Basic Complex Analysis
Basic Complex Analysis
1
this d(f dz) = 0. Moreover df /dx = df /dz, where d/dz = 1/2(d/dx +
(1/i)d/dy). We have f analytic if and only if df /dz = 0. Then df =
(df /dz)dz + (df /dz)dz and we see d(f dz) = 0 iff df /dz = 0 iff f is
holomorphic.
Proof 2: (Goursat), assuming only complex differentiability.
R
6. Analyticity and power series. The fundamental integral γ dz/z. The
P n
fundamental power series 1/(1 − z) = z . Put these together with
Cauchy’s theorem, Z
1 f (ζ)dζ
f (z) = ,
2πi γ ζ − z
to get a power series.
P
Theorem: f (z) = an z n has a singularity (where it cannot be analyti-
cally continued) on its circle of convergence |z| = R = 1/ lim sup |an |1/n .
Q P
7. Infinite products: (1 + an ) converges
Q if |an | < ∞. The proof is in
two steps: first, show that when (1+|an |) converges,Q the differences in
successive
P partial Q products bound P the differences for (1 + an ). Then,
show |an | ≤ (1 + |an |) ≤ exp( |an |).
Q
Example: evaluation of ζ(2) = (1/(1 − 1/p2 )), where p ranges over
the primes, converges (to π 2 /6).
2
recurrence relation. Pisot numbers, the golden ratio, and why are 10:09
and 8:18 such pleasant times.
11. Kronecker’s theorem: one need only check that the determinants of the
matrices ai,i+j , 0 ≤ i, j ≤ n are zero for all n sufficiently large.
Similary, the weighted sum of g(z) over the zeros and poles is given by
multiplying the integrand by g(ζ).
Winding numbers of the topological nature of the argument principle:
if a continuous f : ∆ → C has nonzero winding number on the circle,
then f has a zero in the disk.
15. Rouché’s theorem: if |g| < |f | on ∂Ω, then f + g and g have the same
number of zeros-poles in Ω. Example: z 2 + 15z + 1 has all zeros of
modulus less than 2, but only one of modulus less than 3/2.
3
19. Hadamard’s 3-circles theorem: if f is analytic in an annulus, then
log M(r) is a convex function of log r, where M(r) is the sup of |f | over
|z| = 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.
20. The concept of a Riemann surface; the notion of isomorphism; the three
simply-connected Riemann surface C, C b and H.
23. The complex plane C. The notion of metric ρ(z)|dz|. The automor-
phism group√is solvable. Inducing a metric |dz|/|z| on C∗ . The cone
metric |dz|/| z| giving the quotient by z 2 .
27. Stereographic projection preserves circles and angles. Proof for angles:
given an angle on the sphere, construct a pair of circles through the
north pole meeting at that angle. These circles meet in the same angle
at the pole; on the other hand, each circle is the intersection of the
sphere with a plane. These planes meet C in the same angle they meet
a plane tangent to the sphere at the north pole, QED.
4
b the extended complex plane; the Riemann sphere; the
28. Four views of C:
Riemann surface obtained by gluing together two disks with z 7→ 1/z;
the projective plane for C2 .
29. The spherical metric 2|dz|/(1 + |z|2 ). Derive from the fact “Riemann
circle” and the map x = tan(θ/2), and conformality of stereographic
projection.
30. Some topology of projective spaces: RP2 is the union of a disk and a
Möbius band; the Hopf map S 3 → S 2 .
31. Gauss-Bonnet for spherical triangles: area equals angle defect. Prove
by looking at the three lunes (of Rarea 4θ)R for the three angles of a
triangle. General form: 2πχ(X) = X K + ∂X k.
b →C
32. Dynamical application of Schwarz Lemma: let f : C b be a rational
map. Then the immediate basin of any attracting cycle contains a
critical point.
Cor. The map f has at most 2d − 2 attracting cycles.
5
37. Harmonic functions. A real-value function u(z) is harmonic iff u is
locally the real part ofR an analytic function; indeed, harmonic means
d(∗du) = 0, and v = ∗du. Harmonic functions are preserved under
analytic mappings. Examples: electric potential; fluid flow around a
cylinder.
The mean-value principle (u(0) = average over the circle) follows from
Cauchy’s formula, as does the Poisson integral formula (u(p) = visual
average of u).
