Course
Course
Course
Contents
1 Basic complex analysis . . . . . . . . . . . . . . . . . . . . . . 1
2 The simply-connected Riemann surfaces . . . . . . . . . . . . 27
3 Entire and meromorphic functions . . . . . . . . . . . . . . . 39
4 Conformal mapping . . . . . . . . . . . . . . . . . . . . . . . 56
5 Elliptic functions and elliptic curves . . . . . . . . . . . . . . 79
Forward
4. Lie groups, discrete subgroups and homogeneous spaces (e.g. H/ SL2 (Z);
1
Euclidean plane by z = (x, y) = x + iy. There are two square-roots of −1
in C; the number i is the one with positive imaginary part.
An important role is played by the Galois involution z 7→ z. We define
2
|z| = N (z) =√zz = x2 + y 2 . (Compare the case of a real quadratic field,
where N (a + b d) = a2 − db2 gives an indefinite form.) Compatibility of |z|
with the Euclidean metric justifies the identification of C and R2 . We also
see that z is a field: 1/z = z/|z|.
It is also convenient to describe complex numbers by polar coordinates
2
prime polynomials, with the denominator not identically zero) determine
rational maps f : C → C. The rational functions C(z) are the same as the
field of fractions for the domain C[z]. We set f (z) = ∞ if q(z) = 0; these
points are called the poles of f .
Analytic functions. Let U be an open set in C and f : U → C a function.
We say f is analytic if
f (z + t) − f (z)
f ′ (z) = lim
t→0 t
exists for all z ∈ U . It is crucial here that t approaches zero through arbitrary
values in C. Remarkably, this condition implies that f is a smooth (C ∞ )
function. For example, polynomials are analytic, as are rational functions
away from their poles.
Note that any real linear function φ : C → C has the form φ(v) = av+bv.
The condition of analytic says that Dfz (v) = f ′ (z)v; in other words, the v
part is absent.
To make this point systematically, for a general C 1 function F : U → C
we define
dF 1 dF 1 dF dF 1 dF 1 dF
= + and = − ·
dz 2 dx i dy dz 2 dx i dy
We then have
dF DF
DFz (v) = v+ v.
dz dz
We can also write complex-valued 1-form dF as
dF dF
dF = ∂F + ∂F = dz + dz
dz dz
Thus F is analytic iff ∂F = 0; these are the Cauchy-Riemann equations.
We note that (d/dz)z n = nz n−1 ; a polynomial p(z, z) behaves as if these
variables are independent.
Sources of analytic functions.
Algebraic functions. Beyond the rational and polynomial functions, √ the
analytic functions include algebraic functions suchP that f (z) = z 2 + 1.
A general algebraic function f (z) satisfies P (f ) = N n
0 an (z)f (z) = 0 for
some rational functions an (z); these arise, at least formally, when one forms
algebraic extension of C(z). Such functions are generally multivalued, so we
must choose a particular branch to obtain an analytic function.
Differential equations. Analytic functions also arise when one solves
differential equations. Even equations with constant coefficients, like y ′′ +y =
3
0, can give rise to transcendental functionsRsuch as sin(z), cos(z) and ez . A
special case of course is integration. While (x2 + ax + b) 1/2
R 3 dx can be 1/2given
explicitly in terms of trigonometric functions, already (x + ax + b) dx
leads one into elliptic functions; and higher degree polynomials lead one to
hyperelliptic surfaces of higher genus.
Power series. Analytic
P functions can be given concretely, locally, by
power series such as an z n . Conversely,
P suitable coefficients determine
analytic functions; for example, ez = z n /n!.
Riemann surfaces and automorphy. A third natural source of complex
analytic functions is functions that satisfy invariant properties such as f (z +
λ) = f (z) for all λ ∈ Λ, a lattice in C; or f (g(z)) = f (z) for all g ∈ Γ ⊂
Aut(H).
The elliptic modular functions f : H → C with have the property that
f (z)= f (z + 1) = f (1/z), and hence f ((az + b)/(cz + d)) = f (z) for all
a b ∈ SL (Z).
c d 2
Geometric function theory. Finally a complex analytic function can be
specified by a domain U ⊂ C; we will see that for simply-connected domains
(other than C itself), there is an essentially unique analytic homeomorphism
f : ∆ → U . When U tiles H or C, this is related to automorphic functions;
and when ∂U consists of lines or circular arcs, one can also give a differential
equation for f .
Example. We can also define analytic functions by taking limits of poly-
nomials or other known functions. For example, consider the formula:
ez = lim(1 + z/n)n .
The triangle with vertices 0, 1 and 1 + iθ has a hypotenuse of length 1 +
O(1/n2 ) and an angle at 0 of θ + O(1/n2 ). Thus one finds geometrically that
zn = (1 + iθ/n)n satisfies |zn | → 1 and arg zn → θ; in other words,
eiθ = cos θ + i sin θ.
In particular, eπi = −1 (Euler).
Exponential and trigonometric functions in C. Here are some useful
facts about these familiar functions when extended to C:
| exp(z)| = exp Re z
cos(iz) = cosh(z)
sin(iz) = i sinh(z)
cos(x + iy) = cos(x) cosh(y) − i sin(x) sinh(y)
sin(x + iy) = sin(x) cosh(y) + i cos(x) sinh(y).
4
In particular, the apparent boundedness of sin(z) and cos(z) fails badly as
we move away from the real axis, while |ez | is actually very small in the
halfplane Re z ≪ 0.
Complex integration; Cauchy’s theorem. Now suppose U is a com-
pact, connected, smoothly bounded region in C, f : U → C is continuous
and f : U → C is analytic. We then have:
Theorem
R 1.1 (Cauchy) For any analytic function f : U → C, we have
∂U f (z) dz = 0.
Note that the former depends on a choice of orientation of γ, while the latter
does not.
Proof of Cauchy’s formula: (i) observe that d(f dz) = (∂f )dz dz = 0
and apply Stokes’ theorem. (ii — Goursat). Cut the region U into small
Rsquares, observe that on these squares f (z) ≈ az + b, and use the fact that
∂U (az + b) dz = 0.
5
Aside: distributions. The first proof implicitly assumes f is C 1 , while
the second does not. (To see where C 1 is used, R suppose α = u dx + v dy
and dα = 0 on a square S. In the proof that ∂S α = 0, we integrate vdy
over the vertical sides of S and observe that this is the same as integrating
dv/dx dx dy over the square. But if α is not C 1 , we don’t know that dv/dx
is integrable.)
Notes. More on the Cauchy–Riemann equations and with minimal smooth-
ness assumptions can be found in [GM].
More generally,
R we say a distribution (e.g. an L1 function f ) is (weakly)
analytic if f ∂φ = 0 for every φ ∈ Cc∞ (U ). By convolution with a smooth
function (a mollifier), any weakly analytic function is a limit of C ∞ analytic
functions. We will see below that uniform limits of C ∞ analytic functions
are C ∞ , so even weakly analytic functions are actually smooth.
Cauchy’s integral formula: Differentiability and power series. Because
of Cauchy’s theorem, only one integral has to be explicitly evaluated in
complex analysis (hence the forgetability of the definition of the integral).
Namely, setting γ(t) = eit we find, for any r > 0,
Z
1
dz = 2πi.
S (r) z
1
6
P
In particular, ak (z − p)k has radius of convergence at least R, since
lim sup |ak |1/k < 1/R. This suggests that f is represented by its power
series, and indeed this is the case:
P
Theorem 1.3 If f is analytic on B(p, R), then f (z) = ak (z − p)k on this
ball.
Proof. We can reduce to the case z ∈ B(0, 1). Then for w ∈ S 1 and fixed
z with |z| < 1, we have
1 1
= 1 + (z/w) + (z/w)2 + · · · ,
w−z w
converging uniformly on the circle |w| = 1. We then have:
Z Z
1 f (w) dw X k 1 f (w) dw X
f (z) = = z = ak z k
2πi S 1 w − z 2πi S 1 wk+1
as desired.
Corollary 1.4 An analytic function has at least one singularity on its circle
of convergence.
f (z) = (z + z 2 )f (z) + 1
Theorem 1.5 A power series represents a rational function iff its coeffi-
cients satisfy a recurrence relation.
7
Aside: Pisot numbers. The golden ratio is an example of a Pisot number;
it has the property that d(γ n , Z) → 0 as n → ∞. It is an unsolved problem
to show that if α > 1 satisfies d(αn , Z)
P→ 0,i then α is an algebraic number.
Kronecker’s theorem asserts that ai z is a rational function iff deter-
minants of the matrices ai,i+j , 0 ≤ i, j ≤ n are zero for all n sufficiently
large [Sa, §I.3]
Question: why are 10:09 and 8:18 such pleasant times? [Mon].
P
Isolation of zeros. If f (z) = an (z − p)n vanishes at p but is not
identically zero, then we can factor out the leading term and write:
where g is analytic and g(p) 6= 0. This is the simplest case of the Weier-
strass preparation theorem: it shows germs of analytic functions behave like
polynomials times units (invertible functions).
In particular, we find:
Theorem 1.9 (The mean-value formula) The value of f (p) is the av-
erage of f (z) over S 1 (p, r).
8
Corollary 1.10 (The Maximum Principle) A nonconstant analytic func-
tion does not achieve its maximum in U .
9
Parseval’s theorem. The power series of an analytic function on the ball
B(0, R) also contains
P information about its L2 -norm on the circle |z| = R:
n
namely if f (z) = an z , then we have:
X Z
2 2n 1
|an | R = |f (z)|2 dθ.
2π |z|=R
This comes from the fact that the functions z n are orthogonal in L2 (S 1 ).
It also gives another important perspective on holomorphic functions: they
are the elements in L2 (S 1 ) with positive Fourier coefficients, and hence give
a half–dimensional subspace of this infinite–dimensional space.
Compactness of bounded functions. Cauchy’s bound on a disk also
implies that if f is small, then f ′ is also small, at least if we are not too near
the edge of U .
P n
Laurent series. Using again the basic series 1/(1−z) = z and Cauchy’s
formula over the two boundary components of the annulus A(r, R) = {z :
r < |z| < R}, we find:
10
Theorem 1.19 If f (z) is analytic on A(r, R), then in this region we have
∞
X
f (z) = an z n .
n=−∞
The positive terms converge for |z| < R, and the negative terms converge
for |z| > r.
Corollary 1.20 An analytic function on the annulus r < |z| < R can be
expressed as the sum of a function analytic on |z| < R and a function
analytic on |z| > r.
PU −p, with p n∈
Isolated singularities. As a special case, if f is analytic on
U , then near p we have a Laurent series expansion f (z) = ∞ −∞ an (z − p) .
We say f has an isolated singularity at p.
We write ord(f, p) = n if an 6= 0 but ai = 0 for all i < 0. The values
n = −∞ and +∞ are also allowed. If n ≥ 0, the apparent singularity at p
is removable and f has a zero of order n at p. If −∞ < n < 0, we say f has
a pole of order −n at p. In either of these cases, we can write
near p. The germs of functions at p with finite Laurent tails form a local field,
with ord(f, p) as its discrete valuation. (Compare Qp , where vp (pn a/b) = n.)
If ord(f, p) = −∞ we say f has an essential singularity at p. (Example:
f (z) = sin(−1/z) at z = 0.)
The residue. A critical role is played by the residue of f (z) at p, defined
by Res(f, p) = a−1 . It satisfies
Z
f (z) dz = 2πi Res(f, p)
γ
for any small loop encircling the point p in U . Thus the residue is intrin-
sically an invariant of the 1-form f (z) dz, not the function f (z). (If we
regard f (z) dz as a 1-form, then its residue is invariant under change of
coordinates.)
11
Theorem 1.21 (The residue theorem) Let f : U → C be a function
which is analytic apart from a finite set of isolated singularities. We then
have: Z X
f (z) dz = 2πi Res(f, p).
∂U p∈U
Example. Consider p(z) = z 5 +14z+1. Then all its zeros are inside |z| < 2,
since |z|5 = 32 > |14z + 1| when |z| = 2; but only one inside |x| < 3/2, since
|z 5 + 1| ≤ 1 + (3/2)5 < 9 < |14z| on |z| = 3/2. (Intuitively, the zeros of p(z)
are close to the zeros of z 5 + 14z which are z = 0 and otherwise 4 points on
|z| = 141/4 ≈ 1.93.
