply z1, z2,.. -, zk, with
each number being listed in this sequence as often as it appears in Z. For
1 < n < k we define pn E (0,1] and 0,. E (0,27r], so that zn = 2N+PReie
We do not indicate in the notation the obvious dependence of k, zn, pn,
and On on N. We let
k oo
(14.6) YN(8) _ -2 E E 2j(1+Pn)
n=1 1i=1 I
and

(14.7) hN(0) = RgN(B).

Clearly,

k o0

IhN(0)I :5 2 {2_ci+}Pn)
(14.8) n=1 j=1
< (n(2N+1 , Z) - n(2 N, Z))
Letting

(14.9) fN(O) = hN(O) + 2(n(2 N+1' Z) - n(2N, Z)),

we have

(14.10) 0 < fN(O) < 4(n(2 N+1' Z) - n(2N, Z)).


14. The Miles-Rubel-Taylor Theorem on Quotient Representations 81

We now define sets ZN for all integers N. If ZN # ¢, we have from


(14.10) that
2"
(14.11) 0<1 fN(O) dO < 4(n(2N+1, Z) - n(2 N, Z)).
fo

If [ ] denotes the greatest integer function and LN = [Zx fo"fN(0) do],


we define for 0 < n < LN a monotone increasing sequence 9'n in [0, 27r] by
choosing 9'n to satisfy

rAn
(14.12) 2- oj fN(O) dO = n.

In addition, we let BLN+1 = 2ir. For 0 < n < LN, we let z'n =
and Z' = {2N-lesen : 0 < n < LN}. As before, we do not indicate the
dependence of 61. and en on N. If ZN = 0, we let ZN = 0 and for that
value of N do not define LN or numbers 9n and z;,.
We let Z' = UNZ'N and Z = Z U Z'. From (14.11) and the definition of
Z it follows that n(r, Z') < 4n(4r, Z) and hence that n(r, Z) < 5n(4r, Z)
for all r > 0. From this fact it is immediate that Z satisfies condition (i)
of the lemma.
We now consider a positive integer j and a value of N for which ZN ¢.
Let ZN = {z1, z2,. .. , zk}. From this point until inequality (14.27), we
regard j and N as fixed. Although many of the quantities to be defined
(S, T, P, U0, UE7 V0, and Ve) depend on both j and N, for simplicity we
suppress this dependence from the notation.
A key step in showing Z satisfies condition (ii) of the lemma is to estab-
lish

k
(14.13) < 48j2-i(N-1).

1I

We first observe that


(14.14)
k 1 1 12,,
+ 2if- dO
nZn
k 1 21r k eii(8-e,.)
- (L)j
n=1 2irJo
e-sje2-i(N-1)
n=1
2i(l+v..)
d9

k 1 i k
2-i(N+p.)e-il(O.) = 0.
z
82 Entire and Meromorphic Functions

Thus (14.13) is equivalent to


(14.15)
()I f2lr
-ij9 2 )(N-1) fN(B) d8 < 48j2 j(N-1)
I Z"n I
E xn 21r
1
e

That the quantity on the left side of (14.15) is small can be seen intu-
itively from the observation that the sum is essentially an approximating
R.iemannn sum for the integral. We first estimate
=2t'_, (I\j- 2r r2 e-1792
-7(N-1) fN(O) dB

(14.16)
2x
= 2-j(N-1) cos jOn - 2a I0
fN(B) cos jB dB
112;,1=2N-,

The function cos jB decreases from 1 to -1 on [ix , m i 1 "] for m even,


0 < m < 2j - 1, and increases from -1 to 1 on [ *'!x , m' 1 x 1 for m odd,
O<m<2j-1. Letl,n= (Jx, "`+jlx) forO<m<2j-1. Defines
to be the set of all n, 0 < n` < LN, such that [88, B'+1] is not contained
entirely in some IT1f 0 < m < 2j - 1. Since each point mir/j belongs
to [B'n, B,,+1] for at most two values of n, we see that S has at most 4j
elements. Let T be the set of all n, 0 < n < LN, such that either n or n -1
belongs to S. Thus T has at most 8j elements.
Let P = {0,1, ... , LN} - T. If n E P, it follows that [Bn+l, On] and
[91L1 On+1] belong to the same interval for some m, 0 < m < 2j - 1. Let
U,, be the set of all n E P such that [Bn+l, On] and [On, 0n+1] are contained
in In for some odd m, and let U. be the set of all n -1 such that n E P and
the intervals [B'n_1, On'] and [On, Bn+1] are contained in Im for some even m.
Certainly U,, fl Ue = 0, for if n E Uo fl UE7 then [B'n, Bn+1] is contained in I,,,
for both an odd value of m and an even value of m. Letting U = U,, UUei we
conclude that U and P have the same number of elements. Since 0 V P, we
see that -1 Ue and hence U C {0,1, ... , LN}. Thus, {0,1, ... , LN} - U
has at most 8j elements.
If n E P is such that u E Ua, then cos jB is increasing on [Bn, Bn+1] and
hence, from (14.12),
1
(14.17) coS jAn < I fN(0) cos jB dB.
27r Bn

If n E P is such that n- 1 E Ue1 then cos jB is decreasing on [Bn+l, On'] and


thus

(14.18) cos jOn < 1 r fN(0) cos jB d8.


2w Bn-,
14. The Miles-Rubel-Taylor Theorem on Quotient Representations 83

The definition of U together with (14.17) and (14.18) implies

(14.19)
nEP
cos j0n <
nEU
- j6.+i
1

;,
,

fN (0) cos j0 d8.

It follows that
(14.20)
zx
cos jO'n -
2-
1
f fN(6) cos j9 d8
IYn1=2N-1
0

_ ECosje,+E Cosje",
nEP nET

-
nEU
-(
21r
60

e;
+2
fN(6) cos jO d9 -
O<n<LN
2-
8;.+

Bn
fN(0) cos j9 dO
niU

1 Jf
Bn+l
COSOn - E 27r at.
fN(9) cos jO dO
nET O<n<LN
nU
<8j+8j = 16j.
We now obtain a lower bound for
j21r
(14.21)
IV .
- 2a fN(0) cos j8 d8.

Let Ve be the set of all n E P such that [9n_1, e'n] and [B'., n+1] are
contained in I,,, for some even m, and let Vo be the set of all n - 1 such
that n E P and the intervals [On_ 1, On'] and [en, n+1] are contained in In for
some odd m. Using the same reasoning as before, we see that V. fl Vo = 0
and hence, if V = Ve UV0, then {0, 1, ... , LN} -V has at most 8j elements.
If n E P is such that n E Vef then cos j0 is decreasing on [8n, Bn+1] and

(14.22) cos j8'n > 12a fal.


gn+l fN(9) cos jB dB.

If n E P is such that n -1 E V0, then cos j8 is increasing on [8n_1, On] and

an
(14.23) cosjen > fN(9) cos j9 d8.
aJ Bn-1

NM (14.22), (14.23), and the definition of V,

(14.24) 2
nEP
cos j8'n >-
nEV
1

2A
10.10'
+1
fN(9) cos j8 d8.
84 Entire and Meromorphic Functions

Thus
(14.25)
1 2x
Cos j0, - f fN(0)Cos jO dO
21r o
I z:,1=2 N -'

_ cosjOn + cosjOn
nEP

- f nET

nEY0<n<Lv
E 2J
21r
fN(6) cosjO dO -
n V
A*+ fN(8) cos jO d8
Bn

>
nET
-2
cos jAn -
0<n<LN
1 Bnt:

fon'
fN(0) cos j8 d8

nqY
> -8j - 8j = -16j.
Combining (14.16), (14.20), and (14.25), we conclude that
(14.26)
7 f2-
J e- +je2 -j(N-1) fN(8) dB - 16j2- j(N-1)
<
1=2Nzn
27r

The same discussion applies to the imaginary part of the above quantity.
The only minor modification is that we must divide [0,2w] into 2j + 1
subintervals on which sinjO is alternately increasing and decreasing. Since
2j + 1 < 4j, this causes no difficulty. We obtain

fzx
e-'j''2-j(N-1)fN(0) dO < 32j2-j(N-1).
27r

Combining (14.26) and (14.27), we obtain (14.15), which in turn establishes


(14.13).
We are now in position to show that 2 satisfies condition (ii) of the
lemma. Suppose that s < 8r. We then have trivially for all positive
integers j

1 (i)1<n(sZ) ,

(14.28)
7 r<Iznl<e z'n - 7 rj
< n(8r, Z) < N(8er, Z) < 5A(32er)
jrj - jr.? jr'
Suppose that 8r < s. In this case there exist integers q1 and q2, q1 5
q2 - 3, such that
(14.29) 2q' < r < 2q'+1 < < 2 q' < s < 2 q'+i
14. The Miles-Rubel-Taylor Theorem on Quotient Representations 85

Thus, for
(14.30)
9

T<l<8 'Zn <IZn9, +4 ( Y 'Znq2-ql-1


+
1 J +- 1 1

zn
v=2 2vl+°<Iznl<2q,+..+1 Zn Izn' l=2°l}D-'

1 11 )i + 1 1
1

xn xn
2v2 < zn1<s 3 lz;.i=2q2-.

1
+
I 1.11=2q2
(;:)}
(n')jl
We remark that if Zq,+ = 0 (and hence Z,,,+,, = 0) for some integer v,
then the terms corresponding to that value of v in (14.30) are of course
omitted.
For any positive integer j we have

1 1 3 < n(4r, Z)
77-
(14.31) <
r<Iz, <2 +2 z++ j r1
N(4er, Z) <) (4er)
jr' - jri
From (14.13) we have

q2-ql -1 3
1
- 241+V
Ill
V
zn
+
1

z
(14.32)
<Iz,.1<2Q1+ +l 7 z,
1z' 2qi+.-1 "
q2-qt-1
< 48 E 2-.i(ql+Y-1).
=2

Also from (14.13) we have


(14.33)
I1 1 i 1 1 J

2v2<Iznl<a Zn
+ - E
I41=2°2-1
!
Zn

7
1 1

<Iz..l<<2 2 IZ,.l=2 Un) 7


s<Iz,.l<<2v2+1 Gn)
< (48)2-i(q2-1) +
n ( 2q2+1 Z ) < (48)2-i(q2-1) + ads).
j8i'
86 Entire and Meromorphic Functions

Finally,
(14.34)
1
\j
(1
xn
IZr 1=2v2

i 1 1

2az+'<Iznl<2az+z zn Iz;.1=2°z
z 2az+i<I=,.I<29z+z n Y
(2vz+2, Z) A(4es) .
(48)2-i(92) +n < (48)2-j(9z) +
jsi jsi
Combining (14.30) through (14.34), we have

1
(i)'!
<Ian1<d

(14.35) < L(4+ 9z-9t+1


jrj (48) E
2-j(92+Y-1)+2A(4es)
jsi
< A(4er) + 96 + 2A(4es) < A(4er) + 96A(r) + 2A(4es)
jrj jsi jr' ri jsi
From (14.28) and (14.35), we see that

1 1 < A'A(B'r) A'A(B's)


(14.36)
x rj + Si

for all positive integers j and all 0 < r < s if A' = 100 and B' = 32e. This
completes the proof of the lemma.
15
Canonical Products

We shall suppose that a countable set Z = {zn} of complex numbers is


given. For convenience, we shall suppose that 0 0 Z and that Z has no
finite limit points. More generally, we consider "sets with multiplicity."
This means that some of the zn may be counted multiple times. It is
possible to make this notion rigorous, but at the price of clumsier notation.
For convenience in this chapter, we shall exclude the null function F(z) = 0
from consideration.
Definition. If F is an entire function, we write Z(f) for the set of zeros
(other than the origin) of F.
Definition. If p is a nonnegative integer, we define E(u,p), the Weier-
8tnass primary factor of order p, by
// u2 up/\
E(u,p)=(1-u)explu+ 2 p/
E(u,0)=1-u.
Lemma 15.1. Given any Z, there exist nonnegative integers An such
that the product

f(z) = TIE ( - 'An 1


l
n c1
converges uniformly on compact sets to an entire function f such that
Z(f) = Z.
We omit the easy proof which follows directly from Estimate A later in
this chapter. Any sequence An such that An --i oo will work. The next
theorem is a corollary of Lemma 15.1.
88 Entire and Meromorphic Functions

Weierstrass Factorization Theorem. Given an entire function F, there


exist nonnegative integers An, a nonnegative integer in, and an entire func-
tion g such that

F(z) = zm exp(g(z))00II E ( , An/


n=1 n
Definition. Given any Z, the convergence exponent p1(Z) is defined by

P1(Z) = inf{a : EIznl-" < 00}-

Lemma 15.2. If E z , < oo, then n(t) = o(t"), where n(t) _ 1.


Izn'St
Proof. Choose a large number a, and write
1 > n(t) - n(a)
E IZnI "- t"

Lemma 15.3. E . 1 a < oo if and only if f 0m i, t dt < oo.


Proof. Write
1
= t-" dn(t)
Iz= Iznl" J0
and integrate by parts to prove the lemma.
Corollary. pi(Z) = inf{a : n(t) = O(t")}.
Lemma 15.4. p, (f) <_ p(f)
Proof. We must show that if p = p(f), then for each e > 0, n(t) = 0(t°+`)
But by the Nevanlinna first fundamental theorem, we know that

N(t) = J t n(s) ds = 0(t°+').


0 s

But N(t) > n (2) log 2, by the usual argument, and the result follows.
Definition. The genus of Z, p = p(Z), is defined by p(Z) = inf{q : q is
an integer, E zj ; < oo}.
Definition. The canonical product of genus p over Z is defined by
z
P(z) H E +P
z EZ n
15. Canonical Products 89

Definition. The canonical product PZ over Z is defined by

z
PZ(z) = [J EZn- P
znE Z

where p = p(Z).

Theorem 15.5. The order of Pz is pl(Z).


The next result is a corollary of Theorem 15.5.
The Hadamard Factorization Theorem. Given an entire function f,

f (z) = PZ(z)zm expQ(z),

where m is a nonnegative integer and Q is a polynomial of degree < p(f).


Here, Z=Z(f).
Deflnition. The genus of the entire function f is defined by

P(f) = max(n,p),
where n = deg Q.

Examples. Suppose zn = n2, n = 1, 2,3,... . If f (z) = rj (1 - n ), then


f is of genus 0. If f (z) = ez rj (1 n ), then f is of genus 1.
We first shall show how the Hadamard Factorization Theorem follows
from Theorem 15.5. Without loss of generality, suppose f (O) 96 0. In fact,
assume f (O) = 1. Let
f(z)
9(z) = PZ(z)
Then g is an entire function without zeros, so that there exists an entire
function Q such that
z = exp Q.
We must prove that Q is a polynomial of degree < p(f).
Lemma 15.6. If F and G are meromorphic, then p (c) < max{ p(F), p(G)}
Proof. We have

log T(r, f )
P(f) = lim sup
r_.oo log r

T `r, 1) T(r,f)+O(1)
T(r, fg) = T(r, f)+T(r,g)+D(1),
90 Entire and Meromorphic Functions

from which the lemma is easily proved.


It then follows that exp Q is an entire function of order at most p. Writ-
ing Q = u + iv, we have

Iu(Te'B)I dO = 0(r°)

for each p' > p. Since, with IzI = r and R = 2r, we have

uw
Q(z) - Im Q(O) = u('w) w + z
2i,
it follows that
IQ(z) - Im Q(0)I = 0(r°)
Hence
IQ(z)I = 0(r,"),
and it follows that Q is a polynomial of degree < p.
Proof of Theorem 15.5. Our proof uses the Fourier series method of Chap-
ter 13. It is shorter and less tedious than the standard proofs.
Estimate A. If Iznl > 21z1, then

log E (.!n , p) 1:5 21 IP+1 < I


IP.

n
Let u = z- . Now log E(u, p) E +1 k . Here Jul < a. Hence
00 00
Iulp+1 E 2-k = 2Iut"1.
I log E(u,p)I <_ E Iulk <
P+1 0

It follows from Estimate A that the product

converges uniformly on compact sets to an entire function f, so that


log I f (z) I = E log I E ( p )1We write

00
log If(z)I = E c,, e'-O,
m=-00
15. Canonical Products 91

where Cm = c,,,(r) is the mth Fourier coefficient of log If I. We have seen in


Lemma 13.4.1 how to calculate c,n:

/x 1 r. (xz
\/1k
logE l zn P) = k_=p,+l
k n

so that
00 00 1
z )k)
log f (z) n=0
_ > -k=p+1
> k (zn

Hence
1 0 fork<p
j an
On =
E* fork>p+1.
1 -21
Recall the notation from Chapter 13; near z = 0
00
log 1(z) = Eakzk
k=0

Using the formulas in (13.4.18), we therefore get


W co = >2 log
r
rn

(n)=2m
(r)mlISn I <r
1

u (rI ISnI<r I.. 1<-r


l ifm<p

(m) cm =
-Tm
2mISnI>r
` ()mL>2
1

ISnICr
\rJ
n lm
if m > p + 1.

Choose p' > p. We must prove that for some M = M(p')

rp
Ic,n(r)I <- M form=0,1,2,....
ImI + 1

But this estimate is easily derived, as in Chapter 13, from the fact that
n(r) = O(rP ), once we know that the cm are given by formulas (i), (ii),
and (iii).
The next theorem follows easily from the Miles-Rubel-Taylor Theorem
of Chapter 14 but is included here due to its simple deduction from Theo-
rem 15.5.
Theorem. If f is a meromorphic function in the plane with p = p(f) <
00, then there exist entire functions g and h, with p(g) < p and p(h) < p,
such that f = g/h.
Proof. Let W = {Wn} be the poles of f, let pi = pi (W) be the exponent
of convergence of W, and let p = p(W) be the genus of W. Since N(r, f) <
92 Entire and Meromorphic Functions

T(r, f) = O(rP+E) for each e > 0, we have pl < p. By Theorem 15.5,


p(Pw) = pi. We now let g = f Pw and observe that g is entire. Since
p(g) < max{p(Pw),p(f)} = p, it follows that f = g/Pw is the desired
representation.
It was proved by similar means by Rubel and Taylor that if A(2r)/A(r)
0(1), then every meromorphic function of finite A-type is the quotient of
two entire functions of finite A-type. However, the proof is long and is
subsumed in Miles' proof of the last chapter, so we omit it.
Laguerre's Theorem on Separation Zeros. If f is a nonconstant en-
tire function with only real zeros, has genus 0 or 1, and is real on the real
axis, then the zeros of f' are real and are separated by the zeros of f, and
the zeros off are separated by the zeros of f'.
Proof. We have either
x
f (z) = CzKeaZII 1-
xn
or
x
f(z) = CzKe4zII 1 - z ez/Z",
xn
where the xn are real, and c and a are real (possibly a = 0).
It follows that either
k
f'(z) = + a + 1
f(x) z Z - Zn
or
z
f'(z) = k -F a + E
f(z) z zn(z - zn)'

On writing z = x + iy and recalling that a is real, we get, in both cases,

I.Rz)_-y{
f(z) x2 +
k y2 + (x - zn)2
1 + y2 1
which does not vanish except for y = 0, so that the zeros of f' are real. Since
f is real for real z, the theorems of calculus apply. By Rolle's Theorem,
there is a zero of f' between two consecutive zeros of f, so that the zeros
of f are certainly separated by the zeros of f'. To see that the zeros of f'
are separated by the zeros of f, note that
k
(x) = < 0,
x2 - (x - xn)2

so that /LLI is decreasing in any interval free of zeros of f, so that f'


cannot have two zeros in any such interval. The case of repeated roots is
handled by a suitable convention.
16
Formal Power Series

We consider the formal power series

f(z) = ao + a1z +a2z2 + ...,


which we usually normalize by ao = 0.
Let fn(z : a) = a + a1z + a2z2 + + anzn; fn is a polynomial of
"degree" n (possibly an = 0). We adopt the convention that fn has n zeros
zi, z2, ... , zn, where, if am # 0 but am+1 = am+2 = ... = an = 0, then
2nn+1 = Zm+2 = - - - = Zn = oo. For later use, we shall make the convention
oo/oo = 1.
Let rn(a) = max{jzI : fn(z : a) = 0}, that is, rn(a) is the modulus of
the largest root of fn(z : a) = 0. Note that if an = 0, rn = oo. In a certain
sense, rn(a) measures the disaffinity of f for the value -a.
Theorem 16.1. Given any formal power series f , then
rn(a)
(16.1) lim sup < 2, (b 0 0);
n-oo rn(b)
if b = 0, we have

p p ( 1 1 "-"
lim su rn (a) < 1 + lim suIL
n-.oo rn(0) n-»oo rn(a)
where
f (z) = amzm + am+l zm+l +..., a. # 0.
Note that if the roles of a and b are interchanged, then (16.1) yields
n-.
< lim inf rn (a)
94 Entire and Meromorphic Function

Proof Let al, a2, ... , an be the zeros of fn(z : a) arranged so that (a1 I <
Ia2I < Ianl. Let be the zeros of fn(z : b) arranged
so that IP'I < I02I < < IP.I. Thus fn(z : a) = anH(z - ak) and
fn(z : b) = anH(z - Pk), where we shall take an 0.
Now,

(16.2) fn(z : b) - fn(z : a) = b - a = an[II(z - Pk) - Il(z - ak)]

Suppose there exists a sequence of n for which {an}, {fin} is such that
Ian I = An IPn with An > A > 1. Otherwise, the conclusion of the theorem
is clearly true. Therefore, rn(a) = Anrn(b). Now,

Ian - Pkl -Ianl-IPk1=Anl/nl-IPkl - (An-1)IPkl.


Letting z = an in (16.2) gives b - a = anH(an -,Ok). Hence,

Ib - al >- Ianllrr(an - Pk)l >- IanIfI(An


- 1)IPkI-
Dividing through by b and taking nth roots gives

1- an [i'n(A_1)IflkJ] n

>[
Now suppose we had An - 1 > 1. Then
L
an Ianl
I1-bil>[IbI(1+E)IIIPkI,".
n

But rrlPkI = aa so that supposing An - 1 > 1 implies I1 - e ° > (1 + E);


a contradiction. Hence, lim sup An < 2, so that we have lim sup rn(
r. b) < 2.
n-oo
Suppose now that b = 0. Write f (z) = amzm +an+lzm+1 +..., am # 0.
n-m
We now have fn(z : b) = zman jj (z -,6k). Thus, fn(z : b) - fn(z : a)
k=1
n-m n
-a = anlzm fT (z - Pk) - f1 (z - ak))
k=1 k=1
n-m
We again have Ian-PkI = (An-1)IPkj and I-al = Ianan f1 (an-Pk)I
k=1
n-m
lanl[rn(a)]m fl (An - 1)IQkl:
k=1

(16.3)
_ rr n-m
Llanlfrn(a)ln` H (An -1)IPA; I
^-*"
Ial
L k=1
16. Formal Power Series 95

Now suppose for contradiction that

there exist {En} such that En > 0 and


\l ml
n-m

(The mss-
m

(1+En).
An - 1 = lal(^(laml) (rn(a)
n-m
n-m
Using this expression in (16.3) and that n 113kl = S-I gives
k=1
lal ^ m >lal n'tn (1 + En)n-mi
a contradiction.
Hence,
n m
An-1<lal() 2 1 - 1
laml rn(a)
Consequently,
rn (a)
lim sup < 1 + lira sup 1 ,
rn(0) n-00 rn(a)
which completes the proof. Theorem 16.1 has the following corollary.
Corollary. If f is holomorphic in a neighborhood of 0 (that is, f has a
positive radius of convergence), then

limsuprn(a) < 2
n-.oo rn(b)
for all b.
Definition. Tn (f) _ - log rn (1) is called the discrete characteristic of f.
Theorem (First Fundamental Theorem) 16.2. Tn(f) =Tn(f -a)+O(1)
for all a with at most one exception.
The proof is immediate from the definition of Tn (f) and Theorem 16.1.
Now let us examine the case of our exception more closely,

lim sup rn (1) < 1 + lira sup 1


m
n-.oo rn(0) n-+oo rn(1)
Suppose fn(z : 1) = 1 + a1z + + anzn has zeros l3n arranged in
order of increasing moduli. Thus, anlI(-Qk) = 1 and hence IIlakl =
Accordingly, [rn (1)]" > r , and therefore
^ n-m
(-) n-M --
(-)
1 1
[rn(1)]n-M >
lan1or :5 (lanl^) ^ m
which gives us the next result.
96 Entire and Meromorphic Functions

Corollary. If f is holomorphic, then Tn(f) = Tn(f - a) + 0(1) for all a.


