B Tech Maths
B Tech Maths
DAWEI CHEN
Contents
1. Classical Topics 2
1.1. Complex numbers 2
1.2. Differentiability 2
1.3. Cauchy-Riemann Equations 3
1.4. The Riemann Sphere 4
1.5. Möbius transformations 5
1.6. Integration 6
1.7. Cauchy’s Theorems and Applications 7
1.8. Properties of Analytic Functions 9
1.9. The Winding Number 11
1.10. Singularities 12
1.11. Laurent Series 13
1.12. Residue Theory 15
1.13. Normal Families 18
1.14. The Riemann Mapping Theorem 19
1.15. Schwarz Reflection 21
1.16. Analytic Continuation 23
1.17. Analytic Covering Maps 24
1.18. The Picard Theorems 25
1.19. Infinite Products 27
2. Riemann Surfaces 34
2.1. Preliminaries 35
2.2. Sheaves 37
2.3. Maps between sheaves 38
2.4. Stalks and germs 39
2.5. Sheaf cohomology 40
2.6. Holomorphic vector bundles 45
2.7. Vector bundles and locally free sheaves 46
2.8. Divisors 46
2.9. Line bundles 48
2.10. Sections of a line bundle 50
2.11. The Riemann-Roch formula 52
2.12. The Riemann-Hurwitz formula 53
2.13. Genus formula of plane curves 55
2.14. Base point free and very ample line bundles 56
1. Classical Topics
1.1. Complex numbers.√ A complex number is expressed as z = x + iy, where
x, y ∈ R and i2 = −1. We use C to denote the set of complex numbers, which
geometrically corresponds to the (x, y)-coordinate plane R2 . Define the conjugate
of z to be z = x − iy, and the norm of z satisfies |z|2 = zz = x2 + y 2 .
If z 6= 0, z can also be represented by the polar coordinate z = r(cos θ + i sin θ),
where r = |z|, cos θ = x/r and sin θ = y/r. By comparing power series expansions,
we conclude the Euler formula
eiθ = cos θ + i sin θ
for all θ ∈ R. In this form z = reiθ , which is convenient for multiplication:
z1 z2 = |z1 ||z2 |ei(θ1 +θ2 ) .
Such θ is called the argument of z, which is well-defined modulo multiples of 2π.
Note that z = e(log r)+iθ (where log r is the real log function), hence one can
define log z = log r + iθ, which is a multi-valued function
√ since θ and θ + 2πn for
n ∈ Z are both arguments of z. Similarly for z 6= 0, n z has n different values,
given by
r1/n ei(θ+2kπ)/n
for 0 ≤ k ≤ n − 1.
1.2. Differentiability. Let U be an open subset of R2 . Let f (z) : U → C be a
complex-valued function. We may also write
f (z) = u(z) + iv(z) = u(x, y) + iv(x, y),
2
where u, v : R → R. Recall the difference-quotient definition of derivative
f (w) − f (z)
f 0 (z) = lim .
w→z w−z
Definition 1.1. If f 0 (z) exists and continuous for all z ∈ U , we say that f is
holomorphic on U , and denote it by f ∈ H(U ). A function f ∈ H(C) is called
entire.
Example 1.2. Polynomials and ez are entire functions. A rational function (quo-
tient of two polynomials) is holomorphic away from the zeros of its denominator.
Definition 1.3. We say that f is analytic if f can be represented by a convergent
power series expansion on a neighborhood of every point in U , and we denote it by
f ∈ A(U ).
Later on we will show that H(U ) = A(U ) and the condition that f 0 (z) is con-
tinuous can be dropped from the definition of being holomorphic. Hence one can
freely interchange the notion of being analytic and holomorphic. Below we verify
one direction of their equivalence.
Lemma 1.4. In the above setting, we have A(U ) ⊂ H(U ).
3
converges absolutely and uniformly on (compact subsets of) the disk |z − z0 | < r,
so does the formal derivative series
X∞
nan (z − z0 )n−1 .
n=0
As in calculus, under these conditions differentiation is interchangeable with sum-
mation, hence
∞
X
f 0 (z) = nan (z − z0 )n−1 ,
n=0
which is convergent and continuous for |z − z0 | < r.
By the same argument we have
∞
X
f (k) (z) = n(n − 1) · · · (n − k + 1)an (z − z0 )n−k
n=0
f (n) (z0 )
for |z − z0 | < r, which shows that an = n! for all n ≥ 0.
1.3. Cauchy-Riemann Equations. Abuse notation slightly by writing f (z) =
f (x + iy) = f (x, y). If f 0 (z) exists for some z = x + iy ∈ U , then one can approach
z along the x or y directions, respectively, and obtain that
f (z + h) − f (z) f (x + h, y) − f (x, y) ∂f
f 0 (z) = lim = lim = ,
h∈R
h→0
h h∈R
h→0
h ∂x
f (z + ih) − f (z) f (x, y + h) − f (x, y) ∂f
f 0 (z) = lim = lim −i = −i .
h∈R
h→0
ih h∈R
h→0
h ∂y
Hence we conclude one version of the Cauchy-Riemann equation
∂f ∂f
= −i .
∂x ∂y
If we write f (x, y) = u(x, y) + iv(x, y), then similarly we obtain that
f 0 (z) = ux + ivx = −iuy + vy ,
∂u
where ux = ∂x , etc. The Cauchy-Riemann equation is equivalent to the following
equations
ux = vy , uy = −vx .
If we treat f as a function from R2 to R2 , then the Jacobian matrix of f is
u uy
df = x .
vx vy
In this setting the Cauchy-Riemann equation is equivalent to the property that
df = ρA
for some ρ ≥ 0 and A ∈ SO(2, R). If ρ 6= 0, i.e. if f 0 (z) 6= 0, then df is the
composition of rotation by A and dilation by ρ, hence as a map U → R2 , f preserves
angles between smooth arcs through z and orientation. In general, a map preserving
4
angle and orientation at each point is called conformal. We thus conclude that for
f ∈ H(U ), if f 0 (z) 6= 0 for all z ∈ U , then f is conformal on U .
Conversely if the Cauchy-Riemann equations hold for f = u + iv, we can write
u(x + s, y + t) = u(x, y) + ux (x, y)s + uy (x, y)t + o(|s|, |t|),
v(x + s, y + t) = v(x, y) + vx (x, y)s + vy (x, y)t + o(|s|, |t|).
Using ux = vy , uy = −vx and writing w = s + it, we obtain
f (z + w) − f (z) = (u(x + s, y + t) − u(x, y)) + i(v(x + s, y + t) − v(x, y))
= ux (x, y)s + uy (x, y)t + i(vx (x, y)s + vy (x, y)t) + o(|w|)
= ux (x, y)s − vx (x, y)t + i(vx (x, y)s + ux (x, y)t) + o(|w|)
= (ux (x, y) + ivx (x, y))(s + it) + o(|w|)
= (ux (z) + ivx (z))w + o(|w|).
0
It follows that f (z) exists and equals ux (z) + ivx (z) = vy (z) − iuy (z) as desired.
We have thus proved the following result.
Theorem 1.5. A complex-valued function f (with continuous partials ux , uy , vx , vy )
is holomorphic iff it satisfies the Cauchy-Riemann equation.