Corollary 1.24 Let p(z) = z d + a1 z d−1 + . . . + ad , and suppose |ai | < Ri /d.
Then p(z) has d zeros inside the disk |z| < R.
Proof. Write p(z) = f (z) + g(z) with f (z) = z d , and let U = B(0, R); then
on ∂U , we have |f | = Rd and |g| < Rd ; now apply Rouché’s Theorem.
12
Corollary 1.25 A nonconstant analytic function is an open mapping.
This is nothing more than the winding number of f (∂U ) around p. Ob-
serve that N (f, p) is a locally constant function on C − f (∂U ), zero on the
noncompact component. Summing up:
Theorem 1.27 For any p ∈ C−f (∂U ), the number of solutions to f (z) = p
in U is the same as the winding number of f (∂U ) around p.
Aside: the smooth case. These results also hold for smooth mappings f
once one finds a way to count the number of solutions to f (z) = p correctly.
(Some may count negatively, and the zeros are only isolated for generic
values of p.)
13
Aside: Linking numbers and intersection multiplicities of curves
in C2 . Counting the number of zeros of y = f (x) at x = 0 is the same
as counting the multiplicity of intersection between the curves y = 0 and
y = f (x) in C2 , at (0, 0).
In general, to an analytic curve C defined by F (x, y) = 0 passing through
(0, 0) ∈ C2 , we can associate a knot by taking the intersection of C with the
boundary S 3 (r) of a small ball centered at the origin.
A pair of distinct, irreducible, curves C1 and C2 passing through (0, 0)
(say defined by fi (x, y) = 0, i = 1, 2), have a multiplicity of intersection
m(C1 , C2 ). This can be defined geometrically by intersection with a small
sphere S 3 (r) in C2 centered at (0, 0). Then we get a pair of knots, and their
linking number is the same as this multiplicity.
Examples. It is often simpler the use S 3 = ∂∆2 instead of |x|2 + |y|2 = 1.
Then S 3 is the union of two solid tori, ∆ × S 1 and S 1 × ∆. The axes x = 0
and y = 0 are the core curves of these tori; they have linking number one.
The line y = x is a (1, 1) curve on S 1 × S 1 ; more generally, for gcd(a, b) = 1,
y a = xb is an (a, b) curve, parameterized by (x, y) = eita , eitb ). In particular,
the cusp y 2 = x3 meets S 3 in a trefoil knot. It links x = 0 twice and y = 0
three times.
Problem. Show that the figure-eight knot cannot arise from an analytic
curve.
Multivalued functions. It is useful to have a general discussion of the
sometimes confusing notion of ‘branch cuts’ and ‘multivalued functions’.
Here are 2 typical results.
14
Corollary 1.30 If an = b then there is a unique analytic function R : U →
C such that R(b) = a and R(z)n = z for all z ∈ U .
log(1 + z) = z − z 2 /2 + z 3 /3 − · · ·
and, by the binomial theorem,
15
Corollary 1.33 (Weierstrass-Casorati) If f (z) has an essential singu-
larity at p, then there exist zn → p such that f (zn ) is dense in C.
(Of course we can also compute this integral by factoring Q(x) and using
partial fractions and trig substitutions.)
Example. Where does π come from? It emerges naturally from rational
functions by integration — i.e. it is a period. Namely, we have
Z ∞
dx
2
= 2πi Res(1/(1 + z 2 ), i) = 2πi(−i/2) = π.
−∞ 1 + x
R
Of course this can also be done using the fact that dx/(1+x2 ) = tan−1 (x).
More magically, for f (z) = 1/(1 + z 4 ) we find:
Z ∞ √ √
dx π
4
= 2πi(Res(f, (1 + i)/ 2) + Res(f, (1 + i)/ 2) = √ ·
−∞ 1 + x 2
Both are obtain by closing a large interval [−R, R] with a circular arc in the
upper halfplane, and then taking the limit as R → ∞.
16
We can even compute the general case, f (z) = 1/(1 + z n ), with n even.
For this let ζk = exp(2πi/k), so f (ζ2n ) = 0. Let P be the union of the paths
[0, ∞)ζn and [0, ∞), oriented so P moves positively on the real axis. We can
then integrate over the boundary of this pie-slice to obtain:
Z ∞ Z
dx n−1
(1 − ζn ) n
= f (z) dz = 2πi Res(f, ζ2n ) = 2πi/(nζ2n ),
0 1 + x P
which gives Z ∞
dx 2πi π/n
n
= −1 +1 = ·
0 1+x n(−ζ2n + ζ2n ) sin π/n
Here we have used the fact that ζ2nn = −1. Note that the integral tends to
17
and Res(f, i) = ia /(2i), Res(f, −i) = (−i)a /(−2i) (since xa /(1 + x2 ) =
xa /(x − i)(x + i)). Thus, if we let ia = ω = exp(πia/2), we have
π(ia − (−i)a ) ω − ω3 π π
I(a) = a
= π 4
= −1
= ·
(1 − 1 ) 1−ω ω+ω 2 cos(πa/2)
For example, when a = 1/3 we get
√
I(a) = π/(2 cos(π/6)) = π/ 3.
Residues and infinite sums. The periodic function f (z) = π cot(πz) has
the following convenient properties: (i) It has residues 1 at all the integers;
and (ii) it remains bounded as Im z → ∞. From these facts we can deduce
some remarkable properties: by integrating over a large rectangle S(R), we
find for k ≥ 2 even,
Z X∞
1 f (z) dz k
0 = lim = Res(f (z)/z , 0) + 2 1/nk .
R→∞ 2πi S(R) zk
1
P
Thus we can evaluate the sum 1/n2 using the Laurent series
cos(z) 1 − z 2 /2! + z 4 /4! − · · ·
cot(z) = =
sin(z) z(1 − z 2 /3! + z 4 /5! − · · · )
= z −1 (1 − z 2 /2! + z 4 /4! − · · · )(1 + z 2 /6 + 7z 4 /360 + · · · )
= z −1 − z/3 − z 3 /45 − · · ·
(using the fact thatP(1/(3!)2 − 1/5! = 7/360). This shows Res(f (z)/z 2 , 0) =
−π 2 /3 and hence 1/n2 = π 2 /6. Similarly, 2ζ(2k) = − Res(f (z)/z 2k , 0).
For example, this justifies ζ(0) = 1 + 1 + 1 + · · · = −1/2.
Little is known about ζ(2k + 1). Apéry showed that ζ(3) is irrational,
but it is believed to be transcendental.
P
We note that ζ(s) = 1/ns is analytic for Re s > 1 and extends an-
alytically to C − {1} (with a simple pole at s = 1). Q In particular ζ(0) is
well-defined. Because of the factorization ζ(z) = (1 − 1/ps )−1 , the be-
havior of the zeta function is closely related to the distribution of prime
numbers. The famous Riemann hypothesis states that any zero of ζ(s) with
0 < Re s < 1 satisfies Re s = 1/2. It implies a sharp form of the prime
number theorem, π(x) = x/ log x + O(x1/2+ǫ ).
The zeta function also has trivial zeros at s = −2, −4, −6, . . ..
R
Hardy’s paper on sin(x)/x dx. We claim
Z ∞
sin x dx π
I= = ·
0 x 2
18
Note that this integral is improper, i.e. it does not converge absolutely.
Also, the function f (z) = sin(z)/z has no poles — so how can we apply the
residue calculus?
The trick is to observe that
Z
eix dx
−2iI = lim ·
r→0 r<|x|<1/r x
We now use the fact that |eix+iy | ≤ e−y to close the path in the upper
halfplane, and conclude that
Z
eiz dz
2iI = lim ·
r→0 S 1 (r)
+
z
19
4. Thus any C 2 harmonic function is actually infinitely differentiable.
-4 -2 2 4
-1
-2
-3
20
Harmonic extension. Here is one of the central existence theorems for
harmonic functions.
Theorem 1.34 There is a unique linear map P : C(S 1 ) → C(∆) such that
u = P (u)|S 1 and P (u) is harmonic on ∆.
Poisson kernel. The map P can be given explicit by the Poisson kernel.
For example, u(0) is just the average of u over S 1 . We can also say u(p) is
the expected value of u(z) under a random walk starting at p that exits the
disk at z.
To find the Poisson kernel explicitly, suppose we have a δ-mass at z = 1.
Then it should extend to a positive harmonic function u on ∆ which vanishes
along S 1 except at 1, and has u(0) = 0. In turn, u should be the real part
of an analytic function f : ∆ → C such that f (0) = 1 and Re f |S 1 = 0
and f has a pole at z = 1. Such a function is given simply by the Möbius
transformation f : ∆ → U = {z : Re z > 0}:
1+z
f (z) = ·
1−z
Convolving, we find the analytic function with Re f = u for a given u ∈
C(S 1 ) is given by Z
1
F (z) = f (z/t)u(t)|dt|,
2π S 1
and thus Z
1
u(r, α) = Pr (α − θ)u(θ) dθ,
2π S1
1 − |z|2 1 − r2
Pr (θ) = Re f (z) = 2
= ·
|1 − z| 1 − 2r cos θ + r 2
21
8
-3 -2 -1 0 1 2 3
and then replace z −n by z n to get its extension to the disk. This actually
works, and gives another approach to the Poisson kernel. P
Given f ∈ C(S 1 ), it is not true, in general, that its Fourier series an z n
converges for all z ∈ S 1 . However, it is true that this series converges in
the unit disk and defines a harmonic function there. As we have seen, this
harmonic function provides a continuous extension of f . This shows:
The result above shows the Fourier series of f is Abel summable to the
original function f .
Laplacian as a quadratic form, and physics. Suppose u, v ∈ Cc∞ (C)
– so u and v are smooth, real-valued functions vanishing outside a compact
set. Then, by integration by parts, we have
Z Z Z
h∇u, ∇vi = − hu, ∆vi = − hv, ∆ui.
∆ ∆ ∆
22
To see this using differential forms, note that:
Z Z Z
0= d(u ∗ dv) = (du)(∗dv) + u(d ∗ dv).
∆ ∆ ∆
In particular, we have
Z Z
|∇u|2 = − u∆u.
∆ ∆
1.5
1.0
0.5
0.0
-2 -1 0 1 2
23
It is also easy to argue that |p0 − pT | tends to be small when p0 is close to
S 1 , and hence u(p) is a continuous extension of u|S 1 .
The Poisson kernel (1/2π)Pr (θ)dθ gives the hitting density on S 1 for a
Brownian path starting at (r, 0).
Hyperbolic geometry interpretation. Alternatively, u(z) is the ex-
pected value of u(p) and the endpoint of a random hyperbolic geodesic ray
γ in ∆ with one vertex at z. (The angle of the ray in Tz ∆ is chosen at
random in S 1 .)
Example: fluid flow around a cylinder. We begin by noticing that
f (z) = z + 1/z gives a conformal map from the region U ⊂ H where |z| > 1
to H itself, sending the circular arc to [−2, 2]. Thus the level sets of Im f =
y(1 − 1/(x2 + y 2 )) describe fluid flow around a cylinder. Note that we are
modeling incompressible fluid flow with no rotation, i.e. we are assuming
the curl of the flow is zero. This insures the flow is given by the gradient of
a function.
Harmonic functions and the Schwarz reflection principle. Here is an
application of harmonic functions that will be repeatedly used in geometric
function theory.
Let U ⊂ C be a region invariant under z 7→ z. Suppose f : U → C is
continuous,
f (z) = f (z), (1.2)
and f is analytic on U+ = U ∩ H. We can then conclude that f is analytic
on U . In particular, f is analytic at each point of U ∩ R.
Here is a stronger statement:
24
Remark. One can replace z 7→ z with reflection through any circle, or
more generally with local reflection through a real-analytic arc.
Example. Suppose f (z) is analytic on H and f (z) → 0 along an interval
in R. Then f is identically zero. This extends the ‘isolated zero’ principle
to the boundary of a region.
Additional topics.