38. The Schwarz reflection principle: if U = U ∗ , and f is analytic on U ∩H,
continuous and real on the boundary, then f (z) extends f to all of U.
This is easy from Morera’s theorem. A better version only requires
that Im(f ) → 0 at the real axis, and can be formulated in terms of
harmonic functions (cf. Ahlfors):
If v is harmonic on U ∩ H and vanishes on the real axis, then v(z) =
−v(z) extends v to a harmonic function on U. For the proof, use the
Poisson integral to replace v with a harmonic function on any disk
centered on the real axis; the result coincides with v on the boundary
of the disk and on the diameter (where it vanishes by symmetry), so
by the maximum principle it is v.
39. Reflection gives another proof that all automorphisms of the disk ex-
tend to the sphere.
40. Quotient of the cylinder: s(z) = z+1/z gives the quotient isomorphism,
C∗ /(Z/2) ∼
= C. It has critical points at ±1 and critical values at ±2.
Since z 7→ z n commutes with z 7→ 1/z, there are unique polynomial
pn (z) such that s(z n ) = pn (s(z)). Writing z = eiθ , these are essentially
the Chebyshev polynomials; they satisfy 2 cos(nθ) = pn (2 cos θ).
41. 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 .
42. Solving the cubic. Every cubic polynomial is equivalent to p3 (z) =
z 3 − 3z. But this polynomial arises as a quotient of z 7→ z 3 ; that is,
it satisfies s(z 3 ) = p3 (s(z)). Thus we can solve p3 (z) = w easily: the
solution is s(u), where s(u3) = w.
6
The mapping p3 (z) (and indeed all the maps pn ) preserve the ellipses
and hyperbolas with foci ±2, since z 7→ z n preserves circles centered
at z = 0 and lines through the origin. This facilitates visualizing these
polynomials as mappings of the plane to itself.
2. Weierstrass factor. Inspired by the fact that (1−z) exp log 1/(1−z) = 1
and that log 1/(1 − z) = z + z 2 /2 + z 3 /3 + . . ., we set
z2 zp
Ep (z) = (1 − z) exp z + +···+ .
2 p
7
5. Jensen’s formula. Let f (z) be holomorphic on the disk of radius R
about the origin. Then the average of log |f (z)| over the circle of radius
R is given by:
X R
log |f (0)| + log .
|z|
f (z)=0; |z|<R
8
Therefore |Q(z)| = O(|z|d ) for |z| > 1. By Cauchy’s integral formula,
we find Qd+1 (z) = 0 and thus Q is a polynomial (of degree at most d).
9
P
will assume (1/rn )ρ < ∞. (For the general case, just replace ρ with
ρ + ǫ.)
By construction, the Weierstrasss factor
u2 up
Ep (u) = (1 − u) exp u + +···+
2 p
satisfies
up+1 up+2
log Ep (u) = + +···
p+1 p+2
for |u| < 1. Thus for ‘u small’, meaning |u| < 1/2, we have
10
and thus f has order ρ.
The lower bound works the same way, since equations (2.1) and (2.2)
give bounds for |logEp (u)|. All that remains is to prove that
X
log |1 − z/an | = O(r ρ+ǫ )
r/rn >1/2
and thus N(2r) = O(r ρ). Also when log |1 − z/an | is positive it is no
bigger than O(log r) (the constant depending on the smallest an ), and
so the positive terms contribute at most N(r) log r to the sum, which
is O(r ρ+ǫ).
We now have to worry about the terms where z/an is close P to one. To
ρ ρ
this end, we require that |r − rn | > 1/rn for all n. Since 1/rn < ∞,
a set of r of infinite measure remains. Then for any such r, if z/an is
close to one, we still have
and thus log |1 − a/zn | = O(log rn ) = O(log r). Thus the cases where
z/an is close to one also contribute O(N(r) log r) = O(r ρ+ǫ ).
13. Example:
Y z2
sin(πz) = πz 1− 2 .
n6=0
n
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.
11
14. Alternative proof. We have
X ∞
π2 1 π2 π4 2 1
2 = 2
+ + z + · · · =
sin (πz) z 3 15 −∞
(z − n)2
because both sides are periodic and tend to zero as | Im z|− > ∞.