Another way to record the Cauchy-Riemann equation is as follows. Treat z, z as
independent variables and write x = (z + z)/2, y = (z − z)/(2i). Then we have
∂f ∂f ∂x ∂f ∂y 1 ∂f ∂f
= + = −i ,
∂z ∂x ∂z ∂y ∂z 2 ∂x ∂y
∂f ∂f ∂x ∂f ∂y 1 ∂f ∂f
= + = +i .
∂z ∂x ∂z ∂y ∂z 2 ∂x ∂y
Then the Cauchy-Riemann equation is equivalent to
∂f
= 0,
∂z
which is also equivalent to
∂f ∂f
= .
∂z ∂x
Example 1.6. f (z) = z does not satisfy the Cauchy-Riemann equation, hence it
is not holomorphic.
1.4. The Riemann Sphere. Let C∞ be the one-point compactification of C, by
adding a point at ∞. The neighborhoods of ∞ are the complements of compact
sets in C. It is clear that C∞ is homeomorphic to the 2-sphere S 2 .
A complex structure can be given to C∞ such that it becomes a complex one-
dimensional manifold, i.e., a Riemann surface. Cover C∞ by two charts (C, z) and
(C∞ − {0}, z −1 ), both of which are isomorphic to C. On the overlap C∗ = C − {0},
the transition function is given by the map z 7→ z −1 , which is biholomorphic.
This description is equivalent to saying that C∞ is the complex projective line
CP1 . By definition, CP1 = (C2 − {(0, 0)})/C∗ , with homogeneous coordinates
{[x, y]}, where x, y ∈ C, not both zero, and [x, y] represents the equivalence class
of (x, y) modulo simultaneous scaling. Note that CP1 can be covered by two charts
U = {[x, y]} with y 6= 0 and V = {[x, y]} with x 6= 0. Then the coordinate on U
is u = x/y and on V is v = y/x, with the transition function v = 1/u, which is
exactly the same as above.
5
From now on we will use C∞ , S 2 , or CP1 interchangeably for the notation of the
Riemann sphere.
Using the above charts, one can extend the notion of holomorphic maps from
C → C to C∞ → C∞ . There are three cases. If f (z0 ) = ∞ for some z0 ∈ C, then we
require that 1/f (z) is holomorphic around z0 . To make sense of being holomorphic
at z = ∞, if f (∞) 6= ∞, we require that f (1/z) is holomorphic around z = 0.
Finally if f (∞) = ∞, we require that 1/f (1/z) is holomorphic around z = 0.
Example 1.7. Any rational function on C extends to a holomorphic map from
C∞ → C∞ .
1.5. Möbius transformations. For a matrix
a b
A= ∈ GL(2, C),
c d
define the Möbius transformation associated to A by
az + b
TA (z) =
cz + d
as a map from C∞ → C∞ .
It is easy to check that TA ◦ TB = TA·B and TA−1 = TA−1 . In particular, TA is
biholomorphic and is thus an automorphism of C∞ .
Note that if A e = λA for λ ∈ C∗ , then T e = TA . Conversely if TA = T e for
A A
A, Ae ∈ SL(2, C), then
ad − bc ade − ebe
c
TA0 (z) = = TA0e(z) =
e
2
,
(cz + d) cz + d)2
(e e
hence cz + d = ±(e cz + d)
e since the numerators are 1 by assumption. It follows
that A and Ae are the same matrices in SL(2, C) possibly up to a choice of sign.
Therefore, we conclude the following.
Proposition 1.8. The family of all Möbius transformations is the same as PSL(2, C) =
SL(2, C)/{± Id}.
Lemma 1.9. Every Möbius transformation is the composition of four elementary
maps:
(1) Translation z 7→ z + z0 .
(2) Dilation z 7→ λz, λ > 0.
(3) Rotation z 7→ eiθ z, θ ∈ R.
(4) Inversion z 7→ 1/z.
Proof. If c = 0, then
a b
z+ ,
TA (z) =
d d
which is the composition of translation, dilation and possibly rotation.
If c 6= 0, then
bc − ad 1 a
TA (z) = 2 d
+ .
c z+ c c
Example 1.10. The Möbius transformation z 7→ z−1 z+1 takes the right half-plane
onto the unit disk. In particular, the imaginary axis iR maps onto the unit circle.
6
If we include all lines in R2 as circles (of infinite radius), then it is not hard to
verify the following.
Lemma 1.11. Möbius transformations take circles onto circles.
Proof. This can be checked using the four elementary maps in the composition of
a Möbius transformation.
Note that the equation T (z) = z is (at most) a quadratic equation for any
Möbius transformation T . Hence T can have at most two fixed pointed unless it is
the identity. For instance, z 7→ z + 1 has only one fixed point at ∞, while z 7→ 1/z
has two fixed points at ±1.
Lemma 1.12. A Möbius transformation is determined completely by its action on
three distinct points on C∞ . Moreover, given distinct z1 , z2 , z3 ∈ C∞ , there exists a
unique Möbius transformation T such that T (z1 ) = 0, T (z2 ) = 1, and T (z3 ) = ∞.
Proof. If S and T are two Möbius transformations that agree on three distinct
points, then S · T −1 has at least three fixed points, hence S · T −1 is the identity
transformation, and S = T .
For the other statement, one can first apply Möbius transformations to assume
that zi are not ∞. Then define
z − z1 z2 − z3
T (z) = · ,
z − z3 z2 − z1
which does the job.
Definition 1.13. The cross ratio of four distinct points z1 , z2 , z3 , z4 ∈ C∞ is
defined as
z1 − z3 z1 − z4 z1 − z3 z2 − z4
(z1 , z2 , z3 , z4 ) = : = · .
z2 − z3 z2 − z4 z2 − z3 z1 − z4
Lemma 1.14. The cross ratio is preserved under Möbius transformations. More-
over, four distinct points lie on a circle iff their cross ratio is real.
Proof. For a Möbius transformation T , let wj = T (zj ) for distinct zj for j =
2, 3, 4. Consider the cross ratios S1 (z) = (z, z2 , z3 , z4 ) and S2 ◦ T (z) = S2 (w) =
(w = T (z), w2 , w3 , w4 ) as Möbius transformations. Since S1 (z2 ) = 1, S1 (z3 ) = 0,
S1 (z4 ) = ∞, S2 ◦ T (z2 ) = S2 (w2 ) = 1, S2 (w3 ) = 0, S2 (w4 ) = ∞, we conclude that
S1 = S2 ◦ T , hence
(z1 , z2 , z3 , z4 ) = S1 (z1 ) = S2 ◦ T (z1 ) = (T (z1 ), T (z2 ), T (z3 ), T (z4 )).
for four distinct z1 , z2 , z3 , z4 ∈ C∞ . The other statement follows from taking four
points in the real axis (as a generalized circle).
1.6. Integration. Recall that a smooth path is a C 1 -curve γ : [0, 1] → R2 , i.e., γ
is differentiable with continuous derivatives. In many cases we will also consider
piecewise smooth paths. If γ(0) = γ(1), we say that γ is closed.
Definition 1.15. For any complex-valued continuous function f on U ⊂ C and a
path γ in U , define
Z Z 1
f (z)dz = f (γ(t))γ 0 (t)dt.
