Entire and Meromorphic Functions
Entire and Meromorphic Functions
Walter Bergweiler
CAU Kiel
Winter term 2017/18
3 Jensen’s formula 5
8 Weierstraß products 34
13 Uniqueness theorems 64
References 86
1 AN OVERVIEW 1
1 An overview
A function holomorphic in the complex plane C is called entire. Such a function f
has a power series expansion
∞
X
f (z) = cn z n
n=0
which converges for all z ∈ C.
Examples are
∞
z
X 1 n
f (z) = e = z ,
n=0
n!
1 iz
e − e−iz ,
f (z) = sin z =
2i
or
1 iz
e + e−iz .
f (z) = cos z =
2
A function which is holomorphic in a domain D except for poles is called mero-
morphic in D. (The precise definition and more details will be given later.) In
particular if D = C we just say the function is meromorphic. Functions mero-
morphic in C are precisely those which can be written as a quotient of two entire
functions. (A proof of this will be given later.)
An example is
sin z
tan z = .
cos z
The following result can be seen as the starting point of the mathematical area
we will discuss in this lecture.
Picard’s theorem (1879). Let f be an entire function and let a1 , a2 ∈ C be
distinct. Suppose that f (z) 6= aj for all z ∈ C and all j ∈ {1, 2}. Then f is
constant.
Example. Let f (z) = ez . Then f (z) 6= 0 for all z ∈ C.
The example shows that we need two values a1 and a2 in Picard’s theorem.
Let Cb = C ∪{∞} be the Riemann sphere. A meromorphic function f is thus a
function f : C → C, b with f (z0 ) = ∞ meaning that z0 is a pole. An entire function
f can also be considered as a meromorphic function f : C → C b with f (z) 6= ∞ for
all z ∈ C. This leads to the following result.
Picard’s theorem for meromorphic functions. Let f : C → C b be meromorphic
and a1 , a2 , a3 ∈ C distinct. Suppose that f (z) 6= aj for all z ∈ C and all j ∈
b
{1, 2, 3}. Then f is constant.
The proof is by reduction to the case of entire functions; that is, the case
a3 = ∞. In fact if aj 6= ∞ for all j, then g = 1/(f − a3 ) is entire and
1 1
g(z) = 6= bj :=
f (z) − a3 aj − a3
for j ∈ {1, 2}, so that g and hence f is constant.
2 1 AN OVERVIEW
Borel’s theorem.
log n(r, a) log log M (r, f )
lim sup = lim sup
r→∞ log r r→∞ log r
(iii) f is continuous and for every null-homologous, closed, piecewise smooth curve
γ in D we have Z
f (z)dz = 0,
γ
(iv) f can be developed into a power series around any point in D; that is, for all
z0 ∈ D there exists r > 0 and a sequence (cn )n≥0 in C such that
∞
X
f (z) = cn (z − z0 )n
n=0
for |z − z0 | < r.
A function having one (and hence all) of these properties is called holomorphic.
In (iv) we can take any r > 0 such that
D(z0 , r) := {z ∈ C : |z − z0 | < r} ⊂ D.
The radius of convergence of the power series may be greater than r. The coeffi-
cients have the representations
Z
1 (n) 1 f (ζ)
cn = f (z0 ) = dζ
n! 2πi |ζ−z0 |=ρ (ζ − z0 )n+1
4 2 REVIEW OF COMPLEX FUNCTION THEORY
if 0 < ρ < r.
In (iii), instead of considering all integrals over arbitrary nullhomologous curves,
it
R suffices to consider integrals over the boundaries of triangles in D; that is, if
∂∆
f (z)dz = 0 for every closed triangle ∆ contained in D, then f is holomorphic.
This result is known as Morera’s Theorem.
If K ⊂ C, b not necessarily open, and if f : K → C, then we say that f is
meromorphic (or holomorphic) in K if there exists a domain D ⊃ K such that f
extends to a meromorphic (or holomorphic) function f : D → K.
For K ⊂ C we denote by K the closure of K. For a ∈ C and r > 0 we put
sin(πz) π cos(πz)
f (0) = lim 2
= lim =∞
z→z sin(πz ) z→0 2πz cos(πz 2 )
and
sin(πz) 1
f (1) = lim 2
= .
z→1 sin(πz ) 2
Rational functions (i.e., quotients of polynomials) are meromorphic in C (even
in C). Rational functions of degree 1 are called Möbius transformations (or frac-
b
tional linear transformations). They are bijective maps from C b to C.
b
3 Jensen’s formula
Theorem 3.1. Let r > 0 and let f be holomorphic in D(0, r). Then
Z 2π
1
f (0) = f (reiθ )dθ,
2π 0
Z 2π
1
Re f (0) = Re f (reiθ )dθ,
2π 0
and Z 2π
1
Im f (0) = Im f (reiθ )dθ.
2π 0
Proof. Let
γ : [0, 2π] → C, γ(θ) = reiθ ,
be the standard parametrization of ∂D(0, r). Cauchy’s integral formula yields
Z
1 f (z)
f (0) = dz
2πi γ z
Z 2π
1 f (γ(θ)) 0
= γ (θ)dθ
2πi 0 γ(θ)
Z 2π
1 f (reiθ ) iθ
= ire dθ
2πi 0 reiθ
Z 2π
1
= f (reiθ )dθ.
2π 0
This is the first claim. The second and third claims follow by taking real and
imaginary parts.
Remark. The result is also known as the mean value property.
Theorem 3.2 (Jensen’s formula). Let r > 0 and f meromorphic in D(0, r). Let
a1 , . . . , am be the zeros of f and b1 , . . . , bn the poles of f in D(0, r)\{0}, counted
with multiplicity. Let
X∞
f (z) = ck z k
k=`
6 3 JENSEN’S FORMULA
r(z − a)
ϕ : C → C,
b ϕ(z) = .
r2 − az
Proof of Theorem 3.2. We will start with a special case and then increase the gen-
erality step by step.
(1) Suppose first that f has neither zeros or poles in D(0, r); that is, f is holo-
morphic in D(0, r) and f (z) 6= 0 for z ∈ D(0, r). Then there exists a
function g holomorphic in D(0, r) such that f = exp ◦g. (The existence
of such a function g is standard, but we repeat the argument: First choose
R > r such that f : D(0, R) → C is holomorphic and without zeros. Then
there exists G : D(0, R) → C with G0 = f 0 /f . For example, one may put
R z 0 (ζ)
G(z) = 0 ff (ζ) dζ. We then choose c ∈ C such that g(z) = G(z) + c satisfies
e g(0)
= f (0). The function h defined by h(z) = f (z)/eg(z) = f (z)e−g(z) then
satisfies h(0) = 1 and h0 = f 0 e−g − f g 0 e−g = 0 and thus h(z) = 1 for all
z ∈ D(0, R). Thus f = eg .)
It follows that |f | = eRe g and thus log |f | = Re g. The conclusion now follows
from Theorem 3.1.
(2) We allow zeros and poles in D(0, r)\{0}, but assume that f has no other zeros
and poles in D(0, r); that is, f (0) 6= 0, ∞ and f (z) 6= 0, ∞ for z ∈ ∂D(0, r).
Let ϕ = ϕa be as in Lemma 3.1, that is,
r(z − a)
ϕa (z) = ,
r2 − az
(3) Now we allow that f (0) = 0 or f (z) = ∞, but still assume that f (z) 6= 0, ∞
for z ∈ ∂D(0, r). With ` and c` as in the statement of the theorem we consider
f (z)
h(z) = .
z`
Thus h(0) = c` 6= 0, ∞. Since f (z) = h(z)z ` , part (2) of the proof yields that
Z 2π Z 2π
1 iθ 1
log |f (re )|dθ = log |h(reiθ )|dθ + ` log r
2π 0 2π 0
n m
X r X r
= log |h(0)| − log + log + ` log r,
k=1
|bk | j=1 |aj |
n m
X r X r
= log |c` | − log + log + ` log r
k=1
|bk | j=1 |aj |
as claimed.
(4) We finally consider the general case, i.e., zeros and poles are also allowed on
∂D(0, r). Then
Z 2π
1
log |f (reiθ )|dθ
2π 0
is an improper integral. The existence of this integral is easy to show.
To prove the formula claimed, there are two possibilities:
Proof. Without loss of generality we may assume that 0 < |c1 | ≤ |c2 | ≤ . . . ≤
|cm | ≤ r. Using Z r
r dt
log = log r − log |cj | =
|cj | |cj | t
we find that
m m Z r m Z |cm | Z r
X r X dt X dt dt
log = = +m
j=1
|cj | j=1 |cj | t j=1 |cj | t |cm | t
m m−1
XZ |ck+1 | Z r
X dt dt
= +m
j=1 k=j |ck | t |cm | t
m−1 k Z |ck+1 | Z r
XX dt dt
= +m
k=1 j=1 |ck |
t |cm | t
m−1
X Z |ck+1 | dt Z r
dt
= k +m
k=1 |ck | t |cm | t
m−1
X Z |ck+1 | n(t) Z r
dt
= dt +
|ck | t |cm | t
Zk=1r Z r
n(t) n(t)
= dt = dt.
|c1 | t 0 t
Remark. The above considerations may be elegantly described using the Riemann-
Stieltjes integral. We discuss these integrals in a short excursion:
For a function f : [a, b] → C, a partition of the interval [a, b] given by a = x0 <
x1 < · · · < xn = b and points ξj ∈ [xj−1 , xj ], the corresponding Riemann sum is
given by
n
X
S(f, (xj ), (ξj )) = f (ξj )(xj − xj−1 ).
j=1
If these sums tend to a limit as maxj |xj − xj−1 | → 0, the limit is called Riemann-
Stieltjes integral and denoted by
Z b
f (x)dg(x).
a
For g(t) = n(t) and f (t) = log(r/t) = log r − log t this is precisely the formula from
Lemma 3.2.
