Entire and Meromorphic Functions
Entire and Meromorphic Functions
Entire and Meromorphic Functions
Rubel
with James E. Colliander
ENTIRE AND
MEROMORPHIC
FUNCTIONS
Springer
Universitext
Editors (North America): S. Axler, F.W. Gehring, and P.R. Halmos
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
With 2 illustrations.
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
5. ....................
Elementary Properties of T(r, f) 12
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,
rb rb
/ f da = f(b)a(b) - f(a)a(a) - J a df.
fa a
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),
J f(x)da(x) = E f(xk)ak
a k=1
E akbk =
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
(3.5) E log
r,. <r
! + k+ log r =
fr
+
log t d{n(t) - n(0)} + k+ log r
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(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.
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
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
We may write
Let m(r, f) = i" f ""log+I f (re") I dB. Then we may rewrite Jensen's The-
orem as
(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
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)
+
T (rhJh1) = nT(r, h) E T(r, hj) + n log 2.
0 0
n n
> hjhj = ho + h> hjh'-1
0 1
T(r,fn)=T (r_1fifi).
.
i=0
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
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
For z E E, we have
P
(5.12) log+ > E log+ I I- M
1
If (9(x))I _, 1
g(x) - w
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
f0
0
/3(yt) dt - Jl O(xt) dt =
p
- /3(xt)] dt > 0.
6. The Cartan Formulation of the Characteristic 17
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"°
_, 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)
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
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
( 7.1) u(re'B)
-
The Poisson Formula.
Pu =
1
27r
fir
u(Re"')
R2 - r2
RZ - 2rRcos(B - gyp) r2 dip'
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 '
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
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).
2 f.I=R P log if I.
l
log If (reie) <
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
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
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)
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
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
Choose
R=r 1+ T(r) ) 1
to get
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
(9.6)
v(y) < e(f(x))2.
eff
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
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
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
91
C4
GI
C3
1 2 3 4 5 6
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.
(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
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,
F
11 (r)
r
Rk = log Rk .
But
Ar)
1og =
j log t dp(t), where p(t) _
Rk <t
1.
or
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,
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
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
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:
or
F(r) < B(r) L2µ 1 \\r + A('.) f + 11 .
so that
In this case,
Hence,
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,
Writing F(r) as
n
F(r) Anrn = E AnRn (R)
and using the definition of B(r), we get for R > r:
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
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
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
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
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
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 B r )
k(r) < i
log (1 + og r )
<2logB r+ log rB(r) logB(r).
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
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
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.
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 < _
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.
We have, for ak =k
00
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
log+ log+ M(r) < log+ log+ r + log+ 1og+ M1 (r) + log+ 2
log r - log r
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
J(R) J(R : f)
;R
1
f "/2
x/2 log If (ReiB)I cos 0 d9,
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
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:
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 .
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
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
we get a contradiction.
13
A Fourier Series Method
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
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
n(r, Z) = E 1.
IZ,IST
13. A Fourier Series Method 51
r
E log Ixnl = Jor log (r)t d[n(t, Z)].
Iz+.l<_r
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,.
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
Proof. We have
Zr
n(t Z)
n(r, Z) log 2 < dt < N(2r, Z).
1, t
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
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,
for k = 1,2,3,....
AA(Br)
(13.1.5) Iak + S(r; k : Z)I <
krk
holds, we say that Z is strongly A-poised.
for r > 0 and 1 < k < p(A). For k in this range, we define
A(Bp')
lim = 0.
j moo PjPW
To show that the limit exists, we prove that the sequence {S(pj; k)}, j =
1, 2,..., is a Cauchy sequence. Let
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 <
k
S'(r;k:Z)=k
1Z-1:5r
lrl
Proposition 13.2.2. We have
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
Ck(r;Z:a)= rk {ak+S(r;k:Z)}
(13.2.2) 2
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
A'A(B'r)
Ick(r; Z : a)I <
IkI+1I
and
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
Since c_k(r) = (ck(r)), and since Z has finite upper A-density, the propo-
sition follows immediately.
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(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
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
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
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
A(r) A(2r)
rP-1 - (2r)P-1'
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
(13.3.8)
(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'
tdn(t),
T
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)).
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
Ick(r, f)1:5-r"
)I <- I«k + S(r; k : Z(f)) - S(r; k : W(f))I
(13.4.14) 2
Ick(r,f)I=IZxf x{loglf(Te")I}e-:kedel
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
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,
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
1 1 dO < 1+ e,
27r J xx If (Tese)I011iBr)
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
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.
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
(13.5.9) = E kakzk-1.
r (z) k=1
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
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)
L°
= c Vk zk_1,
k=1
where
f°(z) _ P(z)
00
kakz k 1
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
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
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
IEnI A(O).
<
Iznl
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
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
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
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
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
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
we have
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
k 1 i k
2-i(N+p.)e-il(O.) = 0.
z
82 Entire and Meromorphic Functions
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-,
(14.19)
nEP
cos j0n <
nEU
- j6.+i
1
;,
,
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
an
(14.23) cosjen > fN(9) cos j9 d8.
aJ Bn-1
(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
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
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
2v2<Iznl<a Zn
+ - E
I41=2°2-1
!