γ 0
H
If γ is closed, then we also write γ f (z)dz for this integral.
7
Example 1.17. For n 6= −1, z has an antiderivative (z n+1 )/(n + 1), hence for
n
If n = −1, consider the circle parameterized by γ(t) = re2πit for t ∈ [0, 1]. Then
I Z 1 Z 1
1 1 2πit 0
dz = 2πit
(re ) dt = 2πi dt = 2πi.
γ z 0 re 0
1.7. Cauchy’s Theorems and Applications. The next few results form a corner
stone of complex integration.
Theorem 1.18 (Cauchy’s Integration Theorem). Let γ1 and γ2 be two C 1 -paths
with same endpoints such that they are C 1 -homotopic in U . Then for any f ∈ H(U ),
we have Z Z
f (z)dz = f (z)dz.
γ1 γ2
Proof. We can triangulate the homotopy region into a union of small subregions
with boundary loops ηj such that
Z Z XI
f (z)dz − f (z)dz = f (z)dz.
γ1 γ2 j ηj
H
Hence it suffices to prove that η f (z)dz = 0 for sufficiently small loops η.
Recall that d(f (z)dz) = 0, hence f (z)dz is a closed form. By Poincaré’s lemma,
0
any closed form is locally exact,
H hence on the region bounded by η, f (z)dz = F (z)
for some F , and consequently η f (z)dz = 0.
Alternatively, write f (z) = u(x, y) + iv(x, y). Let V be the region with (positive
oriented) boundary curve η. Then
I I
f (z)dz = f (z)dz
η ∂V
I
= (u + iv)dx + (iu − v)dy
∂V
ZZ
= (−uy − ivy + iux − vx )dxdy
V
= 0,
where we used Green’s theorem and the Cauchy-Riemann equations in the last two
steps.
8
Theorem 1.24 (Uniqueness Theorem). Let f ∈ H(U ). Then the following are
equivalent:
(1) f ≡ 0.
(2) For some z0 ∈ U , f (n) (z0 ) = 0 for all n ≥ 0.
(3) The set {z ∈ U | f (z) = 0} has an accumulation point in U .
In particular, a non-constant holomorphic function has isolated zeros.
10
Proof. The implication (1) =⇒ (2), (3) is clear. Using the power series expansion of
f at z0 , we see that (2) =⇒ (1). Finally suppose (3) holds. Let {zn } be a sequence
of zeros of f in U that converges to z0 as n → ∞. Prove by contradiction. Suppose
N is the minimal nonnegative integer such that f (N ) (z0 ) 6= 0. Then
∞
X
f (z) = an (z − z0 )n = aN (z − z0 )N (1 + O(z − z0 ))
n=0
as z → z0 , where aN 6= 0. It implies that f (z) 6= 0 in a small punctured neighbor-
hood of z0 , contradicting that z0 is a limit of zeros of f . Hence we conclude that
(3) =⇒ (2).
Theorem 1.25 (Open Mapping Theorem). For a non-constant function f ∈ H(U )
and every z0 ∈ U , there exists a positive integer n and a holomorphic function g
locally at z0 such that g(z0 ) 6= 0 and
f (z) = f (z0 ) + (z − z0 )n g(z).
We say that z0 is a zero of order n. In particular, f maps open sets to open sets.
Proof. Take n ≥ 1 to be the first nonzero derivative for f (n) (z0 ) and then we get
the desired expression. Since g(z0 ) 6= 0, locally around z0 one can take a single-
valued holomorphic nth root (not unique) h of g, i.e., g(z) = (h(z))n . Moreover, the
derivative of (z−z0 )h(z) is nonzero at z0 , hence (z−z0 )h(z) is a local diffeomorphism
by the usual inverse function theorem. Then f (z) − f (z0 ) is the composition of this
diffeomorphism with w 7→ wn , which implies that f maps open sets to open sets.
In fact, this proof shows that around a zero z0 of order n, f behaves like a branched
cover of degree n totally ramified at z0 (to be discussed later).
Theorem 1.26 (Maximum Modulus Principle). Let U ⊂ C be a connected open
set and f ∈ H(U ). If there exists z0 ∈ U such that |f (z)| ≤ |f (z0 )| for all z ∈ U ,
then f is constant.
Proof. If f is not constant, then f maps U to an open set, hence f (z0 ) lies in the
interior of f (U ), contradicting that f has maximum modulus at z0 .
Theorem
H 1.27 (Morera’s Theorem). Let f be a continuous function on U . If
γ
f = 0 for all closed paths γ in U , then f ∈ A(U ).
Proof. Fix z0 ∈ U . Define Z z
g(z) = f (w)dw
z0
which is independent of the integration path by the assumption on f . Since g 0 (z) =
f (z), g ∈ H(U ) = A(U ), hence all derivatives of g are analytic on U , including f
as the first derivative of g.
Theorem
R 1.28. If f is continuous on U , then f admits an antiderivative on U iff
γ
f = 0 for all closed paths γ in U .
Proof. If F 0 (z) = f (z) on U , then γ f = 0 by the Fundamental Theorem of Cal-
R
R Rz
culus. Conversely if γ f = 0 for all closed paths γ in U , define F (z) = z0 f (w)dw
for a fixed z0 ∈ U . Then F (z) is well-defined and F 0 (z) = f (z).
Finally we show that the condition of continuous derivatives can be dropped
from the definition of holomorphic functions.
11
Proof. Define
t
γ 0 (s)
Z
g(t) = ds
0 γ(s) − z0
for t ∈ [0, 1]. One checks that
d −g(t)
e (γ(t) − z0 ) = 0,
dt
hence
e−g(t) (γ(t) − z0 ) = e−g(0) (γ(0) − z0 ) = γ(0) − z0 .
Taking t = 1, we have
e−2πin(γ;z0 ) (γ(1) − z0 ) = γ(0) − z0 .
Since γ(1) = γ(0), we conclude that n(γ; z0 ) is an integer. Apparently n(γ; z0 ) is
locally constant, hence constant on each connected component of C − γ. If z0 is
outside of the region R bounded by γ, then 1/(z − z0 ) is holomorphic on R. It
follows that n(γ; z0 ) = 0 on the unbounded component of C − γ.
Corollary 1.31. If γ1 and γ2 are homotopic closed paths in C\{z0 }, then n(γ1 ; z0 ) =
n(γ2 ; z0 ).
Theorem 1.32 (Winding Number Version of Cauchy’s Integral Formula). Let γ be
a closed path homotopic to a point in U , and let f ∈ H(U ). Then for any z0 ∈ U \γ,
we have Z
1 f (z)
n(γ; z0 ) · f (z0 ) = dz.
2πi γ z − z0
Proof. For fixed z0 , define
f (z) − f (z0 )
z 6= z0
g(z) = z − z0
0
f (z0 ) z = z0 .
(3) z0 is essential iff for all 0 < ε 1, the set f (D(z0 , ε)∗ ) is (open) dense in C,
where D(z0 , ε)∗ denotes the punctured disk at z0 contained in U .