Rr
Integrals of the form 0 a(t)
t
dt will occur repeatedly. ItR may be helpful to write
r
them (or at least think of them) as integrals of the form 0 a(t)d log t.
Definition 3.1. Let r > 0 and f meromorphic in D(0, r). For 0 ≤ t ≤ r let n(t, f )
denote the number of poles of f in D(0, t), counted according to multiplicity. Then
Z r
n(t, f ) − n(0, f )
N (r, f ) = dt + n(0, f ) log r
0 t
Remark. (1) The zeros of f are poles of 1/f . For a ∈ C the zeros of f − a are
called a-points of f . They are the poles of 1/(f − a). Thus n(r, 1/(f − a))
and N (r, 1/(f − a)) count the a-points of f .
Sometimes, if the function f considered is clear from the context, we write
n(r, a) and N (r, a) instead of n(r, 1/(f − a)) and N (r, 1/(f − a)). Then we
also write n(r, ∞) and N (r, ∞) instead of n(r, f ) and N (r, f ).
10 3 JENSEN’S FORMULA
Theorem 3.3. Let r > 0 and f meromorphic in D(0, r). Let c` be the first non-
zero coefficient in the Laurent series of f as in Jensen’s formula (Theorem 3.2).
Then Z 2π
1 iθ 1
log |f (re )|dθ = N r, − N (r, f ) + log |c` |.
2π 0 f
and the conclusion follows with Theorem 3.2. The case that f (0) = 0 is analogous.
Here n(0, 1/f ) = `.
Remark. The maximum principle yields that M (r, f ) strictly increases with r if f
is non-constant.
3 JENSEN’S FORMULA 11
Remark. Suppose that log M (r, f ) = O(rρ ) as r → ∞ for some ρ > 0, say
Then
τ Kρ ρ
1 1 1
n r, ≤ N Kr, + O(1) ≤ r + O(1).
f log K f log K
First this follows only if f (0) 6= 0, but considering h(z) = f (z)/z p instead of f if
0 is a zero of f of multiplicity p and noting that n(r, 1/f ) = n(r, h) + p this also
holds if f (0) = 0.
It follows that
1
n r, = O(rρ ) as r → ∞.
f
With K = e1/ρ we actually have
1
n r, ≤ eρτ rρ + O(1)
f
as r → ∞.
More generally,
1
n r, = O(rρ ) as r → ∞ for all a ∈ C .
f −a
So in some sense n(r, 1/(f − a)) does not grow faster than log M (r, f ). We will
see that for all a ∈ C, with at most one exception, n(r, 1/(f − a)) will grow of the
same order of magnitude as log M (r, f ).
n
! n
X X
+
(iii) log xj ≤ log+ xj + log n;
j=1 j=1
1
– N r, is large if f has many a-points;
f −a
1
– m r, is large if f is close to a on some part of ∂D(0, r).
f −a
It follows that
Z r
1 1 1 dt
N r, = N 1, + n t,
f −a f −a 1 f −a t
r
= + O(log r)
π
= T (r, f ) + O(log r).
The first fundamental theorem implies that
1
m r, = O(log r).
f −a
One can show that in fact m(r, 1/(f − a)) = O(1).
Example 4.2. Let f be a rational function, say f (z) = p(z)/q(z) with polynomials
p and q without common zeros. Let
p(z) = a` z ` + . . . + a0 ,
q(z) = bm z m + . . . + b0 ,
with a` 6= 0 and bm 6= 0. By the fundamental theorem of algebra, q has m zeros.
Thus f has m poles, so there exists r0 > 0 with n(r, f ) = m for r ≥ r0 . Hence
Z r
dt
N (r, f ) = N (r0 , f ) + m = m log r + C
r0 t
with a constant C.
We also have
|p(z)| ∼ |a` | · |z|` and |q(z)| ∼ |bm | · |z|m
and hence
|a` | `−m
|f (z)| ∼ |z|
|bm |
as |z| → ∞.
If ` ≤ m, then |f (z)| = O(1) as |z| → ∞ and hence m(r, f ) = O(1) as r → ∞.
Hence T (r, f ) = N (r, f ) + O(1) = m log r + O(1).
If ` > m, then, since
|a` |
log |f (z)| = (` − m) log |z| + log + o(1)
|bm |
as |z| → ∞, we have
m(r, f ) = (` − m) log r + O(1)
as r → ∞. Hence
T (r, f ) = N (r, f ) + m(r, f )
= m log r + (` − m) log r + O(1)
= ` log r + O(1).
5 PROPERTIES OF THE CHARACTERISTIC 17
Had we only aimed at the last equation, we could have reduced the case ` > m to
the case ` ≤ m by passing from f to 1/f and using the first fundamental theorem.
n
! n
X X
(ii) T r, fj ≤ T (r, fj ) + log n.
j=1 j=1
Proof. The corresponding results for m(r, ·) follow directly from Lemma 4.1.
To prove the corresponding estimates for N (r, ·), we note that nj=1 fj can have
Q
a pole only where one of the function fj has a pole. One obtains
n
! n
Y X
n 0, fj ≤ n(0, fj )
j=1 j=1
and ! !
n
Y n
Y n
X
n t, fj − n 0, fj ≤ (n(t, fj ) − n(0, fj )) .
j=1 j=1 j=1
Integration yields !
n
Y n
X
N r, fj ≤ N (r, fj ).
j=1 j=1
Analogously we obtain
n
! n
X X
N r, fj ≤ N (r, fj ).
j=1 j=1
Together with the results for m(r, ·) this yields the conclusion.
Theorem 5.2. Let f be meromorphic and let M be a Möbius transformation (i.e.,
a rational function of degree 1). Then
(ii) z 7→ z + c, c ∈ C (translation)
N (r, P ◦ f ) = d · N (r, f ).
1 3
1 < |ad | · |z|d ≤ |P (z)| ≤ |ad | · |z|d for |z| ≥ r0 .
2 2
Thus
|ad | 3 log d
log + d · log+ |z| ≤ log+ |P (z)| ≤ d · log+ |z| + log+
2 2
for |z| ≥ r0 . This implies that there exists a constant C such that
The order of integration on the left hand side can be interchanged, by Fubini’s
theorem. Here the measurability of
that Z r
s(t, f )
T (r, f ) = dt + log+ |f (0)|.
0 t
x − x1 x2 − x
Φ(x) ≤ Φ(x2 ) + Φ(x1 ).
x2 − x 1 x2 − x 1
Equivalently,
Φ(x) − Φ(x1 ) Φ(x2 ) − Φ(x)
≤ .
x − x1 x2 − x
A differentiable function Φ is convex if and only if Φ0 is non-decreasing. In fact,
if Ψ : I → R is non-decreasing and x0 ∈ I, then Φ : I → R,
Z x
Φ(x) = Ψ(t)dt
x0
Theorem 5.5. Let f be meromorphic. Then N (r, f ) and T (r, f ) are non-de-
creasing and convex in log r.
for a function f holomorphic in D(0, r), and the corresponding formula with f
replaced by Re f .
By Cauchy’s integral formula we have
Z
1 f (ζ)
f (z) = dζ,
2πi |z|=r ζ − z
but taking real parts does not give a formula for Re f (z) in terms of Re f (reiθ ).
Instead we use the following formula.
∂ 2u ∂ 2u
∆u := + = 0.
∂x2 ∂y 2
If f is holomorphic, then Re f is harmonic. This is an immediate consequence of the
Cauchy-Riemann equations. In the opposite direction, if u : D → R is harmonic and
D is simply connected, then there exists a holomorphic function f with u = Re f .
This fails if the domain is not simply connected, an example being u : C \{0} → R,
u(z) = log |z|. But in any simply connected subdomain D of C \{0} we may define
a branch arg : D → R of the argument, and log z = log |z| + i arg z defines a holo-
morphic function in D. Thus the theory of harmonic functions is closely connected
to that of holomorphic functions. We will not consider harmonic functions in detail
and thus have stated Theorem 6.1 for holomorphic functions. The above consider-
ations show, however, that it remains valid for harmonic functions, and in fact the
statement for harmonic functions is equivalent to that for holomorphic functions.
22 6 MAXIMUM MODULUS AND CHARACTERISTIC
r2 (ζ + z)
h : D(0, r) → D(0, r), h(ζ) =
zζ + r2
ζ −z
k(ζ) := h−1 (ζ) = −r2 ,
zζ − r2
and we have h(∂D(0, r)) = ∂D(0, r); see Lemma 3.1 where h(ζ)/r is considered.
We apply the mean value property (Theorem 3.1) to g = f ◦ h and obtain
Jensen’s formula was obtained from the mean value property. Applying similar
arguments to the Poisson integral formula, we obtain the following result.
6 MAXIMUM MODULUS AND CHARACTERISTIC 23
Proof. We restrict, as in the proof of Jensen’s formula (Theorem 3.2) to the case
that f has no zeros and poles on ∂D(0, r), and we proceed as there.