Zn
7
1 1
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
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
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
z
PZ(z) = [J EZn- P
znE Z
where p = p(Z).
P(f) = max(n,p),
where n = deg Q.
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
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
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
00
log If(z)I = E c,, e'-O,
m=-00
15. Canonical Products 91
/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
(n)=2m
(r)mlISn I <r
1
(m) cm =
-Tm
2mISnI>r
` ()mL>2
1
ISnICr
\rJ
n lm
if m > p + 1.
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
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
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,
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,
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
(16.3)
_ rr n-m
Llanlfrn(a)ln` H (An -1)IPA; I
^-*"
Ial
L k=1
16. Formal Power Series 95
(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,
10-1 anRn
la"_2Rn-zl anRn
as Il,.
n < 2 max I an-I. ,...,l
R- ( Rn
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.
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.
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
f xlo If(Pe"°)I(pe;
f()x 27r
If (z)
f xlogIf(pe"°)IdW
f(z) (p? r)2 2w
(r, f
1
log+ + 2 log+ + log+ m(r, f) + log+ m 1 .
f (z) I <_ log+ p p r
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,
Now choose p = r+ logT f)fi and use the Borel lemma on 2 log+ T(p, f)
to find,
Notice that
1 f(z) f'(z) + f'(z)
F(z) -
f(z) f'(z) f(z) f(z) - 1
Thus,
and
T (r, '/=//
T lr,
Therefore, on combining (17.1) and (17.2\\) we obtain
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
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
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)
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
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).
(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:
v=1
(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
,,
q
N(r, f lam N(r,
(4-1) lim
T(r,f) + LI o
from which
(r,T. 1
/ , 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
V=1
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,
n
g) (z) = a(µ)
(17.10) >2 G., (z) z p --1,
o (), ... , m - 1.
V=1 (z)
(17.11) T (m(re9))
Y-1
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
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
(17.13)
f'-1 k'
f-1
1 + ekl - 2ek2
(17.15) f= ek1 - eke
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
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
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.
(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
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
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
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
Y=1 f - a
co(r,9)=N(r,9)-N
\r, 9/
then
/')
(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
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).
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
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).
+
r
21r ,1-
log+ I Bv(ree, ,\v) dO + log+ n(p).
x
l a..l <P
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,
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
- p( -r)[2T(P',f)+C1,
P'(P - r) [2T(P',
r(p' r) f) + C] = 1.
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+
zaD
suupplg(z)I = m xlg(z)I
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
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)
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
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
dµ^(z) = e-zwdA(w).
J
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,
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
(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
27ri
f F(w)41(w)dw =
- f w(-w) dw, = F(1).
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
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
=
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
Ta - (a + h(0)) = o(1).
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
_ 1 r F(z) dz
An 2wi A zn z'
132 Entire and Meromorphic Functions
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
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
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
= -h(8)2 sin E
6 [COS 2 b - cos - --j 1
> 0.
Hence,
2
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
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
Example
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.
so that we take
f(n)z'.
9(z) _ 00
n=0
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
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
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
T; ,f =b,,, n=0,1,2,....
For f E R, write
00
g(z) = E(Tn*f)zn-
0
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
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
(22.1) a(y) - J,(x) < A'(y) - A'(x) + D(1); 0 < x < y < oo.
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
and
p(xo + 0) < A'(xo + 0) - A(xo + 0).
We denote by i (xo) the jump of at xo. Then
We let
A*(t) = [W(t)),
22. On Small Entire Functions of Exponential-Type with Given Zeros 149
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
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
where
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,
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
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),
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
z 0 ys1 dL(s) I.
IVI < 2 sup ( 1
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
but cannot apply Lemma 22.7 to d0 since its support may not be compact.
We truncate the support by defining
/ z1 r
(22.19) flog 1 1 + tJ dAk(t) log I1 - jpk(t) dt.
Now,
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
we have
/
(22.23) Lk(y) = 2 flog f 1 + dA(t).
\\
Hence, by (22.13),
At this point, the idea is to find an entire function F for which the
hypothetical formula
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
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
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
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):
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).
X=- Jr,rs 2
log I1 - t2 I dv(t).
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
where
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'
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,
S(R) = bn1,
l bn I <R
bn jW
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:
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:
a E Q,
lal=QU0<_/3A(fi=aV
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:
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
Note that Sl = {i3} and fl(p) _ /3, so that Tl = {9} also. We now
construct the sequences {fn}, {Sn}, and {T.} so that
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
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].
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:
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
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
To make this clearer, we take the first step toward the formula:
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
[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
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
f (z) = E anz",
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
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
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
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
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
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,
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
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
Weierstrass Factorization
Tsuji, 97 Theorem, 88
Universitext (continued)
10-5
9
ISBN 0-387-94510-5