Theorem (Kakeya) 16.3. Let f be a formal power series not identically
zero and R(f) its radius of convergence. Then R(f) < limn rn < 2R(f
n-oo
Proof. Choose R' < R(f ). The fn have only a finite number of zeros in the
disk DR, and { fn(z)} - { f (z)} uniformly in DR.. Hence, past a certain
no we have, by Hurwitz's Theorem, that the functions fn have the same
number of zeros in DR, as f does. Therefore, the largest zero of fn must
leave DR.. Thus R(f) < lim inf rn, as desired.
n-oo
Alternatively, we have seen that rn > Ian! w and hence liminf rn
n-oo
R(f).
To prove the second inequality in the theorem, we use the following
result.
Lemma. Let P(z) = bo + b1z + - - - + bnzn = 0. Then

IzI < 2max Ibb ll, Ibzn21 ,...,16


(

Proof of Lemma. Clearly, IbnllzIn < Ibol+Ibol+IbiIIzl+---+Ibn-1IIzIn-1.

Writing IzI = r, w e h a v e rn < I b I + l b I r + - - - + I . I rn-1. Suppose

now that r > 2 m a x (I


b^ I ' I
b^ I '...' I I
Then r < 2 r+
2-(n-1)rn + + 2-1rn = rn(a +1 +... + 2;,) < rn, which is impossible
and the lemma is proved.
Now let bk = akRk, P(z) = fn(Rz) = ao + a1Rz + - + anRnzn. Then
the roots of fn(Rz) = 0 are times the roots of fn(z) = 0. By the lemma,

10-1 anRn
la"_2Rn-zl anRn
as Il,.
n < 2 max I an-I. ,...,l
R- ( Rn

Choose R> R(f); it follows that {IanIRn} is unbounded. Therefore, for a


suitable subsequence, IanIRn > IakIRk for all k < n. In that subsequence,
< 2. Hence, lim inf rn < 2R. Therefore, lim inf rn < 2R(f), as was to
n-.oo n-oo
be proved.

Corollary (Okada (28)). A function f is entire if and only ifn-oc


lim Tn(f) =
-00.
16. Formal Power Series 97

log n =1im log n


eorem (Tsuji [45]) 16.4. Let a = o(f) = lim sup sup
n.oo Tn n-.oo log rn
I- en a(f) = P(f)
n log n
proof. We know that R(f) = lim sup from Proposition 11.4. Take
1
I
10g Ian
n log1
pl > p; then, for large n, p' > Hence, IanI < 1-fl and thus
log np,
lanl
to > R > n'' , or rn > n o 1r. Consequently, o < p' for large n;
hence, a < p as desired.
In the other direction, suppose on the contrary that a < p. Choose p'
such that or < p' < p. Then for n large (say n > no ), rn > n or . Choose -
2K
M > 1 so that for K = 1,2,..., no, laid < m (K1) We prove by
2n
induction that for n > no, IanI < m
(n!) °
First,
Ianlrn < Ian-l lrn 1 + ... + Iallrn + 1
since &(z : 1) = 1 + aoz + + anzn = 0. Hence,

IanI < Ian-1I +Ian-2Ir 1 + rn


n
n Ian-2l + ... + tall
1
< Ian-1l + +n°s}
n nT
Therefore,

Ia.l<_M
2n-1
n° [(n- 1)!]°
+ 2n-2
n° [(n- 2)!]
+...+ 2' + si-
n ° (1!)° n° (0!)
20

<M
2n-1 2n-2

(n!)°7 + (n!)°
+ + 2'
(n!) I'
+
20

(n!
,
2n
< M <M (n!)
(n!)° -
as desired. This now leads to a contradiction of the fact that f is of order
p, for we have Ta7 - M 2^° Consequently,

109 I 1an
I ? constant + log n! - n log 2
98 Entire and Meromorphic Functions

so that
n log n c n log n _,p + 0(1)
log constant + p, log n! - n log 2 -

since log n! ti n log n. This implies that p(f) < p' < p = p(f ); a contradic-
tion.
17
Picard's Theorem and the Second
Fundamental Theorem

In this section we shall prove Picard's Theorem, state the second funda-
mental theorem of Nevanlinna (leaving the proof for the next section), and
derive some of its consequences.
We shall use two conventions that will greatly simplify our notation.
First, we shall always work "modulo 0(1)". For example, A = B and
A < B shall mean that A - B is bounded and that A - B is bounded
above, respectively. Second, we shall use a notation of Weyl, writing 11 in
front of a statement to mean that the statement holds with the possible
exception of a set of finite length.

Picard's Theorem. If f is a nonconstant meromorphic function on the


complex plane, then f does not omit three values on the sphere.

It may happen that f omits two values. For example, ez omits 0 and 00.
Our Proof of Picard's Theorem is patterned after our proof of the second
fundamental theorem of Nevanlinna, and illustrates the main features of
the method in a simpler context.

Proof. Without loss of generality, we may assume that f omits 0, 1, and


oo, so that f, 1, and 1 are entire. The next lemma is referred to as the
lemma of the logarithmic derivative.

Lemma. 11 T (r, f) <- K' log r + K" log+ T(r, f) for f omitting 0, 1, and
oo.
100 Entire and Meromorphic Functions

From Poisson's formula (7.1),


I 7r v
logf(z)=-21rI logIf(pei
.,
)I ;±zdco;

f xlo If(Pe"°)I(pe;
f()x 27r

If (z)
f xlogIf(pe"°)IdW
f(z) (p? r)2 2w

log+ I f'(z) I< log+ p + 2log+ p


1 r

+log+ loglf(Pe"°)IdW (modO(1)).


2J
Recalling the definition of m(r, f), we find

(r, f
1
log+ + 2 log+ + log+ m(r, f) + log+ m 1 .
f (z) I <_ log+ p p r

Since f 0 0 or oo, we have m(r,f) = T (r, f) and m (r, f) = T (r, f) .


Moreover, by the first fundamental theorem of Nevanlinna, we have

T(r,f)=T (71) .
Hence,

log+ I
I< log+ p+ 2 log+ p r+ 21og+ T (p, f).
f()
Since the right-hand side is independent of the argument of z, it follows
that the same estimate will hold for the average,

m I r, f I< log+ p+ 2log+ p 1 r+ 2log+ T (p, f ).

Now choose p = r+ logT f)fi and use the Borel lemma on 2 log+ T(p, f)
to find,

T (r, fi ) < log+ r + 2 log+ log+ T (r, f) + 3log+ T (r, f)


< log+ r + 4log+ T (r, f) .
17. Picard's Theorem and the Second Fundamental Theorem 101

This proves the lemma.


Now let
1
F(z)
f(z) + A z) - 1
IF(z) I is large whenever f is near 0 or f is near 1. Of course, the set of values
of z where these occur is disjoint: hence, m(r, F) > m(r, f) + m (r, ff 1 1) .
Therefore,

(17.1) T(r, F) > T I r, 1) +T (r, 1 since f 54 0, 1.


f f-1/ I

Notice that
1 f(z) f'(z) + f'(z)
F(z) -
f(z) f'(z) f(z) f(z) - 1
Thus,

(17.2) T(r,F)<T(r,f)+T(r,f')+T(r, )+T(r, ff'j).


From the first fundamental theorem we have
TIr, f)=T(r,f)=T(r,f-1)=TIr, f1 I

and
T (r, '/=//
T lr,
Therefore, on combining (17.1) and (17.2\\) we obtain

2T(r,f)<T(r,f)+2T(rL) +TI r,fConsequently,

(17.3) T(r,f)<2TIr, f)+T(r,f f ).


The lemma above shows that
TIr, f I < k'logr+k"logT(r, f).
Applying this lemma to f /and f - 1 in (17.3) gives
T(r, f) <k'logr+k' log T(r, f).
Of course f is not a rational function since it omits three values. Conse-
quently, by Theorem 10.2,
T(r'f) -iooas r -ir
log r
But we have
II T(r,f) <k' k" log+T(r,f)
log r +
log r '
which is impossible, and the proof by contradiction is complete.
102 Entire and Meromorphic Functions

Theorem. (Second Fundamental Theorem of Nevanlinna). Suppose f (z)


is a nonconstant meromorphic function in the plane (or in Izi < R, R <
oo). Let a1i a2, ... , a4 be q distinct finite complex numbers. Then

(17.4) m(r,f)+>2rral r, f 1 ) <2T(r, f)-N(r, f)+s(r),


v=1

where
N1 (r, f)=N(r, ' ) +2N(r,f)-N(r,1')
and

S(r)=mlr,f +m r,>
\\ f! i_1
f
f - av

Interpretation. It will be shown that

11 S(r) = O(log T(r, f))+O(logr).

Accordingly, we may think of "S" as standing for "small." To interpret


the function N, (r, f), suppose f (z) = (z - za)Kg(z), K E Z, g(z) # 0
in some neighborhood of zo. We analyze the contribution to N1 due to
the behavior of f at zo by examining the counting function n1(r, f) =
n (r, t, ) + 2n(r, f) - n(r, f'). If K > 0, then f vanishes at zo to order K.
Consequently, n(r, f) and n(r, f') are unaffected by the behavior of f near
zo. The contribution to n1 near z equals the contribution to n (r, Of f).

course, j has a pole of order K - 1 at zo. If K < 0, then f has a pole of


order -K at zo and a similar analysis reveals that n1 "counts" this pole
K - 1 times. The case K = 0 is trivial. In general we find that poles of
order K and zeros of order m are "counted" by n1, respectively, K-1 and
m - 1 times.
The gist of the second fundamental theorem is that for most values of a,
the contribution of m (r, ffl a) to T (r, f 'a) is much smaller than the con-
tribution of N (r, f 1 ) to it. Thus, for most values of a, N (r, f11) makes
the preponderant contribution to T (r, f Q- . Recall that T (r, f 1 a) -
T(r, f) = 0(1) by the first fundamental theorem.
Remarks. The following weaker form of the theorem also gives us the Picard
Theorem:

(17.5) m(r,f)+Emlr, 1 } <2T(r,f)+S(r).


-1 \\\ f -ate/
17. Picard's Theorem and the Second Fundamental Theorem 103

For suppose f omits three values, say 0, 1, and oo. Then, since f is entire,
m(r, f) = T(r, f) and, since f omits 0 and 1, we have

m(r,f T (r,
f ) = T(r, f)
and
(r, J = T (r,
in 1
f-1/J = T(r, f - 1) = T(r, f )
1

Thus (17.5) implies that 3T(r, f) < 2T(r, f) + S(r) and hence T(r, f) <
S(r)-
By a lemma to be proved in the next section we assert that

m (r, f) < 61og+ T (r, f) + 41og+ r.


Hence,
11 S(r) < 6(q + 1) log+ T(r, f) + 4(q + 1) log+ r

and thus
T
T (r, f) < S(r) implies that lim (r' f) < 00,
r-oo 109 r
which is only true, by Theorem 10.2, for rational functions. Since f 96 00,
f is a polynomial and therefore does not omit three values; a contradiction.
Consequences
Definition. Let n(t, f) be the number of poles of f in the closed disk
Dt = {z E C : IxI < t} counted once, no matter what the multiplicity. Let
N(r f) = fro+n(t,f)-n(e,f)dt.
t

1 1
m r, a a
N (r,
Definition. b(a) = limr T(r,
f) - 1 - limr_ 00
T(r f)
6(a) is called the deficiency or defect of the function for the value a.

(r,
f 1 a) -
N (r, f 1 a)
Definition. 9(a) = licnr N
T(r, f)

Definition. The branching index is defined to be 0*(a) where

N r, 1
a
N lr, 1
J
a
@'(a) = 1 - limn..oo T(r, f) = 4M,-. 1 - \T(r f)
104 Entire and Meromorphic Functions

What contributes to 9* (a) is a-points that are taken on by f with multi-


plicity greater than or equal to 2. So a value a that is taken on often with
high multiplicity will have a large branching index.
Remark. 0*(a) > 9(a) + b(a). Then,

f N-1VT 1-NJ
N]
9*(a)-r m I1- NJ
T
= urn

li
00 L r-.°° l
m
lim INTNJ+
> r-oo [1 TN1
= 9(a) + 5(a).

The next result follows from the second fundamental theorem.


Theorem. Let be any finite collection of distinct complex numbers
possibly including oo. Then,

(17.6) 2.

Proof. First suppose f is not a rational function. From the second funda-
mental theorem we have m(r, f) + q m (r, iaV) < 2T (r, f) - Nl (r, f) +
V=1
q
S(r). Adding N(r, f) + > N (r, 7.7t o both sides, we get:

(q + 1)T(r, f) < 2T(r, f) + N(r, f) + EN \(r, 1 ) - N, (r, f) + S(r).


V=1 f - aV
Hence,

(q-1)T(r,f) <EN (r, f-ate


1 -N (r, f,)+N(r,f')-N(r,f)+S(r).

v=1

Wherever f takes the value aV with multiplicity k, n (r, i ) -n (r, f, )


counts 1. Also, at a pole of f, n(r, f') - n(r, f) counts 1. Therefore, we
may write

(q - 1)T(r, f) <
v_1
N (r,fi-/- a+ N(r, f) + S(r).

Hence,

q N (r, ) R (r, f)
(17.7) (q - 1) 5 E T(r, f)y + T(;7) + 7'(rrf)
,.-i
17. Picard's Theorem and the Second Fundamental Theorem 105

Now f is not a rational function, and therefore lim,- , T r = 0 since


IIS(r) = O(log T(r, f)) + O(log r) by Lemma 18.3 in the next section.
Now, using the fact that lim (A + B) < Em A + urn B, we have

,,
q
N(r, f lam N(r,
(4-1) lim
T(r,f) + LI o

from which

(r,T. 1

- rQL lim N T(rf) -lim N(r,


T(rf)
f) < 1 - q.
v=1

Adding q + 1 to both sides gives


q
11- lim NT(r f) +1- < 2.
=1 L
, J p
Hence, E 2 holds if f is not a rational function.
V=1
q
Now if f is a rational function, the same claims hold up to (17.7). We
may suppose If (z) I - IzI', where p is a nonzero integer. Then T(r, f) _
[PI log+ r. Also,

/ , 9 , 4 zy
S(r)=m[r, f+m(r,Ef -a,, =m(r,p)+m(r,t p V
\\ /// \` f ff z ` i=1 z - a
Now m (r,
P)
0. Similarly, for IzI large, we have m (kr, E
= 127rf,, log+ I P ZI dO and, for IzI sufficiently large, m (r, P) _

v=1
follows that lim Tarf = 0, and the rest of the proof follows as before.
Q p-1
= 0. Therefore, it

Corollary. Let be any finite collection of complex numbers possibly


including oo. Then,
v
2.

V=1

Definition. a is called a deficient value of f if b(a) > 0.


If f never takes on the value a, then b(a) = 1. The same is true if f takes
on the value a very infrequently. In general, b(a) measures the tendency of
f to omit the value a and 1 - b(a) measures the tendency of f to take on
the value a.
106 Entire and Meromorphic Functions

Corollary. There are at most countably many deficient values.


Corollary. There are at most two values for which 6 > 3 .
Definition. a is said to be a fully branched value of f if f takes the value
a with muAiplicity either zero or > 2.
Remark 17.1. If f is entire, f can have at most two fully branched values
and this is the best possible, since, for example, ±1 are fully branched
values of sin z. Also, if f is entire, then there are at most two values of a
for which 6(a) > 2.
The reader will find it instructive to work out b(a), 9(a), and 9*(a) for
some specific functions, such as eZ, tan z, and the Weierstrass p function.
Remark. If a is fully branched, then
1
N 1

9* (a) =1im 1 - NTr(r, > lim 1- f -a since T > N.


f) N

N(r, f ia)
1 <
Also, if a is fully branched, then 2, and therefore
N(r,fa)
9*(a) > 2. In general, there are at most four fully branched values for
a meromorphic function.
The next result shows that Nevanlinna theory may be used to study cer-
tain types of exponential identities. As an example, consider the functional
equation
of + e9 = eh,

where f, g, and h are entire functions. Do there exist f, g, and h so that


the identity holds? What is the relationship between f, g, and h? We may
write
of-h + eg-h = 1.
Then the entire function of -h omits 0. But eg-h omits 0 as well, therefore
of -h must also omit 1 for the identity to hold. By Picard's Theorem, ef-h
must be a constant so f = h+ const. Similarly, g = h+ const.
Lemma (Hiromi-Ozawa [15]). Let ao(z), a, (z),... , an(z) be meromorphic
functions and let g1(z), ... , gn(z) be entire functions. Further, suppose that
n `
T(r,ai) = o Em(r,e9°)J j =0,1,...,^
v=1
holds outside a set of finite logarithmic length. If an identity
n
2av(z)e9I(=)
= ao(z)
V=1
17. Picard's Theorem and the Second Fundamental Theorem 107

holds, then we have an identity


n
E c. a.(z)esvW = 0
V=1

where the constants c,,, v = 1, ... , n are not all zero.


Proof. On writing f' = f f, we may use the lemma on the logarithmic
derivative to estimate T(r, f) when f is entire. For notational simplicity,
let Then we have
n
(17.8) E G (z) = ao(z).
V=1

By differentiating both sides we obtain


n
(17.9) G(") (z) = a( 0")(z),

which may be rewritten as

n
g) (z) = a(µ)
(17.10) >2 G., (z) z p --1,
o (), ... , m - 1.
V=1 (z)

We regard this as a system of simultaneous linear equations in the G.


Now we have
G(IJ) (z) = P,(a,,,a;,,...Ia(" ), (µ))eg.,(z)

with a suitable polynomial P, in the indicated functions. Thus we have

(17.11) T (m(re9))
Y-1

outside a set of finite logarithmic length.


Suppose, for the simultaneous equations (17.8) and (17.10), that the
determinant A 34 0. By solving (17.10) with respect to G j = 1, ... , n we
have by Cramers' rule

where
1 ... 1
GI/G1 ... G,1,/Gn
0=
Gin-1)IGn . __ G(n-1)IG.
108 Entire and Meromorphic Functions

and

1 ... 1 a0 1 ... 1
Gj-1 al G .1
G1 ...
C i+1 ...
Gj-1 0 G.
Aj=
Gj_-1_ Gin l1, (n-1) ___ ____
... a0 Gj+t ...
GI GI_' Go

Since

(17.12) T (r, C I=o m(r, e9) f ,

we have
n
T(r,A)=0 m(r,e9°) ,T(r,Aj)=o( m(r,e9°))
=1
for j = 1, ... , n outside a set of finite logarithmic length. Thus we have
m(r,e9") =T(r,e9°) <T(r,a.)+T(r,G.)

o (m(r1e9)
v=1

and hence
r
!2m(r,e) =o (Em(re9))
V=1 v=1
outside a set of finite Lebesgue measure. But this is a contradiction. Con-
sequently, the Wronskian 0 =_ 0 and the result follows.
We say that two meromorphic functions f and g share the value a (a = 00
is allowed) if f (z) = a whenever g(z) = a and also g(z) = a whenever
f (z) = a, counting multiplicities in both cases. A famous theorem of
R. Nevanlinna, which will be proven shortly, implies that if two nonconstant
meromorphic functions f and g on the complex plane share five distinct
finite values (ignoring multiplicity), then it follows that f = g, and the
number 5 cannot be reduced. We consider here the special case g = f', the
derivative of f, and prove the following result.
Theorem. If f is a nonconstant entire function in the finite complex
plane, and if f and f' share two distinct finite values (counting multi-
plicity), then f' = f .
In other words, a derivative is worth two values. We show at the end
of the chapter that the number 2 of the theorem cannot be reduced. We
17. Picard's Theorem and the Second Fundamental Theorem 109

do not now know whether there is a result corresponding to our theorem if


one ignores multiplicities, or if one considers meromorphic instead of entire
functions.
Proof of Theorem. To fix the ideas, we suppose that f and f' share the
values a and b, where a = 1 and b = 2. Other choices of a and b make no
real difference, except if a or b is zero, in which case the analysis becomes
easier, and is left to the reader. We may write then

(17.13)
f'-1 k'
f-1

f' - 1 _ eke '


(17.14)
f-2
where k1 and k2 are entire functions. We solve (17.8) and (17.9) for f to
get

1 + ekl - 2ek2
(17.15) f= ek1 - eke

We now differentiate both sides of (17.10) and substitute (17.8) to get


(17.16)
2e2k' +e 2k2 +(k2-ki-3)e
+ kl+k2 - a2k1+k2 +e k1+2k2 +klek' - k2e
.
= 0. k2

We shall make repeated use of the lemma of Hiromi and Ozawa.


Now, in (17.16), divide by eke to get

(17.17) 2e2k1-k2 + eke + (Jc2 - k' - 3)ek - 22k' + ek'+k2 + k2.

We now apply the lemma to get


(17.18)
ele2kl-k2 + c2ek2 +c3(k' - k' - 3)ekl +c4e2k' +c5ekI k2 +csklek'-k2 = 0,

where c1, ... , cs are constants that are not all zero.
The hypotheses of the Hiromi-Ozawa Lemma are satisfied because, for
example, ki is the logarithmic derivative of ekl and we may use the lemma
of the logarithmic derivative. It follows that T(r, k') = 0(T(r,ek)), outside
of a suitably small exceptional set. At any rate, we divide in (17.17) by ek1
to get

(17.19) clek'-k2 + c2ek2-k' + c4ek' + c5ek2 + csk1e-k2


= c3(k2 ' - k' -3),
and we may use the lemma once more to get
diekl-k2 + d2ek2-kl
(17.20) + daekl + d4ek2 + dskie-k2 = 0
110 Entire and Meromorphic Functions

for suitable dl , ..., d5. Multiply by eke to get

(17.21) dleki + d2e2k2-ki


+ d3ek'+k2 + d4e2k2 = -dsk'

and apply the lemma yet again to get

ulek' + u2e2ki-k2 + u3ekt+k2


(17.22) + u4e2k2 = 0,

where u1i u2, u3, u4 are constants that are not all zero. Now, by successive
applications of the lemma, we reach a contradiction, unless possibly one of
the five following conditions holds for some constant C:

k1=k2+C,12=C,kl=2k2+C,k2=2k1+C,k1=C.
We now rule out these possibilities unless f' = f. First, it is easy to see
that k1 = C (and similarly k2 = C) is consistent with (17.13) and (17.14)
unless d = ec = 1, in which case f' = f. For if (f' - 1)/(f - 1) = d,
then f = (d - 1)/d + bedz for some constant b, and hence f' = bdedz. This
clearly contradicts (17.14) unless d = 1, for we would have

f'-2_bdedz-2--ek2 '
(17.23)
f-2 bedz-2

which is impossible unless d = 1 (remember that f is not constant, so that


b36 0).
Next, we rule out k1 = k2 + C. We go back to (17.16) to get

(17.24) m1e2k1 + rn2e3k' = kim4ek'.