With
r(z − a)
ϕa (z) = 2
r − az
we consider Qn
ϕbk (z)
h(z) = f (z) Qk=1
m .
j=1 ϕaj (z)
Then h is holomorphic in D(0, r) and has no zeros there. Thus log |h| is the real part
of a function holomorphic in D(0, r) and we have |h(ζ)| = |f (ζ)| for ζ ∈ D(0, r).
Hence
Z 2π
1
log f (reiθ ) K(z, r, θ)dθ
2π 0
Z 2π
1
= log h(reiθ ) K(z, r, θ)dθ
2π 0
= log |h(z)|
Xn Xm
= log |f (z)| + log |ϕbk (z)| − log |ϕaj (z)|
k=1 j=1
n 2 m
X r − bk z X r 2 − aj z
= log |f (z)| − log + log
k=1
r(z − bk ) j=1
r(z − aj )
as claimed.
R+r
T (r, f ) ≤ log+ M (r, f ) ≤ T (R, f ).
R−r
Proof. The first inequality is obvious since
Z 2π
1
log+ f reiθ dθ.
T (r, f ) = m(r, f ) =
2π 0
24 6 MAXIMUM MODULUS AND CHARACTERISTIC
Using the functions ϕaj from the previous proof, that is,
R(z − aj )
ϕaj (z) = ,
R 2 − aj z
we have ϕaj (D(0, R)) = D(0, 1) and thus
R 2 − aj z
− log = log |ϕaj (z)| ≤ 0.
R(z − aj )
We also have
R2 − |z|2 R2 − |z|2 R + |z| R+r
0 ≤ K(z, R, θ) = iθ 2
≤ 2
= = .
|Re − z| (R − |z|) R − |z| R−r
It follows that
Z 2π
1
log f Reiθ K(z, R, θ)dθ
log |f (z)| ≤
2π 0
Z 2π
1
log+ f Reiθ K(z, R, θ)dθ
≤
2π 0
Z 2π
1 R+r
≤ log+ f Reiθ dθ
2π 0 R−r
R+r
= T (R, f ).
R−r
The following result is a generalization of Liouville’s theorem that a bounded
entire function is constant.
Theorem 6.4. Let f be an entire function and suppose that
log M (r, f )
L := lim inf < ∞.
r→∞ log r
Then f is a polynomial and deg(f ) ≤ L.
Proof. Let
∞
X
f (z) = cn z n
n=1
and thus
1 M (r, f ) M (r, f )
|cn | ≤ · 2πr n+1 =
2π r rn
for all r > 0 and all n ∈ N. (The last inequality is known as Cauchy’s inequality.)
Given ε > 0 there exist arbitrarily large r with
log M (r, f )
≤L+ε
log r
and thus
M (r, f ) ≤ rL+ε .
It follows that
|cn | ≤ rL+ε−n
for arbitrarily large r and thus that cn = 0 for n > L+ε. The conclusion follows.
The following result is an analogue of Theorem 6.4 for meromorphic functions.
Theorem 6.5. Let f be meromorphic and suppose that
T (r, f )
L := lim inf < ∞.
r→∞ log r
Then f is a rational function (and deg(f ) = L).
Proof. Let C > 1. By Lemma 3.3, applied with R = rC , we have
1
n(r, f ) ≤ C
· N (rC , f )
log(r /r)
C N (rC , f )
= ·
C −1 log rC
C T (rC , f )
≤ ·
C − 1 log rC
C
≤ (L + ε)
C −1
for arbitrarily large r, for any given ε > 0. Thus f has at most L poles. Hence
there exists a polynomial P (of degree at most L) such that g := P f is entire.
Since T (r, P ) ∼ deg(P ) log r as r → ∞ (by Example 2 after Theorem 4.1) we find,
using Theorem 5.1, that
for arbitrarily large r. Applying Theorem 6.3 with R = 2r we thus find that
T (r, f ) ≤ τ rρ
So up to the constant (2 + 1/ρ) ρe the maximum modulus does not grow faster.
The constant (2 + 1/ρ) ρe is not optimal. For ρ > 12 the sharp constant is πρ.
Theorem 6.3 compares log M (r, f ) with T (R, f ) for some R > r. The following
lemma will allow to compare log M (r, f ) with T (r, f ), but not for all values of r.
Lemma 6.1 (Borel). Let x0 , y0 > 0, δ > 0, u : [x0 , ∞) → [y0 , ∞) continuous and
non-decreasing and ϕ : [y0 , ∞) → (0, ∞) continuous and non-increasing. Suppose
that Z ∞
ϕ(x)dx < ∞.
y0
Put
E = {x ∈ [x0 , ∞) : u(x + ϕ(u(x))) ≥ u(x) + δ} .
Then there exist sequences (xk ) and (x0k ) in [x0 , ∞), with xk < x0k ≤ xk+1 for all
k ∈ N, such that
[∞
E⊂ [xk , x0k ]
k=1
and ∞
X
(x0k − xk ) < ∞.
k=1
Remark. The last two conditions imply that E has finite measure. The lemma says
that if x 6∈ E, then u(x + ϕ(u(x))) < u(x) + δ.
6 MAXIMUM MODULUS AND CHARACTERISTIC 27
Proof. We may assume that E is unbounded, since otherwise the result is obvious.
Put x1 = min E. Note that this minimum exists, since u and ϕ are continuous
and hence E is closed. Put x01 = x1 + ϕ(u(x1 )).
We define (xk ) and (x0k ) recursively by xk = min E ∩ [x0k−1 , ∞ ) and x0k =
Finally, if k ≥ 2, then
x0k − xk = ϕ(u(xk ))
≤ ϕ(u(x1 ) + (k − 1)δ)
≤ ϕ(y0 + (k − 1)δ)
Z y0 +(k−1)δ
1
= ϕ(y0 + (k − 1)δ) · dx
δ y0 +(k−2)δ
Z y0 +(k−1)δ
1
≤ ϕ(x)dx
δ y0 +(k−2)δ
and thus ∞ Z ∞
X 1
(x0k − xk ) ≤ ϕ(x)dx < ∞.
k=2
δ y0
if log r 6∈ E.
Let
F = {r : log r ∈ E} .
0
Putting rk = exk and rk0 = exk the condition
∞
X
(x0k − xk ) < ∞
k=1
and we have
∞
[
F ⊂ [rk , rk0 ] .
k=1
In particular, we have Z
dt
< ∞,
F t
corresponding to the condition
Z
dx < ∞.
E
Theorem 6.6. Let f be entire, r0 > 0 and ϕ : [r0 , ∞) → R continuous and de-
creasing with Z ∞
ϕ(t)dt < ∞.
r0
such that
T (r, f )
log M (r, f ) ≤ for r 6∈ F.
ϕ(log T (r, f ))
Proof of Theorem 6.6. We apply Borel’s Lemma 6.1 and the remark following its
proof with w(r) = T (r, f ) and K = 2. Taking R = r(1 + ϕ(log T (r, f ))) in
6 MAXIMUM MODULUS AND CHARACTERISTIC 29
Theorem 5.5 says that T (r, f ) and N (r, f ) are convex in log r (and non-de-
creasing). The following result is an analogue for the maximum modulus.
Theorem 6.7 (Hadamard three circle theorem). Let f be entire, f 6= 0. Then
log M (r, f ) is convex in log r.
Proof. We have to show that if 0 < r1 < r < r2 , then
log r2 − log r log r − log r1
log M (r, f ) ≤ log M (r1 , f ) + log M (r2 , f ).
log r2 − log r1 log r2 − log r1
In order to do so, we consider for α ∈ R the function
φ : C \{0} → R, φ(z) = |z|α |f (z)|.
Then φ has no strict local maxima, since locally φ is the absolute value of a holo-
morphic function. Indeed, in any simply connected subdomain U of C \{0} there
exists a branch log of the logarithm, and
φ(z) = |z α f (z)| = |eα log z f (z)|
for z ∈ U .
We now choose α ∈ R such that
max φ(z) = max φ(z),
|z|=r1 |z|=r2
which is equivalent to
r1α M (r1 , f ) = r2α M (r2 , f ).
Solving the equation for α shows that this holds for
log M (r2 , f ) − log M (r1 , f )
α=− .
log r2 − log r1
Since φ has no local maxima we have
rα M (r, f ) ≤ r1α M (r1 , f ) = r1α M (r2 , f ).
Inserting the value of α found in this equation yields the conclusion.
30 6 MAXIMUM MODULUS AND CHARACTERISTIC
Instead of
M (r, f ) = max |f (z)|
|z|=r
Putting
2π
reiθ − z
Z
1
h(z) = Re f (reiθ ) dθ
2π 0 reiθ + z
we find that
Z 2π
1
Re h(z) = Re f (reiθ )K(z, r, θ)dθ = Re f (z)
2π 0
(2) Let
∞
eiz + e−iz X (−1)k 2k
f (z) = cos z = = z .
2 k=0
(2k)!
Then
er + er
M (r, f ) ≤ = er
2
and
er + e−r er
M (r, f ) ≥ |f (ir)| = ≥ .
2 2
It follows that log M (r, f ) = r + O(1) and hence ρ(f ) = λ(f ) = 1.
7 THE ORDER OF A MEROMORPHIC FUNCTION 33
(3) Let
∞
√ X (−1)k k
f (z) = cos z = z .
k=0
(2k)!
Then √
√ r
r e
M (r, f ) ≤ e and M (r, f ) ≥ f (−r) ≥
2
1
and hence ρ(f ) = λ(f ) = 2
.
The following result is classical, but we omit the proof.