Integrating, we obtain
∞
1 X 1 1
π cot πz = + + ,
z 1
z − n z + n
using the fact that both sides are odd to fix the constant of integration.
On the other hand, the logarithmic derivative of f (z) = sin(πz) is
π cot(πz). Using the fact the logarithmic derivative of 1 − z 2 /n2 is the
same as that of (z − n)(z + n), namely (z − n)−1 + (z + n)−1 we deduce
that g ′ /g = f ′ /f where
∞
Y
z2
g(z) = πz 1− 2 .
1
n
exp(−γz) Y z −1
∞
Γ(z) = 1+ exp(z/n).
z 1
n
Note that this expression contains the reciprocal of the canonical prod-
uct associated to the non-positive integers. The constant γ is chosen
so that Γ(1) = 1; it can be given by:
Xn
γ = lim( 1/k) − log(n + 1).
1
R n+1
This expression is the error in an approximation to 1 dx/x by the
area of n rectangles of base one lying over the graph.
12
Q
16. Let G(z) = ∞ 1 (1 + z/n)e
−z/n
be the canonical product for the set of
negative integers. Evidentally
sin(πz)
G(z)G(−z) = · (2.3)
πz
Moreover, we find G(z − 1) = z exp(az + b)G(z) since both sides have
the same zeros and are of order one. Taking the logarithmic derivative
shows a = 0 and b = γ; that is,
Defining
1
Γ(z) = ,
zeγz G(z)
we obtain Γ(z + 1) = zΓ(z), Γ(0) = Γ(1) = 1, Γ(z) has no zeros, Γ(z)
has simple poles at −1, −2, −3, . . . ; and Γ(n + 1) = n! for n ≥ 0.
Using the fact that
Γ(z)
Γ(z − n) = ,
(z − n) · · · (z − 1)
we find
(−1)n
Res−n (Γ(z)) = ·
n!
17. Sine/Gamma formula: from 2.3 we obtain:
π
= Γ(z)Γ(−z).
z sin(πz)
This is obtained by examining the product formula directly for zeγz G(z) =
1/Γ(z).
Corollary: |Γ(z)| ≤ Re z when Re z > 0.
13
19. Uniqueness Theorem (Wielandt, 1939): If F (z + 1) = zF (z) for Re z >
0, F (1) = 1 and F (z) is bounded on the strip {Re z ∈ [1, 2]}, then
F (z) = Γ(z).
Proof. The functional equation allows one to extend F (z) to a mero-
morphic function on the whole plane, whose poles and their residues
agree with those of Γ(z). Thus G(z) = F (z) − Γ(z) is entire, G(0) =
G(1) = 0 and G(z + 1) = zG(z). Our boundedness assumptions now
imply that G(z) is bounded in the strip S = {Re z ∈ [0, 1]}. Thus
H(z) = G(z)G(1 − z) is also bounded in S. The functional equation
for G implies H(z + 1) = H(z) (as in the sine formula), and thus H is
a constant, which must be zero.
14
23. The Mittag-Leffler theorem. Let an → ∞ in C be a sequence of distinct
points, and let pn (z) be principal parts centered at an . Then there
exists a memorphic function f (z) with poles exactly at an , and with
the prescribed principal parts.
Proof. Let qn (z) be the polynomial given by the power series for pn (z)
about z = 0,P truncated so |pn (z) − qn (z)| < 2−n when |z| < |an |/2.
Then f (z) = (pn (z) − qn (z)) is the desired function.
3 Conformal mapping
1. Normal families: any bounded family of analytic functions is normal,
by Arzela-Ascoli.
15
Corollary: given any quadrilateral Q, the product of the minimum
distances between opposite sides is a lower bound for area(Q).
16
circle and observing that, since |dz|2 = (−i/2)dz dz, the area A of
complement of the image of f is given by:
Z X Z X
i i 2 dz
A=− f df = − (1 − n|bn | ) =π 1− n|bn |2 .
2 S1 2 S1 z
Remark: little is known about optimal bounds for |bn | over Σ. The
area theorem gives |bn | = O(1/n1/2) and it is conjectured that |bn | =
O(1/n3/4 ); see [CJ].