13
Next we show that the integral over γs2 contributes to an for n ≥ 0 and the
integral over γs1 contributes to an for n < 0. Consider the expansion
∞
(z − z0 )n
X
if |w − z0 | > |z − z0 |
(w − z0 )n+1
1 1
n=0
= = ∞
w−z (w − z0 ) − (z − z0 ) X (w − z0 )n
− if |w − z0 | < |z − z0 |.
(z − z0 )n+1
n=0
Since these power series converge absolutely and uniformly on the respective inte-
gration paths, inserting them into the above integral and interchanging the sum-
mation and integration gives the desired expansion as well as the expression of
an .
Alternatively, the value of an can be verified by dividing the Laurent series by
(z − z0 )n+1 and then integrating.
Proof. If f (z) ∈ C for all z ∈ C∞ , then f is entire and bounded, hence constant by
Liouville’s theorem. Suppose f (z0 ) = ∞ for some z0 ∈ C (we may apply Möbius
transformations to assume that f (∞) 6= ∞, which preserve the set of rational
functions). Since limz→z0 f (z) = ∞, z0 is a pole of f . Note that f has finitely many
poles, since the poles are isolated (or equivalently the zeros of 1/f are isolated).
Now subtracting the principal parts of the Laurent series of f at each pole, we
obtain an entire function which is also bounded as its image is a compact subset of
C, hence is a constant. It follows that f is the sum of a constant and those principal
parts, hence f is a rational function.
15
P∞
1.12. Residue Theory. Suppose f (z) = n=−∞ an (z − z0 )n is the Laurent ex-
pansion of f at an isolated singularity z0 . Then the coefficient a−1 is called the
residue of f at z0 , and denote it by Res(f ; z0 ).
g(z)
Example 1.41. Suppose f (z) = z−z0 where g is holomorphic around z0 . Let
g(z) = g(z0 ) + a1 (z − z0 ) + · · ·
by the power series expansion of g at z0 . Then we see that
Res(f ; z0 ) = g(z0 ).
g(z)
Similarly if f (z) = (z−z0 )2 where g is holomorphic around z0 , then
Res(f ; z0 ) = g 0 (z0 ).
Theorem 1.42 (Residue Theorem). Let z1 , . . . , zm ∈ U , f ∈ H(U \ {zi }m
i=1 ), and
γ a closed path homotopic to a point in U such that γ does not pass through any
zi . Then
Z m
1 X
f (z)dz = n(γ; zi ) Res(f ; zi ).
2πi γ i=1
Proof. Each zi is P
an isolated singularity of f . Let fi be the principal part of f at
m
zi . Then g = f − i=1 fi ∈ H(U ), hence by Cauchy’s theorem
I m I
1 X 1
f (z)dz = fi (z)dz.
2πi γ i=1
2πi γ
Note that I
1 1
dz = n(γ; zi )
2πi γ (z − zi )k
for k = 1 and the integral is 0 for k ≥ 2. The desired formula follows from the
definition of residue.
Example 1.43. Let γ be a circle with 1 and 2 inside γ. Then by the Residue
Theorem, Z
z+1
= 2πi(Res(f ; 1) + Res(f ; 2)).
γ (z − 1)(z − 2)
By the previous example, we have
1+1 2+1
Res(f ; 1) = = −2, Res(f ; 2) = = 3.
1−2 2−1
Suppose f is meromorphic at z0 (i.e. z0 is not an essential singularity). Then
f (z) = (z − z0 )m g(z) for some integer m and g(z0 ) 6= 0, where m = ordz0 (f ) is the
zero or pole order of f at z0 (we use −3 for a triple pole, etc in this context). Note
that
f0 m g0
= + ,
f z − z0 g
0
and the set of poles of f /f is contained in the set of zeros and poles of f . In
particular,
Res(f 0 /f ; z0 ) = ordz0 (f ).
Hence the residue theorem implies the following.
16
√
The only pole inside the unit circle is ( 3 − 2)i. Hence by the Residue Theorem
the value of the integral is
2 2π
2πi · √ √ =√ .
( 3 − 2)i − (−2 − 3)i 3
Next we consider improper integrals. Denote by H the closed upper-half plane.
Let us first make the following observation.
Theorem 1.48. Suppose f is holomorphic away from a finite set of singularities,
has no singularities on R, and there exist R, C > 0 and p > 1 such that |f (z)| ≤
C|z|−p for {z ∈ H : |z| > R}, then
Z ∞ Xm
f (x)dx = 2πi Res(f ; zj ),
−∞ j=1
Remark 1.49. In the statement of the above theorem, if we replace the upper-half
plane by the lower-half plane, then the conclusion still holds with the sign changed
to − on the right-hand side, because the lower semicircle used in the proof will have
negative orientation (draw a picture).
Z ∞
1
Example 1.50. Let us evaluate dx. Since the integrand f (z) is
−∞ (1 + x2 ) 2
−4
asymptotically z , the assumption in the above theorem holds. There are two
poles at ±i and only i is contained in the upper-half plane. Write
g(z)
f (z) = ,
(z − i)2
where g(z) = (z + i)−2 . Then
i
Res(f ; i) = g 0 (i) = − .
4
It follows that Z ∞
1 π
dx = 2πi Res(f ; i) = .
−∞ (1 + x2 )2 2
18
for all m, n ≥ N . It proves that {gn } converges uniformly on these closed disks
D(w, r), hence it converges uniformly on every compact subset K of U , as K can
be covered by finitely many such disks.
1.14. The Riemann Mapping Theorem. Let U be a path-connected open sub-
set of C. Recall that U is simply connected if every loop contained in U is con-
tractible (i.e. the fundamental group of U is trivial).
Lemma 1.54. If U is simply connected and f is a nowhere vanishing holomorphic
function on U , then f has a holomorphic logarithm and a holomorphic nth root.
Proof. Suppose f ∈ H(U ) is nowhere zero. Then f 0 /f ∈ H(U ). Since U is simply
connected, for a fixed z0 ∈ U , we have
Z z 0
f (w)
F (z) = dw ∈ H(U ).
z0 f (w)
In particular, F 0 = f 0 /f and F (z0 ) = 0. Since f (z0 ) 6= 0, fix one value of log f (z0 )
(not unique) and define
h(z) = e−F (z)−log f (z0 ) f (z).
Then one checks that h0 (z) = 0 and h(z0 ) = 1, hence h(z) = 1 for all z ∈ U . It
follows that
f (z) = eF (z)+log f (z0 ) .
Hence the exponent can be used as log f . Take g(z) = e(log f )/n . Then g n = f .
A complex valued function is called conformal at z0 (i.e. preserving angles of
two intersecting arcs), if it has a nonzero derivative at z0 .
Theorem 1.55 (Riemann Mapping Theorem). Every proper, simply connected,
non-empty open subset U of C is conformally equivalent to the open unit disk D.
We need some preparations before proving the theorem. First we determine the
conformal automorphisms of D. For w ∈ D, define
z−w
hw (z) =
1 − wz
as a special Möbius transformation, where hw (w) = 0 and hw (0) = −w. If |u| = 1,
then u−1 = u, and
u−w u−w u−w
= −1 = = 1.
1 − wu u −w u−w
It implies that hw maps the unit circle to itself. Since w is inside D and hw (w) = 0,
hw maps D onto itself, hence hw is a conformal automorphism of D.