Apply the lemma again after dividing by ekl to get

(17.25) n1 ek' + n2e2k1 = 01

which implies that k1 = 2k2 + C (and similarly, k2 = 2k1 + C) is impossible


unless f = f. From (17.24) we would get

(17.26) d2e-c + (dlec + d4)e2k2 + d3eceSk2 = -2d5k2

and apply the lemma for the last time to get

(17.27) B1 + 12e2k2 + 13e3k2 = 0,

where P1, £2i t3 are constants that are not all zero. In other words, P(ek2)
0, where P is a cubic polynomial, so eke is a constant, which we already
have ruled out unless f = f'. This completes the proof of the theorem.
IT Picard's Theorem and the Second Fundamental Theorem 111

Finally, it is easy to see that there exists a nontrivial entire function that
does share one value with its derivative. For example,
efZ
=
f( z) = - e)dt
satisfies (f' - 1)/(f - 1) = ez so that f and f' share the value 1. This
shows that the number two of our theorem is the best possible.
Now we come, as an application of the second fundamental theorem,
to a truly beautiful and surprising theorem of Nevanlinna. Recall that,
given two meromorphic functions fl(z) and f2(z), and a complex number
(finite or infinite) w, we say that fl and f2 share the value w (ignoring
multiplicity) if every z for which f1(z) = w also satisfies f2(z) = w, and
vice versa. We use the same terminology if both functions omit the value
w.

Theorem. If two functions, meromorphic in the whole complex plane C,


share five distinct values, then the two functions must be the same.
Note that es and e_2 share 0, oo, 1, -1, so the number five is sharp.
Proof of Theorem. Given a number w, finite or not, let no(r, w) be the
number of common roots of f1 (z) = to and f2 (z) = to contained in the disc
{jz) < r}, each counted only one time. Then put
rr no(t'w) t no(0,w)
No(r,w) = dt+no(0,w)logr
0
and
1
N12(r,'w)=N(r'fi1 w)+IV (r, w)-2No(r,w).
12
Now, taking for wl,... , wq distinct finite complex numbers and applying
the second fundamental theorem to the functions fl and f2i we have, off a
possible exceptional set of finite length,

(q-2)(T(r,fi)+T(r,f2))<E(N(r'f1
q
1
l
)+N(r'f2
iwY))
+O[logrT(r, fl)T(r, f2))]
q q
_ 21: No(r,w,,)
V=1 Z-1
+ O[log rT(r, fi)T(r, f2)].
Now, if the functions f, and f2 are not the same, every common root of the
equations fi = w,,, f2 = w,,, is a pole of the function i z . We deduce
from this that
/ \
Y-1 \ h-f2/J <T(r,fi-f2)+O(1).
112 Entire and Meromorphic Functions

On the other hand,

T(r,fi - f2) <T(r,fl) +T(r,f2) +O(1).


We conclude
q
(q - 4) (T(r, fl) + T (r, f2)) < E N12(r, w.)
V_1
+ O[log(rT(r, fl)T(r, f2))]

off a set of exceptional segments of finite total length.


(If one of the w is infinite, just apply a linear fractional transformation
to fl and
Suppose to begin with that one of the two functions, say fl, is transcen-
dental. Then among the five given values w = w,,, v = 1, 2,. .. , 5, there
are at least three that are taken by fl in an infinite number of points z.
By hypothesis, the same is true for f2. Hence, f2 is also transcendental.
Suppose for a moment that fl and f2 are not the same. Apply the last
inequality above, which shows that for q = 5, the expression N12 vanishing
for the five given values,

T(r, fl) + T(r, f2) < O(log[rT(r,fi)T(r,f2)1)

This implies that

Urn log
Tl)r < oo and lim Tlo' 2) < 00.
r- oo r-co g

This is impossible since fl and f2 are both not rational functions. The
theorem thus is proved by contradiction in the case that one of the given
functions is transcendental.
But the conclusion is obvious if both of the functions are rational, since
a rational function f is uniquely determined by the roots of f (z) = w for
any three distinct values of w.
18
A Proof of the Second
Fundamental Theorem

We now begin the proof of the second fundamental theorem of Nevanlinna.


We continue to use the convention that all equalities are to be read "mod-
ulo 0(1)." There are a number of other proofs of the second fundamental
theorem, some of them leading to generalizations of it.
Lemma 18.1. Let f be a meromorphic function in the plane, and let
al,..., an be distinct complex numbers. Let
n
F(z) =
1(z) - a,,
Then
n
m(r, F) > E m (r

Proof We first remark that the lemma is intuitively obvious, for when f (z)
is "close to" a it contributes to m (r, lam) , but not to any m (r, )
with j # Y.
To prove the lemma in detail, we introduce the following notation. Let
d > 0 be given with 2b < min{ la - a,1 : 1 < v < j < n}. We require
6<1. Let
E. = f{w: Iw - a,,I < b} = {z: If(z) <b}
E,', (r) = E, fl {z : Izl = r}
E,,(r) _ {B : re's E
[0, 2x] - Ea(r).
114 Entire and Meromorphic Functions

In other words, E (r) is the set of all 9 for which If (reie) - a,, < 6, i.e.,
those 9's for which f (reie) contributes significantly to m (r, f la-) . It is
clear that E, (r) and Ej(r) are disjoint provided that v 54 j.
Moreover,
1 " 1
r log+ IF(re`B)IdO >_ 2 r log+ IF(reie)Id8,
21r x J
where f ` is over the set Now

f* log+ IF(reio)Id9
27r J - v-1 27r ,l E.. (rl
n 1
> log+ I d9
2a JE (r) I
f (reie) - a ,

_ 1 log+ do.
27r E.. (r) f(re`e) - aj
j=1
hAm

But
1 1 " + 1
m r,
f-a = 2a ,j 1o g
f (reie) - a
d9

1og+ I
f(reie ) - a
I d9 +27r
1 log+ I ie) I d9.
27r
E.(r) J ca(r) f(re - a

I °IB + O(1).
27r JET r lOg+ I f(re`) - a,,
Also,
n 1

27r
1 log+
L f(reie) - aj
<nal=0(1).
j=1
j6v
Hence we see that
1
log+ d9
m(r, F) > 2a f f(Teie) - a

1 log+ d9
27r E.,(r) f(reie) - aj
j=1
I j #v

Em r, f-°'v \I+0(1),
n
1
v_1
which proves Lemma 18.1.
18. A Proof of the Second Fundamental Theorem 115

Definition. N, (r, f) = N (r, f,) + 2N(r, f) - N(r, f'). Nl (r, f) is inter-


preted in the previous chapter after the statement of the second fundamen-
tal theorem.

Lemma 18.2.
1
> m(r,f-av i <2T(r,f)-Nl(r)-m(r,f)
.1 //
n
+m (r, -f' ll+m r,E
f! V=i f a
Proof. Let F(z) be as in Lemma 18.1 and write
n
z
F(z) 7_(z) . f ) f (z)(x)a"
Then

m(r, F) < m (r,


n
-l +M r, .t + m r, Y_1
1 f'
(18.1)
f/ ` f' I f
Butm(r, )=T(r,f)-N (r,1)andm(r, f)=m(r, f)+N(r, f)-
N (r, .). Hence,
(18.2)
/ // \
m(r,F)< T(r,f)-N(r,f)+mir, f )+N(r,L)-N)lr,

Y=1 f - a

Now observe that if we define

co(r,9)=N(r,9)-N
\r, 9/
then

co(r, 9h) _'P(r, 9) + co(r, h)


Using this observation in (18.2) we see that
(18 . 3)

m(r,F)<T(r,f)-Nl r, f l +m(r, fll +N(r,f')+NI r, fl


-N r, ;1 l -N(r,f)+m "
r, /1).
/ y=1
116 Entire and Meromorphic Functions

If we now add and subtract T(r, f) = m(r, f) + N(r, f) on the right-hand


side, we obtain

/')
(n
m(r, F) < 2T(r, f) - m(r, f) + m r, :c +M (r,
(18.4)
f) `/ L=1 lav

-(2N(r,f)-N(r,f')+N(r,
or

n ,
m(r,F)<2T(r,f)-m(r,f)+mr, f)+m r,E ff av -Ni(r).
\\ v=1

By Lemma 1, m(r, F) > EV_1 m (r, f av ), and Lemma 18.2 follows.


We are now ready to prove the second fundamental theorem of Nevan-
linna.
Theorem. Let f be a meromorphic function in the plane and let
n

f
C f,) v=lf -av
f,
Then

n
m(r, f) + E m (r7 f 1 av) 2T(r, f) - N, (r) + S(r).

Proof. This is just Lemma 18.2.


We saw previously that in order to deduce the Picard Theorem from the
second fundamental theorem, we needed an estimate on the size of S(r).
Such an estimate follows from:
Lemma (Lemma of the Logarithmic Derivative) 18.3. If f is meromor-
phic in the plane and 0 < r < p, then

m (r, f I) 4 log+ p + 3 log+ 1 + 4log+ T (p, f ).


f p-r

Proof. Without loss of generality, we suppose that f (0) 96 0, oo since, if


g(z) = Z' f (z),
m (r,
9) - m \r f 0(1).
18. A Proof of the Second Fundamental Theorem 117

Let z = reie. Then for a suitably defined branch of the logarithm we have
by the Poisson-Jensen formula,

1 r"
logf(x)=2 J_ logIf(pe"°)I pesw
t
!T,
(18.5)
+ E log BP(z, z,) - E log BP(z, iA,
Izv1<P 1W-1<P

where BP(z, a) is the Blaschke factor mapping the disk of radius p onto the
unit disk and a onto 0, A is a real constant, and, as usual, z and w are
the zeros and poles of f, respectively. [Equation (18.1) holds without any
0(1) terms.] Thus, by differentiating (18.5), we obtain
(18.6) I<P(z_'rW,)(p2-ww).
fi(x) x (Pe.P)I p2 - IzzI2
2pe"°
1°g if
f (x) 2w .. A (peiV - x)2 d`p +
(z - zz)
2 - Iw. I2
E
Iw

Taking simple estimates, we have


(18.7)
r I log If (Pe"°)II p2 - IAvI2
IRY (p - r)2 21r
dco +
IA YI<P
Ix - AYIIp2 -
where A now runs over both the zeros and the poles of f . Observe that
p2-Avx p(p2-IAYI2)
I
p(z-A,)I
p(p2 - IAYI2) < 1 p
- I BP(x,A,.)I BP(z,AY)I (p-r)2
-
The last step follows from Ip2 p2 - IAv IIzI p2 - pr =P(P - r).
If we use this estimate in (18.7), take log+ of both sides of (18.7),
and apply the additive inequality satisfied by log+, (log+(a, + . + a,) <
10g+ a, + + log+ a + log n), we then find that
(18.8)

<_ log+ p + 2log+


r + log+ 2 J x Ilog+ If (rei ')I I
1
log+ I f(x) I p
E log+ IBP(z + log+ (n(p, f) + n (p,
Iavi<P AL)I f)).
Since IzI = r < p, B (1, a >_ 1, we see that log+ IB(=,a. = log IBp(z,a-
118 Entire and Meromorphic Functions

Therefore, by Jensen's Theorem,


(18.9)
1 1 1°g log la f if r
21r
J ,,
log+ I
BP(re=B,
1 I dO =
log if r.
la l

To simplify the notation, let n(r) = n(r, f)+n(r, and likewise N(r) _
N(r, f) + N(r,
j)
If we let z = re`B and integrate (18.8) over the circle of
f).

radius r with center at 0, we obtain


(18.10)
f _ - ( log+
r, _
f 21r d
1 7r ff( (re : )
re'e
) ' d8
r
< log+ p + 2 log+ p 1 r+ log+ 2 J x 11og+ If (T e' ') I f dip
x

+
r
21r ,1-
log+ I Bv(ree, ,\v) dO + log+ n(p).
x
l a..l <P

We recognize that i f' Ilog+ If(PQse)IIdcp = m(p, f) + m (p,.I) =


O(2T(P,f)) and

r` +
L 21r log B.(r;`°, A.,) d
la.I<P

_
Ia.,ISP
log (FAP-I '/ L IA.I<r log ( a,.
I I
N(P) - N(r).

Hence,
(18.11)
m(r, f,/ <log+p+2log+p 1 r+log+T(P,f)+log+n(P)+N(P)-N(r).

We now will estimate the term log+ n(p). Choose a number p' with
p' > r. Then

P dt < 1 P n(t) dt
n(p) = n(P)
log(P) fv t log(o) JP t
N(p) 2T(p', f) + C
- -
log(p) log(p)
where C is an appropriate constant. Now

p, (P - P) :5 P tt = log P - p (P - P)-
J
18. A Proof of the Second Fundamental Theorem 119

Hence,
n(P) <
', f)
2T(#', )C

so that
1og+ n(p) < log+ p' + 1og+ p,_P + log+ T(P', f )
1

Hence,

fc/I < log+ p + log+ p' + 2 log+ p 1 r + log+ p, p + 1og+ T (p, f)


M (r,
+ log+T (P , f) + N(p) - N(r).

Therefore, with r < p < p' we have


(18.12)
M 1 r, f \I < 2log+ p' + 2log+ p
///
1 r + log+ p, 1 p + 21og+ T (p!, f)
+ N(p) - N(r).

We now want to make the term N(p) - N(r) small. To do this we use
the logarithmic convexity of N. [N is logarithmically convex, since n(t) is
increasing and N(r) = for ntt dt]. Since r < p < p' and N is logarithmically
convex, we have

N(p) - N(r) - log(e) (N(P') - N(r))


r)
p' -rr) (N (p) - N(r)) < r(d - N(p)
<

- p( -r)[2T(P',f)+C1,

where C is an appropriate constant, C > 1.


Now, considering 2 (PP-r ) [2T(p', f) + C] as a function of p, we see
that it vanishes for p = r and is greater than 1 for p = p'. Since it is a
Continuous function of p, we may choose p so that

P'(P - r) [2T(P',
r(p' r) f) + C] = 1.

For this choice of p, we thus have


N(p) - N(r) < 1 and
(18.13) r(p' - r)
(p - r) =
P'[2T(P',f)+C]
120 Entire and Meromorphic Functions

and
r
(p'-p)=(P -r) 1- p' ]2T (p', f) + C]
Hence,
1
1og+ < log+ p' + log+ T (p', f) + log+
p-r -
1
p' - r
Therefore,
1 1
(18.14) log+ < log+ p' + log+T(p', f) + log+
p-r p' - p.
Also, we see that
1 1
log+ < log+
p' - p - p' - r + log+ 1 r

1 + log+ 1 1
1 -2Tp',f+C
Hence,
1
(18.15) log+ 1 < log+

Using (18.13), (18.14), and (18.15) in (18.12), we have

(18.16) m (r, r) < 4 log+ p' + 4 log+ T (p', f) + 3 log+


p' - r'
which proves thelemma.
We will use the notation 11f (x) < g(x) to mean f (x) < g(x) with Borel
exceptions (i.e., a set of finite length).
Proposition.
I]m(r, 4log+ r + 8log+ T(r, f ).

Proof. Taking p\= r + fl in the lemma, we have

m (r, f, < 4log+ 1 r + log+ 31og+ 1og+ T (r, f)


! 7,(r f) +
+ 41og+ T r+
i 'f
log+ T (r, f)
Hence, by the Borel Lemma, we get

11m (r, f i < 4log+ r + 8log+ T(r, f ).


Corrollary. IIS(r) = O(1og+ T(r, f )) + O(log+ r).
Proof. This follows immediately from the proposition since S(r) is a finite
sum of terms of the form m(r, 4-).
19
"Two Constant" Theorems and the
Phragmen-Lindelof Theorems

The Two Constant Theorem. Suppose that f is holomorphic in D =


1z: IzI < 1} and continuous in D\{1}. Suppose further that if I N in D
and if 1< M in oD\{1}. Then If I < M in D\{1}.
First Proof. Choose a > 0 and let

g(z) _ (Li) u f (z)


f o r a suitable branch of (1 2 1)". Now g is analytic in D and continuous in
D if we define g(1) = 0. By the maximum modulus theorem,

zaD
suupplg(z)I = m xlg(z)I

But Ig(z)I<Mfor zEBID. Hence, Ig(z)I<Mfor zEID,and so


0
If(z)I<MII zI for zED.
Now let a - 0 to get the result.
Second Proof. This proof uses the Cauchy integral formula. Since there ex-
ists a linear fractional transformation U(z) = e'9 taking an arbitrary
Y
Point z0 E D into 0, and mapping I)/{1} conformally onto ID/{1}, we need
only prove that if (0)I < M. But choosing 00 with 0 < 00 < 7r,

if(o)i = I27r
I f(w)dwl
1 w=R w 1 1 If(rei8)Ide+ 27r
21r
l
I
2)
If(Te`B)Id0,
122 Entire and Meromorphic Functions

where R < 1 and

L)fo<1 e1<- J(2) - 1101!50o'

But

(2) Oo N'
and as R-+1-
C 21r2200
1 f(1) < (M+0(1))
Hence, letting R -+ 1-,

If(0)I! oN+Mf 1- 0)

Now letting 00 - 0 we again obtain the desired result.


Third Proof. By Jensen's Theorem,

log If (0)1 <


2,-
f 7r
log If(reze)IdO.

The argument of the second proof works here, too.


Fourth Proof. By a conformal mapping, we may replace the unit disk by
the upper half-plane. We have If I < M on the real axis and if 1:5 N above
the real axis, and must conclude that If I < M above the real axis. Choose
A > 0 and let
g(z) = (ZiA)f(z).
Note that
A I A
z + iA IzI
and that
Ig(x)I < M for all real x.
Hence, if R is large enough, we see by the maximum modulus theorem that
Ig(z)I < M for z in the semicircular region bounded by the real axis and
the are I z I = R, 0 < arg z < 7r. Hence, Ig(z)I < M in the upper half-plane.
Letting A -+ oo, we complete the fourth proof.
19. "Two Constant" Theorems and the Phragmen-Lindelof Theorems 123

A Phragmen-Lindelof Theorem. Let f (z) be holomorphic inside an


angular region of opening xr/a, where a > 1, and continuous on the closure
of the region. Suppose that If (z)I < M on the boundary and that
KeiZI-3

If(z)I <-
in the region, for some constant /3 with Q < a. Then If (z) I < M in the
region.
Proof. Without loss of generality, we may suppose that the region is given
by I arg z I < Za . Choose e > 0, and choose ry with p < y < a. Let
F(z) = e-Ezhf(z),
so that
IF(reie)I = exp{-er7'cos7O}If(z)I
For z = rese in the closed angle, we have exp{-er'1 cosry6} < 1 so that
IF(z) I < If (z) I there. Notice also that by using the estimate on f inside
the region we have
IF(re`B)I < exp{-eR'' cosryO}Kexp{RR},
so that for R large enough we have
IF(Re'o)I < M.

It follows that
If(z)I 5 Mexp{eR1},
and the result follows on letting a - 0.
Taking for simplicity a = 1, the Phragmen-Lindelof Theorem says that
if a function is holomorphic and of order less than one in a half-plane and
bounded on the boundary, then it is bounded inside by the same bound.
The example f (z) = ez in the right half-plane shows that the order 1 is
critical. However, the next result shows that the analogous result holds if
we replace the condition "order less than 1" by "growth at most order 1,
zero exponential type."
Theorem. If the hypothesis is changed to read that for each 6 > 0
If(z)I 5 K6exp{6IzI°}
in the region, then the conclusion follows as before.
Proof. Now let
F(z) = exp{-eza} f (z),
and the same proof works.
Remark. Analogous results hold for functions holomorphic in other regions
and satisfying appropriate growth restrictions. One useful case is the paral-
lel strip. Calderon has used this case in developing a theory of interpolation
of Banach spaces that can be applied in the fields of harmonic analysis and
Partial differential equations.
20
The Polya Representation Theorem

The Polya Representation Theorem plays a central role in the theory of en-
tire functions of exponential-type. We give a somewhat augmented version
of this theorem.
Before proceeding with the theorem, it will be necessary to discuss con-
vex sets. We say that a set E (in the complex plane) is convex if E con-
tains the line segment joining any two points. That is, if z1, z2 E E, then
tzl + (1 - t)z2 E E for all t E [0,1]. The intersection of convex sets is again
convex.
Definition. Given a set A, the intersection of all half-planes that contain
A is called the closed convex hull of A and is denoted by K(A).
Definition. A point is an extreme point of a set if it is not the midpoint
of any line segment contained in the set.
Theorem 20.1. A compact convex set is the convex hull of the set of its
extreme points.
We omit the proof.
Definition. For any set E, k(8) = sup{Re(ze-'B) : z E E} is called the
support function of E.
We note that k(O) measures the directed distance from the origin to the
most remote point of the projection of E on the ray arg z = 0. Note also
that if E is empty, then k = -oo. It is easy to show that, for a given set
E,
K(E) = {z : Re(ze-'B) < k(0) for all 0}.
Remark 20.2. If zo = xo + iyo = roe'Bo and E = {zo}, then k(0) =
TO cos(0 - Bo) = zo cos 0 + yo sin 0.
20. The Polya Representation Theorem 125

Remark 20.3. Let E be a circle with center at 0 and radius R. Then


k(8) = R for all 0.
Remark 20..¢. Let E be the line segment [xo, x1], xo < x1. Then k(8) _
x1 cos 0 if - a < 0 < 2 and k(0) = xo cos 8 if 2 < 0 < s2 . In particular,
if xo = -x1 and x1 = R, then k(0) = RI cos 01. If E is the vertical line
segment [-iR, iR], then k(8) = RI sin 01.
Remark 20.5. If E1 has support function k1 and E2 has support function
k2, then E1 + E2 = {z1 + z2 : z1 E E1, Z2 E E2} has support function
k1 + k2. From Remark 20.3, it follows then that the rectangle with vertices
(±R1i ±iR2) has support function k(0) = R1I cos e] + r21 sin 01.
Remark 20.6. The convex hull of E1 U E2 has support function k =
max{k1, k2 }.

Remark 20.7. If E1 with support function k1 is translated, so that the


point originally at 0 is moved to zp = xo + iyo, then the support function
of the translated set is k(O) + xo cos 0 + yo sin 0.
Definition. Let Ho(oo)be the class of all functions that are holomorphic
near oo and that vanish at oo.
Let f be an entire function of exponential-type and write f (z) _
() zn. The BoreI transform -6 of f, defined by 4'(w) = > anw mar,
belongs to Ho. Each function in Ho is the Borel transform of a unique
f. We have seen (Corollary to Proposition 11.5) that the type of f is the
radius of convergence of the series >2 an war.
In Proposition 11.7, we saw that D(w) = f ow f (t)e- t7°dt, in the sense of
analytic continuation. We also saw that

f (z) = 14'(w)ez'°dw,
27ri , r
where r is a rectifiable curve that winds once around the singularities of
I. We call this the P61ya integral representation formula.
Definition. Let S(4') be the set of singular points of 4', and let k be its
anpporting function. We call S(4') the conjugate indicator diagram of f.
Sometimes this name is used for S*(4'), the closed convex hull of S(4'). We
will write D(f) =
Definition. The indicator function of f is

h(8) = hf(0) = limsup 1 log ]f(reie)1.


r-.oo r
We now state an important part of the Polya Representation Theorem.
126 Entire and Meromorphic Functions

Theorem 20.8. h(0) = k(-0) for all 0.