The Phragmen–Lindelöf Theorems. These theorems address the fol-
lowing question. Suppose f (z) is an analytic function on the horizontal
strip U = {x + iy : a < y < b}, and continuous on U . Can we assert that
supU |f | = sup∂U |f |?
The answer is no, in general. However, the answer is yes if f (x+ iy) does
not grow too rapidly as |x| → ∞. In fact, this is a property of harmonic
functions u(z). The point is that if we truncate the strip to a rectangle by
cutting along the lines where |x| = R, then the harmonic measure of the
ends (as seen from a fixed point z ∈ U ) tends to zero exponentially fast.
Thus if |u(x + iy)| = O(|x|n ) for some n, we get the desired control.
Runge’s theorem. Here is an interesting and perhaps surprising applica-
tion of Cauchy’s formula.
Let K ⊂ C be a compact set, let C(K) denote the Banach space of
continuous functions with the sup-norm, and let A(K), R(K) and P (K) ⊂
C(K) denote the closures of the analytic, rational and polynomial functions.
Note that all three of these subspaces are algebras.
Example. For K = S 1 we R have R(K) = A(K) = C(K) by Fourier series,
but P (K) 6= C(K), since S 1 p(z) dz = 0 for any polynomial. In particular,
1/z 6∈ P (S 1 ).
Remarkably, if we remove a small interval from the circle to obtain an
arc K = exp[0, 2π − ǫ], then 1/z can be approximated by polynomials on K.
Proof. For the first result, suppose f (z) is analytic on a smoothly bounded
neighborhood U of K. Then we can write
Z Z
1 f (t) dt
f (z) = = Fz (t) dt.
2πi ∂U t − z ∂U
Since d(z, ∂U ) ≥ d(K, ∂) > 0, the functions {Fz } range in a compact subset
of C(∂U ). Thus we can replace this integral with a finite sum at the cost
25
of an error that is small independent of z. But the terms f (ti )/(ti − z)
appearing in the sum are rational functions of z, so R(K) = A(K).
The second result is proved by pole–shifting. By what we have just
done, it suffices to show that fp (z) = 1/(z − p) ∈ P (K) for every p 6∈ K.
Let E ⊂ C − K denote the set of p for which this is true.
Clearly E contains all p which are sufficiently large, because then the
power series for fp (z) converges uniformly on K. Also E is closed by defini-
tion. To complete the proof, it suffices to show E is open.
The proof that E is open is by ‘pole shifting’. Suppose p ∈ E, q ∈ B(p, r)
and B(p, r) ∩ K = ∅. Note that fq (z) is analytic on C − B(p, r), and tends to
zero as |z| → ∞. Thus fq (z) can be expressed as a power series in 1/(z − p):
∞
X X
−1
(z − q) = an (z − p)−n = an fp (z)n ,
0
valid for |z| > 1.) Since (z − q)−1 → 0 as |z| → ∞, only terms with n ≥ 0
occur on the right. But fp ∈ A(K) by assumption, and A(K) is an algebra,
so it also contains fpn . Thus fq ∈ A(K) as well.
26
2 The simply-connected Riemann surfaces
Riemann surfaces. A Riemann surface X is a connected complex 1-
manifold. This means X is a Hausdorff topological space equipped with
charts (local homeomorphisms) fi : Ui → C, and the transition functions
fij = fi ◦ fj−1 are analytic where defined.
It then makes sense to discuss analytic functions on X, or on any open
subset of X. Technically we obtain a sheaf of rings OX with OX (U ) con-
sisting of the analytic maps f : U → C.
Aside from C and connected open sets U ⊂ C, the first interesting Rie-
mann surface (and the basic example of a compact Riemann surface) is the
Riemann sphere C. b The map f : C b − {0} → C given by f (z) = 1/z provides
a chart near infinity.
Meromorphic functions. There is also a natural notion of holomorphic
(analytic) maps between Riemann surfaces. A meromorphic function on X
is an analytic map f : X → C b that is not identically ∞.
Since X is connected, the ring K(X) of all meromorphic functions on X
forms a field, and the ring O(X) of all analytic functions forms an integral
domain.
We will see that K(C)b = C(z), while O(C) b = C, so not every meromor-
phic function is a quotient of holomorphic functions. We will see that O(C)
is much wilder (it contains exp(exp z), etc.), and yet K(C) is the field of
fractions of O(C).
Classification. In principle the classification of Riemann surfaces is com-
pleted by the following result:
Theorem 2.1 (The Uniformization Theorem) Every simply-connected
b
Riemann surface is isomorphic to H, C or C.
Then by the theory of covering spaces, an arbitrary Riemann surface sat-
isfies X ∼ b X = C/Γ, or X = C/Γ,
= C, b where Γ is a group of automorphisms.
Such Riemann surfaces are respectively elliptic, parabolic and hyperbolic.
Their natural metrics have curvatures 1, 0 and −1.
In this section we will discuss each of the simply-connected Riemann
surfaces in turn. We will discuss their geometry, their automorphisms, and
their proper endomorphisms.
27
Proper analytic maps f : X → Y are the ‘tamest’ maps between Rie-
mann surfaces. For example, they have the following properties:
Corollary 2.3 The automorphisms of C are given by the affine maps of the
form f (z) = az + b, where a ∈ C∗ and b ∈ C.
Corollary 2.4 Any Riemann surface covered by C has the form X = C/Λ,
where Λ is a discrete subgroup of (C, +).
28
degree of f . In the case X = Y = C, it is the same as the degree of the
polynomial f .
Metrics. A conformal metric on a Riemann surface is given in local coor-
dinates by ρ = ρ(z) |dz|. We will generally assume that ρ(z) ≥ 0 and ρ is
continuous, although metrics with less regularity are also useful.
A conformal metric allows one to measure lengths of arcs, by
Z Z b
L(γ, ρ) = ρ = ρ(γ(t)) |γ ′ (t)| dt;
γ a
|f ′ (z)|ρ2 (f (z)) f ∗ ρ2
kDf k = |f ′ |ρ = = ·
ρ1 (z) ρ1
29
Alternatively, let f (z) = z n . Then we find
if α = 1/n. Thus the case α = 1/n gives the quotient metric on (C, n|dz|)/hζn i,
where ζn = exp(2πi/n).
Orbifold quotients. We can also take the quotient of C by the infinite
dihedral group, D∞ = hz + 1, −zi ⊂ Aut C. The result is C itself, which the
quotient map given by cos(2πz).
Closely related is the important map
π : C∗ → C
given by π(z) = z + 1/z. This degree two map gives the orbifold quotient of
C∗ by z 7→ 1/z. It gives an intermediate covering space to the one above, if
we regard C∗ as C/Z.
The map π sends circles to ellipse and radial lines to hyperboli, with foci
[−2, 2]. In fact all conics with these foci arise in this way. (See Figure 1.)
By attaching cone angles to ±2, one turns C into the (2, 2, ∞) orbifold
X. A loop around a cone point of order n has order n in the orbifold
fundamental group, so π1 (X) = Z/2 ∗ Z/2. In general, by broadening the
scope of covering spaces and deck groups to include orbifolds and maps
with fixed points, we enrich the supply of Euclidean (and other) Riemann
surfaces. For example, the (3, 3, 3) orbifold is also Euclidean — it is the
double of an equilateral triangles.
Chebyshev polynomials. Now let Sn (z) = z n . Since Sn (1/z) = 1/Sn (z),
there is a sequence of polynomials P1 (z) = z, P2 (z) = z 2 −2, P3 (z) = z 3 −3z,
etc. satisfying
Pn (z + 1/z) = (z + 1/z)n ,
or equivalently
Pn (π(z)) = π(Sn (z)).
These Chebyshev polynomials are related to multiple angle formulas for co-
sine, since for z = eiθ we have π(z) = 2 cos θ, and thus:
Pn (2 cos θ) = 2 cos(nθ).
Theorem 2.6 The Chebyshev polynomials preserve the hyperbolas and el-
lipses with foci at ±2.
30
In particular Pn : [−, 2, 2] → [−2, 2] by degree n, and Pnk (z) → ∞ for all
other z ∈ C. (Thus the Julia set of Pn is [−2, 2].)
Solving the cubic. Complex algebra finds its origins in the work of Car-
dano et al on solving cubic polynomial equations. Remarkably, complex
numbers intervene even when the root to be found is real.
One can always make a simple transformation of the form x 7→ x + c to
reduce to the form
x3 + ax + b = 0.
One can further replace x with cx to reduce to the form
x3 − 3x = b.
Thus the solution to the cubic involves inverting a single cubic function
P3 (z) = z 3 − 3z.
But to solve Pn (x) = a, we just write a = y + 1/y (by solving a quadratic
equation), and then we have x = y 1/n + y −1/n . In particular, this method
can be used to solve x3 − 3x = b.
Classification of polynomials. Let us say p(z) is equivalent to q(z) is
there are A, B ∈ Aut(C) such that Bp(Az) = q(z). Then every polynomial
is equivalent to one which is monic and centered (the sum of its roots is
zero). Every quadratic polynomial is equivalent to p(z) = z 2 .
The reasoning above shows, every cubic polynomial with distinct critical
points is equivalent to P3 (z) = z 3 − 3z. Otherwise it is equivalent to z 3 .
But for degree 4 polynomials we are in new territory: the cross-ratio of the
3 critical points, together with infinity, is an invariant.
It is a famous fact (proved using Galois theory) that a general quintic
polynomial (with integral coefficients) cannot be solved by radicals.
given by π(z0 , z1 ) = z0 /z1 ; it records the slope of each line. This gives a
natural identification of Cb with the projective line P1 , i.e. the space of lines
in C2 .
31
Aside: Some topology of projective spaces. The real projective plane RP2 is
the union of a disk and a Möbius band. The natural map p : C2 − {(0, 0)} →
b factors through the Hopf map S 3 → S 2 , whose fibers have linking number
C
one. This map generates π3 (S 2 ).
Möbius transformations. So long as ad − bc 6= 0, the map f (z) =
(az + b)/(cz + d) defines an b
a b
automorphism of C. Its inverse can be found by
inverting the matrix c d . In fact, the map π above transports the linear
b
action of SL2 (C) on C2 to the fractional linear action of PSL2 (C) on C.
b = PSL2 (C).
Theorem 2.7 We have Aut(C)
Note that a Möbius transformation (other than the identity) either has
two simple fixed points, or a single fixed point of multiplicity two. The fixed
points of A correspond to its eigenvectors on C2 .
Note also that all these elements, except for irrational elliptics, generate
b
discrete subgroups of Aut(C).
±1
If A has eigenvalues λ , then tr(A) = λ + 1/λ. Our previous analysis
of this map shows, for example, that the traces of matrices a given value for
|λ| correspond to an ellipse with foci ±2.
32
Cross-ratios. The cross-ratio is an invariant of ordered 4-tuples of points,
characterized by the conditions that [a : b : c : d] = [ga : gb : gc : gd]
b and [0 : 1 : ∞ : λ] = λ ∈ C
for all g ∈ Aut(C), b − {0, 1, ∞}. The cross-
ratio gives an explicit isomorphism between the moduli space M0,4 and the
triply-punctured sphere.
Stereographic projection. There is a geometric identification between
the unit sphere S 2 ⊂ R3 and the Riemann sphere C b in which the north pole
N becomes ∞ and the rest of the sphere is projected linearly to C = R2 ×{0}.
θ/2
1
x
θ
33
θ(N ) = π, then we find x = p(θ) = tan(θ/2). Thus dx = 2 sec2 (θ/2) dθ =
2(1 + x2 ), which gives dθ = 2dx/(1 + x2 ). The case of S 2 follows by confor-
mality and rotation invariance.
34
Area of triangles. The geodesics on S 2 are arcs of great circles. Their
angle sum always exceeds π, and in fact we have:
Theorem 2.14 The area of a spherical triangle is equal to its excess angle,
α + β + γ − π.
Proof. Since the sphere has area 4π, a lune of angle θ has area 4θ. The
three lunes coming from a given triangle T cover the whole sphere, with
points inside T and its antipode triply covered and the rest simply covered.
Thus 4π + 4 area(T ) = 4(α + β + γ).