Theorem 1.56. Every conformal automorphism h of the unit disk is of the form
z−w
h(z) = uhw (z) = u
1 − wz
for some u and w satisfying that |u| = 1 and |w| < 1.
Proof. Suppose h(w) = 0 for some w ∈ D. Consider h ◦ h−1 w , which is a conformal
automorphism of D and takes 0 to 0. By the Schwarz Lemma, h ◦ h−1 w (z) = uz for
some |u| = 1. Replacing z by hw (z), we conclude that h(z) = uhw (z).
20
Proof. The idea is to reduce to the case when C lies in R and use the standard
conjugate z̄. Let h : W → V be a conformal equivalence that maps L onto C,
where L is either a line segment in R or R ∪ {∞}. Replacing W by a connected
component of W ∩ W − which contains L, we may assume that W = W − . Let
κ(z) = z. Then we get a reflection ρ through C on V by setting
ρ = h ◦ κ ◦ h−1 .
Since h is analytic and κ is conjugate analytic, one checks that ρ is conjugate
analytic (exercise). Moreover,
ρ ◦ ρ = κ ◦ κ = Id,
and ρ(z) = z on C because κ(z) = z on L and h−1 maps C to L. Thus ρ is a
reflection through C on V .
If ρ1 is another reflection through C on V1 , then ρ1 ◦ ρ is analytic on V ∩ V1 , since
the composition of two conjugate analytic functions is analytic (exercise). Moreover,
ρ1 ◦ ρ(z) = z on C, hence it must be equal to z on all connected components of
V ∩ V1 containing C, by the Uniqueness Theorem of analytic functions. It implies
that ρ1 = ρ−1 = ρ on this component.
Proof. The proof is the same as the proof of the Schwarz Reflection Principle, where
ρ plays the role of complex conjugate on U . Let
g(z) = f (ρ(z))
on W ∪ C. Then g is continuous on W ∪ C and analytic on W because it is the
composition of two conjugate analytic functions f and ρ. Since ρ fixes points on C
and f is real-valued on C, g and f agree on C, hence they define a single function
h on U such that h is continuous on U and analytic on U \ C. Arguing as before
by Morera’s Theorem, h as an extension of f is analytic on all of U .
Example 1.66. Let us consider the reflection through an arc on the unit circle.
1 eiθ
The map ρ(z) = or ρ(reiθ ) = is defined and conjugate analytic on C \ {0}.
z r
It satisfies ρ ◦ ρ(z) = z. It also fixes each point on the unit circle. Suppose a
domain U meets the unit circle in an arc C and is taken to itself by ρ. Then the
unique reflection through C on U is the map ρ. For such U and C, the previous
theorem says that any function analytic on the part V of U lying on one side of C,
continuous on V ∪ C, and real-valued on C can extend analytically to all of U .
23
where γ(0) = z0 and γ(t) = zt . However, this procedure does not necessarily lead
to a globally defined antiderivative F ∈ A(U ). For example, take f (z) = 1/z and
U = C∗ . If γ is a closed path traversing around 0 one time, then the starting and
ending log branches will differ by 2πi. On the other hand, if U is simply connected,
then by Cauchy’s Integration Theorem such a construction yields a globally defined
F . This holds in general for analytic continuation as follows.
Theorem 1.70 (Monodromy Theorem). Suppose γ0 and γ1 are homotopic curves
in an open set U joining z0 to w, and (f0 , D0 ) an analytic function element at z0 ,
with D0 ⊂ U . If (f0 , D0 ) can be analytically continued along every curve in U , then
the continuations of f0 along γ0 and γ1 are equivalent function elements at w.
Proof. Let {γs } be a one-parameter family of curves from z0 to w as a homotopy
from γ0 to γ1 . Then given any ε > 0, there is a δ > 0 such that ||γs − γr || < ε
whenever |s − r| < δ. Denote by φs the terminal function element for the continu-
ation along γs . We will show that for each r ∈ [0, 1], there is a δ > 0 such that φs
is equivalent to φr whenever |s − r| < δ.
Let 0 = t0 < t1 < · · · < tn+1 = 1 be a partition and (f1 , D1 ), . . . , (fn , Dn ) a
sequence of function elements defining φr = (fn , Dn ) as an analytic continuation
of (f0 , D0 ) along γr . Then γr ([tj , tj+1 ]) ⊂ Dj for j = 0, . . . , n. For each j, let
εj be the distance from γr ([tj , tj+1 ]) to the boundary of Dj . If ||γs − γr || < j ,
then γs ([tj , tj+1 ]) ⊂ Dj . Thus let ε = min{ε0 , . . . , εn }, and choose δ > 0 such
that ||γs − γr || < whenever |s − r| < δ. Then for each s with |s − r| < δ, the
same partition and sequence of function elements (f1 , D1 ), . . . , (fn , Dn ) also defines
24
Then g0 lifts f through h, but only on D0 . The rest of the proof is to show that the
analytic function element (g0 , D0 ) can be analytically continued along any path in
U beginning at z0 . If we can do this, then the Monodromy Theorem implies that
(g0 , D0 ) can be continued to an analytic function g on all of U (since U is simply
connected). Moreover, since f and h ◦ g agree on D0 , then they are the same on all
of U by the Uniqueness Theorem.
Let γ : [0, 1] → U be a path beginning at z0 . Let S be the subset of I = [0, 1]
such that (g0 , D0 ) can be analytically continued along γ as far as to s for s ∈ S.
Clearly if s ∈ S, then t ∈ S for all 0 ≤ t < s. If S = [0, 1], then we are done.
If S = [0, r] for some r < 1, then the terminal element is defined on a disk that
intersects (r, 1], hence S should be larger, a contradiction. The only remaining case
is S = [0, r) with r < 1, which cannot happen as we will prove below.
If S = [0, r), choose a neighborhood A of f (γ(r)) in W such that h−1 (A) is a
disjoint union of open sets Bj , each mapping onto A conformal equivalently under
h. Since f is continuous, we can choose an open disk D ⊂ U such that f (D) ⊂ A
and a point s ∈ [0, r) such that w = f (γ(s)) ∈ f (D) (draw a picture). Since
s ∈ S, the function element (g0 , D0 ) can be analytically continued along γ to
some (gn , Dn ) with γ(s) ∈ Dn . Since f = h ◦ g0 on D0 , again by the Uniqueness
Theorem we have f = h ◦ gn on Dn . In particular, w = f (γ(s)) = h(gn (γ(s))),
hence gn (γ(s)) ∈ h−1 (A), say, gn (γs ) ∈ Bk . Now define a new function element
(gn+1 , Dn+1 ) by setting Dn+1 = D and gn+1 = f ◦ h−1 k , where hk : Bk → A.
Then f = h ◦ gn+1 on Dn+1 . Note that gn (Dn ∩ Dn+1 ) ⊂ Bk , for otherwise the
connected open set Dn ∩ Dn+1 would be separated by gn−1 (Bk ) and the union of the
sets gn−1 (Bj ) for j 6= k. It follows that the two inverse functions of h used in the
definitions of gn and gn+1 agree on f (Dn ∩ Dn+1 ), which implies that gn = gn+1 on
Dn ∩Dn+1 . Since γ(r) ∈ Dn+1 , it means that (g0 , D0 ) can be analytically continued
to γ(r), hence r ∈ S, contradicting that S = [0, r).