The proof of Theorem 20.17 is contained in an appendix at the end of
this chapter. We now make some remarks to illustrate this theorem.
Remark 20.9. If f is of zero-type, then h = 0 so that k = 0, and hence 0
is the only possible singularity of 1.
Remark 20.10. Denoting by r(f) the type off, we have r(f) = max h(0).
Remark 20.11. If f (z) = eaz, where a is real, then

h(0) = lim sup 1 log lea` l = lim sup 1 ar cos 0 = a cos 0.


r.-.oo r r-.oo r
On the other hand, O(w) a" w- r = (w - a)-1 and so S(O) = {a}.
Thus, k(0) = acos0 and we do have h(9) = k(-0).
Remark 20.12. If f (z) = eiz, we have h(0) = - sin O and 4)(w) = (w-i)-1.
Hence, k(0) = sin 0 since S(4,) _ {i}. Note again that h(O) = k(-O).
Remark 20.13. Suppose f (z) _ > a,,ea^z, a finite sum with distinct A,,
and nonzero an. Then O(w) = so that S((D) = {A,.}. Note
that only the extreme points of affect h(0). If the A,, lie on a straight
line, only the endpoints affect h(0).
Remark 20.14. It is easy to see that h f.9 < h f+h9, so that D(fg) C D(f)+
D(g). An interesting problem that has applications in harmonic analysis
is that of finding suitable conditions under which D(fg) = D(f) + D(g).
Let Mo be the class of all Borel measures of compact support.
Definition. If dp E Mo, its Laplace transform dµ^ is defined by

dµ^(z) = e-zwdA(w).
J

It is easy to see that dµ^(z) is an entire function of exponential-type for

dp^(z) = f(_w)e'd/L(w)
dz

(as can be verified on differentiating "by hand") so that d,i' is entire. And

eRjzj
Idi^(x)I < f Iez `Jjdµ(w)I 5 f Idu(w)I,

where R is chosen so that the support of dµ lies in the disk of radius R


centered at 0.
20. The Pblya Representation Theorem 127

Definition. We write dµ - dv to mean that dµ^ = dv^.


It is not hard to show that dp - dv if and only if f f dµ = f fdv for
each entire function f, or for each entire function f of exponential-type. It
is clear that - is an equivalence relation.
If we take dµ = dzjr, where r is a circle, then dp^(z) = fr e-a'dw = 0
by the Cauchy Theorem. Hence, dµ - 0 even though dµ 0 0. We shall
see that for any dµ E mo, dµ - dv, where dv = 4P(-w)(-dw)Ir, where
is the Borel transform of dp^ and IF is any curve that winds once around
S(C.
Definition. [dp] is the class of measures equivalent to dp.
Definition. Mo is the class of all [dp] for dµ E Mo.
Definition. If f is continuous in the plane and dp in Mo, we define the
convolution f * dµ by

(f * dµ) (z) = J f (w - z)dp(w).

Definition. If dp and dv are in Mo, we define the convolution dµ * dv by

f*(dµ*dv)=(f*dµ)*dv.
By means of the Riesz Representation Theorem, it is easy to see that
the above definition defines dµ * dv as a unique measure in Mo. Indeed,
Mo is an algebra over the complex numbers.
Proposition. If dp1 - dp2, then (dpi * dv) - (dµ2 * dv). It thus makes
sense to define [dp] * [dv] = [dµ * dv]. Ma is an algebra over the complex
numbers.
Definition. Let Eo be the algebra of all entire functions of exponential-
type.
Definition. Let E be the space of all entire functions in the topology of
Uniform convergence on compact sets.
Definition. E', the dual of E, is the space of all continuous linear func-
tionals on E.
Now E is a locally convex topological linear vector space. It will appear
that each of the spaces Mo', Eo, Ho(oo) "is" the dual space E.
Definition. For f E E and [dµ] E Mo', define the inner product (Fl, [dµ])
hY

(F1, [dp]) = (F * dp)(0) = f F(-z)dp(z).


128 Entire and Meromorphic Functions

Definition. For F E E and f E Eo, define the inner product (F, f) by

(F,f) = (f(D)F)(0),
where D = dZ. This means that if f (z) = E z", then (F, f)
n F(")(0)
It is not hard to show that, for each f E Eo and F E E, the series defining
(F, f) converges. Indeed, the linear functional A defined by A(F) = (F, f)
is a continuous linear functional on E. The same is true for A(F) = (F, dµ).
Definition. For F E E and fi E H0(oo), define the inner product (F, 4b)
by

(F, _1I
2ri
<P(w)F(w)dw,
r
where r is any curve that winds once around S(4')
Again, A(F) = (F, -6) defines a continuous linear functional on E.
Theorem 20.15. Let [dp] E MO', f E Eo, -0 E H0(oo) be related by
f = dµ^ and P be the Borel transform of f. Then dµ, f, and give rise
to the same functional in E'. And given any functional in E', there is a
unique [du] (or f or t) that gives rise to it.
To say that f gives rise to A is to say that A(F) = (F, f) for each F E E,
with a similar definition for dp or 1b. We omit the proof of the theorem
since much of it is straightforward.
Remark. To illustrate the theorem, take f = ez so that d IA is the unit point
mass at 1 and *1(w) = (w - 1)-1. Then f (D) = eD, so that by Taylor's
theorem

(f (D)F)(0) = F(O) + F'(0) + 1 F"(0) + = F(1).

Note that f F(-z)dp(z) = F(1) also. And, by Cauchy's Theorem,

27ri
f F(w)41(w)dw =
- f w(-w) dw, = F(1).

Summarizing the results of this section so far, we have the following


omnibus theorem. We call it the Pd1ya Representation Theorem, although
the name is not entirely appropriate.
The P61ya Representation Theorem. Let Eo be the algebra of all en-
tire functions of exponential-type and Ma the convolution algebra of equiv-
alence classes of Borel measures of compact support, where du dv means
that f f dµ = f f dv for each entire function f. The Laplace transform is
defined by dµ^(z) = f e-Zwdp(w); dp - dv if and only if dp^ = dv^, so
20. The P61ya Representation Theorem 129

it makes sense to talk of [dp]^ = du^, where [dµ] is the equivalence class
that contains dµ. The Laplace transform is an isomorphism of Mo onto
Fo. To invert the Laplace transform, i.e., given f E E0, to find dp E Mo
such that dµ^ = f, take dp(z) = 4)(-z)(-dz)Ir, where 4) is the Borel
transform off and t winds once around the set S(4)) of singularities of 4).
The indicator diagram of f, D(f), is defined as the convex hull of S(4)). If
h(O) = lim sup '° f t e is the indicator function of f, then h(O) = k(-6),
r-.oo
where k is the support function of D(f ). If E is the space of all entire func-
tions in the topology of uniform convergence on compact sets, then both Eo
and MO' are the dual space of E, where, if f = dp^, then for F E E

(F, f) = (f (D)F) (0) = (F, dµ) = JF(_z)d(z).

We now turn to some applications of this theorem.


[Note. We repeatedly use conventional, but strictly incorrect, phrases
like "r winds once around the set S(4)) of singularities of C" A more
correct wording is "r has winding number -1 around oo, and lies in a
connected and simply connected open set containing co in which 4 is ana-
lytic."]
Definition. For an entire function f and a complex number A, let fa be
the entire function defined by

f.\ (z) = f (z + A) for all z.

Lemma 20.16. For any entire function f, -r(fa) = r(f ).


Proof. With M(r : f) = max{ I f (z) J : IzI = r}, we have

M(r : f'0:5 M(r + IAI : f)

so that

I logM(r : fa) <- 1M(r+ IAI : f) = (1+0(1))r+ ICI logM(r+ IAI : f)-

Hence,
r(A) <- r(f).

Similarly,
-r(f) < T(fa) since
130 Entire and Meromorphic Functions

Lemma 20.17. If f and g are entire functions of exponential-type, then


D(f) = D(g) if and only if
T(f(z)eaz) T(g(z)eaz)

=
for each complex number a.
Proof. Clearly, if D(f) = D(g), then r(f (z)eaz) = r(g(z)e6z). To prove
the converse, it is enough to compute, say, h(O) from r(f (z)eaz). We show
that
h(O) = lien ar(f (z)e6z) - a.
a

Let Da be the indicator diagram of f (z)eaz. Then Da = a + D(f), that


is, D. is D(f) translated by a. Choose B < max (uhf (z) 1, Jhf (-3) 1)
Now when a is large, D. lies to the right of the origin, and Da lies in
the strip {z = x + iy : jyj < B; x < a + h1(0)}. D. also contains the point
(a + h(0), 0). Hence, if T. = T (f (z)eaz), then
Ta > a + h(0)

T < {(a + h(0))2 + B2} j,

Ta - (a + h(0)) = o(1).

Lemma 20.18. If f is an entire function of exponential-type, then for


each complex number A,
D(f) = D(fa)-
Proof. We use Lemma 20.17, with g = fa. Note that
e-aa(ea(z+a)f(z
e6zg(z) = eazf(z + A) = + A)).

So if we let F(z) = e6z f (z), then e6Zg(z) = CFA (z), where C is a nonzero
constant. Since r(F,) = T(f ), we have r(e6z f (z)) = r(e6Z fa (z)), and the
result follows.
The P61ya Theorem has the following corollary.
Proposition. If h (2) < 0 and h (- z) < 0, then f = 0.
Proof. h (2) < 0 implies that D(f) lies above the real axis )<0
implies that D(f) lies below the real axis. Hence, D(f) is ...
consequently, 4 has no singularities.
Since -t(cc) = 0, it follows from Liouville's Theorem that 0 = 0, ano
hence f = 0.
The Pd1ya Representation Theorem provides an alternate proof of Carl-
son's Theorem presented in Chapter 12.
20. The Polya Representation Theorem 131

Theorem (F. Carlson, 1914) 20.19. Suppose f is an entire func-


tion of exponential-type with r(f) < w, and suppose that f (n) = 0,
n = 0, ±1, ±2,.... Then f = 0.
Proof. Consider the function g(z) = f (z)/ sin wz. It is clear that g is entire.
Since
T(r,F/G) <T(r, F) + T(r, G),
g is of exponential type. Now hg (-E) < 0, and hg (a) < 0, so that by the
above proposition, g = 0. Hence f = 0.
Remark 20.20. It is clear from the proof that the hypothesis that r(f) <
x can be relaxed to h f(f7r/2) < w. The condition r(f) < r is sharp,
since sin wz is of type w. The Carlson Theorem says, roughly, that an
entire function of exponential-type must grow fast in a direction at right
angles to its zeros. Later in this section, we shall get a stronger version of
Carlson's Theorem. Chapter 22 is devoted to proving an extremely strong
generalization of it.
Definition. Let k denote the set of all entire functions f of exponential-
type such that h f(±w/2) < w.
Definition. A sequence {An}, n = 0, 1, 2.... is said to be k-admissible
p r o v i d e d that there exists an f E k such that f (n) = An, n = 0,1, 2, ... .
We now give a characterization, due to Buck, of k-admissible sequences.
Theorem (Horseshoe Theorem) 20.21. The sequence {An} is called k-
admissible if and only if F(z) = E Anz" is holomorphic at 0 and oo and
in some neighborhood of the negative real axis as well.
Proof. Suppose first that {An } is k-admissible, so that An = f (n) for some
f E k. Then, if z is small,

F(z) _ f(n)zn = fb(w)enwdw I zn (z) dw.


(2wi 2wi 1

We choose 1 to be a rectangular path that winds around S(4') such that


r lies in the strip
S. = {z = x + iy : lyl < w}.
It follows that F(z) can be analytically continued in the complement of the
image of S(4') under the mapping a-", and that F(oo) = 0.
To prove the other half of the theorem, suppose that F satisfies the
requirements of the theorem. We may write

_ 1 r F(z) dz
An 2wi A zn z'
132 Entire and Meromorphic Functions

where A is a circle around 0. But A is homotopic, in the complement of


S(F), the set of singularities of F, to the path AR illustrated in the figure
below, where A' is a circle of radius R.

T U Horseshoe Contour

Hence,
1
An _27ri r F(z) dz
JAR zn z
Now
-+ 0 as R --+ oo, since F(oo) = 0.
I R
Hence,
An
_ 1 /' F(z) dz
27ri A. z" z'
where A* is any curve that winds around S(F) with multiplicity -1. We
thus are led to define
F(z) dz
f (W)
= 217ri J z- z

for an appropriate branch of zw on A*. It is easy to check that f E k, and


the proof is complete.
We now may prove the following extension of Carlson's Theorem.
20. The PGlya Representation Theorem 133

Theorem 20.22. Suppose that f E k and that f(n) = 0 for n e


0,1,2,.... Then f = 0.

Proof. We have

F(z)
_ Ef
(n)zn
= 2ri
J(1 - zew)-1-t(w)dw.

By hypothesis, F = 0. Hence,

zF(z) = 1 J z i(w)dw = 0.
2iri r 1- ze'
On letting z --, oo, we have

' J ew-b(w)dw = 0.
ri J

Since f (z) = -I- Jr we have f(- 1) = 0.


Now we use the fact that f E k; thus f_1 E k, where f_1(z) = f(z- 1).
We may apply the same argument to f -I to conclude that f_1(-1) = 0,
that is, f(-2) = 0. Repeating the argument, we get f (n) = 0 for n =
0, so that f = 0 by Carlson's Theorem.
We may prove several theorems about k-admissible sequences by means
of the above characterization using classical theorems of function theory
that relate the properties of a function to its power series coefficients.
Theorem 20.23. If {An} is k-admissible and JAnI1"" -* 0, then An= 0
for each n.
The proof is trivial.
Theorem (Pringsheim). Let f (z) _ E anz" have nonnegative coefficients
and radius of convergence R. Then z = R is a singular point of f.
Theorem 20.24. If {An} is admissible and (-1)"An > 0 for n =
0,1,2,..., then An=0 fore=0,1,2,....
Proof. F(-z) = E(-1)"Anzn has nonnegative coefficients, but no positive
real number is a singularity of F.
Hadamard Gap Theorem. If AZ) = E anz" with an = 0 except for
n = nk, where lim 1, then every point of the circle of convergence
off is a singular point of f.
Theorem 20.25. If {An} is k-admissible and An = 0 except for n = nk
with lim nk+1 > 1, then An = 0 for n = 0, 1, 2, ... .
nk
This is a simple consequence of the Hadamard Gap Theorem.
134 Entire and Meromorphic Functions

Fabry Gap Theorem. If f (z) = > anzn with an = 0 except for n = nk


and k-'nk -+ oc, then every point on the circle of convergence of f is a
singular point of f.
Theorem (Szego). Suppose that f (z) _ anzn, where the an lie in some
finite set. Either IzJ = 1 is a natural boundary of f or f is a rational
function, and the an are eventually periodic.
As a corollary we have
Theorem 20.26. If {An } is k-admissible and the An lie in some finite
set, then the An are eventually periodic.
Appendix
The proof that h(0) = k(-0) is presented in this section. The actual
proof of this assertion is fairly simple, but we prefer to give some of the
background concerning supporting functions of convex sets. First, we give
a simple necessary and sufficient condition that a function h(0) should be
the supporting function of a nonempty compact convex set. The condition
is that the function should be "subsinusoidal." Next, we prove that if h(0)
is the indicator function of an entire function of exponential type, then
h(0) is subsinusoidal. Finally, we show that h(0) is the supporting function
of the conjugate of the conjugate indicator diagram. From now on, when
we speak of a "function of 0," we mean a function that is 27r-periodic; and
when we speak of a "supporting function," we mean a supporting function
of a nonempty compact convex set.
Our treatment is a combination of the treatments in Pd1ya [31] and Boas
[5].

Definition. A function H(0) is a sinusoid if it has the form H(O) _


acos0+bsinO.
Remark. Given 01 $ 02 and real numbers h1 and h2, there is a unique sin
soid H such that H(01) = hi and H(02) = h2. We call H the interpolating
sinusoid: It is given by (0 < 02 - 01 < 7r)

sin(02 - 0) sin(0 - 01)


(20.1) H(0) =
hlsin(02 - 01) + h2sin(02 - 01)

Definition. Given a function h(0) and 01 # 02, we call H the interpolating


sinusoid of h if H is given by (20.1) with h1 = h(01) and h2 = h(02).
Definition. A function h(0) is subsinusoidal if it is majorized by each of
its interpolating sinusoids, that is,
h(o1)sin(03 - 02) + sin(02 - 01)
(20.2) h(02) < h(03)
sin(03 - 01) sin(03 - 01)

whenever 01 < 02 < 03 with 0 < 03 - 01 < ir.


20. The P61ya Representation Theorem 135

Remark. The theory of subsinusoidal functions has some similarity to the


theory of convex functions.
Remark. If h is subsinusoidal, and if H is sinusoidal and H(81) > h(81),
H(82)>h(02),then H(8)>h(8) if81 <0<82with 0<82-81 <7r.
Remark. The sum of two subsinusoidal functions is subsinusoidal.
Remark. That h is subsinusoidal is equivalent to the assertion that the
point h(0)eie does not lie outside the circle that passes through the points
0, h(01)e",, and h(02)e`°2, where 81, 02, and 0 are in the specified range.
This geometric interpretation can be used to supply geometric proofs of
some of the subsequent results.
Problem. Suppose that h1 is subsinusoidal, h2 is supersinusoidal, and
h1(0) > h2(8) for all 8. Does there exist a sinusoidal function H such that
h1(0) > H(O) > h2(9) for all 0?

Lemma 20.27. Suppose that h is subsinusoidal, that 01 < 02 < 83, that
0 < 03 - 0 < 7r, and that H(8) is a sinusoid such that h(01) < H(81) and
h(02) > H(02). Then h(03) > H(03).

Proof. Suppose b > 0 and h(83) < H(03) - b. Let H6 be the sinusoid such
that
H6(01) = H(01), H6(83) = H(03) - 6.
Then H5(02) < H(02), since
sin(0 - 01)
H5(8) = H(8) - 6
sin(83 - 0&
Since h is subsinusoidal and we have
h(01) < H5(01), h(02) 5 H6(02),

it follows that
h(02) < H6(02) < H(02),
which is impossible.
Lemma 20.28. A function h(8) is subsinusoidal if and only if
(20.3) h(91)sin(03 - 02) + h(92)sin(01 - 03) + h(03) sin(02 - 01) > 0
whenever 01 < 02 < 83, 02 - 01 < 7r; 03 - 02 < 7r.
Proof. Clearly, (20.3) is equivalent to (20.2) if 83 - 01 < 7r. To prove
(20.3) in general, choose 04 so that 02 < 04 < 81 + 7r and let H(8) be the
interpolating sinusoidal for h at 01, 82. By Lemma 20.27, h(84) > H(04).
Repeating this argument with 02, 04, 03 we get h(03) > H(03). Now
h(81)sin(03 - 02) + h(02)sin(81 - 03) + H(03) sin(02 - 01) = 0,

but sin(02 - 81) > 0 and h(03) > H(03), so the result follows.
136 Entire and Meromorphic Functions

Lemma 20.29. If h is subsinusoidal, then it is continuous and even has


left and right derivatives at each point. The left derivative is never greater
than the right derivative.
Proof. Choose 0 and suppose without loss of generality that h(8) < 0.
Otherwise, consider h - H where H is a sinusoid such that H(8) > h(8).
Choose c > 0, 6 > 0 with c + b < ir. Applying (20.2) in turn to the following
triples cp1, V2, 03:
(i) 0 -E-6, 0-E, 0
(ii) 0-E, 0, 0+E,
(iii) 0, 0 + E, 0 + E + 6,
we eventually obtain
h(8) - h(8 - E - 5) h(0) - h(0 - E)
<
sin(E + 6) sin f
(2U 4)
< h(8 + E) - h(8) < h(0 + E + 6) - h(8)
sin f sin (E + 6)

For example, considering the triple (i), we get


h(8 - E - 6) sin E - h(8 - E)sin(E + 6) > -h(8) sin 6
so that
[h(8 - E - 6) - h(0)] sin c - [h(0 - E) - h(8)] sin(E + 6)
> h(0)[sin(E + 6) - sin c - sin 6]

= -h(8)2 sin E
6 [COS 2 b - cos - --j 1
> 0.

Hence,
2

[h(0 - E - 6) - h(0)] sin c > [h(0 - E) - h(9)] sin(c + 6),


which proves the first inequality of (20.4). The others are proved in a
similar way. Now (20.4) is precisely the assertion that h(8+ex)-h(8) is an
increasing function of x when x is small, from which all of the assertions
of the lemma follow easily.

Lemma 20.30. If h'(8) denotes the nght derivative of h(0), then


h(0) cos(W - 0) + h'(0) sin(V - 0) - h(V) < 0
for each pair 0, cp with 0 - ir < ' < 0 + ir.
Proof. Choose c so that 0 < E < it and apply (20.2) to the following triples
WP1, cp2, V3: (i) gyp, 8, 0 + E (ii) 0, 0 + E, W. This use of (20.2) is admissible if
either 8-ir<8<.Xor0+E<V<E+7r. We get
(20.5) h(8) sin(<p - 0 - E) + h(0 + E) sin(O - cp) + h(w) sin c > 0,
20. The Pdlya Representation Theorem 137

which may be rewritten as


(20.6)
0) - sin(cp - 0 - e) h(0 + e) - h(O)
h(9) sin(cp - 0) - h(V) < 0.
sin e + sine
Now let e - 0 to get the assertion of the lemma.
Theorem 20.31. The function h(9) is the supporting function of some
nonempty compact convex set K if and only if h(0) is subsinusoidal.
Proof. Suppose first that h supports K, so that for each z = x + iy E K
and any 01, 03 we have
(20.7) xcos01 +ysin01 < h(91)
(20.8) xcos93+ysin93 <h(83).
If we now choose 01, 02, 03 with 01 < 02 < 03, 02 - 01 < ir, and 03 - 02 < 9r,
we may multiply (20.7) by the positive quantity sin(03 - 02) and (20.8) by
the positive quantity sin(02 - 91) and then add to get
h(91)sin(03 -02)+h(03)sin(02 -01) > (xcos02+ysin02)sin(03-01);
but x cos 0 + y sin 0 = h(0) for some z = x + iy in K, and it follows that h
is subsinusoidal.
To prove that if h is subsinusoidal, then it is supporting, let
Ke={z=x+iy:xcos0+ysin9<h(9)}
and then let
K=nK9.
0
Each Ke is a closed half-plane in the direction 0. It follows that K is closed
and convex. Further, K is bounded since it is contained in the rectangle
KO n K z n K. n Kg f . We now prove that each K9 contains a point of K
on its boundary, thus completing the proof of the theorem. Given 0, then,
we must prove that there exists a point ze = xe + iye E K for which
(20.9) xe cos 9 + ye sin 9 = h(0).
By the theory of envelopes, we also want
(20.10) xe sin 9 - ye cos 9 = -h'(0),
where we interpret h'(9) as the right derivative.
Following this heuristic idea, we let xei ye be the simultaneous solution
of (20.9) and (20.10). But on applying Lemma 20.30, we get for all cp
xe cos cp + yo sin cp < h(cp).
It follows that zB E K, and since
h(9) = max{x cos 0 + y sin O : x + iy E K},
the result is proved.
138 Entire and Meromorphic Functions

Theorem 20.32. If f is an entire function of exponential type and


h(0)=hf(0)=r rlogIf(rei8)I,
then h(0) is subsinusoidal, and h(0) is consequently a supporting function.
Proof. The proof is a simple application of the Phragmen-Lindelof princi-
ple. Suppose that 0 < 02 - 01 < x, and let h1 = h(01), h2 = h(02), and
for 6 > 0 let H6 be the interpolating sinusoid for h1 + 6, h2 + 6 at 01i 02.
Let A = A6 be the complex number for which H6(0) = R{Ae-'B}, and let
F(z) = f (z)e Az so that if z = reie, then
IF(z)I = If(z)I exp{-rH6(0)}.
Now F(z) is bounded on the rays z = reie', z = re'e2, and of order 1 in
the angle between. By the Phragmen-Lindelof Theorem, F(z) is bounded
in the angle 01 < 0 < 02, and it follows that h(0) < H(0) for each 0 in this
range. Since llm H6(0) = Ho(0), the result is proved.
6-0
Theorem 20.33. If f is an entire function of exponential type and h(0) _
h(0) = Um I log If (re") I, then h(0) = k(-0) where k is the supporting
r-00 I'
function for the conjugate indicator diagram.
Proof. Let I' = Of be the Borel transform off and let D = D f, the conju-
gate indicator diagram of f , be the convex hull of the set of singularities of
0. Let C be a rectifiable curve that winds once around D and that stays
within an E-neighborhood of D, where e > 0.
Now
f (x) =
if
I -b(w) exp(zw) dw,
2i c
so that if z = refe, then
If (reie)I < A mEax I exp(zw) I,

where A is a constant. Hence,


h(0) < max R(weie).
WEC

If we now let c - 0, we see that h(0) < k(-0).