Then
n log n
ρ(f ) = lim sup .
n→∞ − log |cn |
Here we put n log n/(− log |cn |) = 0 if cn = 0.
Remark. (1) The result gives an easy possibility to construct functions with
ρ(f ) = µ for any preassigned µ ≥ 1. E.g. for 0 < µ < ∞ we can take
cn = n−n/ρ .
(2) Taking the lower limit in the formula in Theorem 7.2 does in general not give
the lower order. However, we have
nk log nk−1
λ(f ) = max lim inf ,
(nk ) k→∞ − log |cnk |
Proof. We only consider the sum f = f1 + f2 , the other cases are analogous.
By Theorem 5.1 we have
T (r, f ) ≤ T (r, f1 ) + T (r, f2 ) + log 2 ≤ 2 max {T (r, f1 ), T (r, f2 )} + log 2
and thus
log T (r, f ) ≤ log+ max {T (r, f1 ), T (r, f2 )} + O(1)
which implies that ρ(f ) ≤ max{ρ(f1 ), ρ(f2 )}. Suppose that ρ(f1 ) 6= ρ(f2 ), say
ρ(f1 ) < ρ(f2 ). Then
ρ(f2 ) = ρ(f − f1 ) ≤ max{ρ(f1 ), ρ(f2 )} ≤ ρ(f )
since ρ(f1 ) < ρ(f2 ). Thus ρ(f ) = ρ(f2 ) = max{ρ(f1 ), ρ(f2 )}.
8 Weierstraß products
The fundamental theorem of algebra says that a polynomial p(z) = ad z d + · · · + a0
with a0 , ad 6= 0 has a factorization
d d
Y Y z
p(z) = ad (z − zj ) = a0 1− ,
j=1 j=1
zj
∞
Y k
Y
aj := a1 · a2 · . . . · aN −1 · lim aj .
k→∞
j=1 j=N
so that the absolute convergence of the series implies that of the infinite product.
Suppose now that the infinite product converges absolutely. Then cj → 0 and thus
there exists N with |cj | ≤ 1 for j ≥ N . Since x ≤ 2 log(1 + x) for 0 ≤ x ≤ 1 we
deduce that
n n n
!
X X Y
|cj | ≤ 2 log(1 + |cj |) = 2 log (1 + |cj |)
j=N j=N j=N
Theorem 8.2. Let D be a domain and let (uj ) be a sequence of functions holo-
morphic in D. Suppose that ∞
P
j=1 |u j | converges locally uniformly in D. Then
∞
Y
f (z) := (1 + uj (z))
j=1
E(u, 0) = 1 − u
and
q
!
X uk
E(u, q) = (1 − u) exp
j=1
k
for q ∈ N are called Weierstraß primary factors.
Lemma 8.1. Let q ∈ N0 and u ∈ D(0, 1). Then
|E(u, q) − 1| ≤ |u|q+1 .
Since all coefficients of the Taylor series expansion of the right hand side are neg-
ative we conclude that ∞
X
E(u, q) = 1 − αj uj
j=q+1
8 WEIERSTRASS PRODUCTS 37
Hence
∞
X
|E(u, q) − 1| ≤ αj |u|j
j=q+1
∞
X
= |u|q+1 αj |u|j−q−1
j=q+1
X∞
≤ |u|q+1 αj
j=q+1
q+1
= |u| .
Theorem 8.3. Let (zj ) be a sequence in C \{0} with limj→∞ |zj | = ∞. Then there
exists a sequence (qj ) in N0 such that
∞ qj +1
X z
j=1
zj
converges locally uniformly in C and represents an entire function whose zeros are
precisely the zj .
Here multiplicities are counted in the sense that if a ∈ C appears m times in
the sequence (zj ), i.e., card{j ∈ N : zj = a} = m, then a is a zero of multiplicity m.
Proof. The first claim is satisfied for qj = j. To see this let R > 0. Then there
exists j0 ∈ N with |zj | ≥ 2R for j ≥ j0 . For z ∈ D(0, R) and j ≥ j0 we thus have
q +1 j+1 j+1
z j z R 1
= ≤ = j+1 .
zj zj 2R 2
P∞ qj +1
This implies that the sequence j=1 |z/zj | converges uniformly in D(0, R).
Since R > 0 was arbitrary this yields locally uniform convergence in C.
Let now (qj ) be a sequence such that this series converges locally uniformly.
Theorem 8.2 now yields that
∞
Y z
E ,q
j=1
zj
is entire without zeros and thus has the form eg with an entire function g.
Remark. If f has only finitely many zeros, we have the same result, except that
the product occurring is finite.
Theorem 8.5. Let f be meromorphic in C. Then there are entire functions g and
h such that f = g/h.
Proof. Let (zj ) be the (finite or infinite) sequence of poles of f in C \{0}, counted
according to multiplicity. By Theorem 8.3 there exists an entire function k which
has precisely the zj as zeros.
If 0 is a pole of f , let m be its multiplicity and put m = 0 otherwise. Define h
by h(z) = z m k(z). Then h is entire and so is g = f h. The conclusion follows.
is called the exponent of convergence of the sequence (zj ), with σ = ∞ if the series
diverges for all µ.
If σ < ∞, then
( ∞
)
X 1
q := q((zj )) := min m ∈ N : m+1
<∞
j=1
|zj |
Remark. 1. Clearly q ≤ σ ≤ q + 1.
9 THE HADAMARD FACTORIZATION THEOREM 39
Theorem 9.1. Let (zj ) be as in Definition 9.1. Let n(r) be the number of zj in
D(0, r) and put Z r
n(t)
N (r) = dt.
0 t
Then
log n(r) log N (r)
σ = lim sup = lim sup .
r→∞ log r r→∞ log r
In the proof we will use the following lemmas.
Lemma 9.1. Let (zj ) and n(r) be as in Theorem 9.1, and let µ > 0. Then
X 1 Z r
n(t) n(r)
µ
=µ µ+1
dt + µ .
|zj | 0 t r
|zj |≤r
Lemma 9.2. Let (zj ) and n(r) be as in Theorem 9.1, and let µ > 0. Then
∞
X 1
µ
<∞
j=1
|zj |
40 9 THE HADAMARD FACTORIZATION THEOREM
if and only if Z ∞
n(t)
dt < ∞.
0 tµ+1
In this case
n(r)
lim = 0.
r→∞ r µ
Proof. Suppose first that the series converges. Lemma 9.1 implies that
R ∞the integral
µ
also converges and the limit limr→∞ n(r)/r = 0 exists. But since 1 dt/t = ∞,
the existence of the limit and the comparison test imply that the limit is actually
equal to 0.
Suppose now that the integral converges. Using the comparison test as before we
see that lim inf r→∞ n(r)/rµ = 0. Lemma 9.1 now yields that the series converges,
and as before this yields that in fact limr→∞ n(r)/rµ = 0.
Proof of Theorem 9.1. Suppose that σ < ∞ and choose µ with σ < µ < ∞. Then
∞
X 1
µ
< ∞.
j=1
|zj |
We thus have n(r) ≤ rµ and hence log n(r) ≤ µ log r for large r and hence
log n(r)
lim sup ≤ µ.
r→∞ log r
log n(r)
lim sup ≤σ
r→∞ log r
This is the first inequality claimed. To prove the second one, let q ≥ 1 and |u| ≥ 1.
Then
q q q
!
X 1 j X 1 q X 1
log |E(u, p)| ≤ log |1 − u| + |u| ≤ |u| + |u| ≤ |u|q 1 + .
j=1
j j=1
j j=1
j
Since q Z q
X 1 dt
≤ = log q
j=2
j 1 t
=: Σ1 + Σ2 .
Since (zj ) has genus q, Lemma 9.2 yields that the integral
Z ∞
n(t)
dt
r tq+2
converges and
n(R)
lim = 0.
R→∞ Rq+1
It follows that Z ∞
X 1 n(t) n(r)
= (q + 1) dt − q+1
|zj |q+1 r tq+2 r
|zj |>r
9 THE HADAMARD FACTORIZATION THEOREM 43
and thus Z ∞
q+1 n(t)
Σ2 ≤ (q + 1)r dt − n(r).
r tq+2
If q = 0, we have
X z
Σ1 = log 1 −
zj
|zj |≤r
X r
≤ log 1 +
|zj |
|zj |≤r
X r
≤ log 2
|zj |
|zj |≤r
X X r
= log 2 + log
|zj |
|zj |≤r |zj |≤r
Z r
n(t)
= n(r) log 2 + dt
0 t
Z r
n(t)
≤ n(r) + dt
0 t
by Lemma 3.2. Combing the estimates for Σ1 and Σ2 gives the conclusion in this
case.
If q ≥ 1, then, by Lemmas 9.3 and 9.1,
X r q
Σ1 ≤ M (q)
|zj |
|zj |≤r
X 1
= M (q)rq
|zj |q
|zj |≤r
Z r
q n(t) n(r)
= M (q)r q q+1
dt + q
0 t r
Z r
n(t)
= M (q)q rq q+1
dt + M (q)n(r).
0 t
Adding this to the estimate for Σ1 yields the conclusion, since we may replace
M (q)q also by M (q)(q + 1) before the first integral.
Theorem 9.4. Let (zj ), q and P be as in Theorem 9.3; that is, P is the Weierstraß
product formed with the zj , of genus q. Then ρ(P ) = σ((zj )).