1.18. The Picard Theorems. We will apply the previously developed techniques
to prove the following two amazing theorems.
Theorem 1.77 (Little Picard Theorem). A non-constant entire function takes on
every value in C except possibly one.
Theorem 1.78 (Great Picard Theorem). An analytic function with an essential
singularity at z0 takes on every complex value, except possibly one, infinitely often
in every neighborhood of z0 .
The upshot for proving the Picard theorems is to construct an analytic covering
map from the unit disk to C with two points removed. We begin with some facts
about reflection through a circle.
Suppose C1 and C2 are two circles in the plane that intersect at two points. If
the tangents to the two circles are perpendicular at the two points, then we say
that the circles meet at right angles (draw a picture). Indeed if the tangents are
perpendicular at one intersection point, then they are perpendicular at the other
point (by symmetry to the line connecting their centers).
Theorem 1.79. If two circles C1 and C2 meet at right angles, and A is the arc of
C1 lying inside C2 , then reflection through A maps C2 onto itself (draw a picture).
Proof. There is a Möbius transformation h that maps C1 onto R ∪ ∞ and maps
the inside of C1 onto the upper-half plane H. We can further choose which point
26
Theorem 1.80. With C1 , C2 as above, the reflection through C1 takes any other
circle C, which meets C2 at right angles, to another circle which meets C2 at right
angles or a line through the center of C2 .
Proof. As discussed before a reflection through a circle is a Möbius transformation
followed by conjugation, hence it takes a circle C to either a circle or a line. The
previous theorem says that the reflection through C1 takes the inside of C2 onto
itself, hence the image of C under this reflection still meets C2 in two points. Since a
conformal map preserves angles and conjugation reverses angles, their composition
is also angle reversing. It follows that if two curves meet perpendicularly, then
the same holds for their image under a reflection. Therefore, reflection through C1
takes C to a circle or line which also meets C2 at right angles, and the only way
the image of C can be a line is when it passed through the center of C2 .
Now we construct an analytic covering map from the open unit disk D onto
C \ {0, 1}. First note that the three points −1, eiπ/3 , and e−iπ/3 are equidistant
points on the unit circle C, hence we can join each pair of these points with an arc
of a circle that meets C at right angles. These three arcs bound an open curvilinear
triangle V0 (draw a picture), which is simply connected. By the Riemann Mapping
Theorem, there is a conformal equivalence h : V0 → H where H is the upper-half
plane. As remarked before, h extends to a continuous map V 0 → H ⊂ S 2 which
takes the boundary of V0 to the boundary of H in S 2 . By composing with a suitable
Möbius transformation, we may assume that h takes −1, eiπ/3 , and e−iπ/3 to ∞,
1, and 0, respectively.
Since h is real-valued on the three edges of V0 , by the circle version of the
Schwarz Reflection Principle, h has an analytic continuation by reflection across
each of these edges. This makes h being defined and analytic in the set V1 (draw a
picture). By the previous two theorems, each of the three new cells is contained in
D and bounded by three circles that meet C at right angles. Moreover, as h maps
V0 onto the upper-half plane, it maps each of the new cells of V1 onto the lower
half-plane.
We can now analytically continue h to larger and larger domains {Vn } by repeat-
ing the above process (draw a picture). One checks that the union of these Vn is D,
so we obtain an analytic function h which maps D onto a subset of C containing
the upper and lower open half-planes and the intervals (−∞, 0), (0, 1), and (1, ∞)
on R. The points ∞, 1, and 0 are not in the image of h, because every triangular
cell has all of its vertices on the unit circle coming from reflections of the vertices
of V0 that map to ∞, 1, and 0.
Theorem 1.81. The map h : D → C\{0, 1} described above is an analytic covering
map.
27
Proof. Use the famous “Angels and Demons” illustration (print out the picture and
distribute).
The proof of the Great Picard Theorem relies on the following technical result.
Theorem 1.83. Let U be a simply connected subset of C. Then a sequence of
analytic functions on U with values in C \ {0, 1} is either a normal family or they
converge uniformly to ∞.
Proof. See [T, Theorem 7.4.5] for a proof.
1.19. Infinite Products. First, recall that for a half-open half-closed interval I of
length 2π, the log function associated to I is
logI z = log |z| + i argI z,
where argI z is the unique argument of z lying in I for a nonzero complex number z.
If I = (−π, π], we call it the principal branch of the log function and simply denote it
by log. In particular, if u and v both have positive real part, then im log u, im log v ∈
(−π/2, π/2), and im log uv ∈ (−π, π), hence in this case log u + log v = log uv, since
they cannot differ by a nonzero multiple of 2πi.
Qn
Definition 1.85. If {uk } is a sequence
Q∞ of complex numbers and pn = k=1 uk ,
we say that the infinite product k=1 uk converges to a complex number p if
limn→∞ pn = p.
28
Theorem 1.86. The P∞ infinite product converges to a non-zero number p if and only
if the infinite sums k=1 log uk converges to a number λ, where p = eλ . Moreover,
if the infinite series converges absolutely, then the infinite product is unchanged by
any rearrangement of the factors.
Proof. Define pn as above. First suppose p = limn→∞ pn exists and is nonzero.
Since the principal branch of log is continuous at 1, we have
lim log(pn /p) = 0.
n→∞
be its nth partial sum. Then limn→∞ λn = λ. Since pn = eλn and the exponential
function is continuous, the sequence {pn } converges to eλ .
Finally if the series converges absolutely, then each of its rearrangement con-
verges to the same number. It follows that each rearrangement of the infinite
product also converges to the same number.
Now consider a sequence of functions {uk (z)} defined on a set S. We say that
the infinite product of {uk (z)} converges uniformly on S if the sequence {pn (z)} of
the partial products converges uniformly on S.
Theorem 1.87. Let {uk (z)} be a sequence of complex-valued functions defined and
bounded uniformly on a set S. If the series
∞
X
log uk (z)
k=1
Proof. As before, let λn (z) be the nth partial sum of log uk (z), and pn (z) the nth
partial product of uk (z). By assumption, λn (z) − λ(z) converges uniformly to 0 on
S. Since the exponential function is continuous, it implies that
pn (z)
= eλn (z)−λ(z)
p(z)
converges uniformly to 1.
Since the pn (z) are bounded uniformly on S, then <(λn (z)) = log |pn (z)| are
bounded uniformly on S, and hence <(λ) is bounded on S. It follows that p(z) =
eλ(z) is bounded on S. Hence pn = (pn /p)p converges uniformly to p on S.
converges uniformly
P∞ on S if and only if the other one does.P∞
Therefore, if k=1 |ak (z)| converges uniformly, then k=K log(1 + ak (z)) con-
verges uniformlyQand absolutely. By the previous theorems, this implies the uniform
∞
convergence
Q∞ of k=K (1 + ak (z)) to a function on S with no zeros. It follows that
k=1 (1+ak (z)) converges uniformly on S, the limit is unaffected by rearrangement
of the factors, and each of its zeros is a zero, with the same order, of the product
of the factors 1 + ak (z) for k < K.
converges uniformly on D(0, R). Hence by the above theorem, the infinite product
also converges uniformly on D(0, R) for each R, and hence on each bounded subset
of C.