In the other direction, it is enough to prove that h(0) > k(O) since, if
we replace f (z) by g(z) = f (ze'`°), the general case follows from hg(0) >_
kg(0), since hg(0) = h f(0 + gyp), 4Ds(w) = e"P4bf(we-"°) Dg = e"'Df and
kg(0) = k(0 - cp). Now, asrwe have seen,

for w > h(0),


-O(w) =
J0 co f (t)e-*w dt
so that fi has no singularity to the right of the line x = h(0) and the
inequality h(0) > k(O) follows.
21
Integer-Valued Entire Functions

An integer-valued entire function f is one such that f (n) is an integer for


n = 0, 1, 2, . Some examples are
(i) sin irz
(ii) 2Z
(iii) any polynomial with integer coefficients.
In this section, we shall mainly follow a paper of Buck [7]. In outline,
a certain construction generates a special class Ri of integer-valued entire
functions. We will be concerned with finding growth conditions on an
integer-valued entire function f that imply f E R1. The three examples
above belong to R1.
Definition. We say that an algebraic number a is an algebraic integer if
it satisfies a polynomial equation:
(21.1)

where aJ E Z, j = 0, I,_ , n - 1. Notice that the coefficient of z" is 1.


Examples. Any n E Z satisfies z - n = 0. And ±i satisfies z2 + 1 = 0.
It is not hard to prove that the algebraic integers form a ring. If the
integer n in (21.1) is minimal, the other roots are called the conjugates of
Q. The collection of all of the roots of a minimal polynomial is called a
complete set of algebraic conjugates. It is not hard to show that if a is a
root of P (where P is not necessarily minimal), then each conjugate of a
is also a root.
Consider now the polynomial

Q(x) = 1±q1x+g2z2 +... +qnz",


140 Entire and Meromorphic Functions

where qj E Z f o r j = 1, 2, ... , n. We can write


n
Q(x) _ fl (1 -$jz),
j=1

where the f3j run over one or more complete sets of algebraic integers. This
may be seen from the fact that the f3j are the roots of the polynomial

R(x) =
znQ ( 2) = xn + Q1xn-1 + ... + qn

Now let P be any polynomial with integer coefficients and Q as above.


Then
P(x)
= E bnxn, where bn E Z.
Q(z)
This follows since we can write
1 1 4;(,)
1
= 1 +Q`(z) + [Q*(z)]2 + .. .
Q(x) 1 + qlx + + gnxn 1-
= 1+B1x+B2x2+ .
The BI are clearly integers. To express the bj in terms of the ,6j, let us
first suppose for simplicity that the f3j are distinct, and that P = 1. Using
partial fractions and writing
00
= r fjnxn'
1 - #jx L1
n=1

we have
E.
M 00
1 = L.r
j=1 1 - f3jx
=j=1Ln=O
L
11(1 - Qjx)
1

where the Ej are the coefficients in the partial fraction expansion. Hence,
n
L
bn = [ Ejf31",

j=1

so that it is natural to take


m
f(z) = Ej/,,
j=1

where Q. is suitably defined.


21. Integer-Valued Entire Functions 141

Example

Q(z) = 1 + z2 = (1 + iz)(1 - iz)

2{1+iz+i2z2 } }
Q(z) 21 iiz+21+iz
+
[in + (-i)n]
2
1,0,-1,0,1,0,-1,0,...
f (z) = 2 [iZ + (-i)Z] = cos 2 Z.

In case Q has repeated roots, or if P is not constant, some minor modifi-


cations must be made, and in general we have
m
b = Ef'.i(n),

so that we take

(21.2) f(z) _ Pj(z)Q,,


i=t

where the P3 are suitable polynomials (not necessarily integer-valued).


Definition. Let Rl be the class of functions f constructed above.
Definition. Let R be the class of integer-valued entire functions f of
exponential-type for which h f(±ir/2) < 7r.
Our problem is to find additional growth conditions on f that imply
f E R1 if f E R. The conditions will be phrased in terms of the "mapping
radius" of certain sets associated with the indicator diagram of f.
Definition. Let S be a simply connected open set containing 0 such that
the complement of S contains at least two points. Let cp be the function
(whose existence and uniqueness is guaranteed by the Riemann Mapping
Theorem) that maps S conformally one-one onto the unit disk D = {z :
IzI < 1} and such that w (O) = 0 and cp'(0) > 0. Then cp is the normalized
mapping function of S and p(S) = ) is called the mapping radius of S.
We shall need the following deep theorem of P61ya, which we state with-
out proof. The proof may be found in [8].
142 Entire and Meromorphic Functions

Theorem. If g(z) = >J 0 b,, E Z, and if g is analytic in a region


S containing 0, with p(S) > 1, then there exist polynomials P and Q, with
integer coefficients, Q(0) = 1, such that
P
9=Q
We also require a simple lemma on polynomials, whose proof we leave to
the reader.
Lemma. If P and Q are polynomials with integer coefficients and Q(0) =
1, then there exist polynomials P1 and Q1 with integer coefficients Q1(0) =
I such that P1 and Q1 have no common factors and PI /Q1 = P/Q.
Now given f E R, let

f(n)z'.

9(z) _ 00
n=0

As we have seen earlier,


1 f 1
9(z) = 2ri 1 - zew -b(w)dw
r
for any curve t that winds once around D(f) = S`(4). Now let S be
the complement of the image of D(f) under the mapping e''', i.e., S =
C\ exp(-D(f)).
Now g is analytic in S, so that if p(S) > 1, then g = , where P and
Q are polynomials with integer coefficients and Q(0) = 1. By the lemma
above, we may suppose that P and Q have no common factors. By the
construction that characterizes R1, we can find a function fl E R1 such
that f, (n) = f (n) for n = 0, 1, 2,.... By Carlson's Theorem, if we know
that h f, (f 2) < ir, then we have f = fl, so that f E R. To prove that
h f, (±M) < r, we write
M
fl (z) = E P, (z)f; .
=1

By construction, the fji 1 are the roots of Q, so that the ,QJ 1 are the
singularities of g, and hence the 131 1 are in the complement of S. Hence,
we may write f3j = exp(-,uj), where µj E D(f), so that
m
fi (z) _ Pi (z) exp(-FUiz),
j=1

and it follows that hf, (f 2) < a since D(f) is interior to the strip Iyj < 7r
We therefore have proved the next theorem.
21. Integer-Valued Entire Functions 143

Theorem. If f E R, and if the complement of the image of D(f) under


the map a-' has mapping radius exceeding 1, then f E Rl.
For applications, a variant of the foregoing procedure gives a more useful
result. We let A be the difference operator defined by

(Af)(z) = f(z + 1) - f(z),

defining
A°f = f
and
An+lf = A(Anf)
Now, using Taylor's Theorem, we may write

A=eD-1
An = (eD - 1)n

where
D
dz'
We define the functionals Tn and Tn by

Tnf = f (n)
T* f
= (Anf)(0)
To illustrate,
T = f (0)
Tif =f(1)-f(0)
T2*f=f(2)-2f(1)+f(0)
T3 *f = f(3)-3f(2)+3f(1)- f(0).
It is easy to show that

T,a = (-1)n >(-1)k (n)


0 Tk.

Tn=E(k)Tk.
For example, the first identity is proved on writing
n
An = (eD - 1)n = ()e)(_1y1_1c.
k0
144 Entire and Meromorphic Functions

Definition. To say that a sequence {bn }, n = 0, 1,2,..., is K *-admissible


is to say that there is an f E R such that

T; ,f =b,,, n=0,1,2,....
For f E R, write
00

g(z) = E(Tn*f)zn-
0

It is easily seen that

g(z) = 27ri -D(d)dw'


Jr 1 - z((w)
where ((w) = e' - 1 and I' is a curve that winds once around D(f). We
see that g is analytic outside the image of D(f) under the map (e'° -1)-1.
The argument may be reversed to prove the next result.
Theorem. A sequence {bn} is K`-admissible if and only ifEbnzn is
analytic on the segment [-1,01.
We may now prove the main result of this section.
Theorem. If f E R1, let E be the complement of the image of D(f)
under the mapping (e' - 1)-1 and let E" be the image of D(f) under the
mapping eL - 1. If p(>) > 1, then f E R1 and f (z) = F, Pk(z)(1 + ak)z,
where the Pk are polynomials and the Qk run through the complete sets of
conjugate algebraic integers lying in `.
Proof. Let
g(z) = E(Tk f)zk,

and let
F(z) = E(Tkf)zk
As we have seen,
1 1
9(z) = dw
21ri Jr(w) 1 - z(e"' - 1)
F(z) ID(w)1 dw.
27ri J r - zew
Now,
1 1 / 1
9(z) = 27ri 1 + z (w)1 - 1+z ew dw
r
so that
9(z)= 1+zF(1+z}
21. Integer-Valued Entire Functions 145

Similarly,
F(w)= Iwg+
I 1 ww )-
Since p(E) > 1, we see by the P61ya Theorem that g = 11, where P
and Q are polynomials with integer coefficients and Q(O) = 1. Thus,

1 P(1 w)
w) (1-w)'+1Q(1-WW) Q*(w)'

where we choose N > max(deg P, deg Q). Now, P* and Q* are polynomials
with integer coefficients and Q* (0) = 1. Thus, f E R1. As before, we see
that
f (z) = E A(z)'Y; "
where the yi are the reciprocals of the roots of Q* and the Pi are polyno-
mials. If we write yi = 1 +- pi, we see that Qs 1 is a root of Q, and since the
roots of Q are the singularities of g, the theorem is proved.
Using this theorem and some facts about algebraic numbers, the next two
results can be proved easily. We state them without proof, as illustrative
applications. For details, see the paper of Buck [7].
Theorem. If f is an integer-valued function of exponential type such that
hf(7r/2) = hf(-ir/2) = 0 (that is, the indicator diagram off is a horizontal
line segment), and if L = exp h f(0) - exp h f(7r) < 4, then f E R1.
Theorem. If, in addition, L < V5-, then for some polynomials
Po, Pi,... , Pn, we have
f (z) = Po(z) + Pi (z)2z + + P,n (z)nz.
22
On Small Entire Functions of
Exponential-Type with Given Zeros

This chapter is extracted from a paper of the same name by P. Malliavin


and L. A. Rubel [22]. We obtain here a result that considerably generalizes
Carlson's Theorem presented in Chapter 20.
For a sequence A of positive real numbers, we denote by F(A) the ideal,
in the ring of all entire functions, of those entire functions that vanish at
least on A. (We exclude once and for all the null function f = 0 and the
ideal containing only the null function.) We introduce an order relation
in this system of ideals, F(A) < F(A'), meaning that for each g E F(A'),
there is an f E F(A) such that If (iy)I < Ig(iy) I for every real y. Crudely
stated, F(A) < F(A') if it is easier to construct small entire functions that
vanish on A than those that vanish on A'.
The major problem is to decide, by elementary computations on A and
A', whether F(A) < F(A'); we solve this problem here. By specialization,
then, we prove as a corollary the following result.
Theorem 22.1. There exists a function f E F(A) such that
If (iy)I < exp Trblyl if and only if

)(y)-)(x) <blog(x)+0(1), x<y,


where )(t) is the sum of the reciprocals of the elements of A that do not
exceed t.
Remark. Carlson's Theorem deals with the case b = 1 and A = {1, 2,3,...
The main innovation of our method is to give our entire functions suitable
zeros on the imaginary axis, in addition to the required real zeros.
22. On Small Entire Functions of Exponential-Type with Given Zeros 147

We proceed now to the body of the exposition. We study sequences


A = {an} of positive real numbers,
A:0<Ao<A, <...
and define
A(t) = >i an i
t
A(t) >2 1=J sd.1(s).
An <t 0

Definition. A(t) is called the characteristic logarithm of A, and A(t) is


called the counting function of A.
For simplicity, we suppose that A is an infinite sequence and that
A(t)
D(A) = lim sup < oo,
t-.oo t

since the problem is trivial if A is finite or if D(A) = oo. The function


W(z) = W(z : A) belonging to F(A) is called the Weierstrass product
(over A) and is defined by
! z).
W(z:A)=H 1-z2
\\ n
We may write

log IW(z A)l =


j log I 1-

where z = reie. For 0 # 0, 7r we have, on integrating by parts,


ezi8 l dA(t),

log IW(reie)I = r J
0
P(t,0)A(rt) I dt,

where
1-t2 cos20
P(t, 6) =
21 - 2t2 cos 20 + t4'
We define, for 0 < b < oo, the arithmetic progression Ab by
2 3
Ab b'b,b,...
1

and observe that

Ab(t) = [bt] = bt + O(1)


Ab(t) = blogt+O(1) (for t>1)
sin 7rbz
W(z:Ab)_
irbz
hwb (0) = irbl sin 61.
We write A C A' to indicate that A is a subsequence of A', and remark
that A C A' if and only if A(y) -A(x) < A'(y) - A'(x) for x < y.
148 Entire and Meromorphic Functions

Definition. A is equivalent to A', written A - A', shall mean that A'(x) -


A(x) = 0(1).
Definition. A' > A shall mean that there exists a sequence A", A" D A,
such that A" - A.
Definition. A < A' shall mean that there exists a sequence A"', A' C A',
such that A"' - A.
Although A < A' and A' > A mean two different things, the first corol-
lary of the next lemma resolves this notational difficulty.
Lemma 22.2. A > A' if and only if

(22.1) a(y) - J,(x) < A'(y) - A'(x) + D(1); 0 < x < y < oo.

Likewise, A < A' holds if and only if (22.1) is satisfied.


Corollary 22.3. A < A' if and only if A'> A.
Corollary 22.4. If A - A1i A' - Ai, and A < A', then Al < A'1.
Corollary 22.5. If Al < A2 and A2 < A3i then Al < A3.
Corollary 22.6. If A < A' and A' < A, then A - A'.
Thus, < is a well-defined partial ordering of equivalence classes under

Proof of Lemma 22.2. That A' > A and A < A' each imply (22.1) is trivial.
To show that (22.1) implies that A' > A, we define

W(x) = inf{A'(s) - A(s) : s > x}.

It follows from (22.1) that V(x) > -K for some constant K.


Now V(x) is constant except for possible jumps at the jumps of A(x).
Let xo be a point of discontinuity of W. Then,

W(xo - 0) = A'(xo - 0) - A(xo - 0)

and
p(xo + 0) < A'(xo + 0) - A(xo + 0).
We denote by i (xo) the jump of at xo. Then

(22.2) L 4 (xo) < AA'(xo) - AA(xo) < AA'(xo)

We let
A*(t) = [W(t)),
22. On Small Entire Functions of Exponential-Type with Given Zeros 149

where [a] denotes the integral part of a and


t
-D(t) = f s dcp(s),
U

and let A* (t) be the characteristic logarithm of that sequence A* whose


counting function is A*(t). The function A*(t) is constant except possibly
at the jumps of ap(t), and we have

AA* (X0) < +i (xo)


o
Using (22.2), we get

(22.3) t1A*(xo) < 1 + AA, (X0).


xo
Furthermore, xo0)*(xo) and xo0)'(xo) must be integers, so that (22.3)
implies
(22.4) AA*(xo) < AA'(xo),
and this means that A* is a subsequence of A'.
We now define
A"(x) =A(X) + A*(x),
so that A"(x) is the characteristic logarithm of some sequence A" D A. To
prove that A" - A', we must prove that 6(x) = 0(1), where
6(x) = A(x) + A* (X) - A, (X).

Now,
t
W(t) - (P(0) = fo d4i(s)
s
and

A*(t) = f o
t 1 d('F(s)]
An integration by parts shows that

A* (t) - W(t) = 0 (i),


so that it is enough to prove that 9(x) = 0(1), where
0(x) = )(x) + W(x) - )'(x) = )(x) - A'(x) + inf{A'(s) - A($) : s > x}.
But it is clear that 9(x) < 0, and (22.1) is simply another way of saying
that 9(x) > 0(1).
To prove that (22.1) implies that A < A', we put
(22.5) AY"(x) = )'(x) - A*(x).
Since by (22.4) A* is a subsequence of A', there is a subsequence A" defined
by (22.5) and A" is a subsequence of A'. Since we already have shown that
A"' A, i.e., 6(x) = 0(1), the proof is complete.
We now state the main result.
150 Entire and Meromorphic Functions

Main Theorem. Given A and A', the following three statement are equiv-
alent
(i) F(A) < F(A').
(ii) A < A'.
(iii) There exists a single pair, fo, go with fo E F(A), go E F(A'),
Ifo(iy)I < Igo(iy) I for all real y and such that the only zeros of go in the
open right half-plane belong to A.
Theorem 22.1 is a direct corollary of this result. Given A and b, choose
A' = Ab and go(z) = Since Igo(iy)I - e'"bl't, the equivalence of (ii)
and (iii) proves Theorem 22.1.
Proof of the Main Theorem. We leave the proof that (ii) implies (i) for
later. It is clear that (i) implies (iii); a suitable choice for go(z) in (iii) is
the Weierstrass product W(z : A'). We now prove that (iii) implies (ii).
We write f and g instead of fo and go. Now we choose p with 0 < p < ao
so that all the zeros, z, = r,,eie^, of f in the right half-plane (assuming for
convenience that f has no zeros on z = iy) satisfy rn > p, and write one
form of Carleman's Theorem (Chapter 12), taking y > x > p as

(22.6) E(y) - E(x) = I(y) - I(x) + J(y) - J(x) + 0(1),

where

E(R)=E(R: f)= n - Rz )Cos0,,,


R
I(R) = I(R f) = 1 1 1
) log If (it)f (-it) I dt,
:
21r ( t2 R2

J(R) = J(R : f) _ 7rR I log If(Rese) I cos a d8.


J
Now

(22.7) R2 cos0 = O(1)

since
rn
R2
cos0, < rn =
R2
1 L n- 1 n(R) = 0(1),
R-R
where n(r) counts the number of zeros of f whose modulus does not exceed
r.
Also,

(22.8) J(R) = 0(1)


22. On Small Entire Functions of Exponential-Type with Given Zeros 151

since

1
j a l og If (Reie)I cos Bd8 < log If (Reie)II dO
R 2
RJ x

5RO(R) = 0(1),
since

1 Iloglf(Re'B)II d0=rn(R,f)+m(R, 1) <2T(R,f)


27r j x

and f is of exponential-type.
From (22.7), we obtain
(22.9) E(y) - E(x) > A(y) - A(x) + 0(1),
and using (22.8) and (22.9) in (22.6) we get
(22.10) A(y) - A(x) < 1(y) - I(x) + 0(1).
Now, since If (iy)I < Ig(iy)I, we see that
(22.11) I(y:f)-I(x:f) :5 I(y:9)-I(x:g).
On the other hand, applying Carleman's Theorem now to g, whose only
zeros in the right half-plane are the A' , we see that
(22.12) I(y : g) - I(x : 9) = A'(y) - A'(x) + 0(1).
Combining (22.12) with (22.11) and (22.10), we get
A(y) - A(x) < A'(y) - A' (X) + 0(1),

and the proof is complete.


To prove now that (ii) implies (i), we suppose that A(y)-A(x) < A'(y)-
A'(x) + 0(1), and we are given g E F(A'); we must construct a function
f E F(A) with I f (iy)I 5 I9(iy)I for all y. By Lemma 23.1, we may suppose
that A(t) = Y(t) + 0(1), since A is a subsequence of a sequence A" for
which this is true, and F(A") C F(A).
By the Hadamard Factorization Theorem we may write

9(z) = 91(z)92(z),

where

9i(z)=H(1- 3 Iexp (z
),
92(x) = CzI`ea2rf 1 - Sn
x zl
exp
Can
152 Entire and Meromorphic Functions

where the (n 54 0 are the zeros of g that are not counted in A'.
Writing log 1gl(iy)I as a sum of logarithms, and that sum as a Stieltjes
integral, we get

(22.13) log Igl(iy)I = 2 J log (1 + t2 J t dA'(t).


o \ /
The next lemma provides the main tool of our construction; it will enable
us to "move the zeros" from the real axis to the imaginary axis.
Lemma 22.7. Let dL be a measure with compact support contained in
an interval f f, e-1) for some small e > 0. Then there exists a function V(t)
defined on (0, oo) such that

(22.14) flog (1 + t2 d0(t) = r log I1 - e I w(t) dA(t)


and

z 0 ys1 dL(s) I.
IVI < 2 sup ( 1

Proof. By a contour integration, it is easy to see that


roo 2
log (1+x21_ log I1-
J0
tt

By Fubini's theorem,
(22.15)
fiog(i+E) dd(r) = fiogii_i{Jz1t21 do(t) } dw
We therefore are led to define

2 f t2 d0(t)
(22.16) W(w) _
i w2 + t2 t

and (22.14) asserts (22.13) in another form. The bound on 'p follows from
integrating by parts in (22.15):

2
(22.17) 'P(w) _
J0 fJx 0 dx \x2 +w 2 )
Hence,
)
1W(W)1 1
I.
1+ J0 1 dx ( + w2 I STp I J x da(t)
22. On Small Entire Functions of Exponential-Type with Given Zeros 153
z
since-xr+w is increasing, and the lemma is proved.
We now choose

(22.18) 6(t) = 2 {A'(t) - A(t)}, d0(t) = t d6(t)

but cannot apply Lemma 22.7 to d0 since its support may not be compact.
We truncate the support by defining

6k(t)- 6(t) ift<k


{ 6(k) if t > k
Ok(t) = t d6k(t)

with the same convention for A(t) and A'(t).


We now apply Lemma 22.7 to dOk and conclude that there exist func-
tions Wk(t) such that

/ z1 r
(22.19) flog 1 1 + tJ dAk(t) log I1 - jpk(t) dt.