2
11. Proof that |a2 | ≤ 2: first, apply
pthe area theorem to3 conclude |a2 −
2
a3 | ≤ 1. Then consider g(z) = f (z ) = z + (a2 /2)z + · for f (z) =
z + a2 z 2 + · · · ∈ S
17
We may finally conclude that S is compact, and indeed we can obtain
by integration the upper and lower bounds:
r r
2
≤ |f (z)| ≤ ·
(1 + r) (1 − r)2
where f (qi ) = pi .
Proof: Notice that g(z) = (f (z) − pi )αi extends by Schwarz reflection
across qi , where αi = 1/(1 − µi ). This implies
18
and the
P formula results by integration. Note that the residue theorem
gives πµi = 2π which is also consistent geometrically (the sum of
the exterior angles in π.)
R
19. Examples of Schwarz-Christoffel: log(z) =p dζ/ζ (maps to a bigon
R
with external angles of π); sin−1 (z) = dζ/ 1 − ζ 2 (maps to a triangle
with external angles π/2, π/2 and π.)
21. Bloch’s Theorem: there exists a universal R > 0 such that for any
f : ∆ → C with |f ′ (0)| = 1, not necessarily univalent, there is an open
set U ⊂ ∆ (perhaps a tiny set near S 1 ) such that f maps U univalently
to a ball of radius r.
It
√ is known that the best value of R, Bloch’s constant, satisfies 0.433 <
3/4 ≤ R < 0.473. The best-known upper bound comes from the Rie-
mann surface branched with order 2 over the vertices of the hexagonal
lattice.
23. These apparently unrelated theorems can both be proved using the
same idea. (Cf. [BD] and references therein.)
We first consider Bloch’s theorem. Given f : ∆ → C, let
denote the norm of the derivative from the hyperbolic metric to the
Euclidean 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. We can
also shrink and translate f so 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 anayltic
19
functions. Thus the new f has an unramified disk of definite radius;
but then the old f does as well.
b we now let
24. Given g : C → C,
|g ′(z)|
kg ′(z)k∞ = .
(1 + |g(z)|2)
This is the norm of the derivative from the Euclidean metric |dz| to the
spherical metric of curvature 4. Note that g(z) = exp(z) has kg ′k∞ =
1/2; a function with bounded derivative can be rather wild.
Theorem. Let fn : ∆ → C b be a sequence of analytic functions such
′
that kfn (0)k∞ → ∞. There after passing to a subsequence, there is
a nonconstant entire function g : C → C such that g = lim fn ◦ rn .
Moreover, sup kg ′ (z)k∞ < ∞.
Here rn are Möbius transformations such that any compact set K ⊂ C
satisfies rn (K) ⊂ ∆ for all n ≫ 0.
b the norm
Proof. The main idea is to consider, for g : ∆ → C,
|g ′(z)|(1 − |z|2 )
kg ′ (z)k1 =
(1 + |g(z)|2)
|g ′(z)|(1 − |z/R|2 )
kg ′ (z)kR =
(1 + |g(z)|2)
20
Now for any compact set K, we have ρRn → |dz| uniformly on K. Thus
kgn′ (z)k∞ is uniformly bounded on K. Thus by Arzela-Ascolia, we can
extract a uniformly convergent subsequence. This gives the desired
limit g(z).
26. Classical Proof of Little Picard: The key fact is 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 quadratilateral 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.
21
that are covering maps to their image. Using the uniformization of the
triply-punctured shere, 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.
29. Proof of Great Picard. Consider a loop γ around the puncture of the
disk. If f sends γ to a contractible loop on the triply-punctured sphere,
then f lifts to a map into the universal cover H, which implies by
Riemann’s removability theorem that f extends holomorphically over
the origin.
Otherwise, by the Schwarz lemma, f (γ) is a homotopy class that can
be represented by an arbitrarily short loop. Thus it corresponds to a
puncture, which we can normalize to be z = 0 (rather than 1 or ∞).
It follows that f is bounded near z = 0 so again the singularity is not
essential.
4 Elliptic curves
1. Theorem: Any discrete subgroup of (C, +) is isomorphic to Zn with
n = 0, 1 or 2. In the last case, the generators are linearly independent
over R.
22
nontrivial then either Λ ∼
= Z or Λ is a lattice, and in both cases we can
ask to construct all functions with the given periods.
For Λ = 2πiZ the answer is: f (t) = g(exp(t)), where g is meromorphic
on C∗ . For Λ a lattice we will find f (t) = g(℘(t), ℘′ (t)), where g is a
rational function of two variables.