Definition 1.90. Let f ∈ A(U ), not identically zero. Then the meromorphic
function f 0 /f is called the log derivative of f on U .
This definition arises formally from (log f )0 = f 0 /f .
Theorem 1.91. Let {fn } be a sequence of analytic functions on a connected open
set U . If {fn } converges uniformly to f (not identically zero) on U , then the
sequence {fn0 /fn } converges uniformly to f 0 /f on compact subsets of U \ S, where
S is the set of zeros of f .
Proof. Similar to one of the homework problems.
Corollary 1.92. Let {uk } be a sequence of analytic functions on a connected open
set U . If the infinite product
∞
Y
f (z) = uk (z)
k=1
converges uniformly on compact subsets of U to a function f which is not identically
zero, then the infinite sum
∞
X u0k (z)
uk (z)
k=1
converges uniformly to f 0 /f on compact subsets of U \ S, where S is the set of zeros
of f .
Example 1.93. Consider the function
∞
z2
Y
f (z) = πz 1− 2
k
k=1
where the infinite product converges uniformly on each bounded subset of C. Its
log derivative can be written as
∞
f 0 (z) 1 X 2z
= + .
f (z) z z 2 − k2
k=1
Since
2z 1 1
= + ,
z 2 − k2 z−k z+k
we obtain that
n
f 0 (z) X 1
= lim .
f (z) n→∞ z−k
k=−n
Indeed with some more work one can show that
f (z) = sin(πz).
Given a sequence of points in U such that they do not have any accumulation
point in U , can one construct an analytic function on U such that it has exactly
the points of this sequence as its zeros, with each zero order equal to the number
of times it appears in the sequence? In order to answer this question, we first
introduce the following functions.
31
Given f ∈ A(U ) and a sequence {zk } ⊂ U , we say that {zk } is a list of zeros
of f (counting with multiplicity) if each zk is a zero of f and each zero w of f
occurs ordw (f ) times in this sequence. Now we can define an infinite product
called Weierstrass product to answer the previous question.
Theorem 1.95. If {zk } is a sequence of non-zero complex numbers and {pk } is a
sequence of nonnegative integers such that
∞ pk +1
X z
<∞
zk
k=1
Proof. Without loss of generality, suppose exactly m terms in the sequence are 0.
We may arrange them to be the first m terms. Then {zk }∞ k=m+1 is a sequence of
nonzero complex numbers. Given any R > 0, since zk → ∞, there is K > m such
that |zk | > 2R for all k ≥ K. Then the series
X z k
zk
k=m+1
Proof. Choose {pk } such that the hypotheses of the previous theorem hold (say,
taking pk = k − 1 always works). Then the resulting function
∞
Y
g(z) = z m Epk (z/zk )
k=1
33
converges uniformly on compact sets and has the same zeros as f with the same
multiplicities. Therefore, f g −1 is an entire function with no zeros, hence f g −1 = eh
for some entire function h.
Example 1.99. Consider f (z) = sin(πz). It has a simple zero at each integer and
no other zeros. Since
∞
X 1
< ∞,
k2
k=1
we can take pk = 1 for all k. Then the above theorem implies that
Y Y
sin(πz) = eh(z) z E1 (z/k) = eh(z) z (1 − z/k)ez/k .
k6=0 k6=0
Note that
(1 − z/k)ez/k (1 + z/k)e−z/k = 1 − z 2 /k 2 .
Hence we can write
∞
Y
sin(πz) = eh(z) z (1 − z 2 /k 2 ).
k=1
Indeed with some more work one can show that
Y∞
sin(πz) = πz (1 − z 2 /k 2 ).
k=1
z − wk
k=1
Proof. Let h ∈ M(U ) and let {zk } be a sequence of the poles of h, where each
zk appears as many times as its pole order. By the previous theorem, there exists
g ∈ A(U ) such that g has {zk } as a list of zeros. It implies that f = gh ∈ A(U ),
and h = f /g as desired.
is defined as a meromorphic function on D(0, rm ) and has the required poles and
principal parts at the points of S contained in D(0, rm ). Since limm→∞ rm = R, we
thus conclude that f (z) is meromorphic on the entire D(0, R) and has the desired
poles and principal parts.
2. Riemann Surfaces
In this section we carry out complex analysis on Riemann surfaces, introduce
fundamental results about Riemann surfaces, and interpret them from a modern
viewpoint.
35
2.1. Preliminaries. We go over some of the previous concepts and results in the
context of Riemann surfaces.
Definition 2.1 (Riemann surface). A Riemann surface X is a one-dimensional
complex manifold, i.e., X is a real surface with a complex atlas of charts
{φi : Ui → Vi ⊂ C},
where {Ui } is an open covering of X, φi is a homeomorphism, and
φj ◦ φ−1
i : φi (Ui ∩ Uj ) → φj (Ui ∩ Uj )
is a conformal equivalence. We say that such a complex atlas of charts is a complex
structure on the underlying real surface.
Example 2.2. (a) Recall that P1 = C ∪ {∞}, homeomorphic to the real sphere.
Take
U1 = P1 \{∞} = C,
U2 = P1 \{0} = C∗ ∪ {∞}.
Define φ1 (z) = z, φ2 (z) = 1/z for z 6= ∞ and φ2 (∞) = 0. Then φ2 ◦ φ−1 1 :
C∗ → C∗ is given by z 7→ 1/z, which is a conformal equivalence. Therefore, P1
is a Riemann surface, called the Riemann sphere or the (complex) projective
line.
(b) Suppose ω1 , ω2 ∈ C are linearly independent over R. Define
Γ = Zω1 + Zω2 = {nω1 + mω2 | n, m ∈ Z},
called the lattice spanned by ω1 and ω2 . Define an equivalence relation z1 ∼ z2
on C if z1 − z2 ∈ Γ. Denote by E the set of equivalence classes, i.e., E = C/Γ.
Equip E with the quotient topology, i.e., U ⊂ E is open if and only if π −1 (U )
is open for the projection π : C → E. Then E is homeomorphic to a torus.
One can enable E a complex structure as follows. Let V ⊂ C be an open
subset such that no two points in V are equivalent with respect to Γ. Then
U = π(V ) is open in E, and let φ : U → V be the inverse of π, which forms a
complex chart on E.
Exercise 2.3. Prove that the above complex charts form a complex structure on
E.
Definition 2.4 (Holomorphic function). Let X be a Riemann surface. A function
f : X → C is called holomorphic, if for every chart φ : U → V ⊂ C on X the
function
f ◦ φ−1 : V → C
is holomorphic. The set of all holomorphic functions on X is denoted by O(X)
(new notation, previously we used H(X)).
Definition 2.5 (Holomorphic map). Let X and Y be Riemann surfaces. A con-
tinuous map f : X → Y is a called holomorphic, if for every chart φ : U1 → V1 ⊂ C
on X and ψ : U2 → V2 ⊂ C on Y with f (U1 ) ⊂ U2 the function
ψ ◦ f ◦ φ−1 : V1 → V2
is holomorphic. If f is biholomorphic, then X and Y are called isomorphic.
36
for every p ∈ Σ, then f is called meromorphic, and the points in Σ are poles of f .
The set of all meromorphic functions on X is denoted by M (X).