The general formula can be deduced from the constant curvature formulas
(for K = ±1) by subdivision.
Corollary 2.16 Aut(H) corresponds to the subgroup SL2 (R) ⊂ SL2 (C).
35
Corollary 2.17 Aut(∆) corresponds to the subgroup SU(1, 1) of isometries
of the form |Z|2 = |Z0 |2 − |Z1 |2 .
Corollary 2.18 The automorphism group of the disk is given by the sub-
group ( ! )
a b 2 2
Aut(∆) = : |a| − |b| = 1 ⊂ PSL2 (C).
b a
Note that these matrices satisfy AQA∗ = Q, where Q = 1 0
0 −1 .
Hyperbolic geometry. The hyperbolic metric on H is given by ρ =
|dz|/ Im z = |dz|/y. On the unit disk ∆, the hyperbolic metric becomes
ρ = 2|dz|/(1 − |z|2 ).
The geodesics for the hyperbolic metric are given by circles orthogonal
to the boundary (of H or ∆). For example, in the case of the imaginary
axis γ = iR+ ⊂ H, it is easy to see that the projection π : H → γ given
by π(x, y) = (0, y) is distance-decreasing, and so γ gives a shortest path
between any two of its points. The general case follows from this one, since
any circle orthogonal to the boundary of H is equivalent, under Aut(H), to
γ.
These geodesics satisfy all of Euclid’s postulates except the fifth. Thus
if we declare them to be straight lines, we find:
Theorem 2.20 Euclid’s fifth postulate cannot be deduced from the other
axioms of geometry.
36
Triangles. Gauss-Bonnet for hyperbolic triangles reads:
area(T ) = π − α − β − γ.
For example, an ideal triangle in H has vertices at infinity and internal angles
of 0. Its area is π.
Curvature. We remark that the Gauss curvature of a conformal metric
ρ(z) |dz| is readily computed in terms of the Laplacian of ρ: we have
This invariant is useful for the classification of isometries. Here are the
possibilities:
1. (Elliptic) τ (f ) = 0, achieved. Then f is conjugate to a rotation fixing
i, of the form ! !
a b cos θ − sin θ
= .
c d sin θ cos θ
(Note that when θ = π above, the matrix is −I and so f (z) = z.) In
the disk model, we can simply conjugate so f (z) = e2iθ z.
37
Proof. The maximum principle implies that the analytic function g(z) =
f (z)/z satisfies |g(z)| ≤ 1. If equality holds, then g is constant.
Proof. Pass to the universal covers of the domain and range and apply the
usual Schwarz lemma.
with zeros satisfying |ai | < 1. Using the fact that 1 = zz on S 1 , it is easy
to show that |B(z)| = 1 on S 1 , and thus B : ∆ → ∆ is proper. Conversely,
we have:
38
Theorem 2.24 Every proper analytic map f : ∆ → ∆ is a Blaschke prod-
uct.
Proof. A proper map is surjective, so f has at least one zero, say a. Let
M (z) = (z − a)/(1 − az). Then g = f (z)/M (z) : ∆ → ∆ is analytic, with
one fewer zero than f . The proof now follows by induction on the degree of
f.
Theorem 2.25 Let f : C b→C b be a rational map of degree d > 1. Then the
immediate basin of any attracting cycle contains a critical point.
39
P Q
Theorem 3.1 If 1/|an | is finite, then f (z) = (1 − z/an ) defines an
entire function with zeros exactly at these points.
with ak ≥ 0 for all k. Integrating term by term and using the fact that
Ep (0) = 1, we find
∞
X
p+1
1 − Ep (z) = z bk z k
0
40
P
with bk ≥ 0. We also have bk = 1 − Ep (1) = 1, and hence for |z| < 1 we
have X
|1 − Ep (z)| ≤ |z|p+1 bk = |z|p+1 .
Proof. The previous estimate yields convergence of the tail of the series:
for all z ∈ B(0, R), we have:
X ∞
X ∞
X
|1 − En (z/an )| ≤ (|z|/|an |)n+1 < (R/2R)n+1 < ∞.
|an |>2R 1 1
X ∞
X X
|1 − Ep (z/an )| (|z|/|an |)p+1 < ∞. ≤ Rp+1 1/|an |p+1 < ∞.
|an |>2R 1
P
≥ 0 is the least integer such that
If p Q 1/|an |p+1 < ∞, then we say
P (z) = Ep (z/an ) is the canonical product associated to (an ).
The counting function. It is also useful to introduce the counting function
−β
P(r) = |{n β: |an | < r}|. Then r N (r) is a rough approximation to
N
|an |<r 1/|an | . Consequently we have:
log N (r)
α = lim sup ·
r→∞ log r
41
In more detail: suppose N (r) ≤ r β ; then we can collect the points an into
groups where 2n < |an | ≤ 2n+1 ; then
X X X
|an |−α ≤ (2n )−α N (2n+1 ) ≤ 2β 2n(β−α) < ∞
Examples: Polynomials have order 0; sin(z), cos(z) and exp(z) have order
√
1; cos( z) has order 1/2; Ep (z) has order p; exp(exp(z)) has infinite order.
Hadamard’s 3-circles theorem. Here is useful general property of the
function M (r) = sup|z|=r |f (z)|. The following result applies not just to
entire functions, but also to functions analytic in an annulus of the form
r1 < |z| < r2 .
Theorem 3.5 For any analytic function f (z), the quantity log M (r) is a
convex function of log r.
Proof. A function φ(s) of one real variable is convex if and only if φ(s) +
ar satisfies the maximum principle for any constant a. This holds for
log M (exp(s)) by considering f (z)z a locally.
42
√ p
Corollary 3.6 We have M ( rs ≤ M (r)M (s).
The convex functions satisfying F (log r) = log M (r) look roughly linear
for polynomials, e.g. like F (x) = deg(f )x + c, and look roughly exponential
for functions of finite order, e.g. F (x) = exp(ρ(f )x).
Hadamard’s factorization theorem. We can now state a formula that
describes every entire function of finite order in terms of its zeros and an
additional polynomial.
Corollary 3.8 Suppose f (z) and g(z) are entire functions of order ρ with
the same zeros, and f ′ /f = g ′ /g at ⌊ρ⌋ distinct points (where neither func-
tion vanishes). Then f is a constant multiple of g.
Theorem 3.9 Let f (z) be an entire function of finite order with no zeros.
Then f (z) = eQ(z) , where Q(z) is a polynomial of degree d, and ρ(f ) = d.
43
Lemma 3.10 Let Q(z) be an entire function satisfying Re Q(z) ≤ A|z|d +B
for some A, B > 0. Then Q is a polynomial of degree at most d.
Proof. There is a constant C > 0 such that for R > 1, Q maps ∆(2R) into
the half-plane U (R) = {z : Re z < CRd }. By the Schwarz Lemma, Q is
distance-decreasing from the hyperbolic metric on ∆(2R) to the hyperbolic
metric on U (R). Since ∆(R) ⊂ ∆(2R) has bounded hyperbolic diameter,
the same is true for Q(∆(R)) ⊂ U (R). Therefore in the Euclidean metric,
This shows |Q(z)| = O(|z|d ) for |z| > 1, and hence Q is a polynomial of
degree at most d.
Proof of Theorem 3.9. Since f has no zeros, f (z) = eQ(z) for some entire
d
function Q(z). Since f has finite order, |f (z)| = O(e|z| ) for some d, and
thus Re Q(z) ≤ |z|d + O(1); now apply the Lemma above.
II. Functions with zeros. Next we analyze entire functions with zeros.
We will show:
Theorem
P 3.11ρ+ǫLet f (z) be an entire function of order ρ with zeros (an ).
Then 1/|an | < ∞ for all ǫ > 0.
Jensen’s formula. Informally, the result above says that if f has many
zeros, then M (r) must grow rapidly. The relation between the size of f and
its zeros is made precise by the following important result:
44
Proof. We first note that if f has no zeros, then log |f (z)| is harmonic and
the formula holds. Moreover, if the formula holds for f and g, then it holds
for f g; and the case of general R follows from the case R = 1.
Next we verify that the formula holds when f (z) = (z − a)/(1 − az) on
the unit disk, with |a| < 1. Indeed, in this case log |f (z)| = 0 on the unit
circle, and log |f (0)| + log(1/|a|) = log |a/a| = 0 as well.
The general case now follows, since a general
Q function f (z) on the unit
disk can be written in the form f (z) = g(z) (z − ai )/(1 − ai z), where g(z)
has no zeros.
and hence α = lim sup log N (r)/ log r ≤ lim sup log M (2r)/ log(2r) = ρ.
Of course functions can also grow rapidly without having any zeros. But
then Jensen’s formula shows the average of log |f | is constant over every
circle |z| = R; so if f is large over much of the circle, it must also be close
to zero somewhere on the same circle. (For example, ez is very large over
half the circle |z| = R, and very small over the rest.)
III. Canonical products. Our next task is to determine the order of a
canonical product. It will also be useful, to complete the proof of Hadamard’s
theorem, to obtain lower bounds for such a product.
Proof.
P Let rp+1n = |an | and r = |z|. Recall that
P p pis the least integer such
that (1/rn ) < ∞. So we also have 1/rn = +∞. This implies
45
P
p ≤ α ≤ p + 1. For convenience we will assume (1/rn )α < ∞. (For the
general case, just replace α with α + ǫ.)
As we have seen, the Weierstrass factor
z2 zp
Ep (z) = (1 − z) exp z + + ··· +
2 p
satisfies, for ‘small z’ meaning |z| < 1/2, the inequality |1−Ep (z)| ≤ |z|p+1 ≤
1/2, and hence also the inequality
On the other hand, for ‘large z’, meaning |z| ≥ 1/2, we have
46
The minimum modulus. To control the result of division by a canonical
product, we now estimate its minimum modulus. As an example — for
sin(z), we have m(r) ≍ 1 for infinitely many r.
Theorem 3.15 The minimum modulus of the canonical product P (z) above
satisfies m(r) ≥ exp(−r α+ǫ ) for arbitrarily large r.
Proof. We must show | log m(r)| = O(r α+ǫ ). The follows the same lines as
the bound log M (r) = O(r α ) just obtained, since (3.1) and (3.2) give bounds
for | log Ep (z)|. The only nuance is that we cannot ignore the logarithmic
term in (3.2). We must also decide which values of r to choose, since m(r) =
0 whenever r = |an |.
To this end, we fix ǫ > 0 and exclude from consideration the balls Bn
defined by |z − an | < 1/rnα+ǫ . Since the sum of the radii of the S excluded
balls in finite, there are plenty of large circles |z| = r which avoid Bn .
To complete the proof, it suffices to show that for z on such circles, we
have X
| log(1 − z/an )| = O(r α+ǫ ).
|z−an |<rn
Note that the number of terms in the sum above is at most N (2r) = O(r α ).
Because we have kept z away from an , we have
Consequently
X
| log(1 − z/an )| = O(N (2r) log r) = O(r α+ǫ ),
|z−an |<rn
as desired.
47
Trigonometric functions. As a first example, we determine the canonical
factorization of the sine function:
Proof. Indeed, the right hand side is a canonical product, and sin(πz) has
order one, so the formula is correct up to a factor exp Q(z) where Q(z) has
degree one. But since sin(πz) is odd, we conclude Q has degree zero, and
by checking the derivative at z = 0 of both sides we get Q = 0.
Sine, cotangent and zeta. The product formula above gives, under log-
arithmic differentiation,
∞
(sin(πz))′ 1 X 1 1
= π cot(πz) = + + ·
sin(πz) z z−n z+n
1
This formula is useful in its own right: it shows π cot(πz) has simple poles at
all points of Z withPresidue2kone. This property can be used, for example, to
evaluate ζ(2k) = ∞ 1 1/n and other similar sums by the residue calculus
— see the examples in Chapter 1.
We note that the product formula for sin(z) can also be used to prove
ζ(2) = π 2 /6, by looking at the coefficient
P of z 2 on bothPsides of the equa-
tion. The sine formula also shows a<b 1/(ab) = π /5!, a<b<c 1/(abc)2 =
2 4
π 6 /7!, etc., so with some more work it can be used to evaluate ζ(2k). For
example, we have
X X !