To see ℘(z) is elliptic, use the fact that ℘(z) = ℘(−z) and ℘′ (z) =
−2ζ3 (z) (which implies ℘(z + λ) = ℘(z) + Aλ ).
23
Other ways to construct elliptic functions of degree two: if Λ is gener-
ated by (λ1 , λ2 ), first sum over one period to get:
∞
X 1 π2
f1 (z) = = ;
−∞
(z − nλ1 )2 λ21 sin(πz/λ1 )2
then ∞
X
f (z) = f1 (z − nλ2 )
−∞
where X′ 1
Gn = Gn (Λ) = .
Λ
λ2n
By a straightforward calculation (using the fact that an elliptic function
with no pole is constant) we have:
24
7. Geometry of the Weierstrass map ℘ : X → C. b We have ℘(x) = ℘(y)
iff x = ±y. Thus the critical points of ℘, i.e. the zeros of ℘′ , are the
points of order two: {0, c1 , c2 , c3 }.
Since ℘ has degree two, all its critical points are simple and its critical
values, ∞ and ℘(ci ) = ei are distinct. Here ri are the roots of the cubic
equation p(x) = 4x3 − g2 x − g3 .
The map z 7→ (℘(z), ℘′ (z)) sends X − {0} bijectively to the affine cubic
curve y 2 = p(x). Thus ℘ “uniformizes” this plane curve of genus 1.
8. Real case. Now suppose Λ = Zλ1 ⊕ Zλ2 with λ1 > 0 and λ2 ∈ iR+ .
Then Λ is invariant under both negation and complex conjugation.
Thus we have
℘(z) = ℘(−z) = ℘(z).
It follows that ℘ must be real on the locus R ⊂ X which is fixed
under z 7→ z or z 7→ −z. This locus is covered in C by the vertical
and horizontal lines through the points of (1/2)Λ. In particular it the
rectangle S with vertices the points of order two, {0, c1, c2 , c3 }.
It is now easy to check that R = ℘−1 (R ∪ {∞}), and (using the fact
that ℘(z) = 1/z 2 + · · · ) that ℘ is decreasing along [0, λ1 /2] and hence
maps S homeomorphically to −H.
The inverse map on the upper halfplane, f : H → e1 + S, satisfies
f (℘(z)) = z and thus f ′ (℘(z))℘′ (z) = 1. Consequently:
1
f ′ (℘) = (℘′ )−1/2 = p ,
4℘3 − g2 ℘ − g3
25
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(X) is of degree exactly two over C(℘), and ℘ is
transcendental over C. The first assertion is obvious, and if the second
fails we would have K(X) = C(℘), which is impossible because ℘ is
even and ℘′ is odd.
11. 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 λ 2λ2
Then (log σ)′′ = −℘(z), from which it follows that σ(z+λ) = σ(z) exp(aλ +
bλ z). From this it is easy to see that
σ(z − a1 ) . . . σ(z − an )
σ(z − p1 ) . . . σ(z − pn )
26
P P
defines an elliptic function whenever ai = pi , and therefore this
is the only condition imposed on the zeros and poles of an elliptic
function.
12. The moduli space of complex tori, M1 , classifies Riemann surfaces of
genus 1 up to isomorphism.
Since any elliptic curve has the form X = C/Λ, we see that M1 is
the same as the lattice in C modulo the action of C∗ . If a lattice
Λ = Zλ1 ⊕ Zλ2 is given an oriented basis, then it is similar to Z ⊕ Zτ
where τ ∈ H. Thus H is the space of similarity classes of framed
lattices.
Any two oriented bases of Λ ∼
= Z2 are related by an element of SL2 (Z).
That is, if
Z ⊕ Zτ1 = α(Z ⊕ Zτ2 )
with α ∈ C∗ , then we can write
! !
τ1 aτ2 + b
=α
1 cτ2 + d
for some g = ( ac db ) in SL2 (Z), and thus
τ1 = (aτ + 2 + b)/(cτ2 + d).
(The fact that ad − bc = 1 instead of −1 insures that Im τ1 > 0 if
Im τ2 > 0.) Conversely, any two points of H in the same orbit of
SL2 (Z) give similar lattices and hence isomorphic complex tori. This
shows:
27
Theorem 4.2 The region | Re τ | ≤ 1/2, |τ | > 1 in H is a fundamental
domain for SL2 (Z).