44 9 THE HADAMARD FACTORIZATION THEOREM
It follows that ρ(P ) ≤ σ + ε, and this yields ρ(P ) ≤ σ since ε can be chosen
arbitrarily small.
h(z) = eQ(z)
Now ρ(z −m ) = 0 and denoting by (zj ) and (pj ) the sequences of zeros and poles in
C \{0} we have
ρ(P0 ) = σ((zj )) ≤ ρ(f )
and
ρ(P∞ ) = σ((pj )) ≤ ρ(f ).
Hence
ρ(h) ≤ ρ(f ).
It follows that for ε > 0 we have
log M (r, f ) = log max eQ(z) = log max eRe Q(z) = A(r, Q)
|z|=r |z|=r
we thus have
A(r, Q) ≤ rρ(f )+ε
for large r. Theorem 6.10 yields that Q is a polynomial and deg Q ≤ ρ(f ) + ε. The
conclusion follows since ε can be taken arbitrarily small.
Example 9.1. Let f (z) = sin πz. The zeros are 0, ±1, ±2, . . . . Since
X 1
<∞ for µ>1
|k|µ
k∈Z\{0}
while X 1
=∞ for µ ≤ 1,
|k|µ
k∈Z\{0}
the exponent of convergence and the genus of the sequence of zeros are both equal
to 1. Moreover, ρ(f ) = 1 and f has a simple zero at 0. By Hadamard’s factorization
theorem there exits a polynomial Q of degree at most 1 such that
Q(z)
Y z
f (z) = ze E ,1 .
zj
j∈Z\{0}
46 9 THE HADAMARD FACTORIZATION THEOREM
We conclude that
eaz+b = ea(−z)+b = e−az+b
for all z ∈ C and thus a = 0. Hence
sin πz 1
eb = Q∞ 2 2
z j=1 (1 − z /j )
for all z ∈ C, where “for all z ∈ C” is understood to mean that if the expression
on the right hand side has a removable singularity, then it has to be replaced by
the appropriate value. It follows that
sin πz 1 sin πz
eb = lim · Q∞ 2 2
= lim = π.
z j=1 (1 − z /j ) z
z→0 z→0
1
f (z) = z −1 eaz+b Q∞ z −z/j
j=1 (1 + j )e
eaz+b
= Q∞
z j=1 (1 + zj )e−z/j
∞
eaz+b Y jez/j
= ·
z j=1
j+z
with a, b ∈ C.
The functions zf (z) and f (z + 1) have the same poles (in −1, −2, −3, . . . ) and
no zeros. Their quotient zf (z)/f (z +1) is thus entire and without zeros. Moreover,
it has order at most 1 and is thus of the form ecz+d with c, d ∈ C. More precisely,
we have
∞
zf (z) eaz+d (z + 1) Y j j + z + 1 ez/j
=
f (z + 1) ea(z+1)+b j=1 j + z j e(z+1)/j
n
−a
Y j+z+1
= e (z + 1) lim e−1/j
n→∞
j=1
j+z
n
!
X 1
= e−a lim (z + n + 1) exp −
n→∞
j=1
j
n
!
z+n+1 X1
= e−a lim exp log n −
n→∞ n j=1
j
n
!
X 1
= e−a lim exp log n −
n→∞
j=1
j
= e−a eγ
where !
n
X 1
γ = lim log n − .
n→∞
j=1
j
The limit γ = 0.57721 . . . is called the Euler-Mascheroni constant (or Euler’s con-
stant).
For a detailed proof that the limit exists one may write
n Z n
X 1 1 1 1
− log n = − dx +
j=1
j 1 [x] x n
48 10 PREPARATIONS
1 1 x − [x] 1 2
0≤ − = ≤ ≤ 2 for x ≥ 1
[x] x x[x] x[x] x
the integral Z ∞
1 1
− dx
1 [x] x
converges.
Choosing a = −γ in the definition of f we thus have f (z +1) = zf (z). Choosing
b = 0 we have
e−γz
f (1) = lim zf (z) = lim Q∞ −z/j
= 1.
j=1 (1 + z/j)e
z→0 z→0
j=1
j
As noted above, we have Γ(n) = (n − 1)! for n ∈ N and Γ(z + 1) = zΓ(z) for
z ∈ C \{0, −1, −2, . . . }.
Remark. For Re z > 0 the Gamma function is often defined by
Z ∞
Γ(z) = e−t tz−1 dt.
0
It can be shown that this agrees with our definition. The functional equation
Γ(z + 1) = zΓ(z) follows from this via integration by parts.
as r → ∞, for all a ∈ C. So if T (r, f ) is large, then one of the terms N (r, 1/(f − a))
or m(r, 1/(f − a)) must be large, meaning that f has many a-points or f is close
to a on some part of the circle ∂D(0, r).
10 PREPARATIONS 49
The second fundamental theorem will say that for “most” values of a the first
alternative will hold. More precisely, we will see that if a1 , . . . , aq ∈ C, then
q
X 1
m r, + m(r, f ) ≤ 2T (r, f ) + . . .
j=1
f − aj
where the dots indicate a “small error term”. Together with the first fundamental
theorem this yields
q
X 1
(q − 1)T (r, f ) ≤ N r, + N (r, f ) + . . . .
j=1
f − aj
where
1
N1 (r) = N r, 0 + 2N (r, f ) − N (r, f 0 ) ≥ 0
f
for r ≥ 1 and
q
f0 f0
X
S(r) ≤ m r, + m r, + O(1)
f j=1
f − aj
as r → ∞.
Since q
1 X 1
N r, = N r,
P ◦f j=1
f − aj
the first fundamental theorem yields that
1 1
m r, = T (r, P ◦ f ) − N r, + O(1)
P ◦f P ◦f
q
X 1
= q · T (r, f ) − N r, + O(1)
j=1
f − a j
q
X 1
= T (r, f ) − N r, + O(1)
j=1
f − a j
q
X 1
= m r, + O(1).
j=1
f − aj
with constants cj ∈ C \{0}. (In fact, we have cj = 1/P 0 (aj ) for 1 ≤ j ≤ q.) It
follows, using Lemma 4.1, that
f0
1 1
m r, = m r, · + O(1)
P ◦f P ◦ f f0
f0
1
≤ m r, + m r, 0 + O(1)
P ◦f f
q
!
f0
X 1
= m r, cj + m r, 0 + O(1)
j=1
f − aj f
q q
f0
X
X
+ 1
≤ m r, + log |cj | + log q + m r, 0 + O(1).
j=1
f − a j j=1
f
Adding m(r, f ) on both sides and noting that T (r, f ) = m(r, f ) + N (r, f ) by
definition yields the conclusion.
This yields the conclusion if f has no zeros and poles, since then f may be written in
the form f = eg with a holomorphic function f . Then f 0 /f = g and log |f | = Re g.
In the general case we assume again for simplicity that f has no zeros and poles
on ∂D(0, r) and, as in the proof of Theorem 3.2 and 6.2, consider
Qn
j=1 ϕbj (z)
h(z) = f (z) Qm
j=1 ϕaj (z)
with
r(z − a)
ϕa (z) = .
r2 − az
Then h has no zeros and poles in D(0, r) and |h(reiθ )| = |f (reiθ )| for θ ∈ R. Using
the result already proved we conclude that
Z 2π
1 2r h0 (z)
log f reiθ
dθ =
2π 0 (reiθ − z)2 h(z)
n 0 m 0
f 0 (z) X ϕbj (z) X ϕaj (z)
= + −
f (z) j=1
ϕbj (z) j=1 ϕaj (z)
n
f 0 (z) X
1 bj
= + +
f (z) j=1
z − bj r 2 − bj z
m
X 1 aj
− +
j=1
z − aj r 2 − aj z
as claimed.
Remark. In the last equation we have used that if F = f1 · f2 and G = f1 /f2 , then
F0 f0 f0 G0 f0 f0
= 1+ 2 and = 1 − 2.
F f1 f2 G f1 f2
These rules can be proved (and remembered) by noting that locally we may write
log F = log f1 + log f2 with suitably chosen branches of the logarithm so that
F0 f0 f0
= (log F )0 = (log f1 )0 + (log f2 )0 = 1 + 2 ,
F f1 f2
with an analogous reasoning for G0 /G. Of course, the above rules can also be
checked with the product and quotient rules.
n
! n
X X
f λj xj ≤ λj f (xj ).
j=1 j=1
11 THE LEMMA ON THE LOGARITHMIC DERIVATIVE 53
Proof. The conclusion is trivial if all xj are equal. Suppose that this is not the
case. Then n
X
ξ := λj xj
j=1
is in the interior of I. Since f is convex, there exists an affine function g such that
g(ξ) = f (ξ) and g(t) ≤ f (t) for all t ∈ I. Here by affine we mean that g has the
form g(t) = at + b for certain a, b ∈ R. It follows that
n
!
X
f (ξ) = g(ξ) = aξ + b = a λj xj + b
j=1
n
! n
X X
=a λ j xj +b λj
j=1 j=1
n
X
= λj (axj + b)
j=1
n
X
= λj g(xj )
j=1
n
X
≤ λj f (xj ).
j=1
The conclusion follows by taking the limit through a sequence of partitions with
maximal interval length tending to 0, using also that f is continuous.
Remark. The result can be extended to Lebesgue integrable function, e.g. by ap-
proximation by continuous functions. Using f = exp and g = log h one obtains the
following result.