1 1 X 1 π4 π4 π4
ζ(4) = − 2 = − = ·
a2 b2 (ab)2 36 60 90
a<b
48
Translation and duplication formulas. Many of the basic properties
of the sine and cosine functions can be derived from the point of view of
the uniqueness of an odd entire function with zeros at Zπ. For example,
equations sin(z + π) = sin(z) and sin(2z) = 2 sin(z) sin(z + π/2) hold up to
a factor of exp(az + b) as a consequence of the fact that both sides have the
same zero sets.
It has half the terms that appear in the factorization of sin(πz); indeed, we
have
sin(πz)
G(z)G(−z) = · (3.3)
πz
Euler’s constant. We have G(0) = 1, but what is G(1)? To answer this,
we define Euler’s constant by:
This
R n+1 expression gives the limiting error obtained when one approximates
1 dx/x by n unit rectangles lying above the graph of y = 1/x. (The sum
is finite because these areas can all be slid horizontally to lie disjointly inside
a fixed rectangle of base one.)
Then, using the fact that (1 + 1)(1 + 1/2) · · · (1 + 1/n) = (n + 1), we find
G(1) = exp(−γ).
Functional equation. The functions G(z − 1) and zG(z) have the same
zeros. How are they related? By Hadamard’s theorem, we have G(z − 1) =
z exp(az + b)G(z) for a, b. In fact:
49
while the logarithmic derivative of G(z − 1) is just
∞
X X 1 ∞
1 1 1 1
− = −1+ − .
1
z+n−1 n z 1
z+n n+1
P∞
Since 1 (1/n − 1/(n + 1)) telescopes to 1, these series are equal, and hence
G(z − 1) and zG(z) agree up to a constant. The value of this constant is
determined by setting z = 1.
10
-3 -2 -1 1 2 3
-5
-10
2. Γ(n + 1) = n! for n ≥ 0;
50
Using the fact that
Γ(z + 1)
Γ(z − n) = ,
(z − n) · · · (z − 1)z
we find
(−1)n
Res−n (Γ(z)) = ·
n!
From (3.3) we obtain Euler’s supplementj
π
= Γ(z)Γ(1 − z),
sin(πz)
which implies, for example:
√
Γ(1/2) = π.
Using the fact that Γ(z) = Γ(z), we find from this last that
π
|Γ(iy)|2 = ·
y sinh(πy)
We also note that by the functional equation, Γ(z) has constant sign on
each interval of the form (n, n + 1); in fact the sign is positive for n > 0 and
(−1)n for n < 0.
Gauss’s formula. The formula above for the Γ function would not be easy
to discover from scratch. Here is a formula that could also be taken as a
simpler definition of the Gamma function.
−1
As motivation, we note that for integers z we have z! = lim nz z+n n ,
which is consistent with the formula above if we multiply both sides by z.
Indeed, formally we have
z+n (z + 1) · · · (z + n) (n + 1) · · · (n + z) nz
= = ≈ ,
n n! z! z!
51
and solving for z! and taking the limit gives the formula above.
Proof. By definition, we have:
n
1 γz
Y z −z/k
= lim ze 1+ e .
Γ(z) n→∞ k
k=1
Proof. This inequality holds for each term in the product above.
52
A duplication formula. As an application one can prove ‘multiple angle’
formulas for Γ(nz). The simplest is:
√
Corollary 3.21 We have 2 πΓ(2z) = 22z Γ(z)Γ(z + 1/2).
53
which is itself proved using the identity
Z ∞ 2 Z
−x2 /2 2
e dx = e−r /2 r dr dθ = 2π.
−∞
R∞
To this end we rewrite Γ(s) = 0 eφ(t) dt, where φ(t) = −t + (s − 1) log t.
Then φ′ (t) = −1+(s−1)/t vanishes at t0 = s−1, with φ′′ (t0 ) = −(s−1)/t20 =
−1/(s − 1). So we have
54
Proof. Let qn (z) be the polynomial given by truncating the power series
−n when |z| < |a |/2. Then f (z) =
P pn (z) so that |pn (z) − qn (z)| < 2
for n
(pn (z) − qn (z)) is the desired function.
...this topsy–turvy way of doing things (die Dinge auf dem Kopf
zu stellen) should be not sanctioned by anyone who sees mathe-
matics as something other than a disordered heap of mathemat-
ical results.
—A. Pringsheim, 1915.
55
4 Conformal mapping
We now turn to the theory of analytic functions as mappings. Here the
dominant operation is composition, rather than addition or multiplication.
The same result holds for any simply-connected region U ⊂ C b such that
b b
|C−U | ≥ 1. Note that π1 (U ) = (1) iff C−U is connected, so the complement
is either empty, one point or a nontrivial continuum.
Since Aut(∆) is large, the Riemann map is not unique. It can be made
unique by picking a point p ∈ ∆ and requiring f (p) = 0 and f ′ (p) > 0. The
value of f ′ (p) is an interesting invariant of (U, p); its reciprocal is called the
conformal radius of U with center p.
Examples of Riemann maps.
1. We first note that ∆ ∼= H, e.g. by the Möbius transformation A(z) =
−i(z −1)/(z +1). We also remark that A : S 1 → R b becomes in angular
coordinates,
2 eiθ/2 − e−iθ/2
A(θ) = = tan(θ/2).
2i eiθ/2 + e−iθ/2
This is the change of variables used in calculus to integrate rational
functions of sine and cosine.
56
send L to H. (This illustrates the idea that the Riemann map U ∼
= ∆ is
often constructed as a composition of several simple transformations.)
C−∆→C−K
57
Let us say f : U → C is univalent if f is injective and analytic. By
injectivity, its inverse is analytic, and f provides a homeomorphism between
U and V = f (U ). (It need not provide a homeomorphism between their
closures.) We then have:
g = B ◦ s−1 ◦ A ◦ f : U → ∆
58
satisfies g(p) = 0 and g ′ (p) > 0. Then g ∈ F as well.
Now notice that f = (A ◦ s ◦ B) ◦ g = h ◦ g. By construction, h : ∆ → ∆
is a proper map of degree two, with h(0) = 0. Thus |h′ (0)| < 1 by the
Schwarz lemma. But g′ (0) = h′ (0)f ′ (0), so g′ (0) > f ′ (0) = M , contrary to
the definition of M since g ∈ F.
Thus f must have been surjective after all, so it provides the desired
conformal map between U and ∆.
In other words,
p the length of the image of some horizontal line is bounded
above by area(U )a/b.
Corollary 4.7 If area(U ) = area(R(a, b)), then the image of some horizon-
tal line is shorter in U (or at least no longer).
59
Jordan curves. Here is a basic fact about a Jordan curve J ⊂ C. Given
a, b ∈ J, let [a, b] ⊂ J be the subarc joining these two points with smallest
diameter. Then diam[a, b] → 0 if |a − b| → 0.
Proof of Theorem 4.5. Given a point z ∈ ∂∆, map ∆ to an infinite strip,
sending z to one end. Then there is a sequence of disjoint squares in the strip
tending towards that end. The images of these squares have areas tending to
zero, so there are cross-cuts γn ⊂ ∆ enclosing z such that length(f (γn )) → 0.
Thus the endpoints of f (γn ) converge to points an , bn ∈ J, with z ∈ [an , bn ]
for n ≫ 0. Since diam[an , bn ] → 0, the disk bounded by f (γn ) ∪ [an , bn ]
shrinks to z, and this implies f (zn ) → f (z) whenever zn → z in ∆.
Consequently f extends to a continuous map S 1 → J. Since f |∆ is a
homeomorphism, f |S 1 is monotone; thus it is injective unless it is constant
on an arc. But it cannot be constant on an arc, since f is nonconstant.
Local connectivity. The key point in the proof above is the following
property of a Jordan domain U: if a, b ∈ ∂U are joined by a short arc
α ⊂ U , then the disk cut off by α has small diameter. This principle is
called ‘short dam, small lake’.
Recall that a compact set K is locally connected if for every open set
U ⊂ K and x ∈ U there is a connected open set with x ∈ V ⊂ U .
Exercise. If there exists a continuous surjective map f : S 1 → K, then K
is locally connected.
The argument in the proof of Theorem 4.5 furnishes short cross-cuts for
any Riemann map, so it also shows:
60
More generally, it is known that any bounded analytic function has radial
limits a.e. By [Ko], not much more can be said — given a Gδσ set A ⊂ S 1 of
measure zero, there exists a bounded function that fails to have radial limits
exactly on the set A. For Riemann mappings, radial limits exist except
outside a very small set — the exceptional set A has capacity zero and in
particular H. dim(A) = 0.
Harmonic functions and boundary values. Recalling that if u is har-
monic and f is analytic then u◦f is also harmonic, it is now a simple matter
to solve the Dirichlet problem (at least implicitly) on any Jordan domain.
Theorem 4.11 For any Jordan domain U ⊂ C, there exists a unique map
P : C(∂U ) → C(U ) such that P u|∂U = u and P u is harmonic in U .
b is isomorphic
Theorem 4.12 The universal cover of any annulus U ⊂ C
to C or H.
61
Proof. Let g : V → V denote a generator for π1 (U ) ∼ = Z acting on its
universal cover. If V ∼= C then g is conjugate to g(z) = z + 1 and we have
U ∼= C/hgi ∼ = C∗ by the map π(z) = exp(2πiz). The same reasoning shows
U∼ = ∆∗ if V ∼ = H and g is parabolic.
Otherwise V ∼ = H and we can assume g(z) = λz, λ > 1. This means the
core curve of U is a geodesic of length L = log λ. Then π(z) = z α maps V
to A(R) if we choose α correctly. We want π to map [1, λ] onto the unit
circle, so we want π(λ) = λα = exp(αL) = 1; so we take α = −2πi/L. Note
that this is a purely imaginary number, so π sends R+ to the unit circle and
R− to a circle of radius R = (−1)α = exp(απi) = exp(2π 2 /L). (We put a
minus sign in α so that R > 1.)
f (z) = z + a2 z 2 + z3 z 3 · · ·
62
their Laurent series has the form:
X∞
bn
F (z) = z + ·
n=1
zn
Proof. Integrate F dF over the unit circle and observe that, since |dz|2 =
(−i/2)dz dz, the area A of K(F ) is given by:
Z Z dz
i i X
2
X
A=− F dF = − 1 − n|bn | =π 1− n|bn |2 .
2 S1 2 S1 z
P
Corollary 4.16 We have n|bn |2 ≤ 1. In particular, we have |bn | ≤
√
1/ n.
Proof. This follows from the fact that |bn | ≤ n−1/2 . So if we fix R > 1,
then for |z| ≥ R we have
X
|F (z) − z| ≤ n|bn |/Rn+1 = C(R) < ∞.
63
Proof. We first note that F (z) = 1/f (1/z) + a2 is in Σ. Indeed, we find:
z
1/f (1/z) =
1 + a2 /z + a3 /z 2 + · · ·
= z 1 − (a2 /z + a3 /z 2 + · · · ) + (a2 /z + a3 /z 2 + · · · )2 − · · ·
a22 − a3
= z − a2 + + ···
z
and thus F (z) has b1 = a22 − a3 . So we find |a22 − a3 | ≤ 1, which is sort of
interesting, but not what we are aiming for yet.
Next we use a nice trick to make p f (∆) more symmetrical: we consider,
instead of f (z), the new map g(z) = f (z 2 ). This map is also in S and it
is given by
p
g(z) = z 1 + a2 z 2 + a3 z 2 + · · · = z + (a2 /2)z 3 + · · ·
Theorem 4.20 (Koebe 1/4 theorem) For any f ∈ S we have B(0, 1/4) ⊂
f (∆).
Proof. Suppose p 6∈ f (∆). Note that A(z) = z/(1 − z/p) has A(0) = 0 and
A′ (0) = 1. Thus
f (z)
g(z) = = (z + a2 z 2 + · · · )(1 + z/p + · · · ) = z + (a2 + 1/p)z 2 + · · ·
1 − f (z)/p
64
From the Koebe theorem we get an important comparison between the
hyperbolic metric ρ and the ‘1/d’ metric δ = δ(z)|dz| = |dz|/d(z, ∂U ).