Now,

(22.20) Iwk(t)f < B,

where B is a constant that is independent of k, namely, from the bound on


IV(t)l in the lemma and the equivalence of A and A',

B=2supIA(t)-A'(t)1.
On putting
/ ya \ 1 y2
(22.21) Lk(y) = fiog 1 1 + Via) t dAk(t) + floe1 1 - t2 I dlbk(t),

where

(22.22) d4ik(t) = cak(t) dt,

we have
/
(22.23) Lk(y) = 2 flog f 1 + dA(t).
\\

Hence, by (22.13),

(22.24) kl oo Lk(y) = log I91(iy)I


154 Entire and Meromorphic Functions

At this point, the idea is to find an entire function F for which the
hypothetical formula

log I F(iy)l =k f log 11 - zR!I dfik(t)

holds in some appropriate sense. First, however, the limit need not ex-
ist, but a simple argument with normal families will handle this difficulty.
Also, the measures d4k(t) = cpk(t)dt are unsuitable since they need not be
positive and cannot be discrete. [It is easy to see that all the d4k(t) are
positive only in case A C A', a trivial case.] But first we show that adding
a constant to Yak, in order to make d<bk(t) positive, does not change L.
Then we show that the resulting measure may be made discrete with little
loss of precision.
Resuming the construction, we define

(22.25) PA; (t) = B + cpk(t)

and by (22.20) conclude that


(22.26) 1/ok(t) > 0 for all t.

A contour integration shows that

fioghi_hhl
(22.27) E! dt = 0,

so that
a z
(22.28) Lk(y) = f log (i + ta) tt dak(t) + flog I1 - t I

where d'Yk(t) = 1k(t) dt. Now let Wk(t) = [Wk(t)], the integral part of
WYk(t), and define

(22.29) Lk(y) = f log (1 + M!)


t2)
t
2
z
d)tk(t) + flog I1- tz I dlk(t).

Lemma 22.8. There is a constant fl, independent of k, such that for all
y>1
(22.30) f log 11 - to I d`l`k(t) < flog 11 - t I dWk(t) +,Qlog IyI
Proof. We apply the next lemma with %Pk(t) = v(t) and '1 (t) = n(t). ,0
is independent of k because dand ITkt) -'pk()i are bounded
independently of k.
22, On Small Entire Functions of Exponential-Type with Given Zeros 155

Lemma 22.9. Suppose that v(r) is a continuously differentiable function


for 0 < r < oo, that 0 < V(r) < B < oo, that n(r) is nondecreasing, and
that for some constant C

v(r) > n(r) > v(r) - C.

Then
log 11 - t2zI dn(t) < r log I1- e l dv(t) + 0(logy)
J
asy-->oc. J
Proof. For fixed r, we write L(t) = log 1 1 - 2 I and point out that L is
Lebesgue integrable on (0, oo):

L(0+) = +oo, L(r-) = L(r+) = -oo, L(oo) = 0

and that L(t) is decreasing and continuous for t E (0, r) and increasing
and continuous for t E (r, oo). We must compare Y = fo L(t)dn(t) and
Z = f °O L(t)dv(t). We will prove that Y < Z + O(log r). We assume that
v'(t) > p > 0. This involves no loss of generality since, if we replace v(t)
by v(t) + t and n(t) by n(t) + t, we change Y and Z not at all, because
f0 L(t)dt = 0. We may suppose, without loss of generality, that v(0) = 0,
since suitably redefining v on the interval [0,11 changes Z only by O (1),
which is small compared to the allowed discrepancy O(log r).
With each large r we associate the numbers r1 and r2 such that

v(rj)=n(r)=v(r2)-C.
Since v'(t) > p, we will have r - rl < r2 - r1 < v . It is easy to see that
the following inequalities hold:
If
r L(t) dn(t) < r1 L(t) dv(t),
0 J0

Jr
L(t) dn(t) < Jrz
L(t) dv(t).

It follows that Y = X + Z, where

X=- Jr,rs 2
log I1 - t2 I dv(t).

We shall prove that X < O(log r). Clearly,


rrs _
X<-! log
r,
It t r I dv(t).
156 Entire and Meromorphic Functions

Since r2 - rl < p and v'(t) < B, we have

X<-B
jr2
t-r I dt<B(rz-rl)logrz-B Jrrs log-It-rldt
log-I t
i ddd r1

so that
X< pClog(r+C)+2B.
`\
We now consider polynomials Pk defined by
tzz

log IPk(z)I = flog! 1 + I d''k(t).

Lemma 22.10. There is a constant B', independent of k, such that for


all z

(22.31) log IPk(z)I < B'Iz .

Proof. Putting z = x + iy, we see that


flog 11 tz
fiogii+..i dWk(t) < + 2 I dW(t)
But since 'It(t) = [ fo {B + cpk(s)}ds], and since Icpk(s)I < B by (22.20),
the proof is immediate on integrating by parts.
Since the family {Pk(z)} is therefore uniformly bounded in each disk
DR = {z : IzI < R}, it is consequently a normal family, and we may
extract a subsequence {Pk,(z)} that converges uniformly on compact sets
to an entire function F:

(22.32) F(z) = kJim


1-00 Pk.(z).

Because Pk(0) = 1 for each k, it follows that F(0) = 1. By Lemma 22.10,


we conclude that F is of exponential-type. Since Pk has only imaginary
zeros, so has F. Furthermore, F is an even function since each Pk is.
Let i1' = {iryn} be the zeros of F on the positive imaginary axis. Since
F is of exponential-type, 1' has finite upper density. Thus,
z
(22.33) F(z) = II (1 + z2 }
'Yn

At last we can define f (z). Let


(22.34) f (z) = fl(z)F(z)gz(z),
22. On Small Entire Functions of Exponential-Type with Given Zeros 157

where

(22.35) f1(z)=11(1- Zj exp z 1As

a consequence of the estimates (22.24) and (22.30) we see that


(22.36) log If (iy) I < log I g(iy) I + O(log IYD.

Now (22.36) is as good for our purposes as If (iy)I 5 Ig(iy)I, since we could
otherwise consider f *(z) = f (z)a{(iz)-1 sin(iz)}b for a suitable choice of a
and b.
It is not obvious, though, that f is of exponential-type, since fl and g2
need not be of exponential-type, although they are certainly of order 1.
To prove that f is of exponential-type we appeal to Lindelof's Theorem
proven in Chapter 13.
Let us denote by a and b the zeros, other than the origin, of f and g,
respectively. Then we see that

Ibnl<R
bn fA'

bn' + A(R) + (.\'(R) - A(R))


Ibnl<R
bn ¢A'

bn' + A(R) + 0(1)


lbnI<R
bn4A'

since A'(R) - A(R) = 0(1) by hypothesis. Now the zeros of f, other than
the origin, fall into three categories: (i) those bn not counted in A', (ii) the
elements of A, and (iii) the zeros of F. If we consider

S(R) = E b.-1,

the zeros of F contribute nothing to S(R) since F is even. Hence,

S(R) = bn1,
l bn I <R
bn jW

and it follows that


S(R) = O(1).
This, together with the obvious fact that the zeros of f have finite upper
density, implies (via Lindelof's Theorem) that f is of exponential-type.
The proof is complete.
23
The First-Order Theory of the
Ring of All Entire Functions

The material of this chapter is drawn from the paper [3], "First-Order
Conformal Invariants." Let E denote the ring of all entire functions as an
abstract ring. Much information about the theory of entire functions is
present in the theory of C. For example, an entire function f omits the
value 7 if there exists an entire function g such that (f - 7)g = 1.
We show that even using a restricted logic (first-order logic), a great
deal can be expressed in the theory of E. We shall show, indeed, that all
of classical function theory can be so expressed. By the ring language we
mean the first-order formal language appropriate to the structure E. This
language has basic symbols for addition and multiplication of entire func-
tions, as well as the usual logical symbols: A ("and"), V ("or), -, ("not")
and = ("implies") as well as quantifier symbols d ("for all") and 3 ("there
exists") together with variables that range over E. For convenience we also
include in the ring language a constant symbol which is a name for the
constant function i = vr---I. (Otherwise, there would be no way to distin-
guish between the two solutions of f2 + 1 = 0.) Formulas and sentences
in this language are finite combinations of these basic symbols, arranged
according to the obvious formal rules of grammar. A key restriction is that
the language is first-order, which means that we can only use quantifiers
over elements of E and not over subsets, ideals, relations, etc. Also, the
expressions in a first-order language are always finite in length. (See [39]
for a general treatment.)
The algebra language is appropriate to E as an algebra.. This is formed by
adding to the ring language a 1-place predicate Const. In E (as an algebra)
we interpret Const(f) to mean that f is a constant function. (Thus we are
23. The First-Order Theory of the Ring of All Entire Functions 159

identifying C with the field of constant functions in E and are using the
ordinary addition and multiplication of functions in E to play the role of
addition of constants and the scalar multiplication in the algebra.)
But in E it is easy to say that f is a constant (either f = 0, f = 1, or
f omits 0 and 1) so that the ring language and the algebra language are
equivalent.
In dealing with rings and algebras, the expressive power of first-order
sentences is reasonably well understood. For example, to say that a ring
is commutative is first-order (VxVy(xy = yx)) but, at least superficially,
to say that a ring is simple is not first-order, since it seems to require
quantification over subsets (there does not exist a proper two-sided ideal),
and in fact does so require. It is first-order to say that a ring has at least
two elements (3x3y(x # y)), but it is not first-order to say that a ring
is infinite, since this requires a sentence of infinite length, as one might
suspect.
Nearly all of the results in this chapter depend on the fundamental
definability results for the algebra E. We show that there are formulas in
the algebra language which define in t the set N of nonnegative integers,
the set Z of all the integers, the field of rational numbers Q, the field of real
numbers IR, the ordering relation < on 1[t, and the absolute value function
on C. We also show how to interpret in the first-order language of 6 the
quantifiers that range over countable sets and sequences of constants. (It is
striking that second-order concepts can be represented within the restricted
first-order language.)
An immediate consequence of this is that second-order number theory is
interpretable in the first-order theory of E. Other recursive undecidability
results are treated later. For example, we show that the first-order theory
of the ring of entire functions is recursively isomorphic to second-order
number theory. (This improves on a result of Robinson [33], who showed
how to interpret first-order number theory in the first-order theory of entire
functions.)
Later we extend these definability results even further. We give enough
examples to suggest that all of classical analytic function theory on C can
be interpreted in the first-order theory of E. This is quite surprising, given
the apparent limitations of the first-order algebra language. We also show
how to interpret in the first-order language of E the quantifiers that range
over countable subsets and sequences in E itself. That is, the first-order
language is already as expressive in this context as the restricted second-
order language.
The Ring Language
In this section we will begin to explore the expressive power of the first-
order theory of the ring E. To begin, note that the constant functions 0 and
1 are definable using their first-order properties in E. Also, the property
that f is a unit in E is first-order expressible: 3g(fg = 1). Thus we can
160 Entire and Meromorphic Functions

express, for any definable constant c, the condition that f omits the value
c on C by saying that f - c is a unit in £. Since 1 and i are definable in £,
so is each Gaussian rational number. This means that for each Gaussian
rational q there is a formula Fg(x) in the algebra language such that, for
any function f in £, Fq (f) holds in £ if and only if f equals the constant
function q.
In this section we present a detailed study of certain basic definability
questions for the algebra e. These matters are fundamental to the general
content of this chapter and are of interest in their own right.
We call a function f E £ a point function if it has a unique zero on C,
of multiplicity one. These will be used to represent the points of C within
C. It is easy to construct a formula P(x) in the algebra language such that
P(f) holds in £ if and only if f is a point function. That is, P(f) should
express

f is notaunit,and
dgVh[f = gh = g is a unit or h is a unit].
Using this we can, for example, construct an algebra language formula
A(x) such that A(f) holds in .6 if and only if f is a 1- i conformal mapping
on C. That is, A(f) should express the condition:
For any constant a and any point functions g and h,
if both g and h divide f - a, then g divides h.
[To say that a point function g divides f - a is equivalent to saying that
f has value a at the unique point of C where g is zero. To say that one
point function divides another is equivalent to asserting that they are zero
at the same point of C. Of course the conformal maps of C are exactly the
functions f (z) = az + b, a A 0.]
Next we discuss how to code arbitrary countable or finite sets of con-
stants in C using first-order formulas. It is very striking that we can rep-
resent second-order mathematical concepts in a first-order language.
Given f, g E C, define V(a; f, g) to mean that a is a constant, g 34 0,
and there exists zo E C such that g(zo) = 0 and f (zo) = a. This can be
represented by an algebra language formula, since V(a; f,g) is equivalent
to the existence of a point function h such that h divides g and h divides
f - a (together with the other conditions, that a is constant and g # 0).
We think of the pair (f, g) as coding the set of constants
E = {aECI V(a; g) holds in 61.
If g = 0, then this set is empty. If g 0, then it is a countable or finite
subset of C. Moreover, by taking g to have an infinite sequence of zeros
tending to oc and by letting f vary over £, then we obtain for E all possible
countable or finite subsets of C [38].
23. The First-Order Theory of the Ring of All Entire Functions 161

We may use this idea, for example, to find a formula IZ(x) in the algebra
language such that IZ(g) holds in £ if and only if the zero set of g in C is
infinite while g # 0. Namely, let IZ(g) be the formula:

g # 0 A 3f [V (0; f, 9) A da(V (a; f, 9) = V (a + 1; f, 9))]

This condition asserts the existence of a function f such that the set E
N. By the well-known interpolation theorem for entire functions this can
happen exactly when g has infinitely many zeros.
Theorem 23.1. The following sets and relations are all definable in £
by formulas in the algebra language:
The set Z of positive and negative) integers,
That set N of nonnegative integers,
The set Q of rational numbers,
The set Q(i) of Gaussian nationals,
The ordering relations < on Z, N, Q, and
The absolute value function on Q.
Proof. This is immediate from the discussion above. To get N we use the
Peano axioms: N is the smallest countable subset of C that contains 0 and
contains a+1 whenever it contains a. Precisely, the formula N(a) defining
Nis
a is constant A
bfd9[{V(0;f,9) AVQ(V(/3;f,9)) V(/3+ 1,f,9)}
V(a; f, 9)]
Once having defined N, Z and Q are trivial:

aEZ4--* aENV -aEN,


a EZAyENAy j4 QAay=p].
To define the Gaussian rationals Q(i) we note that

aEQ(i),# 3Q3ry[f EQA7EQAa= fi+yi].


To define <, on Q say, note that

a E Q,
lal=QU0<_/3A(fi=aV

Note that since N is definable in E, there is an effective procedure for


interpreting the first-order theory of (N, +, ) into the first-order theory of
162 Entire and Meromorphic Functions

E. (The operations + and are just the natural operations on £.) Actually,
this interpretation extends to the second-order theory of (N, +, ), since we
have a way to discuss countable sets of constants in the first-order language
of C. It follows that the first-order theory of £ is very undecidable in the
sense of recursive function theory. This will be discussed fully below.
We caution the reader that at this stage we are only able to define the
field of rational constants in E. It is possible to define the real field R, as
we show below, but there does not seem to be any easy way to do it.
Once we have a first-order definition of R in £, no matter what it is, we
get immediately as a bonus a first-order defuiition of < on R and of the
absolute value on C:

IaI =&3 373S[7ERA6ERAa=7+Si


A/iERAO<QA/32=72+62].

We now turn to the construction of an algebra language formula R(x)


that defines the field of real constants in E. The construction of R(x) is
complicated, but it is based on an elementary fact: A constant a is a real
number if and only if there is a Cauchy sequence S of rationals such that
every analytic function that carries S to an equivalent Cauchy sequence
of rationals also preserves a. Our problem has two parts: to prove that
this equivalence is correct and to construct an algebra language formula
that expresses the right side of it. To prove the equivalence we need the
following fact:
Lemma A. If / E R and a E C\R, then there is an entire function f
such that f (/3) = /3, f (a) 54 a, and f maps Q onto Q.
Proof. This argument is a simple modification of the proof given in [27].
Fix Q E R and a E C\R. If ,6 E Q, then f is trivial to find, so we may
assume /3 V Q also. We take p(z) = e1, so that p(z) > elzl for all z E C.
Take S = T = Q U {,0} and choose fo E EM so that

fo(3) = 2/3,I3`(fo(a) -a)I -I! 2p(IaI),

and let 6 = 2. (Here, EM is the set of all entire functions whose restriction
to R is a real, monotonically nondecreasing function.) Let us take an
enumeration of S (and T) with xl = (3 (in particular, xl 54 0), and let
6 > 0 be such that

fo(xi) + 6x1 E T and ISzI < 2p(Izl) for all z E C,


where fo is any function chosen from EM. We define
f1(z) = fo(z) + Sz, S1 = {xt} and T1 ={fl(at)}.
23. The First-Order Theory of the Ring of All Entire Functions 163

Note that Sl = {i3} and fl(p) _ /3, so that Tl = {9} also. We now
construct the sequences {fn}, {Sn}, and {T.} so that

fn(x) (2-1 + 2-")b and fn(Sn) = Tn


for all n E N, X E R. Suppose that fn, Sn, and Tn have been constructed
and choose a polynomial g with real coefficients such that
(1) g(z) = 04= z E S. (z E C)
(2) Ig(z)I 5 2 n-1P(IzI) (z E C)
(3) g'(x) > -2-n-16 (x E R).
[Any polynomial of odd degree with positive leading coefficient for which
(1) is valid will also obey (2) and (3) after it has been multiplied by a small
enough positive constant. The degree can be chosen odd by adjusting the
multiplicity of one of the zeros.]
For each M E [0, 1] we have

(fn + Mg)'(x) = fn(x) + Mg'(x) > (2-1 - 2-n-1)b (x E ]R),

so that In + Mg is strictly increasing on R.


Moreover, if we let M vary in [0, 1], then for x V S,,, (fn + Mg)(x)
varies in an interval of R that contains points of T\Tn; and for y V Tn,
(fn + Mg)-1(y) varies in an interval of JR that contains points of S\Sn.
Now for n odd, let x be the point of S\Sn with smallest index and let
M E [0,1] be such that (fn + Mg)(x) E T. We define

fn+1 = fn + Mg, Sn+1 = S. U {x}


and
Tn+1 =Tn U {fn+1(x)}.
For n even, let y be the point of T\Tn with smallest index and let
M E [0, 1] be such that { fn + Mg}-1(y) E S. We define

fn+1 = In + Mg, Tn+1 = Tn U {y}

and
Sn+1 = Sn U {fn+l(y)}.
The following properties of the constructed sequences are easily verified:
(a) Ifn(z) - fn-1(z)I <- 2-np(IzI) (n E N,z E C)
00 00
(b) U Sn = S, U Tn = T, and
n=1 n=1

fm (Sn) = Tn, (m, n E N, m > n).

From (a) it follows that fn converges pointwise to a function f for which

1 f (z) - fo(z)I <- P(Iz) (z E C).


164 Entire and Merornorphic Functions

Now p(IzI) is a function of z that is bounded on compact subsets of C, so


the convergence of {f, (z)} is uniform on such sets. From this we conclude
that f is an entire function.
For each n E N we have fn(x) > 16, (x c ]l2), so the same is true for f.
Hence, f is strictly increasing on R. From (b), it follows that f (Sn) = Tn
for each n, and so f(S) = T. Moreover, we have insured that fn (p) = /3
for all n E N and therefore f (/3) = /3. This implies f (Q) = Q also.
Finally we have

I(f(a) - a)I I(fo(a) - a)I - I(fo(a) - f(a))I


> 2p(I al) - p(IaI) > 0,
which means f (a) 0 a. This completes the proof.
Now it is easy, using Lemma A, to show that our characterization of
the real numbers is correct. In one direction, the equivalence is trivial: If
a E P., we need only take S to be a Cauchy sequence of rationale that
converges to a. For the converse direction, suppose that a E C\R and S is
any Cauchy sequence from Q. Let # be the limit of S in R Use Lemma A
to get an entire function f such that f (/j) = ,Q, f (a) # a, and f (Q) C Q.
Then f maps S to a Cauchy sequence of rationals that is equivalent to S
(since both converge to Q) yet f does not preserve a. That is, a E C\R
implies that the right side of our equivalence is false. This completes the
proof that our characterization of R is correct. Now we must express it
formally within the algebra language.
Theorem 23.2. There are formulas R(x), L(x, y), and M(x, y) in the
algebra language such that for any a, ,0 E E:
(i) a E HIS R(a) holds in e;
(ii) For a, Q E P, a < fl = L(a, /3) holds in E;
(iii) For a, /3 E C, j al = /3-#=:>=:a, 0) holds in E.
Proof. We begin by building some machinery for discussing sequences of
constants within the first-order language of E. (Earlier we did the same
for countable sets of constants.) This is done using a triple of functions
(f, g, h); g has infinitely many zeros on C, and on that zero set h takes on
the values 0, 1, 2,... . In effect, h lists the zeros of g. Then the sequence
coded by (f, g, h) is (a,,), where a,, is the value f (z,,) at the zero of g
where h take the value n. First let Basis (g, h) be a formula in the algebra
language which expresses the fact that h "lists" the zeros of g in the manner
discussed above:

Basis(g, h)t=da[V(a; h, g)t=a E NJ


A daVpVq[{P(p) A P(q) A p divides q
A p divides h - a A q divides g A q divides h - a}
= pdivides q].
23. The First-Order Theory of the Ring of All Entire Functions 165

Now we construct an algebra language formula Seq(a, n; f, g, h) which


expresses that a is the nth term of the sequence of constants coded by the
triple (f, g, h):

Seq(a,n; E NABasis(f,g,h)
A a E C A 3p[P(p) A p divides g A p divides h - n
A p divides f - a].

Using the defining formulas described in Theorem 24.1, we now can


say, using a formula in the algebra language, when the sequence coded by
(f,g, h) is a Cauchy sequence of rational numbers and when this sequence
of rationals converges to 0. (Note that since we have the absolute value
function only on Q at this point, there is no hope of discussing converging
or Cauchy sequences outside Q. This is precisely our difficulty in this entire
discussion.) For the first of these:

Cauchy Rat Seq(f, g, h)4-- Basis(g, h)


A VaVn(Seq(a, n, f, g, h) a E Q)
AV6 EQ+3mENdi,j ENVa,Q EQ[{m <i
A m < j A Seq(a, i, f, g, h) A Seq(,Q, j, f, g, h) }
* Ia-pI<_6].
Here we simply have written down in formal terms the usual version of the
concept to be defined. (Q+ is the set {q E Q 10 < q}, so it is defined by
an algebra language formula.)
For rational sequences converging to 0, we have an entirely similar for-
mula:

Zero Rat Seq (f, g, h)t Basis(g, h)


A dadn(Seq(a, n, f, g, h)a E Q)
Ad6 E Q+ 3m E NVi E lWa E Q[{m <_ i A Seq(a,i, f,gh)}
= Jal < 6].
Consider a pair (g, h) for which Basis (g, h) holds and suppose that fl,
f2 are functions so that (fl, g, h) and (f2i g, h) code Cauchy sequences of
rational numbers, say (an) and (,3n), respectively. Then it is clear that
(h - f2, g, h) codes the sequence (an - Qn). [Here it is essential that the
same (g, h) be used.] Therefore, (an) and (On) are equivalent if Zero Rat
Seq(fl - f2i g, h) holds.
Next we face the second major difficulty in formalizing our charactriza-
tion of R in the algebra language: We have no direct means of expressing
that "the value of f at a equals fl," where f E E and a, p E C. We
approach this indirectly by introducing, as a parameter, a 1 - 1 conformal
166 Entire and Meromorphic Functions

mapping a on C. Of course a is an element of E and, as noted above, the


property of being a 1 - 1 conformal mapping is expressible by an algebra
language formula A(u). Now we define a formula in the algebra language,
which we will abbreviate by writing f (a) = /3, by
Cr

f (a) = /3=a is a 1 - 1 conformal map


or

Aa E range(a) A f (a-1(a)) = /3
e=A(a) A a - a is not a unit
Aa, /3 E C A a- a divides f -/3.
This formula, which will be quite important in later sections as well, here
enables us to express the condition that the composition f o a-1 carries
one sequence coded by (f1, g, h) to a second sequence coded by (f2, g, h).
The formula that expresses this is the following:

'dra E NVa, /3 E C[Seq(a, n, fl, g, h) A Seq(/3, n, f2, g, h) f (a)


o 91.