Mord b
0,4 = C − {0, 1, ∞}
M0,4 = Mord
0,4 /S3 .
Here Mord
0,4 is the moduli space of ordered quadruples of distinct points
b up to the action of Aut(C).
on C, b Any such quadruple has a unique
representative of the form (∞, 0, 1, λ), giving a natural coordinate for
this moduli space.
If we reorder the quadruple, the cross-ratio changes, ranging among the
six values
(There is a natural action of S4 , but the Klein 4-group Z/2 × Z/2 acts
trivially.) There are 5 points fixed under the action of S3 : the points
−1, 1/2 and 2 each have stabilizer Z/2, and correspond to the vertices
of a square; while the points
√
±ω = 1/2 ± −3/2
28
b Let Z/n act
14. Orbifolds. Here are some other examples of quotients of C.
on Cb by z 7→ exp(2πik/n)z. Extend this to an action of the dihedral
group Dn by adding in z 7→ 1/z. Then the quotient in both cases
b again: the (n, n)-orbifold and the (2, 2, n)-orbifold respectively.
is C
The quotient map can be given by F (z) = z n in the first case and by
f (z) = z n + z −n in the second.
15. The modular function. We now define a map J : M1 → M0,4 by asso-
ciating to any complex torus, the four critical values of the Weierstrass
℘-function. (Note: any degree two map f : X = C/Λ → C b is equiva-
lent, 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 = ℘(pi ).
Then their cross-ratio (together with the critical value at infinity) is
given by:
e3 − e2
λ(τ ) = ·
e1 − e2
This depends on an ordering of the points of order two. Now SL2 (Z) =
Aut(Λ) acts on the three points of order two through the natural quo-
tient
0 → Γ(2) → SL2 (Z) → SL2 (Z/2) → 0.
In particular, λ is invariant under the subgroup Γ(2) of matrices equiv-
alent to the identity modulo two.
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:
b ) ∈ Γ(2) then −a2 − bc = 1 implies −a2 = 1 mod 4 which is
if g = ( ac −a
impossible.
Summing up, we have:
29
Theorem 4.4 The map
b − {0, 1, ∞})/S3
J : M1 = H/ SL2 (Z) → M0,4 = (C
1 X′ 1 1 π2
℘(z) = 2 + − + ǫ(z) = + C + ǫ(z),
z (z − n)2 n2 sin2 (πz)
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)
30
16. Picard’s theorem revisited. It is now straightforward, by the theory of
covering spaces, to show:
λ : H/Γ(2) → Mord b
2 = C − {0, 1, ∞}
This gives the one line proof of the Little Picard Theorem: ‘consider
λ−1 ◦ f : C → H.’ Cf. [Bol, p.39–40]:
17. Periods. The lattice Λ associated to 4 points can also be covered ex-
plicitly (up to similarity) as follows.
Let f : X = C/Λ → C b be any degree two map, with branch locus
B ⊂ C, b such that f (z) = f (−z). Note that the quadratic differential
2
dz on X = C/Λ is invariant under z 7→ −z. Thus it descends to
b It has poles at the branch
a well-defined differential q(w) dw 2 on C.
points B, and these four points uniquely determine q up to scale.
In the case w = ℘(z), we have dw 2 = ℘′ (z)2 dz 2 and hence
dw 2 dw 2
q(w) dw 2 = = ·
℘′ (z)2 4w 3 − g2 w − g3
31
Now we have a map π1 (C b − B) → Z/2 defining the covering X. Let K
√
be the kernel of this map. Then q has a single-valued branch along
any [γ] ∈ K, and [γ] lifts to a loop δ ∈ π1 (X). Consequently:
Z Z
√
q = ± dz ∈ Λ.
γ δ
All elements of Λ arise in this way, and thus Λ can be recovered as the
√
periods of q.
18. Modular forms: algebraic perspective. We now wish to classify holo-
morphic k-forms ω = ω(z) dz k on the orbifold
J
= M0,4 ∼
M1 ∼ = C,
which are ‘holomorphic’ at the orbifold and cusp points. The set of all
such ω forms the finite-dimensional vector space Mk , called the space
of modular forms of weight 2k.