54 11 THE LEMMA ON THE LOGARITHMIC DERIVATIVE
Theorem 11.4. Let f be meromorphic with f (0) = 1 and let 0 < r < R and
0 < α < 1. Then
1+α !
f0
1 24 R T (R, f )
m r, ≤ log 1 + .
f α 1−α R−r rα
Proof. Let r < ρ < R and z ∈ D(0, ρ). Theorem 11.1 yields that if f (z) 6= 0, ∞,
then
Z 2π
f 0 (z) 1 2ρ X 1 aj
log f ρeiθ
≤ 2
dθ + + 2
f (z) 2π 0 (ρ − |z|) z − aj ρ − aj z
|aj |<ρ
X 1 bj
+ + 2 .
z − bj ρ − bj z
|bj |<ρ
Here the aj and bj are the zeros and poles of f . We combine the sequences of zeros
and poles into one sequence (cj ). Noting that
1
| log |w|| = log+ |w| + log+
|w|
f 0 (z)
X
2ρ 1 1 cj
≤ m(ρ, f ) + m ρ, + + .
f (z) (ρ − |z|)2 f z − cj ρ 2 − cj z
|cj |<ρ
f 0 (z) 4ρ X 1
≤ 2
T (ρ, f ) + 2 .
f (z) (ρ − |z|) |z − cj |
|cj |<ρ
f 0 (z) f 0 (z)
+
log ≤ log 1 +
f (z) f (z)
α
f 0 (z)
1
= log 1 +
α f (z)
α
1 (4ρ) X 1
≤ log1 + 2α
T (ρ, f )α + 2α .
α (ρ − |z|) |z − cj |α
|cj |<ρ
Integrating this, using Jensen’s inequality, and also noting that 4α < 4 and 2α < 2
we obtain
0
Z 2π α
f 1 1 4ρ X 1
m r, ≤ log1 + 2α
T (ρ, f )α + 2 dθ
f α 2π 0 (ρ − r) |re − cj |α
iθ
|cj |<ρ
α Z 2π
1 4ρ X 1 dθ
≤ log1 + 2α
T (ρ, f )α + 2 .
α (ρ − r) 2π 0 |re − cj |α
iθ
|cj |<ρ
f0 4ρα
α 2
m r, ≤ log 1 + T (ρ, f ) + α (n(ρ, f ) + n (ρ, 1/f )) .
f (ρ − r)2α r (1 − α)
1 1
We now choose ρ = (R + r). Then ρ ≤ R and ρ − r = (R − r) and thus
2 2
4ρα α 16Rα
T (ρ, f ) ≤ T (R, f ).
(ρ − r)2α (R − r)α
and also
1 2R
n ρ, ≤ T (R, f ).
f R−r
Altogether we thus have
f0 16Rα
1 8R
m r, ≤ log 1 + T (R, f ) + T (R, f ) .
f α (R − r)2α (R − r)rα (1 − α)
11 THE LEMMA ON THE LOGARITHMIC DERIVATIVE 57
for x 6∈ E.
f0
m r, ≤ (1 + ε) (log T (r, f ) + log r)
f
for r 6∈ E.
Proof. Without loss of generality we may assume that f (0) = 1, since otherwise
we can replace f by g(z) = cz p f (z) with a suitable c ∈ C \{0} and p ∈ Z. Then
0
f0
g
m r, = m r, + O(1)
g f
and
T (r, g) = T (r, f ) + O(log r) = (1 + o(1))T (r, f )
58 11 THE LEMMA ON THE LOGARITHMIC DERIVATIVE
f0
24 1+α 2+2α T (R, f )
α · m r, ≤ log 1 + R (log T (r, f ))
f 1−α rα
24 1+α 2+2α
≤ log 1 + 2 r(log T (r, f )) T (R, f )
1−α
24 1+α
≤ log 2 + log 2 + log r
1−α
+ (2 + 2α) log log T (r, f ) + log T (R, f )
24 1+α
≤ log 2 + log 2 + log r
1−α
+ (2 + 2α) log log T (r, f ) + log T (r, f ) + log K
f0
ε
α · m r, ≤ 1+ log T (r, f ) + log r
f 2
f 0 (z)
1
=O
f (z) |z|
as z → ∞ and thus
f0
m r, =0
f
for all large r. Therefore it is no restriction to consider only transcendental func-
tions in Theorem 11.5.
If f has finite order ρ(f ), then Theorem 11.5 gives
f0
m r, ≤ (ρ(f ) + 1 + ε) log r for r 6∈ E.
f
The following result improves the constant on the right hand side and – more
importantly – says that in this case no exceptional set E is required.
12 THE SECOND FUNDAMENTAL THEOREM 59
Let ε > 0. Then T (2r, f ) ≤ (2r)ρ(f )+ε for large r and thus
f0
1 24 1+α ρ(f )+ε ρ(f )+ε−α
m r, ≤ log 1 + 2 2 r
f α 1−α
1 + 24 1+α+ρ(f )+ε + ρ(f )+ε−α
≤ log 2 + log r + log 2
α 1−α
max{ρ(f ) + ε − α, 0}
≤ log r + O(1)
α
as r → ∞. Considering the limits as α → 1 and ε → 0 we obtain the conclusion.
where
1
N1 (r) := N r, 0 + 2N (r, f ) − N (r, f 0 )
f
and the term S(r, f ) satisfies the following:
(i) There exists a subset E of [0, ∞) of finite measure such that
as r → ∞, r 6∈ E.
as r → ∞, r 6∈ E. The conclusions (ii) and (iii) yield that this holds without the
exceptional set E if f has finite order.
Proof of Theorem 12.1. Theorem 10.1 says that the conclusion holds with
q
f0 f0
X
S(r, f ) = m r, + m r, + O(1).
f j=1
f − aj
for all a ∈ C.
b
In analogy to the terminology n(r, a) we also put
1
m(r, a) = m r,
f −a
if a ∈ C and m(r, ∞) = m(r, f ). The first fundamental theorem thus takes the
form
N (r, a) + m(r, a) = T (r, f ) + O(1)
for all a ∈ C.
b This implies that
m(r, a)
δ(a, f ) = lim inf .
r→∞ T (r, f )
and, in particular, X
δ(a, f ) ≤ 2.
a∈C
b
and thus
q
X
Θ(aj , f ) ≤ 2.
j=1
N (r, aj ) ≥ mj N (r, aj ).
and thus
N (r, aj ) 1
Θ(aj , f ) = 1 − lim sup ≥1− .
r→∞ T (r, f ) mj
The conclusion now follows from Theorem 12.4.
Remark. If f does not take a value aj at all, that is, f (z) 6= aj for all z ∈ C, then
the hypothesis of Theorem 12.5 is satisfied for any mj ∈ N.
Taking the limit as mj → ∞ we may put 1/mj = 0 in Theorem 12.5 then.
Example. Let f (z) = cos z, a1 = 1, a2 = −1, a3 = ∞. We can take m1 = m2 = 2
and, in the sense of the previous remark, m3 = ∞ and 1/m3 = 0. Then we have
equality in Theorem 12.5.
13 Uniqueness theorems
Our first application of the second fundamental theorem are some uniqueness the-
orems.
Definition 13.1. Let f and g be meromorphic and a ∈ C. b Then f and g are said
−1 −1
to share the value a if f (a) = g (a); that is, for z ∈ C we have f (z) = a if and
only if g(z) = a.
If, in addition, for every a-point z0 of f and hence g the multiplicity of z0 as
an a-point of f is the same as the multiplicity as an a-point of g, then f and g are
said to share the value a with multiplicity.
(2) Let
ez + 1 (ez + 1)2
f (z) = and g(z) =
(ez − 1)2 8(ez − 1)
Then f and g share 0, ∞, 1 and − 81 , but without multiplicity. This is clear
for the values 0 and ∞. Moreover, we have
and
g(z) = 1 ⇔ (ez + 1)2 = 8(ez − 1) ⇔ e2z + 2ez + 1 = 8ez − 8
⇔ e2z − 6ez + 9 = 0 ⇔ (ez − 3)2 = 0 ⇔ ez = 3.
Thus f and g share 1. An analogous argument shows that they also share − 81 .
13 UNIQUENESS THEOREMS 65
On the other hand, by the second fundamental theorem (more precisely, Theo-
rem 12.3) we have
5
X
3T (r, f ) ≤ N (r, aj ) + S(r, f )
j=1
Remark. (1) The examples given before Theorem 13.1 show that the number 5
in Theorem 13.1 cannot be replaced by 4.
(2) If f and g share 4 values a1 , . . . , a4 , the arguments in the proof of Theo-
rem 13.1 yield that
4
X
N (r, aj ) ≤ T (r, f ) + T (r, g) + O(1),
j=1
X4
1
T (r, f ) ≤ + o(1) N (r, aj ) for r ∈
/ Eg
2 j=1
and X4
1
T (r, g) ≤ + o(1) N (r, aj ) for r ∈
/ Eg .
2 j=1
as well as
4
X
N (r, aj ) ∼ 2T (r, f ) ∼ 2T (r, g) for r ∈
/ E.
j=1
So at least two of the shared values do not have this property, and we may assume
without loss of generality that this is the case for 0 and ∞; that is,
1
N r,
f N (r, f )
lim sup > 0 and lim sup >0
r→∞ T (r, f ) r→∞ T (r, f )
r∈E
/ r∈E
/
with α = 1/c, β = 1/(1 − c) and γ = 1/(c(c − 1)). The lemma on the logarithmic
derivative (Theorem 11.5) implies that
f0
m = o(T (r, f )) for r ∈
/ E.
f (f − 1)(f − c)
Analogously we have
g0
m = o(T (r, g)) for r ∈
/ E.
g(g − 1)(g − c)
As remarked before Theorem 13.2, we have
Next we show that H is entire. In order to do this we note that poles of H can
only occur at points where f and g take one of the shared values 0, 1, c and ∞.