Note that for a fixed 1-1 conformal mapping a, as f ranges over E the
composition mapping f o a-1 ranges over E.
We now are ready to give the formula in the algebra language which
defines R in E. (It is convenient to define C\R instead.) We see that
a E C\R,# a E C and there is a 1- 1 conformal mapping a on C, and
for any Cauchy sequence (an) of rationale that is f E E and a Cauchy
Sequence (/3n) of rationals with the properties
(1) f (an) = /3n for all n E N);
(ii) (an) and (/3n) are equivalent Cauchy sequences of rationals;
(iii) f (a) = a is false.
a
It remains only to show that this equivalence is correct. Earlier we
proved the equivalence a E C\R=a E C, and for any Cauchy sequence
(an) of rationals there is an entire function g and a Cauchy sequence (/3n)
of rationals with the properties
(i') g(an) = Qn for all n;
(ii') (an) and (/3n) are equivalent Cauchy sequences of rationals;
(iii') g(a) 0 a.
Clearly, if a exists for a as above, and if (an), (/3n) and f satisfy (i), (ii),
and (iii), then we need only set g equal to f o a-1. Conversely, the range
of every 1 - 1 conformal mapping or on C includes a and Q. Given such
a a and g satisfying (i'), (ii'), and (iii'), just take f to be the composition
g0a.
This finally completes the proof that R is definable in E. As was dis-
cussed earlier, from this we get formulas defining < on R and the absolute
value on C. Thus the proof of Theorem 23.2 is complete.
Theorem 23.2 is fundamental to nearly all of our other results.
23. The First-Order Theory of the Ring of All Entire Functions 167

More on the Algebra Language


In this section we present a variety of results which, when taken together,
show that the expressive power of the algebra language is strong enough
to include all of the classical mathematical theory of entire functions. It
is quite striking that this should be possible, since the first-order algebra
language seems to be so limited. These results are not used in any other
part of this chapter, and we have not made a great effort to be complete
nor to give more than sketchy proofs. Our goal is to give examples that
show what is possible.
We discussed how to deal with sequences of constants within the first-
order theory of E. Now we will improve this to code sequences of functions.
The idea is to fix a point zo in C and a sequence {zn } in C that converges to
zo. By the uniqueness of analytic continuation, a function f E E is uniquely
determined by the sequence of constants if (zn)}. Hence, a sequence if. I
of functions can be coded by an infinite matrix of constants If. (z.) m, n E
N}. Now this is not quite enough, since we cannot refer to the points of
C in a direct way. We overcame this earlier by introducing a parameter
a that is a 1 - 1 conformal mapping on C. We replace the points zn by
the constants an = a(zn) and refer to the values Ym,n = fn,(zn) by using
the equivalent, definable relationship fm(an) =#,",n. Finally, we code the
sequence (an) and the matrix (f3,,,n) into a single matrix that has (a - n)
as the top row.
To code an infinite matrix of constants we proceed as earlier, but use a
basis triple (g, hl, h2) instead of a pair (g, h). Here we require that g(# 0)
have an infinite zero set in C, as before, and that (hl, h2) map this zero set
bijectively onto N x N. It is routine to construct a formula M Basis(x, y, z)
in the algebra language such that M Basis(g, hl, h2) holds in E if and only
if this basis condition is satisfied.
When M Basis(g, hl, h2) is true, any function f in E determines a matrix
of constants (a,nn) by taking a,nn to be the value of f at the unique zero
z,,,n of g for which hl(z,,,n) = m and h2(z,,,n) = n. Earlier we expressed
this relation using point functions. Also, by the interpolation theorem for
entire functions, every matrix (a,,,n) is coded in this way by some f, no
matter which basis triple (g, h1, h2) is used.
The matrices (a nn) that arise in coding sequences of functions are not
arbitrary, of course. First, the sequence (a,nn) from the top row must be a
Cauchy sequence in C. Finally, for each m > 0 there must exist a function
f,n in E such that

fm(aon) = amn for all n E N.

Evidently, f,,, is uniquely determined by this condition. It is the nth


function in the sequence coded by (a,nn) and a. As was discussed pre-
viously, we take the matrix (amn) to be coded by some (f, g, h1, h2), where
M Basis(g, hl, h2), holds.
168 Entire and Meroniorphic Functions

We can construct an algebra formula Code(f, g, hi, h2, v) that holds in


£ if and only if a is a 1 - 1 conformal mapping on G and M Basis(g, h1, h2)
holds and the matrix (amn) coded by f using the basis (g, h,, h2) sat-
isfies the conditions above. [That is, the sequence aon and its limit he
in the range of a and, for each m > 0, there is an fn in £ so that
f,n(ao,n) = amn for all n E N.] Then we formulate an algebra formula
F Seq(k, m, f, g, h1 i h2, a) that holds in ,6 if and only if Code (f, g, h,, h2, v)
holds, m E N, and k is the (unique) function that satisfies k(aon) = amn
for all n E N. That is, F Seq(k, m, f, g, hl, h2, a) holds if and only if k is
the mth function in the sequence coded by (f, g, h1, h2, o).
Using the absolute value on complex constants and the "evaluation" of
functions f via a (that is, in the form f (a) _ 6), we now may obtain algebra
formulas that express various types of convergence of the coded sequence of
functions. This can be done for pointwise convergence, uniform convergence
on C, or even uniform convergence on compact subsets of C. For this last
type of convergence it is not necessary to quantify over arbitrary compact
sets, but rather to use a particular exhaustion of C by compact sets. For
example, suppose Code (f, g, h1i h2, v) holds in £ and we wish to discuss
convergence of the coded sequence of functions (f,,,). Consider the sets
G' C range(a) defined by

Gn={aECIlaI<n}.
These sets are definable using an algebra formula from the parameters n
and a. Also, (f,n) converges uniformly on compact subsets of C if and only
if (f,,, o or-') converges uniformly on each of the compact sets G' . This
can be expressed by an algebra formula using F Seq at the end to replace
mention of (fn).
This method of representing sequences of functions within t enables us
to define many specific sequences-for example, the sequence of powers
(f'n) of a particular function. FYom this we can find an algebra formula
which expresses that g is equal to a polynomial in f.
Next we discuss a method for interpreting in £ the lattice of open subsets
of C. (This procedure is also used later, where it is discussed in greater
detail.) This is done by associating each open set O with the set R(O) of
all pairs (q, r), where q is a Gaussian rational, r is a rational > 0, and the
disc {a E C I ja - qj < r} is contained in O. Since 0 is the union of this
family of discs, we see that 0 54 £) implies R(O) # 1Z(D) for open sets 0,
D. The set R(O) can be coded by a quadruple (f1, f2, g, h) by first writing
R(O) as a sequence (qn, rn) for n E N and then taking (qn, rn) to be the
value of (f, (z), f2(z)) at the unique z E C for which g(z) = 0 and h(z) = n.
Here, (g, h) satisfies the Basis formula described above.
We first obtain an algebra formula 0 Basis(fi, f2, g, h) that is true in £
if and only if there is an open set 0 such that R(O) = {an, /3n) I n E N},
23. The First-Order Theory of the Ring of All Entire Functions 169

where (an, Jan) is the sequence of pairs coded by (fl, f2) using the basis pair
(g, h). This must express that every a,, is in Q(i), every fl, is in Q+; also,
it must express that if q E Q(i), r E Q+ and if the disc {ry E C I I'y - qi < r}
is contained in the union of the discs {ry E C I I' - anI < 8,, 1, then (q, r)
is in the set {(an, Jan) n E N}. Of course, 0 is just the union of the
discs {-y E C I Iry - anI < On } for n E N. Thus we can obtain an algebra
formula Open(a, f1, f2, q, h) which expresses that 0 Basis(f 1, f2, q, h) holds
and that a is in the open set 0 for which R.(0) = {(an, fn) I n E N}.
Evidently we also can represent sequences of open sets by using matrix
pairs (a,nn, Ann) of constants, each row of which codes an open subset of
C. This gives us another way of referring to an exhaustion (Gn) of C by
compact sets, referring instead to the sequence CGn) of open sets. Also,
we can develop a way of representing all analytic subsets of C (including
all Borel sets) using the Souslin operation applied to sequences of sets. In
addition, we can construct an algebra formula that expresses the value of
Lebesgue measure applied to these analytic sets using approximation by
open and closed sets. These matters are more complicated, and we omit
the details.
Finally, we discuss how to express integration of functions and the
"Nevanlinna characteristic" applied to entire functions using algebra formu-
las. First we consider integration. Fix f E e and let o be a 1-1 conformal
mapping on C We then can express by using algebra formulas how to
integrate f o or-' along a circular curve -y = {z I Iz - al = r} contained in
the range of a. That is, we have an algebra formula Int(f, Q, a, r, J3) which
holds in E if and only if or, a, r are as above and fr f dz = P. This can
be done by considering a sequence of successively finer subdivisions of ry
and by evaluation of f o a-1 on points of ry using our expressions f (17) = 6
as above. We then get upper and lower sums and evaluate the integral by
taking limits. (Lebesgue integration can be handled also, but it is more
complicated. In any case, our functions are highly continuous.) By a simi-
lar procedure we can also treat integration over the inside of the curve 'y,
with respect to area measure.
Now we consider the "Nevanlinna characteristic" on entire functions.
For an entire function f this is defined by
2,K
T(r,f) = r log, If(re'BIdO.
J0

We also consider the order a of f defined by


log T(r, f )
a = lim sup
r-00 log r

Note that when we apply the above approach to integration, the I - 1


conformal mapping o that appears as a parameter must be affine. Also,
170 Entire and Meromorphic Functions

the order of f is equal to the order of f o a-1, no matter which affine or


we choose. This shows that there is an algebra formula Ord(f, a) which is
true in the algebra of entire functions if and only if the order of f equals
a.
Obviously there is no completely natural point at which to stop this
general discussion of how to express the mathematics of entire functions in
the first-order theory of E. As far as we can see, essentially all of the classi-
cal theory of entire functions, including topological and measure-theoretic
aspects, can be so represented.
Derivatives and Definable Constants
We will call a constant a E C definable if there is a formula D(x) in the
algebra language such that, for every function f E E,

D(f) holds in E4=#,f equals the constant a.

Evidently, i is a definable constant, since we have included a name for it


in the algebra language. Also, it is clear that a is definable if and only if
both the real part of a and the imaginary part of a are definable.
One way to obtain a large number of definable constants is via infinite
series and the device for coding sequences of constants that was discussed
earlier. Recall that such a sequence is coded by a triple of functions (f, g, h):
a is the nth term of the sequence if f (z) = a, where z is the (unique) zero
of g for which h(z) = n. It is now a routine matter, given the defin-
ability results obtained earlier, to construct an algebra language formula
Series(x, y, z, w) such that, for any a, f, g, h E 6,

Series(a, f, g, h) holds in E if and only if (f, g, h) codes


a sequence (an) and E an converges to a.

To make this clearer, we take the first step toward the formula:

Series(a, f, g, h)= Basis(g, h) A a E C


A (3k)[Vn E NV/3, -y, 6 E C(Seq(/3, n, k, g, h)
ASeq(ry,n+1, f,g,h) =Seq(/3+ry,n+1,k,g,h))
A d/3 E C(Seq (Q, 0, f, g, h) = Seq(/3, 0, k, g, h) )
AYbER+3nEN`dmENd/3EC(n <mASeq(l,m,k,g,h)
= I/j - al 6)l.

The middle two clauses of this formula assert that the sequence (/3n) coded
by the triple (k, g, h) is the sequence of partial sums of the sequence (an)
coded by (f, g, h). The third clause asserts that (fan) converges to a.
Now we will show, as an example, that e is a definable constant. This fol-
lows because we can say in a first-order way that (f, g, h) codes a sequence
23. The First-Order Theory of the Ring of All Entire Functions 171

(an) which satisfies the recurrence relation ao = 1, an+1 = an/(n + 1),


forcing an = 1/n! and so e = an. Thus the defining formula D(x) for e
is obtained from:

D(a) 4= f 3g3h[Series(a, f, g, h) A Seq(1, 0, f, g, h)


A bn E N V,3 E gSeq(,0, n, f, g, h)) Seq(/3/n + 1., n + 1, f, g, h)].

[Strictly speaking, we cannot use division, but this is easily eliminated from
D(x).]
Clearly this same device can be used to obtain most of the familiar
transcendental real numbers, as well as many others, for example, Liouville
numbers such as F, 10'x'. All that is required is that the number be the
limit of a series whose terms are generated using some recurrence formula.
Theorem 23.3. Let F be the set of all definable constants from C. Then
F is a countable algebraically closed field that contains e and ir.
Proof. It is clear that F is a subfield of C. For example, if D(x) defines
a = 0, then the formula

3y(D(y) A xy = 1)

defines 1/a. Also, F is countable, since there are only countably many
formulas in the algebra language. It is easy to show that F is algebraically
closed. For example, suppose aj is definable by D3 (x) for j = 0, 1, 2, 3.
We will show how to define a root of the polynomial p(z) = aoz3 + a1z2 +
a2z + a3. Consider the linear ordering on C defined by taking

a < Qt=*X(a) <_ R(f3) or


(R(a) = R(/3) and 3`(a) < 3`(Q)).

This ordering is definable by an algebra formula B(x, y) in the sense that


for any a, 0 E C
B(a, /3) holds in Era < /3.
Using this we can define the "smallest" root of p(z) by a formula D(x).
Namely, D(a) is equivalent to

3ao3a13a23a3[Do(ao) A Dj(aj) A D2(a2)


AD3(a3)AaECAaoa3+a1a2+a2a+a3=0
AV/3(/3 E RAa0Q3+a112+a2f3+a3) = 0 B(a,Q)].

Other interesting definability results for the constants come from an


indirect treatment of derivatives, which we now discuss. Of course, there
is no hope of obtaining a first-order definition of the relation between a
172 Entire and Meromorphic Functions

function and its derivative, since this relation is not conformally invariant.
Our indirect approach comes via the use of a parameter a, which is a
1 - 1 conformal mapping, as was used earlier to obtain the definition of
R. Given such a a and given f, g E E, we will show that the relation
(goo-1) = (f oa'1)' is a first-order property of the triple (f, g, a). Namely,
this relation is equivalent to the condition:
Va E C 3h3k3/3 E C 3y E Cq [if a - a is a nonunit, then
g=/3+(a-a)h and f =y+f3(a-a)+(a-a)2k].
Proof. Let a = a(z) for some z E C and suppose the equations
g = +(a-rr)h and f = y+/3(a-a)+(a-a)2k holdover C. Substitute
01-1 in the second equation and get
foa-1(w) = y + /3(w - a) + (w - a)2k(a 1(w)) for all to E C.
Differentiating with respect to w and setting w = a yield (f oa-1)'(a) = /3.
Since (g o or-') (a) = J3, we have the desired equation.
(==>) Suppose that (goo r-') = (foa-1)' and consider a E C. Expanding
g o a-1 and f o a-1 about the point a yields
(g o or-')(w) = /3 + (w - a)h(w),
(f o a-1)(w) =,y +/3(w - a) + (w - a)2k(w).
Substituting to = a(z) yields the equations needed.
The exponential function is uniquely determined on any neighborhood
of 0 by its functional equation f = f and f (0) = 1. This leads to a certain
definability result for the exponential function:
Theorem 23.4. There is a formula E(x, y) in the algebra language such
that for any a, /3 E E
E(a, f3) holds in E if and only if a, /3 E C and e° = /3.
Proof. Given a, /3 E C, consider the following condition:
there exists a 1 - 1 conformal mapping a and a
function f, both in E, such that
(i) f(0) =1,
°
(n) (foa-' )' _ (foa-'),
() f(a) =°
From our previous discussion it is clear that this can be expressed by an
algebra language formula. Parts (i) and (ii) imply that f o a-1 is equal to
the exponential function, and part (iii) therefore says that e" Thus,
E(a, /3) implies e" = /3.
Conversely, given that e" = /3, to satisfy E(a, /3) we need only set or
equal to any 1 - 1 conformal mapping on C and then set f (z) = e°(z) for
z E C.
23. The First-Order Theory of the Ring of All Entire Functions 173

Corollary 23.5. Let Fo be the field of definable constants. Then Fc is


closed under the exponential function.
Note that Corollary 23.5 gives an alternate proof that F0 contains e.
Also, it yields that F0 has infinite transcendence degree. Indeed, by the
Hernlite-Lindemann Theorem, if al, ... , an are algebraic numbers (hence
in FO) that are linearly independent over Q, then en',...,ea^ are alge-
braically independent elements of F0. (See [21].)
We close this section by noting that the other elementary functions,
such as sin(z) and cos(z), can be treated in a way that is similar to our
discussion of el.
Recursive Undecidahility
Earlier there was presented a way of defining N in F and a method for
coding countable sets of constants. This yields an effective interpretation
of second-order number theory in the first-order theory of F.
This interpretation is an example of a 1 - 1 reduction of one "problem"
or set to another. If 61, 62 are sets of sentences in formal languages L1,
L2, respectively, we say 61 is 1-1 reducible to 62, and write 61 <1 62, if
there is an effectively computable 1-1 function W (i.e., a recursive function)
such that for any sentence S of L1,
S E 614=(S)E62.
We say that C51 and 62 are recursively isomorphic if there is an effectively
computable function cp that maps the sentences of L1 bijectively onto the
sentences of L2 and which satisfies '(61) = 652. Evidently this means
that the problems of deciding membership in 61 and in 62 are effectively
equivalent in a strong way. It is a well-known fact [34] due to MybM that
if 61 _<1 62 and 652 <1 61 , then 61, 62 are recursively isomorphic. (The
converse is obvious.)
As an example of what type of undecidability theorem can be proved
using the results above, we consider the ring of entire functions F. Robinson
(33] showed that the first-order theory of this ring is undecidable by showing
that first-order number theory can be interpreted in it. The following result
is a substantial improvement of this.
Theorem 23.6. The first-order theory of the ring of entire functions is
recursively isomorphic to second-order number theory.
Proof. Let 61 denote the first-order theory of 6 and let 62 denote the
second-order theory of (N, +, ) We will show 61 <1 62 and 62 <1 E51 -

Let 6, denote the first-order theory of the algebra of entire functions F.


Earlier it was shown that the field of constants is definable in F, which
yields a direct interpretation of 6; into 61. In particular, 6 <j 61. As
was discussed earlier in this section, we also have a direct interpretation
of second-order number theory in 6. This implies 62 <1 61, so that
62 <1 61 has been proved.
174 Entire and Meromorphic Functions

To show 61 <1 62i we sketch how to give an interpretation of the first-


order theory of E in second-order number theory. Each entire function f
has a power series representation centered at 0,

f (z) = E anz",

with an infinite radius of convergence. Each coefficient a" = x" + y"i


can be identified with two sequences of integers, say {rn,j I j E N} and
{8",j I j E N}, where r",o is the integer part of x" and r, ,j is the jth decimal
digit of xn, and similarly for sn,j and y". Therefore, f can ultimately be
identified with an infinite matrix of integers M = (mij I i, j E N) obtained
by setting mi,2j = r j and rn1,2j+1 = sfj for all t, j E N. Finally, M can be
converted to a set of integers M# _ {2i 3j 5"'' 1 i, j E N}. Evidently we
can recover the function f from the set M#. To show that this provides
the desired interpretation, one should prove that the collection of all sets of
the form M# is definable in the second-order theory of (N, +, ) and prove
the same for the relations on these sets which correspond to addition and
multiplication of entire functions. The details are tedious and routine, and
we choose to omit them.
Corollary 23.7. Second-order number theory is 1 - 1 reducible to To.
24
Identities of Exponential Functions

In this chapter, which is based on [13], we take up some questions prompted


by mathematical logic, notably Tarski's "High School Algebra Problem."
We study identities between certain functions of many variables that are
constructed by using the elementary functions of addition z + y, multi-
plication x . y, and one-place exponentiation ex, starting out with all the
complex constants and the independent variables z1, ... , z,,. We show that
every true identity in this class follows from the natural set of 11 axioms
of High School Algebra. The major tool in our proofs is the Nevanlinna
theory of entire functions of n complex variables, of which we give a brief
sketch. It is entirely parallel to the one-variable theory presented in detail
earlier in this book. The timid reader can take n = 1, at least for a first
reading.
Tarski's conjecture for a more extended class of terms than those we
consider here has been shown to be false by Wilkie (see [50]). The largest
class that we are aware of for which Tarski's axioms have been shown to
be complete was studied in [11].
We briefly recapitulate the basic Nevanlinna theory.
Consider first the case of one variable, n = 1. For meromorphic functions
f of one variable defined on the complex plane C, the characteristic function
is defined for 0<r< cc by

T(r,f) = rn(r,f) + N(r, f):


this is a sum of two terms: the proximity function
n
m(r,f) log+If(re'B)IdO,
27r
176 Entire and Meromorphic Functions

which measures how close, on the average, f is to oo, and the average
counting function
N(r,f)= I-
f rn( f) dt,
U

where n(r) is the number of poles of f in the disc IzI < r. Here, r ranges
over the interval 0 < r < oc, and appropriate modifications must be made
in the definitions of N(r, f) if f (0) = 0 or if f (0) = oc. The function
log+(t) is defined by setting log+(t) = log(t) for t > 1 and log+(t) = 0 for
0 < t < 1. The growth of the characteristic T(r, f) as r --, oo gives a very
useful measure of the growth of f. The basic properties that we shall use
are listed below.
(C2.0) T(r, f) is a nondecreasing function of r and a convex function of
log r.
(C2.1) T(r, f + g) < T(r, f) + T(r, g) + 0(1).
(C2.2) T(r, fg) <T(r, f) +T(r,g).
(C2.3) T(r,1/(f - a)) = T(r, f) + 0(1) for any complex constant a.
(C2.4) T (r, f 1g) 5 T (r, f) + T(r, g) + 0(1).
(C2.5) T(r, e9)/T(r,g) -i oo as r - oo if T(r, g) is unbounded.
The other basic fact we need about the characteristic T(r, f) is the
Lemma of the Logarithmic Derivative (LLD):
(C2.6)
m(f, f'/f) 0 (log(T(r, f))+logr),
except possibly for r lying in a set E of finite length.
When it comes to several variables, the theory is substantially the same
and the basic properties we need are still expressed in the same form (C2.0)-
(C2.6). (See [42] for the details of proofs.) Alternatively, one could use
a characteristic based on the exhaustion of C" by balls, rather than by
polydisks DPti . (See [48] and [101.) For a meromorphic function f (z1 i ... , z,.,)
defined on C", we define
fn W
log+ I f
(re`0',
... re`s,. )jdqi ... din ,

where 0 < r < oo. Also, N(r, f) is defined much as before, as an averaged
counting function of the poles of f. Note that if f is a holomorphic function
on C", then T(r, f)=m(r,f).
(In [42], a characteristic T(T, f) is developed for a vector variable f =
(r1,. .. , rn), but we use only the diagonal case r1 = . = rn = r.)
The basic properties above, including LLD, are shown to hold in [42] or
follow exactly as in the case of one variable (e.g., (C2.5)). One thing which
needs explanation is the derivative f' that occurs in the LLD. When n > 2,
we shall take f to stand for the Euler operator
of of
f'=Df=zlaz1 +.+znaxn.
24. Identities of Exponential Functions 177

This has the useful property that D f = 0 if and only if f is identically


constant. (This is because f may be expanded as a nicely converging sum
of homogeneous polynomials and because DP = mP for any homogeneous
polynomial of degree m.)
Our basic tool is a lemma proved in one dimension by Hiromi and
Ozawa-see Chapter 17 of this book. The proof in n dimensions is similar
to the proof in 1 dimension and depends only on the properties (C2.0)-
(C2.6) we have listed of the characteristic function T(r, f ). (Carlos Beren-
stein has recently pointed out that the proof in [13] is incomplete. The
authors of [13] are preparing a complete proof. More varied Wronskians
are needed to establish linear dependence in the N-dimensional version of
the Hiromi-Ozawa lemmal. See [4] for a correct version of the proof for the
ball-characteristic.)
Lemma 24.1. (Hiromi-Ozawa). Let ao(z),... , a,(z) be meromorphic
functions and let gl (z),... , gn(z) be holomorphic functions defined on the
domain CN. Suppose that these functions satisfy

(a) T (r, aj) = o (m(r,e1))


s-1
for each j = 0,1, ... , n; and

(b) T(r, ell) 0 O(log r)

for at least one i = 1, 2, ... , n. Undf. these hypotheses, if the identity

1: aj(z)e9i(=) = ao(z)

j=1

holds for z E CN, then there exists constants c1, ... , cn (not all 0) so that
n
Cj . aj (z)e91(Z) = 0
jj=11

for all z E CN.