By removable singularities, ω can be considered an ordinary rational
b The condition that it is holomorphic at an orbifold point
k-form on C.
then becomes the following.
32
Definition. A rational form ω(z) dz k is a modular form of weight 2k
on M0,4 ∼
= C if it has poles of order at most k, k/2 and 2k/3 at z = ∞,
z = 1 and z = 0.
We let Mk denote the finite-dimensinoal space of all modular forms of
weight 2k. The total number of poles of ω as a classical form is 2k.
The allowed orders of zeros and poles, and the dimensions of Mk , are
given in Table 1 for small values of k.
Examples:
b with just a single pole, so
(a) There is no meromorphic 1-form on C
dim M1 = 0.
(b) The quadratic differential F2 = dz 2 /(z(z − 1)) spans M2 . Note
that this differential has a simple zero, in the orbifold sense, at
z = 0. It is covered by the differential on the triply-punctured
spheres with double poles of equal residue at each puncture.
33
(c) The cubic differential F3 = dz 3 /(z 2 (z − 1)) spans M3 . This time
there is a zero of order 2 at z = 1.
(d) The products F22 and F3 F2 span M4 and M5 .
(e) The forms F23 and F32 span M6 , which is two-dimensional.
(f) The discriminant
dz 6
D6 = F23 − F32 = 4 ∈ M6
z (z − 1)3
vanishes at infinity, but has no others zeros or poles (in the orbifold
sense).
(g) Ratios of forms of the same weight give all rational functions.
Indeed, we have
z = F23 /D6 ,
and thus any rational function of degree d can be expressed as a
ratio of modular forms of degree k = 6d.
19. Cusp forms. We let Mk0 ⊂ Mk denote the space of cusp forms vanishing
at infinity. Every space Mk , k ≥ 2 contains a form of the type F2iF3j
which is not a cusp form; thus Mk ∼
= C ⊕Mk0 . By inspection, dim Mk0 =
0 for k ≤ 5. On the other hand, any cusp form is divisible by D6 . This
shows:
Corollary 4.10 The forms F2i F3j with 2i + 3j = k form a basis for
Mk .
34
Corollary 4.11 The ring of modular forms M is isomorphic to the
polynomial ring C[F2 , F3 ].
Proof. The preceding results shows the natural map from the graded
ring C[F2 , F3 ] to M is bijective on each graded piece.
F (−1/τ ) = f (Z ⊕ τ −1 )Z) = f (τ −1 (Z ⊕ τ Z) = τ 2k F (τ ).
35
(b) In particular, for k ≥ 2 the Eisenstein series
X′
Gk (τ ) = (n + mτ )−2k
are modular forms of weight 2k. Evidentally Gk converges to
2Z(2k) as Im τ → ∞, so these are holomorphic at infinity but not
cusp forms.
(c) Similarly for the normalized functions g2 = 60G2 , g3 = 140G3 .
Theorem 4.12 Under the isomorphism M1 ∼ = M0,4 , an analytic mod-
ular form is the same as an algebraic modular form.
Corollary 4.13 The ring of modular forms is isomorphic to C[g2 , g3 ].
21. 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 = ω
the zeros are arrayed like the cube roots of unity and hence g3 (ω) = 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) = π 6 /945.
This gives the values g2 (∞) = 60G2 (∞) = 4π 4 /3 and g3 (∞) = 140G3 (∞) =
(4/27)π 6, and thus
(g23 /g32)(∞) = 27.
22. The cusp form ∆ and the modular function J. By the preceding cal-
culation, the discriminant
∆(τ ) = g23 (τ ) − 27g3(τ )2
is a cusp form of weight 12. We have seen such a form is unique up to
a scalar multiple, and is nonvanishing everywhere except for a simple
zero at infinity.
Theorem 4.14 We have J(τ ) = g23(τ )/∆(τ ).
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(ω) = 0 since g2 (ω) = 0, and J(i) = 1 since g3 (i) = 0.
36
23. The product expansion: for q = exp(2πiτ ), we have
∞
Y
∆(q) = q (1 − q n )24 .
1
References
[BD] F. Berteloot and J. Duval. Sur l’hyperbolicité de certain
complèmentaires. Enseign. Math. 47(2001), 253–267.
37