These maps are equivalent: F (z) = 1/f (1/z) + 2. We have K(F ) = [−2, 2]
and K(f ) = (−∞, −1/4]. The map f (z) shows the Bieberbach conjecture
is sharp. We also note the problem with trying to find a ‘Bieberbach con-
jecture’ for Σ: there is no map which simultaneously maximizes all the bn ’s.
Indeed, by the area theorem, if b1 = 1 then the rest of the bn ’s are zero.
The distortion theorems. Given f ∈ S, think of f (∆) as a splattered
egg; then one finds that no matter what, the yolk f (B(0, r)), r < 1, is still
good (not too distorted). For example, the curvature of f (S 1 (r)) is bounded
by a constant Kr independent of f . Also, f (S 1 (r)) is convex if r is small
enough.
Qualitative theorems of this type can be easily deduced from compact-
ness of S. The Koebe distortion theorems make these results more precise.
They state:
(1 − r) (1 + r)
3
≤ |f ′ (z)| ≤
(1 + r) (1 − r)3
and
r r
2
≤ |f (z)| ≤ · (4.1)
(1 + r) (1 − r)2
65
The proof can be made rather conceptual. Let U ⊂ C be a proper
simply-connected region, and let f : U → C be a univalent map. We will
use the hyperbolic metric ρ(z) |dz| on U and the Euclidean metric |dz| on
C. It is then natural to study how these metrics compare under f . To this
end we define
δ(z) = log(|f ′ (z)|/ρ(z))
Note that if we replace f (z) with af (z) + b, it only changes δ(z) to δ(z) +
log |a|.
(1 − r)2 ρ(r) (1 − r)
|f ′ (z)| ≥ 2
· = ·
(1 + r) ρ(0) (1 + r)3
66
These results also show that S is compact.
Multiply-connected regions. We briefly mention one of the standard
forms for a region that has ‘more than two holes’.
Suppose U = C − (K1 ∪ · · · ∪ Kn ), where the Ki are disjoint compact
connected sets, none of which is a single point. We then have:
2. For any pair of maps defined near ∞ by z+b1 /z+. . ., we have b1 (f ◦g) =
b1 (f ) + b1 (g).
3. Thus the complement is also true in the case n = 1. For in this case
we can assume (after translation) that U is the image of g ∈ Σ. Then
F = f ◦ g −1 , and Re b1 (F ) = Re b1 (f ) − b1 (g) = 1 − Re b1 (g).
But even more is true! Unless U is already a slit domain, Re b1 (F ) > 0,
by the area theorem applied to g — since |b1 (g)| ≤ 1.
67
5. We should also check uniqueness. For this it is useful to think of
U as the interior of a compact, smoothly bounded domain in C, and
b as normalized maps of the form fi (z) = 1/(z −p)+gi (z)
f1 , f2 : U → C
whose images are horizontal slit domains. Then the bounded analytic
function h(z) = f1 (z) − f2 (z) on U also sends each component of ∂U
into a horizontal line. Thus h(∂U ) has winding number zero about
points not on these lines, so h must be constant.
Number of moduli.
S The number of n-slit regions in C, i.e. those of the
form S = C − n1 [zi , zi + ai ] with ai ∈ R, has real dimension 3n. We can
normalize by an affine transformation so the first slit is, say, [−2, 2]; so up
to isomorphism, the number of such slit regions is 3n − 3.
Now to take an n-connected region U and produce a slit region, we
need to choose a point p ∈ U to send to infinity and we need to choose a
‘horizontal’ direction at p. That gives 3 more real parameters. But of course
these choices are only relevant up to the action of the automorphism group
of U , which has dimension 3, 1 and 0 for n = 1, 2 and n ≥ 3. Altogether we
find:
68
Let f : X → C b be a meromorphic function on a Riemann surface. Then
′
ω = df = f (z) dz is naturally a 1-form. That is, if we change coordinates
by setting z = φ(w), then in these new coordinates we have
Let’s check this in an example: if we take ω(z) = dz/P (z) where P (z)
is a polynomial of degree d ≥ 2, then it says:
X 1
= 0.
P ′ (z)
P (z)=0
P
ThisP can be verified using partial fractions: we have P (z) = ai /(z − bi )
and ai = 0 because |P (z)| = O(|z|−2 ) for large z. It can also be proved
by integrating dz/P (z) around a large circle and taking the limit.
If ω = ω(z) dz is a meromorphic 1-form on the sphere, we can set π(z) =
1/z and form π ∗ ω to find:
Theorem 4.26 Every meromorphic 1-form on the sphere has 2 more poles
than zeros.
Proof. This is true for ω = dz/z, which has a simple poles at 0 and ∞
and no zeros. It then follows for any other 1-form η, since f = η/ω is a
meromorphic function, which must have the same number of poles as zeros.
69
Corollary 4.27 A holomorphic 1-form on C b must be zero. In particular,
if ω1 and ω2 are 1-forms with the same principal parts, then ω1 = ω2 .
Remark: one forms and vector fields in higher genus. The first
assertion can also be seen by integrating ω to obtain a global holomorphic
b On a Riemann surface of genus g, a meromorphic 1-form has
function on C.
2g − 2 more zeros than poles, and dim Ω(X) = g. The space of holomorphic
b = 3, dim Θ(E) = 1 on a
vector fields, on the other hand, satisfies dim Θ(C)
torus, and dim Θ(X) = 0 in higher genus. This is because a meromorphic
vector field has 2g − 2 more poles than zeros.
One can use the vanishing of Θ(X) is see an n-connected plane region U
has a zero-dimensional automorphism group for n 6= 3; if not, there would
be a boundary-parallel holomorphic vector field on U , but then the double
X of U would satisfy dim Θ(X) > 0; while its genus is given by g = n − 1.
Measuring distortion: the 3 great cocycles. Now let C(f ) be a dif-
ferential operator that sends meromorphic functions to meromorphic forms.
We say C is a cocycle if it satisfies:
Their values are functions, 1-forms and 2-forms respectively. The groups
they annihilate are transformations of the form f (z) = z + a, f (z) = az + b
and f (z) = (az + b)/(cz + d).
Let us check this for Sf : if f (z) = (az + b)/(cz + d) and ad − bc = 1,
then f ′ (z) = 1/(cz + d)2 , hence N f (z) = −2c/(cz + d), which satisfies
2c2 1
(N f )′ = 2
= (N f )2 .
(cz + d) 2
70
cannot be cocycles, because the rational maps of degree d > 1 do not form
a group. P
To get some insight into Sf (and
P T d ), note that if f (z) = ai z i =
(az + b)/(cz + d), then (cz + d) ai zi = az + b and hence there must
be a linear relation between most of the coefficients of zf (z) and f (z); in
particular, we must have
!
a1 a2
det = a22 − a1 a3 = 0,
a2 a3
which just says that f ′ f ′′′ − (3/2)(f ′′ )2 vanishes. This expression is also the
denominator of Sf . Simiarly a rational map of degree two is characterized
by the property that
a1 a2 a3
det
a2 a 3 a4
= 0.
a3 a4 a5
where f (qi ) = pi .
71
maps: gj = Aij ◦ gi , and thus N (gi ) = N (gj ). It remains to show N f has
simple poles with the given residues at the pi , and at infinity.
Let αi = 1 − µi . The idea of the proof is that f (z) behaves like (z − pi )αi
near pi , and thus N f has the same residue, which is αi − 1 = −µi . To check
this, we note that g(z) = (f (z)−pi )1/αi extends by Schwarz reflection across
qi , to give a conformal map near qi . Thus we can locally write
g(z)αi = f (z) − pi .
Taking the nonlinearity of both sides, and using the fact that the residue is
preserved under pullback, we find Resqi (N f ) = 1 − αi = −µi .
Near infinity, f (z) behaves like 1/z which has nonlinearity −2dz/z and
hence residue
P 2 at infinity. This is consistent with the residue theorem,
because πµi = 2π. Since a 1-form is determined by its singularities, we
have justified the formula for N f .
Integrating, we find
X
log f ′ = −µi log(z − qi ) + C;
72
R
Theorem 4.29 The map f (z) = (1 − z n )−2/n dz maps the unit disk to
a regular n-gon, sending 0 to its center and the nth roots of unity to its
vertices.
Note that f ′ (0) = 1, and the distance from the center of the polygon to
one of its vertices is given by
Z 1
Rn = (1 − z n )−2/n dz.
0
73
In particular, the conformal radius of the unit square (sides of length
2) is 2/S4 = 4Γ[3/4]2 /π 3/2 = 1.078705 . . .. (It
√ is clear that the conformal
radius of the unit square lies between 1 and 2.)
Quadrilaterals. A quadrilateral Q is a Jordan domain with 4 distinguished
points on its boundary. A conformal map between quadrilaterals is required
to preserve these points as a set.
(N f )2 1 − α2 dz 2
Sf (z) = (N f )′ − = ·
2 2 z2
The quantity 1 − α2 is a version of the residue for the quadratic differen-
tial Sf (z) dz 2 . Now given α, β, γ there is a unique quadratic differential
with double poles at 0, 1, ∞ with the corresponding residues, and no other
singularities. It is given by
1 1 − α2 1 − β2 1 − γ 2 − (2 − α2 − β 2 )
Q(α, β, γ) = + + dz 2 .
2 z2 (z − 1)2 z(z − 1)
74
Theorem 4.31 Let P be a circular triangle with interior angles πα, πβ and
πγ. Then the Riemann mapping f : H → P sending 0, 1 and ∞ to these
vertices satisfies Sf = Q(α, β, γ).
u′′ + (Q/2)u = 0.
75
Theorem 4.34 (Great Picard Theorem) An analytic function f : U →
C takes on every value in C, with at most one exception, in every neighbor-
hood of an essential singularity p.
The first theorem follows from the second by considering the essential
singularity at z = 0 of f (1/z). These results generalize Liouville’s theorem
and the Weierstrass-Casorati theorem respectively.
Rescaling arguments. Here is a third result that initially seems unrelated
to the first two.
Corollary 4.36 For any analytic map on ∆, the image f (∆) contains a
ball of radius R|f ′ (0)|.
Note that the ball usually cannot be centered at f (0); for example,
f (z) = exp(nz)/n satisfies f ′ (0) = 1 but the largest ball about f (0) = 1/n
in f (∆) ⊂ C∗ has radius 1/n.
√ The optimal value of R is known as Bloch’s constant. It satisfies 0.433 <
3/4 ≤ R < 0.473. The best-known upper bound comes from the Riemann
surface branched with order 2 over the vertices of the hexagonal lattice.
These apparently unrelated theorems can both be proved using the same
idea. (Cf. [BD] and references therein; for another way to relate these
theorems, see [Re, Ch. 10].)
Proof of Bloch’s theorem. Given f : ∆ → C, let
kf ′ (z)k = kf ′ (z)k∆,C = (1/2)|f ′ (z)|(1 − |z|2 )
denote the norm of the derivative from the hyperbolic metric to the Eu-
clidean metric. By assumption, kf ′ (0)k = 1/2. We can assume (using
f (rz)) that f is smooth on S 1 ; then kf ′ (z)k → 0 as |z| → 1, and thus
sup |kf ′ (z)k is achieved at some p ∈ ∆.
Now replace f with f ◦ r where r ∈ Aut(∆) moves p to zero. Replacing
f with af + b with |a| < 1, we can also arrange that f (0) = 0 and kf ′ (0)k =
1; this will only decrease the size of its unramified disk. Then kf ′ (z)k ≤
kf ′ (0)k = 1, and thus f |∆(1/2) ranges in a compact family of nonconstant
analytic functions. Thus the new f has an unramified disk of definite radius;
but then the old f does as well.
76
Rescaling proof of Picard’s theorem. The proof will use the following
remarkably general rescaling theorem. This argument is related to Bloch’s
proof, to Brody’s reparameterization theorem and to other results in com-
plex analysis.
|g′ (z)|
kg′ (z)k∞ = ,
(1 + |g(z)|2 )
and kg′ k∞ = sup kg ′ (z)k over all z ∈ C. This is the norm of the derivative
from the scaled Euclidean metric ρ∞ = 2|dz| to the spherical metric. Note
that g(z) = exp(z) has kg′ k∞ = 1/2; a function with bounded derivative
can be rather wild.