Our use of the Nevanlinna theory tools previously discussed comes en-
tirely through this Hiromi-Ozawa Lemma. Indeed, we use it only in cases
where 91,.. . , g. are holomorphic functions, so that T (r, e91) = m(r, e9')
(and the prohibition that the function not take the value 0 at the origin
is satisfied), and in cases where ao, a1,.. . , an are slowly growing func-
tions. Here we consider expressions that are built up from variables and
complex constants using addition, multiplication, and the 1-variable expo-
nential function es (where e is the usual base of the natural logarithm).
178 Entire and Meromorphic Functions

We prove a version of Tarski's High School Algebra Conjecture for these


expressions. (This result was proved independently by van den Dries [47]
and, for terms containing just one variable, by Wilkie [49]. Their methods
are quite different from ours.) We also settle positively a conjecture, due
to Schanuel, which asserts that if f is a function on C' which is defined
by an expression of this type, and if f is nowhere equal to 0, then f = e4
for a function g on C' which is also defined by an expression of the kind
considered here.
Definition 24.2. E is the smallest class of terms which contains the
variables x1, x2, ... and a constant for each complex number, and which
contains the terms s + t, s t and exp(t) for each s, t E F_
Here we interpret exp(t) to stand for et. We note that if t E E and the
variables of t are among x1,.. . , x,,, then t defines a holomorphic function
on all of C". Ifs E E also has its variables among x1i ... , xn, we write t - s
to mean that t and s define the same function on C". (Various equivalent
formulations of this definition are possible in special cases because of the
uniqueness of holomorphic functions. For example, if t and s contain only
real constants, we may be interested only in the functions they define on
R". But t - s will hold as long as t and s define the same function on IR",
or even on S" where S C C is any set with a limit point in C.)
One has the additional useful fact that a holomorphic function f on
C" has the small characteristic T(r, f) = O(log(r)) if and only if f is a
polynomial. Hence, if f is a polynomial and g is any nonconstant holo-
morphic function on C", then T(r, f) = o(T(r,eg)) (which is necessary as
part of the hypotheses of the Hiromi-Ozawa Lemma as we apply it). See
[17, Proposition 4.4ff]. For holomorphic functions this can be proved by
estimating the Poisson integral for log if I to show that f is of polynomial
growth as a function of x3 (when xi, i 54 j, are held fixed), for each j = 1,
2, ... , n. By the Liouville Theorem in one variable, then, f is a polynomial
separately in each x3. That f is globally a polynomial now follows from
[30]. (There must be many other proofs of our assertion in the literature.)
Theorem 24.3. (Tarski's Conjecture for E). If t, s are any two terms
in E and t =_ s, then the identity t = s is probable from the axioms

x + (y + z) = (x + y) + z, x(yz) = (xy)z,

x+y=y+x, xy=yx,
x+0=x,
x(y+z)=xy+xz,
exp(x + y) = exp(x) - exp(y),
together with all axioms giving the facts of addition, multiplication, and
exponentiation for constants from C.
24. Identities of Exponential Functions 179

Proof. Because we have included here a constant for -1, the operation
of subtraction is available and we need only consider the case where s is
0. That is, if t E E, then we must show that t = 0 is formally derivable
whenever t - 0.
Moreover, it is easy to show that for any term t E E there are terms
s1, ... , sk E E and polynomials Pi,. . , pk in n variables, with coefficients
in C (also realized as terms in E) so that the identity

t = p1 exp(si) + ... + Pk - exp(ak)

is provable from the permitted axioms. We will prove the theorem by induc-
tion on the total cumber of symbols in the sequence s1i ... , Sk, showing th
Pl,... pk are polynomials, sl,... , sk E E and pi exp(s1)+ . +pk exp(sk) -
0, then p, exp(sl) + +pk exp(ak) = 0 is formally derivable. (Note that
we allow sj to be 0.)
First suppose that k = 1: If p1 exp(si) - 0, then p1 - 0. It is well
known that p1 = 0 is provable from the admitted axioms, since p1 is a
polynomial. Hence, p1 exp(s1) = 0 also is provable.
From now on assume k > 1. Assume p1, ... , pk are polynomials,
81, ... , sk E E and pl exp(s 1) + +pk exp(sk) - 0. For 1 < j < k, let arj
be the function on C' defined by exp(s,). (Choose n so that all variables in
each pi and sj are included among x1,... , x,,.) Note that we may assume
each w1 is nowhere equal to 0 on C.
After dividing by wk we have

Pl(lr1/lrk) +"' +pk-1(lrk-1/7rk) - -Pk


Suppose first that we can apply the Hiromi-Ozawa Lemma. In this setting,
this means T(r, 7r;/lrk) 54 O(log(r)) for each 1 < i < k -1. If so, then there
exist constants c1,. .. , ck_1 (not all 0) so that

c1P1(lr1/xk) + ... + ck-lPk-1(rk-1/irk) = 0,

which gives us an identity with k-1 exponentials after multiplying through


by Irk. By the induction hypothesis, the formal identity

C1P1 exp(sl) + ... + ck_1Pk-1 exp(sk-1) = 0

is derivable in the allowed system. Now we can use this identity to solve
for one of the expressions p. exp(s,,) (1 < j < k - 1) and eliminate it
from the original expression P1 exp(s1) + + pk exp(sk). The resulting
identity (setting this expression = 0) has at most k - 1 exponentials, so it
is derivable. From this one deduces the desired identity

Pl exp(sl) + ... + Pk exp(sk) = 0.


180 Entire and Meromorphic Functions

On the other hand, it may happen that for some i(1 < i < k - 1),
T(r,1ri/lrk) = O(log r). Since it , irk are nowhere 0, it follows that iri - clrk
for some constant c. [By (C2.3) the same kind of "big-0" estimate holds
for Irk/7ri, and hence both 7ri/Irk and irk/iri are polynomials.] That is,
exp(si) - c - exp(sk), so that for some constant d E C, c = ed and
Si - sk - d. Using the induction hypothesis, we therefore get a formal
derivation of si - sk - d = 0 and, hence, also of exp(si) = c - exp(sk). This
allows us to reduce the original identity to one involving only exp(sj) for
i < j < k - 1, which will be derivable by the induction hypothesis. Again
this yields a derivation of the identity pl exp(sl) + +pk exp(sk) = 0 and
completes the proof.
Theorem 24.3 has an interesting corollary for trigonometric functions,
which we present next. Consider terms in a language with constants for
a l l the complex numbers, variables x1, x2, ... , and function symbols for
addition, multiplication, and for sin and cos. Let E* be the set of all these
terms.
Corollary 24.4. If t, s are any two terms in E* and t = s, then the
identity t = s is provable from the axioms

x + (y + z) _ (x + y) + z, x(yz) _ (xy)z,
x+y=y+x, xy=yx,
x+0=x, 1-x=x,
x(y+z)=xy+xz, 0-x=0,
sin(x + y) = sin(x) cos(y) + coa(x) sin(y),
sin(-1 - x) = -1 - sin(x)
together with all axioms giving the facts of addition, multiplication, sin,
and cos for constants from C.
Proof. We use the fact that in the context of the complex plane, ex is
interdefinable with sin and cos. Note that since the allowed axioms include
the identities sin(a/2) = 1 and cos(ir/2) = 0, we can prove cos(x) =
sin(x + it/2). This in turn allows us to derive the other addition identity,

cos(x + y) = cos(x) cos(y) - sin(x) sin(y).

In E*, let EXP(x) be an abbreviation for the term


It is easy to verify that from the allowed identities in E* one can prove the
exponential identity

EXP(x + y) = EXP(x) . EXP(y)

as well as all the numerical facts involving EXP.


24. Identities of Exponential Functions 181

Given any term t in E*, we define a term t# in E by replacing (induc-


tively) each term of the form sin(s) by
-.5i (exp(i s) - exp(-i s)),
and cos(s) by
.5. (exp(i s) - exp(-i s)).
If t is any term in E we define t* in E* by replacing (inductively) each
term of the form exp(s) by EXP(s). Note that if t E E*, then the identity
t = (t#)* is provable from the axioms allowed in Corollary 24.4.
Now suppose t, s E E* and t - s. Then t# = s#, so the identity
t# = s# is provable from the axioms allowed in Theorem 24.3. Hence,
(t#)* = (s#)* is provable in the system of Corollary 24.4. It follows that
t = s is also provable in that system, completing the proof.
Remark. Suppose t, s E E* and t, s only contain real constants. We do not
know if there is a proof of the identity t = s in the system of Corollary 24.4
in which only real constants appear.
Next we settle positively a conjecture of Schanuel.
Theorem 24.5. Let t E E and suppose the function represented by t is
nowhere equal to 0. Then log(t) is in E, in the sense that t - e$ for some
3EE.
Proof. Let 7r be the function (on C' say) defined by t. There is some
holomorphic function G on Cn so that 7r - eG. We may suppose t is a
term of the form pl exp(s1) + + Pk exp(sk), and we argue by induction
on the number of symbols in sl,... , sk as in the proof of Theorem 24.3.
Clearly we are done if k = 1. Assume k > 1, and for 1 < j < k let 7r5
be the function on C' defined by exp(s,). Then we have puns + +
Pkxk ° eG so that pl(lrle-G) + . + pk(irke-G) - 1. We may assume
the functions plinl a-G, ... , pkake-G are linearly independent (otherwise,
we could replace t by a simpler term to which the induction hypothesis
would apply). Hence, the Hiromi-Ozawa Lemma cannot apply. It follows
as argued in the proof of Theorem 24.3 that there must exist 1 < i < j < k
so that 7ri/irk is identically constant. Again this permits us to reduce the
complexity of t and to apply the induction hypothesis. This completes the
proof.
We conclude with a related problem.
Problem. Suppose that f is an entire function for which there exists a
term t E E such that the function represented by t is equal to f2. Then
must there exist a term s E E such that the function represented by s is f ?
Put more simply (but not as correctly), if f is entire and f2 E E, must
f E E? Even if one assumes that f 2 and f 3 belong to E (and hence fn e E
for n = 2, 3, 4, 5, ... ), does it follow that f (assumed to be entire) lies in
E?
References

1. Apostol, T., Mathematical Analysis, second edition, Addison-Wesley,


Reading, MA, 1974.
2. Beck, W., Efficient quotient representations quotient representations
of meromorphic functions in the disk, Ph.D. Thesis, University of
Illinois, Urbana, IL, 1970.
3. Becker, J., Henson, C. W., and Rubel, L. A., Annals of Math. First-
order conformal invariants, 112 (1980), 123-178.
4. Berenstein, C., Chang, D: C., and Li, B. Q., Complex Variables A
note on Wronskians and linear dependence of entire functions in C,
24 (1993), 131-144.
5. Boss, Entire Functions, Academic Press, New York, 1954.
6. Bucholtz, J. D. and Shaw, J. K., Thins. AMS Zeros of partial sums
and remainders of power series, 166 (1972), 269-184.
7. Buck, R. C., Duke Math J. Integral valued entire functions, 15 (1948),
879-891.
8. Dienes, The Taylor Series, Dover, 1957.
9. Edrei, A., and Fuchs, W. H. J, Trans. Amer. Math. Soc. Meromorphic
functions with several deficient values, 93 (1959), 292-328.
10. Gauthier, P. M., and Hengartner, W., Annals of Math. The value
distribution of most functions of one or several complex variables,
96(2) (1972), 31-52.
11. Gurevic, R. H., Trans. Amer. Math. Soc. Detecting Algebraic (In)De-
pendence of Explicitly Presented Functions (Some Applications of
Nevanlinna Theory to Mathematical Logic), 336 (1) (1993), 1-67.
12. Hayman, W. K., Meromorphic functions, Oxford, at the Clarendon
Press, 1964 (Oxford Mathematical Monographs).
13. Henson, C. W., and Rubel, L. A., Trans. Amer. Math. Soc. Some
applications of Nevanlinna theory to mathematical logic: identities of
References 183

exponential functions, 282(1) (1984), 1- 32 (and Correction 294 (1)


(1986), 381).
14. Hille, E., Ordinary Differential Equations in the Complex Domain,
John Wiley and Sons, New York, 1976.
15. Hiromi, G., and Ozawa, M., Kodai Math. Sem. Report, On the ex-
istence of analytic mappings between two ultrahyperelliptic surfaces,
17 (1965), 281-306.
16. Kujala, R. 0., Bull. Amer. Math. Soc. Functions of finite A-type in
several complex variables, 75 (1969), 104-107.
17. Kujala, R. 0., Trans. Amer. Math. Soc. Functions of finite A-type in
several complex variables, 161 (1971), 327-258.
18. Laine, I., Nevanlinna Theory and Complex Differential Equations, W.
de Gruyter, Berlin, 1993.
19. Lang, S., and Cherry, W., Topics in Nevanlinna Theory, Lecture
Notes in Math. 1443, Springer-Verlag, New York, 1980.
20. Lindelof, E., Ann. Scient. Ec. Norm. Sup. Fonctions entieres d'ordre
entier, 41 (1905), 369-395.
21. Mahler, K., Lectures on Transcendental Numbers, Lectures Notes in
Math. 356, Springer-Verlag, Berlin, 1976.
22. Malliavin, P., and Rubel, L. A., Bull. Soc. Math. France On small
entire functions of exponential type with given zeros, 89 (1961), 175-
206.
23. Markusevic, A. I., Entire Functions, American Elsevier Publishing
Company, New York, 1966.
24. Miles, J., J. Analyse Math. Quotient Representations of Meromorphic
Functions, 25 (1972), 371-388.
25. Miles, J., Bull. Amer. Math. Soc. Representing a meromorphic func-
tion as the quotient of two entire functions of small characteristic, 71
(1970), 1308-1309.
26. Nevanlinna, R., Le Theoreme de Picard-Borel lemma et la Theorie
des Fonctions Meromorphis, second edition, Chelsea Publishing Com-
pany, New York, 1974.
27. Nienhuys, J. W., and Thieman, J. G. F., Proc. Dutch Academy of Sci-
ence, Ser. A. On the existence of entire functions mapping countable
dense sets on each other, 79 (1976), 331-334.
28. Okada, Y., Science Rep. of the Tohoku Imperial University Note on
Power Series, 11 (1922), 43-50.
29. Painlev6, P.. Lecons sur la Theorie Analytique des Equations Differ-
entielles, Professes a Stockholm, Hermann, Paris, 1897.
30. Palais, R. S., Amer. J. Math. Some analogues of Hartogs' theorem in
an algebraic setting, 10 (1978), 387-405.
31. P61ya, G., Math. Zeitschrift Untersuchungen fiber Lucken and Singu-
laritaten von Potenzreihen, 29 (1929), 549-640.
32. Reinhart, G., Schanuel Functions and Algebraic Differential Equa-
tions, Ph.D. Thesis, University of Illinois, Urbana, IL, 1993.
184 Entire and Meromorphic Functions

33. Robinson, R., Trans. Amer. Math. Soc. Undecidable rings, 70 (1951),
137-159.
34. Rogers, H., Theory of Recursive Functions and Effective Computabil-
ity, McGraw-Hill, New York, 1967.
35. Rubes, L. A., Duke Math. J., A Fourier series method for entire func-
tions, 30 (1963), 437-442.
36. Rubel, L. A., and Taylor, B. A., Bull. Soc. Math. France A Fourier
series method for meromorphic and entire functions, 96 (1968), 53-96.
37. Rubel, L. A., and Yang, C. C., Values shared by an entire function and
its derivative, Lecture Notes in Mathematics, Springer-Verlag, Berlin,
1977, pp. 101-103; in Complex Analysis Conference in Lexington,
Kentucky, 1976.
38. Rudin, W., Real and Complex Analysis, McGraw-Hill, New York,
1974.
39. Schonfield, J. R., Mathematical Logic, Addison-Wesley, Reading, MA,
1967.
40. Sodin, M., Ad. Soviet Math. Value Distribution of Sequences of Ra-
tional Functions, 11 (1992).
41. Stoll, W., Proc. Sympos. Pure Math. About entire and meromorphic
functions of exponential type, 11 (1968), Amer. Math. Soc., Provi-
dence, RI, 392-430.
42. Stoll, W., Internat. J. Math. Sci. Value distribution and the lemma
of the logarithmic derivative in polydisks, (4) (1983), 617-669.
43. Taylor, B. A., Duality and entire functions, Thesis, University of Illi-
nois, 1965.
44. Taylor, B. A., Proc. Sympos. Pure Math. The fields of quotients of
some rings of entire functions, 11 (1968), Amer. Math. Soc., Provi-
dence, RI, 468-474.
45. Tsuji, M., Japanese J. Math. On the distribution of zero points of
sections of a power series, 1 (1924), 109-140.
46. Tsuji, M., Japanese J. Math. On the distribution of zero points of
sections of a power series III, 1 (1926), 49-52.
47. van den Dries, L., Pacific J. Math. Exponential rings, exponential
polynomials, and exponential functions, 113(1) (1984), 51-66.
48. Vitter, A., Duke Math. J. The lemma of the logarithmic derivative in
several complex variables, 44 (1977), 89-104.
49. Wilkie, A., On exponentiation-a solution to Tarski's High School
Algebra Problem, preprint, ca. 1982, but never published.
50. Wilkie, A., private communication.
51. Zygmund, A., Trigonometrical Series, 2nd. Ed., Cambridge, 1988.
Index

algebraic integer, 139 defect, 103


deficiency, 103
Boas, 47 deficient value, 105
Borel Lemma, 26, 28, 29 d(a), 103
Borel transform, 43, 125
B(r), 30 eff

branching index, 103 in words, A(r) approaches L


Buck, 131, 139 effectively, 28
effectively, 185
Calderon, 123 28
eff'
canonical products, i, 87-89, 91 , 28
eff

Carleman's Theorem, i, 45, 47 exponential-type, 41


Carlson's Theorem, 47, 130 extreme point, 124
characteristic logarithm, 147
characteristic, Ahlfors-Shimizu, Fabry Gap Theorem, 134
18, 19 finite A-density, 51
characteristic, Cartan, i, 16-19 finite A-type, 49
characteristic, Nevanlinna, 10, 18 finite (A)-type, 13, 40
Clunie's Theorem, 14 first fundamental theorem, i, 9-
11, 95
conjugate indicator diagram, 125
formal power series, 93
convex, 16 Fourier coefficient, 49
convex hull, 124 Fourier coefficients associated
convolution, 127 with Z, 56
corrected ratios, 32 Fourier series method, 49, 90
counting function, 7, 50, 147 fully branched value, 106
186 Entire and Meromorphic Functions

Gauss Mean Value Theorem, 6, Malliavin, 146


21 Malliavin-Rubel Theorem, 47
genus, 88 mapping radius, 141
growth function, 51 maximum term, 30
Miles, 50
Miles-Rubel-Taylor Theorem, i,
Hadamard factorization theorem, 78, 91
89, 151 m(r, f), 10
Hadamard gap theorem, 133
Hahn-Banach Theorem, 48 Nevanlinna, 108, 111
Hausdorff-Young Theorem, 69 Newton polygon, 31
N(r, f), 1U
Hayman, 14
n(r, f ), 7
12, §1.5-1.6, 18
N(r,f),7
Hiromi-Ozawa Lemma, 106, 109
n(t, f), 103
Ho(oo), 125
Horseshoe Theorem, 131 Okada, 96
h(8), 125 order, 40, 41

Phragmen-Lindelof Theorems, i,
indicator function, 125
121, 123
integer valued entire function, 139 Picard's Theorem, i, 99, 101, 103,
interpolating sinusoid, 134 105, 107, 109, 111
mr(f), 31
Jensen's Theorem, 6 Poisson Kernel, 20
Poisson-Jensen Formula, i, 20, 21
P61ya Representation Theorem,
K`-admissible, 144 124, 128
k-admissible, 131 principal indices, 31
Kakeya, 96 Pringsheim's Theorem, 133

A-admissible, 55
quotient representations, i, 78
Laguerre's Theorem, 92 rank of maximum term, 30
Laplace transform, 43, 126 regular, 58
A-balanced, 52, 58
Lindelof, 50, 74 sampling theorem, 47
Lindelof's Theorem, 157 Second Fundamental Theorem, i,
99, 101, 103, 105, 107, 109, 111
Liouville's Theorem, 23, 130
Second Fundamental Theorem of
logarithmic convexification, 32 Nevanlinna, 102
Logarithmic Derivative, 116 share the value a, 108
logarithmic length, 28 sinusoid, 134
logarithmically convex, 17 S(r), 102
A-poised, 53 Stone-Weierstrass Theorem, 48
Index 187

strongly A-balanced, 52 Two Constant Theorem, i, 121,


strongly A-poised, 53 123
subsinusoidal, 134 type, 41
supporting function, 124
Szego 134 , 99

Weierstrass Factorization
Tsuji, 97 Theorem, 88
Universitext (continued)

Sagan: Space-Filling Curves


Samelson: Notes on Lie Algebras
Schiff: Normal Families
Shapiro: Composition Operators and Classical Function Theory
Smith: Power Series From a Computational Point of View
Smorynski: Self-Reference and Modal Logic
Stillwell: Geometry of Surfaces
Stroock: An Introduction to the Theory of Large Deviations
Sunder: An Invitation to von Neumann Algebras
Tondeur: Fuhations on Ricm.Lnudn I'Aanifolds
Universitext
The book is an introduction to the theory of entire and mero-
morphic functions intended for advanced graduate students in
mathematics and for professional mathematicians. The book
provides a clear treatment of the Nevanlinna theory of value
distribution of meromorphic functions, starting from scratch. It
contains the first book-form presentation of the Rubel-Taylor
Fourier series method for meromorphic functions and the Miles
theorem on efficient quotient representation. It has a concise
but complete treatment of the Polya theory of the Borel trans-
form and the conjugate indicator diagram. It contains some of
Buck's results on integer-valued entire functions, and the
Malliavin-Rubel uniqueness theorem. The book closes with
applications to mathematical logic. In particular, the first-order
theory of the ring of entire functions is developed and ques-
tions concerning identities of exponential functions are studied
as in Tarski's "High School Algebra Problems." The approach
of the book gets to the heart of the matter without excessive
scholarly detours. It prepares the reader for further study of the
vast literature on the subject. which is one of the cornerstones
of complex analysis.

10-5

9
ISBN 0-387-94510-5

You might also like