Using the partial fraction decomposition mentioned above we have
0
g0
0
g0
0
g0
f f f
H=α − +β − +γ − ,
f g f −1 g−1 f −c g−c
function f 0 /(f (f − 1)(f − c)) has a zero (of multiplicity 2p − 1) at z0 , and so does
g 0 /(g(g − 1)(g − c)). It follows that H has a zero at z0 . In particular, H is entire
as claimed.
Moreover, the argument shows that
1
N (r, f ) ≤ N r, .
H
Together with the first fundamental theorem we thus have
N (r, f ) ≤ T (r, H) + O(1).
Since H is entire we have N (r, H) = 0 and thus the estimate for m(r, H) obtained
above yields
T (r, H) = m(r, H) = o(T (r, f )) for r ∈
/ E.
Combining the last two equations we have
N (r, f ) = o(T (r, f )) for r ∈
/ E.
However, we assumed that
N (r, f )
lim sup > 0.
r→∞ T (r, f )
r∈E
/
Remark. The second example after Definition 13.1 shows that the conclusion of
Theorem 13.2 need not hold if f and g share the values without multiplicities.
However, a result of Gundersen (1983) says that it suffices to assume that 2 of the
4 values are shared counting multiplicity. It is an open question whether this also
holds if only one of the 4 values is shared counting multiplicity.
Theorem 14.1. Let f be an entire function which is not if the form f (z) = z + c
for some c ∈ C. Then f ◦ f has a fixed point; that is, there exists z0 ∈ C with
f (f (z0 )) = z0 .
Proof. Suppose that f ◦ f has no fixed point. Then f has no fixed point. Thus
f (z) 6= z for all z ∈ C and this implies that f (f (z)) 6= f (z) for all z ∈ C. We
deduce that the function h defined by
f (f (z)) − z
h(z) =
f (z) − z
is entire and satisfies f (z) 6= 0 and f (z) 6= 1 for all z ∈ C. By Picard’s theorem, h
is constant, say h(z) = c for all z ∈ C, with c ∈ C \{0, 1}. Hence
and thus
(f 0 (f (z)) − c)f 0 (z) = 1 − c.
Since c 6= 1 this yields that f 0 (z) 6= 0 for all z ∈ C. This implies that f 0 (f (z)) 6= 0
for all z ∈ C. Moreover, the last equation implies that f 0 (f (z)) 6= c for all z ∈ C.
Picard’s theorem implies that f 0 ◦ f is constant. Hence f 0 is constant. (Note that
the assumption that f has no fixed point implies that f is non-constant.) Thus
f has the form f (z) = az + b with a, b ∈ C. The hypothesis that f has no fixed
points no implies that a = 1.
However, Nevanlinna theory can be used to show that h has “few” 1-points;
that is, N (r, 1/h) = o(T (r, h)). The proof can then be completed as before.
We will actually prove a more general result involving the iterates f n defined
by f 1 = f and f n = f ◦ f n−1 for n ≥ 2.
T (r, f ◦ g)
lim
r→∞
=∞
r∈E
/
T (r, g)
Proof of Theorem 14.2. Let b ∈ C be such that f has infinitely many b-points
a1 , a2 , a3 , . . . . Then g(z) = aj implies that f (g(z)) = b. For q ∈ N and r ≥ 1 we
thus obtain X q
1 1
N r, ≥ N r, .
f ◦g−b j=1
g − a j
(In fact, this also holds with N (r, ·) replaced by N (r, ·), but we will not need this.)
By the second fundamental theorem we have
q
X 1
N r, ≥ (q − 2 − o(1))T (r, g) for r ∈
/ E,
j=1
g − aj
so that
T (r, f ◦ g)
lim inf
r→∞
≥ q − 2.
r∈E
/
T (r, g)
Since q ∈ N was arbitrary, the conclusion follows.
Remark. The conclusion of Theorem 14.2 and the Corollary actually holds without
any exceptional set E.
14 AN APPLICATION TO ITERATION THEORY 71
Proof. Suppose that there are two integers m, n ∈ N such that f has only finitely
many periodic points of periods m and n. Without loss of generality we may assume
that m < n. Let a1 , . . . , ap be the periodic points of period m and b1 , . . . , bq be the
periodic points of period n.
We put ` = n − m and consider the function h defined by
f n (z) − z
h(z) = .
f ` (z) − z
As in the proof of Theorem 14.1 we consider the zeros, 1-points and poles of h.
First we note that by the first fundamental theorem and Theorem 14.2 we have
1
N (r, h) ≤ N r, ` ≤ T (r, f ` (z) − z) + O(1)
f (z) − z
≤ T (r, f ` ) + O(log r)
= o(T (r, f n ))
as r → ∞, r ∈ / E.
Let now z0 ∈ C with h(z0 ) = 1. Then f ` (z0 ) = f n (z0 ) = f m (f ` (z0 )) so that
f ` (z0 ) is a fixed point of f m . Thus f ` (z0 ) = aj for some j ∈ {1, . . . , p} or there
exists k ∈ N with 1 ≤ k < m such that f k+` (z0 ) = f k (f ` (z0 )) = f ` (z0 ). This
implies that
p m−1
1 X 1 X 1
N r, ≤ N r, `
+ N r, k+` .
h−1 j=1
f (z) − aj k=1
f (z) − f ` (z)
72 14 AN APPLICATION TO ITERATION THEORY
as r → ∞, r ∈
/ E, for some subset E of [0, ∞) of finite measure.
Proof. We consider the meromorphic function h defined by
a3 (z) − a2 (z) f (z) − a1 (z)
h(z) = · .
a3 (z) − a1 (z) f (z) − a2 (z)
The hypothesis that T (r, aj ) = o(T (r, f )) and the first fundamental theorem yield
that
1 1 1
N r, ≤ N r, + N r, + N (r, a1 ) + N (r, a2 ) + N (r, a3 )
h f − a1 a3 − a1
1
≤ N r, + o(T (r, f )).
f − a1
15 DIFFERENTIAL EQUATIONS 73
T (r, h) ∼ T (r, f ).
The conclusion now follows from the second fundamental theorem applied to h.
Remark. Theorem 14.4 has been extended to the case of q functions a1 , . . . , aq
satisfying T (r, aj ) = o(T (r, f )) by Yamanoi in 2004: given ε > 0 we have
q
X 1
(q − 2 − ε)T (r, f ) ≤ N r,
j=1
f − aj
for r outside a set of finite measure. With N (r, ·) replaced by N (r, ·) this had been
proved before by Steinmetz in 1986.
for r ∈
/ E. Together with the assumption that (ii) holds for k we deduce that (ii)
also holds with k replaced by k + 1. Together with (i) this implies that (iii) also
holds with k replaced by k + 1.
As a simple but instructive example of how Nevanlinna theory is applied to
differential equations is the following result.
Theorem 15.2. Let f be meromorphic and P a polynomial. If f satisfies the
differential equation f 0 = P (f ), then deg(P ) ≤ 2.
Proof. With d = deg(P ) we deduce from Theorems 5.3 and 15.1 that
as r → ∞, r ∈
/ E. This implies that d ≤ 2.
Remark. A solution of f 0 = 1 + f 2 is given by f (z) = tan z.
Of course, the differential equation f 0 = P (f ) can be integrated directly:
Z z Z z Z f (z)
f 0 (ζ) dw
ζ= ζ= .
z0 z0 P (f (ζ)) f (z0 ) P (w)
Proof. Choose r0 > 0 such that all zeros and poles of the coefficients a0 , . . . , ad are
contained in D(0, r0 ). Then a pole z0 of f of order m is a pole of P (z, f (z)) of
order dm. This implies that there exists M ∈ N such that
for r ≥ r0 . Hence
aj (z) ∼ cj · z mj
as |z| → ∞. Hence
aj (z) cj
≤ (1 + o(1)) · |z|mj −md
ad (z) cd
as |z| → ∞. Thus there exist R > 0 and L > 0 such that
aj (z) 1
≤ |z|L and ≤ |ad (z)| ≤ |z|L
ad (z) |z|d
so that Z
1
log+ |f (reiθ )|dθ ≤ (L + 1) log r
2π [0,2π]\J
and
Z
1
log+ |P (reiθ , f (reiθ ))|dθ ≤ (2L + (L + 1)d) log r + log(d + 1).
2π [0,2π]\J
Together with the estimate for N (r, ·) obtained already this yields
Let
d
X
P (z, w) = ak (z)wk ,
k=0
factor (with respect to w), then d := max{m, n} is called the degree of R with
respect to w and denoted by degw (R).
In elementary number theory, Bezout’s lemma says that the greatest common
divisor gcd(p, q) of two integers p and q has a representation ap + bq = gcd(p, q)
with integers a and b. (The proof is obtained from the Euclidean algorithm.) In
particular, if p and q are coprime (German: teilerfremd), then there exist a and b
such that ap + bq = 1. This result actually holds not only for the ring of integers,
but in any principal ideal domain (German: Hauptidealring). In particular, this is
true for the ring of polynomials in one variable over some field. We shall use this
for the field of rational functions.
Lemma 15.1. Let P (z, w) and Q(z, w) be two coprime polynomials in w with
rational functions in z as coefficients. Then there exist polynomials A(z, w) and
B(z, w) in w with rational functions in z as coefficients such that
A(z, w)P (z, w) + B(z, w)Q(z, w) = 1.