Similarly, for g : ∆(R) → C,b we define
1 − |z/R|2
kg′ (z)kR = |g′ (z)| ·
1 + |g(z)|2
This is the derivative from a suitably rescaled hyperbolic metric ρR on ∆(R)
b Clearly ρR → ρ∞ uniformly on compact sets. Its key property is that
to C.
b stabilizing ∆(R).
k(g ◦ A)′ kR = kg ′ kR for all A ∈ Aut(C)
We also note that the set of maps with uniformly bounded derivatives
in one of these norms is compact.
Proof of Theorem 4.37. Let us first consider an arbitrary nonconstant
analytic function f (z) and a radius R > 0. We claim there exists an S ≥ R
and an A ∈ Aut(C) b such that g = f ◦ A is analytic on ∆(S), and
77
Now just let g(z) = (f ◦ B)(z/M ), and S = RM .
Applying this claim to fn and a sequence Rn → ∞, we obtain Sn → ∞
and maps gn = fn ◦ An with kgn′ (0)k∞ = 1 and kgn′ kSn ≤ 1. Now pass to a
convergent subsequence.
Classical proof. The classical proof of the Little Picard Theorem is based
on the fact that the universal cover of C − {0, 1} can be identified with the
upper halfplane.
To see this, it is useful to start by considering the subgroup Γ0 ⊂ Isom(∆)
generated by reflections in the sides of the ideal triangle T with vertices
{1, i, −1}. For example, z 7→ z is one such reflection, sending T to −T . By
considering billiards in T , one can see that its translates tile the disk and
thus T is a fundamental domain for Γ0 . Thus the quadrilateral F = T ∪(−T )
is a fundamental domain for the orientation-preserving subgroup Γ ⊂ Γ0 ,
and the edges of −T are glued to the edges of T to give a topological triply-
punctured sphere as quotient.
Now let π : T → H be the Riemann mapping sending T to H and its
vertices to {0, 1, ∞}. Developing in both the domain and range by Schwarz
reflection, we obtain a covering map π : ∆ → C b − {0, 1, ∞}.
Given this fact, we lift an entire function f : C → C − {0, 1} to a map
fe : C → H, which is constant by Liouville’s theorem.
Uniformization of planar regions. Once we know that C b − {0, 1, ∞} is
uniformized by the disk, it is straightforward to prove:
b with |C
Theorem 4.39 The universal cover of any region U ⊂ C b − U| ≥ 3
is isomorphic to the unit disk.
78
Sketch of the proof. Consider a basepoint p in the abstract universal
e → U , and let F be the family of all holomorphic maps
cover π : U
e , p) → (∆, 0)
f : (U
that are covering maps to their image. Using the uniformization of the
triply-punctured sphere, we have that F is nonempty. It is also a closed,
normal family of functions in O(U e ); and by the classical square-root trick,
it contains a surjective function (which maximizes |f ′ (p)|). By the theory
of covering spaces, this extremal map must be bijective.
y 2 = x3 + ax + b.
79
A special feature of the case of cubics is that the solutions (in projective
space) form a group.
f (z + α) = f (z + β) = f (z)
This sum does not quite converge for k = 2; it can be made to converge by
writing
1 X′ 1 1
℘(z) = 2 + 2
− 2·
z (z − λ) λ
Λ
The critical properties of ℘(z) are that is it even, and it has a unique pole
of order 2 on E.
80
Figure 6. The Weierstrass ℘-function for the hexagonal lattice.
81
1. The sum of the residues of f (z) dz over E, or over points in P , is zero.
2. The function f has the same number of zeros as poles. The number
of each is called the degree d = deg(f ).
P P
3. If f has poles p1 , . . . , pd and zeros a1 , . . . , ad in E, then pi = ai
in the group law on E.
Proofs. (1) This follows applying Stokes’ theorem to the closed form f (z) dz
on E − {p1 , . . . , pd }, or by integrating f (z) dz around the boundary of P (we
may assume f has no poles on ∂P .)
(2) This is a general property of proper maps between Riemann surfaces.
For a direct proof, one can also apply the residue theorem to df /f .
(3) This property is special to elliptic curves. Let P = [0, α] × [0, β], and
assume f has no zeros or poles in P . Then we have
X X Z
−1 zf ′ (z) dz
ai − pi = (2πi) ·
∂P f (z)
We wish to show that this quantity lies in Λ. The R integrals over opposite
edges cancel, up to a terms of the form λ(2πi)−1 e f ′ (z)/f (z) dz with λ ∈ Λ.
Since the f is periodic, it has an integral winding number N (e) on each edge,
and these terms have the form N (e)λ ∈ Λ.
We will later see that we may construct an elliptic function with given
zeros and poles subject only to constraint (3).
Pushforward. Here is another way to see property (3). Let f : E → C b
b
be a nonconstant meromorphic function. Then f∗ (dz) = 0, since Ω(C) = 0.
Now choose a path C ⊂ C b running from 0 to ∞ and avoiding the critical
values of f . Then C e = f −1 (C) ⊂ E gives a collection of arcs connecting
(ai ) and (pi ) in pairs, which we can assume have the same indices. We then
have Z Z X
0= f∗ ω = dz = pi − ai mod Λ.
C e
C
82
Let ei = ℘(ci ).
We note that the critical values ei are distinct; indeed, since ℘′ (zi ) = 0,
the function ℘(z) − ei has a double zero at zi , so it cannot vanish anywhere
else. This shows:
(Note: it is not at all easy to say where the two zeros of ℘(z) lie, except in
the case of symmetric lattices.)
Proof. From what we have seen above, the two sides are doubly-periodic
functions with the same zeros and poles. Thus they are multiples of one
another. They are equal since near zero they are both asymptotic to 4/z 6 .
Corollary 5.6 The map π(z) = (℘(z), ℘′ (z)) maps E = C/Λ bijectively to
the smooth projective cubic curve E defined by
Proof. Since E is compact, the function x = ℘(z) maps E onto C; b and for
a given x, the two possible values of y solving the equation above are given
by y = ℘′ (±z).
83
A B
A B
Power series expansion. Let O(d) ⊂ K(E) be the vector space of doubly–
periodic functions with a pole of order d at z = 0 and no other poles.
where X′ 1
Gn = Gn (Λ) = .
λ2n
Λ
84
Proof. Neglecting terms of order O(z 2 ), we have:
1 9G2
℘(z)3 = 6
+ 2 + 15G3 and
z z
2
−2 4 24G2
℘′ (z)2 = 3
3
+ 6G2 z + 20G3 z + · · · = 6 − 2 − 80G3 .
z z z
P
Corollary 5.10 We have ei = 0.
85
Theorem 5.11 The value ℘(z) is real if and only if z lies on one of the
vertical or horizontal lines through (1/2)Λ.
Corollary 5.12 The map ℘|S gives the unique conformal map from S to
−H such that ℘(ci ) = ei for i = 1, 2, 3 and ℘(0) = ∞.
β α+β
c3
c2
0 c1 α
86
Proof. If ζ = ℘(z) then f (ζ) = z, and hence f ′ (℘(z))℘′ (z) = 1. Conse-
quently
1 1 1
f ′ (ζ) = =p =p ·
℘′ (z) 4℘(z)3 − g2 ℘(z) − g3 4ζ 3 − g2 ζ − g3
Thus the theory of elliptic functions emerges as the special case of the
Schwarz-Christoffel formula giving the Riemann map for a rectangle.
4
-2 -1 1 2
-2
-4
Location of roots. Note that the argument shows the roots of 4x3 − g2 x −
g3 = 0 are real, and satisfy e2 < e3 < e1 . We also note that ℘′ (z) ∈ R
along the horizontal lines forming ℘−1 (R), and ℘′ (z) ∈ iR along the vertical
lines, since ℘ most rotate the latter by 90◦ to make them real. Thus only
the horizontal lines map to the real points of the cubic y 2 = 4x3 − g2 x − g3 .
Periods. This picture makes clear the close relationship between the gen-
erators α, β of Λ and the cubic polynomial
namely, we have
Z ∞ Z e3
α 3 −1/2
= (4x − g2 x − g3 ) dx = (4x3 − g2 x − g3 )−1/2 dx,
2 e1 e2
and
Z e2 Z e1
β 3 −1/2
= (g3 + g2 x − 4x ) dx = (g3 + g2 x − 4x3 )−1/2 dx.
2i −∞ e3
87
In all 4 integrals we take the positive square-root; thus, these are simply
integrals of |(℘′ )−1 (x)|.
In each case, the fact that the two integrals agree follows from Cauchy’s
theorem (applied to two homotopic loops in C − {e1 , e2 , e3 }), or via a change
of variables coming from a Möbius transformation that swaps one interval
for the other.
More generally, we have:
The condition on the roots insures that the integrand can be defined
continuously on a neighborhood of γ.
Evaluation of g2 and g3 . The preceding formula sometimes permits the
evaluation of g2 and/or g3 . For example, we have
X′ Z ∞ 6
−6 dx 256π 3 Γ[7/6]6
140 λ =4 √ = ·
1 x3 − 1 Γ[2/3]6
Z[ρ]
This follows from the fact that for Λ = Z[ρ], we have g2 = 0 and g3 > 0
satisfies Z ∞
1 dx
= p ,
2 e1 4x3 − g3
where e1 = (g3 /4)1/3 . Similarly, we have
X′ 64π 2 Γ[5/4]4
60 λ−4 = ·
Γ[3/4]4
λ∈Z[i]
88
Proof. To see that ℘ and ℘′ generate K(E) is easy. Any even function
f :E→C b factors through ℘: f (z) = F (℘(z)), and so lies in C(℘). Any odd
function becomes even when multiplied by ℘′ ; and any function is a sum of
one even and one odd.
To see that the field is exactly that given is also easy. It amounts to
showing that K(E) is of degree exactly two over C(℘), and ℘ is transcen-
dental over C. The first assertion is obvious (else ℘ would be constant), and
if the second fails we would have K(E) = C(℘), which is impossible because
℘ is even and ℘′ is odd.
1 1 1
℘(−z) ℘(−w) ℘(z + w) = 0.
℘′ (−z) ℘′ (−w) ℘(z + w)
Proof. Then the line passes through ∞ which is the origin of E, consistent
with the equation p + (−p) + 0 = 0. (Alternatively, observe that x = ℘(z)
is even and y = ℘(z) is odd.)
89
Figure 10. Image of 2−5 Λ under ℘, for Λ = Z[i] and Z[ρ].
Corollary 5.20 The point 2a can be constructed by taking the line tangent
to E at a, finding its other point of intersection with E, and then negating
its y coordinate.
℘(2z) = f (℘(z)).
The preimages of the critical points of f lie along the images of the horizontal
and vertical lines in C under ℘, so they give a way of visualizing the ℘
function. For more details, see [Mil].
An elliptic function with given poles and zeros. It is natural to try
to construct an elliptic function by forming the Weierstrass product for the
lattice Λ:
Y′ z z z2
σ(z) = z 1− exp + .
λ λ 2λ2
Λ
90
5.2 Aside: Conics and singly-periodic functions
As a point of reference, we describe the parallel theory for conics and rank
one lattices in C.
Let Λ ⊂ C be a discrete subgroup isomorphic to Z. Then we can rescale
so Λ = Z, and form, for k ≥ 2, the singly-periodic functions
∞
X 1
Zk (z) =
−∞
(z − n)k
Just as we did for elliptic functions, we next note that the power series
1 π2z
P (z) = − + ···
z 3
gives
1 π2 1 2π 2
−P ′ (z) = + + ··· and P (z)2 = − + ···
z2 3 z2 3
yielding the differential equation
−P ′ (z) = P (z)2 + π 2 .
Here we have used the fact that both P (z) and P ′ (z) are bounded as
| Im z| → ∞. Put differently, we have:
91
Theorem 5.23 The map π : C → P2 given by
π2
−P ′ (z) = ·
sin2 πz
The fundamental period of Λ can now be expressed as
Z ∞ Z ∞ Z 1
dx dx
= 2 2
= dz = 1.