Theorem 15.4. Let f be meromorphic and let R(z, w) be rational in both variables.
Then
T (r, R(z, f (z))) = degw (R) · T (r, f ) + O(log r).
Proof. We write
m n
P (z, w) X
k
X
R(z, w) = with P (z, w) = ak (z)w and Q(z, w) = bk (z)wk
Q(z, w) k=0 k=0
Pd
b (z)wk
= k=0 k
Pd−1 ad (z)
k=0 ak (z) − bk (z) wk
bd (z)
78 15 DIFFERENTIAL EQUATIONS
where
m−n P`
X
k dk (z)f (z)k
α(z) = ck (z)f (z) and β(z) = Pk=0
n k
.
k=0 k=0 bk (z)f (z)
By Theorem 15.3 we have T (r, α) = (m − n)T (r, f ) + O(log r) and by the induction
hypothesis we have T (r, β) ≤ n · T (r, f ) + O(log r). Combining the two estimates
we obtain the desired upper bound for T (r, R(z, f (z))).
Now we turn to the proof of the lower bound for T (r, R(z, f (z))); that is, we
prove that
T (r, R(z, f (z)) ≥ degw (R) · T (r, f ) + O(log r).
Again we may assume that m > n. Since the case n = 0 is covered by Theorem 15.3
we may thus assume that 1 ≤ n < m. Choose A and B according to Lemma 15.1
so that
A(z, w)P (z, w) + B(z, w)Q(z, w) = 1.
With s = degw (A) and t = degw (B) we have s + m = t + n. Since m > n this
yields t > s. We put C = A/B and write A∗ (z) = A(z, f (z)), B ∗ (z, f (z)), etc. In
particular, R∗ (z) = R(z, f (z)). Then
∗ 1 A∗ Q∗ 1
C + ∗ = ∗+ ∗ = ∗ ∗
R B P B Q
and hence
∗ 1 1
T r, C + ∗ = T r, ∗ ∗
R B Q
= T (r, B ∗ Q∗ ) + O(1)
= degw (BQ) log r + O(log r)
= (t + m)T (r, f ) + O(log r)
15 DIFFERENTIAL EQUATIONS 79
by the upper bound proved already. Combining the last two estimates we have
1
(t + m)T (r, f ) = T r, C + ∗ + O(log r) ≤ t · T (r, f ) + T (r, R∗ ) + O(log r)
∗
R
and hence
degw (R) · T (r, f ) = m · T (r, f ) ≤ T (r, R∗ ) + O(log r) = T (r, R(z, f (z))) + O(log r)
as claimed.
Theorem 15.5 (Malmquist’s theorem). Let f be a transcendental meromorphic
function and let R(z, w) be rational in both variables. Suppose that f satisfies the
differential equation
f 0 (z) = R(z, f (z)).
Then R is a polynomial in w and degw (R) ≤ 2.
Proof. By Theorem 15.1 we have
for r ∈
/ E while Theorem 15.4 says that
u0 = a1 u + a0 v
v 0 = −a2 u
Standard techniques show that for every linear system of differential equations
with entire coefficients the solutions are also entire. Thus all solutions u and v
of the above system are entire. The quotient f = u/v of two solutions is thus
meromorphic, and we have
0
u 0 u0 v − uv 0 (a1 u + a0 v)v − u(−a2 u)
f = = 2
= = a0 + a1 f + a2 f 2 .
v v v2
The following result is proved in the same way as the previous theorem.
Theorem 15.6 (Malmquist-Yosida theorem). Let f be a transcendental meromor-
phic function, R(z, w) be rational in both variables and n ∈ N. Suppose that f
satisfies the differential equation
There the τj are distinct constants and a, b and c are rational functions.
Another generalization of Theorem 15.6 due to Eremenko says that if
as r → ∞, r ∈
/ E, for some subset E of [0, ∞) of finite measure.
as r → ∞, r ∈
/ E, with
1
N1 (r, g) = N r, 0 + 2N (r, g) − N (r, g 0 ).
g
Here and in the following E always denotes a subset of [0, ∞) of finite measure.
The first fundamental theorem implies that
1 1
2T (r, g) − N1 (r, g) = m(r, g) + N (r, g) + m r, + N r,
g−1 g−1
1
− N r, 0 + 2N (r, g) − N (r, g 0 ) + O(1)
g
1 1 1
= m(r, g) + m r, + N r, − N r, 0
g−1 g−1 g
0
+ N (r, g ) − N (r, g) + O(1).
Since
1 1 1
N r, − N r, 0 = N r, − N0 (r)
g−1 g g−1
and
N (r, g 0 ) − N (r, g) = N (r, g)
this yields
1 1
2T (r, g)−N1 (r, g) = m(r, g)+m r, +N r, −N0 (r)+N (r, g)+O(1).
g−1 g−1
82 16 EXCEPTIONAL VALUES OF DERIVATIVES
Together with the inequality obtained at the beginning of the proof this yields that
1 1
m r, ≤ N r, − N0 (r) + N (r, g) + o(T (r, g))
g g−1
as r → ∞, r ∈ / E.
By Theorem 15.1 we have T (r, g) = T (r, f (k) ) ≤ (k + 1)T (r, f ) as r → ∞,
r∈/ E. This implies that the o(T (r, g))-terms occurring above can be replaced by
o(T (r, f )). The lemma on the logarithmic derivative implies that
f (k)
1 1
m r, = m r, (k) · (k−1)
f (k−1) f f
f (k)
1
≤ m r, (k) + m r, (k−1)
f f
1
≤ m r, (k) + o(T (r, f (k−1) )
f
1
= m r, (k) + o(T (r, f )
f
as r → ∞, r ∈
/ E. Thus
1 1
m r, ≤ N r, + N (r, g) − N0 (r) + o(T (r, g))
f g−1
and hence
1 1
T (r, f ) = N r, + m r,
f f
1 1
≤ N r, + N r, + N (r, g) − N0 (r) + o(T (r, g))
f g−1
1 1
= N r, + N r, (k) + N (r, f ) − N0 (r) + o(T (r, f ))
f f −1
as r → ∞, r ∈
/ E.
One may replace the 1-points of f (k) by the a-points of f (k) for any a ∈ C \{0}.
This follows by considering f /a instead of f . However, one may not replace the
1-points of f (k) by the zeros of f (k) . In fact, f (z) = ez has no zeros and poles, and
f (k) (z) 6= 0 for all z ∈ C and all k ∈ N. And for f (z) = ez + 1 we have f (z) 6= 1
and f (k) (z) 6= 0 for all z ∈ C and all k ∈ N.
We will now prove a result of Hayman which implies that in order to conclude
that a meromorphic function f is rational it suffices to assume that f has only
finitely many zeros and f (k) has only finitely many 1-points for some k ∈ N. So no
hypothesis on the poles is required.
Theorem 16.2. Let f be a transcendental meromorphic function and k ∈ N. Then
1 1 2 1
T (r, f ) ≤ 2 + N r, + 2+ N r, (k) + o(T (r, f ))
k f k f −1
as r → ∞, r ∈
/ E.
The idea is to estimate the term N (r, f ) occurring on the right hand side of
the inequality in Theorem 16.1 in terms of the other counting functions occuring
there. In order to do so we write
where N s (r, f ) denotes the counting function of the simple poles and N m (r, f )
denotes the counting function of the multiple poles, counted without multiplicity.
First we prove the following lemma.
Lemma 16.1. Let f be a transcendental meromorphic function and let N0 (r) be
as in Theorem 16.1. Then
1
kN s (r, f ) ≤ N m (r, f ) + N r, (k) + N0 (r) + o(T (r, f ))
f −1
as r → ∞, r ∈
/ E.
Proof. Let g = f (k) be as in the proof of Theorem 16.1 and put
k+1
g 0 (z)k+1 f (k+1) (z)
h(z) = = k+2
.
(1 − g(z))k+2 (1 − f (k) (z))
Let z0 be a simple pole of f . Then f (k) and hence 1 − f (k) have a pole of order
k + 1 at z0 . In fact,
a a
1 − f (k) (z) = 1 + O (z − z0 )k+1
k+1
+ O(1) = k+1
(z − z0 ) (z − z0 )
as z → z0 for some non-zero constant a (depending on z0 ). It follows that
(k + 1)a (k + 1)a
f (k+1) (z) = k+2
+ O(1) = 1 + O (z − z0 )
(z − z0 )k+2 (z − z0 )k+2
84 16 EXCEPTIONAL VALUES OF DERIVATIVES
as z → z0 . Hence
as r → ∞, r ∈ / E.
Next we note that if z0 is a pole of f of multiplicity p ≥ 2, then h has a zero of
multiplicity (p + k)(k + 2) − (p + k + 1)(k + 1) = p − 1 at z0 . Further zeros of h
can arise only from zeros of f (k+1) which are not 1-points of f (k) so that
1
N r, ≤ N m (r, f ) + N0 (r).
h
References
[1] Anatoly A. Goldberg, Iossif V. Ostrovskii: Value distribution of meromorphic
functions. American Mathematical Society, Providence, RI, 2008.
[3] Gerhard Jank, Lutz Volkmann: Einführung in die Theorie der ganzen und mero-
morphen Funktionen mit Anwendungen auf Differentialgleichungen. Birkhäuser
Verlag, Basel, 1985.
The above is only a small selection of the many books on the theory of entire and
meromorphic functions.