−∞ y −∞ π + x 0
I.e. the change of variables x = P (z) transforms the integrand into the
standard form dz on C/Z.
Remark. Any smooth conic in P2 is equivalent to the parabola above, so
we have uniformized all conics. The familiar conics x2 + y 2 = 1, x2 − y 2 = 1
and xy = 1 are isomorphic to C/2πZ or C/2πiZ, and are uniformized by
(cos(t), sin(t)), (cosh(t), sinh(t)) and (et , e−t ) respectively. Note that all 3
curves have, over C, two asymptotes, corresponding to the ends of C∗ .
Theorem 5.24 Suppose the cubic polynomial 4x3 +ax+b has distinct roots.
Then there exists a lattice Λ such that (x, y) = (℘(z), ℘′ (z)) satisfies y 2 =
4x3 + ax + b.
92
As a set, we let
π(τ ) = [Z ⊕ Zτ ] ∈ M1 .
φ : Z2 → Λ,
coming from the basis (τ, 1). All other bases for Λ with thesame orientation
as this one are given by (aτ + b, cτ + d), where g = ac db ∈ SL2 (Z). The
marked lattice with this new basis is equivalent to (g(τ ), 1) where
aτ + b
g(τ ) = ·
cτ + d
Thus forgetting the marking altogether is the same as taking the quotient
of H by the action of all g ∈ SL2 (Z), and hence:
T : Zn → Λ ⊂ Rn
with T ∈ SLn (R). Two such lattices are similar iff they differ by a rotation,
i.e. T1 = R ◦ T2 where R ∈ SOn (R). Thus the Teichmüller space of lattices
in Rn is the homogeneous space
93
For n = 2 we have H ∼ = H because SO2 (R) is the stabilizer of τ = i for the
usual action of SL2 (R) on H.
Finally T1 (Zn ) = T2 (Zn ) iff T1 ◦T2−1 gives an isomorphism of Zn to itself,
which shows
Theorem 5.27 The moduli space of lattices in Rn is isomorphic to
94
which each have stabilizer Z/2 and correspond to the vertices of a square;
and the points
√
{ρ, ρ} = {1/2 ± −3/2} = zeros of z 2 − z + 1,
satisfying
(The last fact explains the 4/27.) We should really think of the image as
M0,4 and in particular remember the orbifold structure: Z/2 at F = 1 and
Z/3 at F = 0.
The modular function. We now define a map J : M1 → M0,4 by
associating to any complex torus, the four critical values of the Weierstrass
℘-function. (Note: any degree two map f : X = C/Λ → C b is equivalent, up
to automorphisms of domain and range, to the Weierstrass ℘ function, so
the associated point in M0,4 is canonically determined by X.)
More concretely, given τ ∈ H we define the half-integral points (c1 , c2 , c3 ) =
(1/2, τ /2, (1 + τ )/2) and the corresponding critical values by ei = ℘(ci ).
Then their cross-ratio (together with the critical value at infinity) is given
by:
e3 − e2
λ(τ ) = ·
e1 − e2
For example, we have seen that if τ = iy ∈ iR+ then e2 < e3 < e1 , so
λ(iy) ∈ (0, 1); moreover e3 = 0 and e2 = −e1 for τ = i, and thus λ(i) = 1/2.
The value of λ(τ ) depends only on the ordering of E(2)∗ , the three
nontrivial points of order two on E. Now SL2 (Z) = Aut(Λ) acts on E(2) ∼ =
(Z/2)2 through the natural quotient
95
Now any elliptic element in SL2 (Z) has trace −1, 0 or 1, while the trace
of any element
in Γ(2) must be even. Moreover, trace zero cannot arise:
if g = ac −a
b ∈ Γ(2) then −a2 − bc = 1 implies −a2 = 1 mod 4 which is
impossible.
By assembling 6 copies of the fundamental domain for SL2 (Z) (some cut
into two pieces), one can then show:
We may now state the main result relating lattices and cross-ratios.
Theorem 5.31 The function λ is real on the orbit of the imaginary axis
under SL2 (Z).
This orbit gives the edges of a tiling of H by ideal triangles; see Figure
11.
Corollary 5.32 The map λ gives an explicit Riemann map from the ideal
triangle spanned by 0, 1 and ∞ to H, fixing these three points.
96
Figure 11. Tiling for Γ(2).
4 (λ2 − λ + 1)3
J(τ ) = F (λ(τ )) = ·
27 λ2 (1 − λ)2
The S3 -equivariance of λ then implies:
97
isomorphism class of E = C/(Z ⊕ Zτ ); we have J(τ1 ) = J(τ2 ) iff E1 ∼= E2 ;
and every complex number arises as J(τ ) for some τ .
Classical proof. Here is the classical argument that J is a bijection. The
value of J(τ ) determines a quadruple B ⊂ C b which in turn determines
b
a unique Riemann surface X → C of degree two, branched over B, with
X∼ = C/Z ⊕ τ Z. Thus J(τ1 ) = J(τ2 ) iff the corresponding complex tori are
isomorphic iff τ1 = g(τ2 ) for some g ∈ SL2 (Z). Thus J is injective.
To see it is surjective, we first observe that J(τ + 1) = J(τ ). Now if
Im τ = y → ∞, then on the region | Im z| < y/2 we have
1 X′ 1 1 π2
℘(z) = + − + ǫ(z) = + C + ǫ(z),
z2 (z − n)2 n2 sin2 (πz)
where ǫ(z) → 0 as y → ∞. (In fact we have
X′ π2
ǫ(z) = = O(e−πy ).
n
sin2 (π(z + nτ ))
We really only need that it tends to zero; and we will not need the exact
value of the constant C = −π 2 /3).
Since sin(z) grows rapidly as | Im z| grows, we have
(e1 , e2 , e3 ) = (π 2 + C, C, C) + O(ǫ)
2. f (−1/τ ) = τ 2k f (τ ); and
98
The vector space of all such forms will be denoted by Mk . The product
forms of weight 2k and 2ℓ has weight 2(k + ℓ), so ⊕Mk forms a graded ring.
The first two properties imply that
f (τ ) = F (Z ⊕ τ Z)
satisfies f (τ + 1) = f (τ ) and
f (−1/τ ) = F (Z ⊕ τ −1 Z) = F (τ −1 (Z ⊕ τ Z) = τ 2k f (τ ).
99
then the first two conditions above just say that g∗ ω = ω for all g ∈ SL2 (Z).
(Recall that g′ (z) = (cz +d)−2 .) Since dz = (2πi)−1 dq/q, the third condition
says that ω descends to a form
k
−k dq
ω = (2πi) f (q)
q
with poles of order ≤ k at 0, 1 and ∞. The converse is also true. This shows:
(i) The zeros Z(ω) are invariant under S3 and satisfy |Z(ω)| ≤ k,
when counted with multiplicity, and determine ω up to constant
multiple.
(ii) The poles of ω at 0, 1, ∞ are all of the same order, which can
be k, k − 2, k − 4, etc. We have ω ∈ Sk iff the order of pole is
k − 2 or less.
Part (i) comes from the fact that ω has at most 3k poles, hence at most
3k − 2k = k zeros. The parity constraint in (ii) comes from the fact that
g(z) = 1 − z leaves ω invariant, and g ′ (∞) = −1.
Examples.
100
3. We have dim M2 = 1. The only possibility is Z(ω) = {ρ, ρ}; thus the
quadratic differential
(z 2 − z + 1) dz 2
F2 =
z 2 (z − 1)2
spans M2 . (We have met this differential before in the study of Schwarz
triangle functions.)
4. Similarly, dim M3 = 1; it is spanned by the cubic differential
(z − 2)(z − 1/2)(z + 1) dz 3
F3 = ,
z 3 (z − 1)3
which has zeros at −1, 1/2, 2.
5. The products F22 and F3 F2 span M4 and M5 . This is because S3 has
unique invariant sets with |Z| = 4 and 5.
6. We have dim M6 = 2; it is spanned by F23 and F32 . These two forms
are linearly independent because they have different zero sets.
7. The discriminant
4 dz 6
D6 = (F23 − F32 ) = 4
27 z (z − 1)4
is the first nontrivial cusp form; it has poles of order 4 at (0, 1, ∞) and
no zeros. It spans S6 .
8. Ratios of forms of the same weight give all S3 -invariant rational func-
tions. Indeed, as we have seen, an isomorphism C/S b 3 ∼ = Cb sending
(∞, ρ, 2) to (∞, 0, 1),is given by
F23 (z) 4 (z 2 − z + 1)3
F (z) = = ·
F2 (z)3 − F3 (z)2 27 z 2 (z − 1)2
More generally, any S3 -invariant rational function of degree d can be
expressed as a ratio of modular forms of degree d.
b 3 by substituting
9. If desired, we can regard F2 and F3 as forms on C/S
w = J(z). They then come:
dw2 dw3
F2 = and F3 = ·
4w(w − 1) 8w2 (w − 1)
We also find D6 = (1/432)(w−4 (w − 1)−3 ) dw6 .
101
Cusp forms. We let Sk ⊂ Mk denote the space of cusp forms with poles
of order ≤ k − 2 at 0, 1, ∞.
Every space Mk , k ≥ 2 contains a form of the type F2i F3j which is not a
cusp form; thus Mk ∼
= C ⊕ Sk . By inspection, dim Sk = 0 for k ≤ 5. On the
other hand, any cusp form is divisible by D6 . This shows:
Corollary 5.38 The forms F2i F3j with 2i + 3j = k form a basis for Mk .
Proof. By the preceding Corollary these forms span Mk , and the number
of them agrees with dim Mk as computed above.
Proof. The preceding results shows the natural map from the graded ring
C[F2 , F3 ] to M is bijective on each graded piece.
102
Values of g2 and g3 . We now note that for τ = i, the zeros of 4℘3 −g2 ℘−g3
must look like (−1, 0, 1) and thus g3 (i) = 0. Similarly for τ = ρ the zeros
are arrayed like the cube roots of unity and hence g3 (ρ) = 0.
To determine the values at infinity, we observe that as τ → ∞ we have
G2 (τ ) → 2ζ(4) = π 4 /45
and
G3 (τ ) → 2ζ(6) = 4π 6 /945.
This gives the values g2 (∞) = 60G2 (∞) = (4/3)π 4 and g3 (∞) = 140G3 (∞) =
(8/27)π 6 , and thus
(g23 /g32 )(i∞) = 27.
Proof. First note that this is a ratio of forms of weight 12 and hence a
modular function, i.e. it is invariant under SL2 (Z). Since g2 (∞) 6= 0, it
has a simple pole at infinity, and thus has degree one on M1 . We also have
J(ρ) = 0 since g2 (ρ) = 0, and J(i) = 1 since g3 (i) = 0.
Remarks. One can use residues to determine the exact relationship be-
tween F2 and g2 . Namely, F2 ∼ (1/4) dw2 /w2 , while for q = exp(2πiτ ) we
have dq = 2πiq dτ , and hence
4π 4 dq 2 π 2 dq 2
g2 = 60G2 (τ ) dτ 2 ∼ = − ·
3 (2πiq)2 3 q2
103
If we let j(τ ) = 1728J(τ ), then
X ∞
1
j(τ ) = + 744 + an q n
q 1
with an ∈ Z.
Connections with additive and multiplicative number theory. Incredibly,
we have
∞
Y
∆(q) = (2πi) q (1 − q n )24 .
12
This gives a close connection between the theory of modular forms and the
partition function p(n), since
Y X
(1 − q n )−1 = p(n)q n .
References
[BD] F. Berteloot and J. Duval. Sur l’hyperbolicité de certain
complèmentaires. Enseign. Math. 47(2001), 253–267.
104
[Ko] S. V. Kolesnikov. On the sets of nonexistence of radial limits of
bounded analytic functions. Russian Acad. Sci. Sb. Math. 81(1995),
477–485.
[Mon] M. G. Monzingo. Why are 8:18 and 10:09 such pleasant times? Fi-
bonacci. Quart. 2(1983), 107–110.
105