Remark 2.7. In a neighborhood of a pole p, let z be a suitable coordinate with
z(p) = 0. Then as we saw before the Laurent series expansion of f is
∞
X
f= ai z i
i=−k
Theorem 2.9. Suppose X is a Riemann surface and f ∈ M (X). For each pole p
of f , define f (p) = ∞. Then f : X → P1 is a holomorphic map.
Proof. Take the Laurent series expansion
∞
X
f (z) = ai z i = z −k g(z)
i=−k
locally at a pole p with a suitable coordinate z(p) = 0, where k is the pole order of
p and g(p) 6= 0. Then 1/f = z k (1/g) is holomorphic at p.
Exercise 2.10. Prove that every holomorphic function on a compact Riemann
surface is constant. Hint: a non-constant holomorphic function is open.
Definition 2.11 (Covering, branch and ramification). Suppose X and Y are Rie-
mann surfaces and π : Y → X is a non-constant holomorphic map. Then π is called
a (possibly branched) covering.
A point y ∈ Y is called a ramification point of π, if there is no neighborhood
V of y such that π|V is injective. A point x ∈ X is called a branch point of π, if
π −1 (x) contains a ramification point. The map π is called unramified if it has no
branch point (i.e., no ramification point).
For p ∈ Y and q = π(p) ∈ X, by properties of non-constant holomorphic func-
tions, there exist neighborhoods U of p and V of q as well as suitable coordinates
x and y such that x(p) = y(q) = 0 and f |U (x) = xk for some k ∈ Z+ . If k = 1,
then p is not a ramification point of π. If k > 1, then p is a ramification point. We
say that the multiplicity of π at p is k, denoted by multp (π), and the ramification
order of π at p is k − 1, denoted by ordp (π). In particular, multp (π) − 1 = ordp (π).
Example 2.12. (a) Let π : C → C be π(z) = z k for an integer k ≥ 2. Then 0 is
the only ramification point.
(b) Consider the exponential map exp : C → C∗ by z 7→ ez . Then exp is unramified.
In particular, if V ⊂ C does not contain two points that differ by an integer
multiple of 2πi, then exp |V is injective.
(c) Let Γ be a lattice in C, and E = C/Γ the associated torus. Then the projection
C → E is unramified.
37
Show that the quotient sheaf Op is isomorphic to the skyscraper sheaf with stalk C
at p.
It is more convenient to verify injectivity and surjectivity for maps of sheaves by
using stalks.
Proposition 2.30. Let φ : E → F be a map for sheaves E and F on a topological
space X.
(1) φ is injective if and only if the induced map φp : Ep → Fp is injective for the
stalks at every point p.
(2) φ is surjective if and only if the induced map φp : Ep → Fp is surjective for
the stalks at every point p.
(3) φ is an isomorphism if and only if the induced map φp : Ep → Fp is an
isomorphism for the stalks at every point p.
Proof. The claim (3) follows from (1) and (2). Let us prove (1) only, and one can
find the proof of (2) in many standard textbooks.
Suppose φ is injective. Take a section σ ∈ E (U ) on an open subset U . If φ([σ]) =
0 ∈ Fp , there exists a smaller open subset V ⊂ U such that φV (σ) = 0 ∈ F (V ),
hence σ|V = 0 ∈ E (V ). Consequently the equivalence class [σ] = 0 ∈ Ep and we
conclude that φp is injective.
Conversely, suppose φp is injective for every point p. Take a section σ ∈ E (U ).
If φ(σ) = 0 ∈ F (U ), then for every point p ∈ U , [φ(σ)] = 0 ∈ Fp . Since φp is
injective, it implies that [σ] = 0 ∈ Ep i.e. there exists an open subset Up 3 p such
that σ|Up = 0 ∈ E (Up ). Applying the gluing property to the open covering {Up } of
U , we conclude that σ = 0 ∈ E (U ).
Remark 2.31. The image of φ does not automatically form a sheaf. In general,
it is only a presheaf. If the sheafification of Im(φ) equals F , we say that φ is
surjective. In particular, it does not mean E (U ) → F (U ) is surjective for every
open set U . Sometimes one has to pass to a refined open covering in order to obtain
a surjection between sections.
Example 2.32. Consider the exponential map exp : O → O ∗ on the punctured
plane C\{0}. As a map of sheaves it is surjective, but the section z over C\{0}
does not have an inverse. Nevertheless, it does have an inverse over any contractible
open subset.
2.5. Sheaf cohomology. Let F be a sheaf on a topological space X. Take an
open covering U = {Uα } of X. Define the k-th cochain group
Y
C k (U , F ) := F (Uα0 ∩ · · · ∩ Uαk ).
α0 ,...,αk
which is independent of the choice of φ. Finally, we define the k-th (Čech) coho-
mology group by passing to the direct limit:
H k (X, F ) := lim H k (U , F ).
−→
The definition involves direct limit, which is inconvenient to use in practice.
Nevertheless, we can simplify the situation once the open covering U is fine enough.
We say that U = {Ui }i∈I is acyclic respect to F , if for any k > 0 and i1 , . . . , il ∈ I
we have
H k (Ui1 ∩ · · · ∩ Uil , F ) = 0.
Theorem 2.37 (Leray’s Theorem). If the open covering U is acyclic respect to F ,
then H ∗ (U , F ) ∼
= H ∗ (X, F ).
Remark 2.38. In the context of complex manifolds, if Ui ’s are contractible, then
U is acyclic respect to the sheaves we will consider. While for algebraic varieties,
if Ui ’s are affine, then U is acyclic.
Example 2.39. Let us compute the cohomology of the structure sheaf O on P1 .
It is clear that H 0 (P1 , O) = C, since any global holomorphic function on P1 is
constant. For higher cohomology, use [X, Y ] to denote the homogeneous coordinates
of P1 . Take the standard open covering U = {[X, Y ] : X 6= 0} and V = {[X, Y ] :
Y 6= 0}. It is acyclic respect to the structure sheaf O (morally because U, V ∼ =C
is contractible). Let s = Y /X and t = X/Y as affine coordinates of U and V ,
respectively. Suppose h is an element in C 1 ({U, V }, O), i.e. h ∈ O(U ∩ V ). We can
write
X∞
h= ai si .
i=−∞
Now take
∞
X
f =− ai si ∈ O(U ),
i=0
−1
X −1
X
g= ai si = ai t−i ∈ O(V ).
i=−∞ i=−∞
which is holomorphic if and only if ai = 0 for all i. Hence w is the zero one-form
and H 0 (P1 , Ω) = 0. Now take ω ∈ C 1 ({U, V }, Ω), i.e. ω ∈ Ω(U ∩ V ) = Ω(C∗ ), we
express it as
X ∞
ω= ai ti dt.
i=−∞
Note that any α ∈ Ω(U ) and β ∈ Ω(V ) can be written as
X ∞
α= bi si ds,
i=0
∞
X
β= ci ti dt.
i=0
Hence on U ∩ V we have
∞
X ∞
X
δ((α, β)) = β − α = − bi t−i−2 dt + ci ti dt.
i=0 i=0
−1
Note that only the term t is missing from the expression. We conclude that
H 1 (P1 , Ω) = {a−1 t−1 dt} ∼
= C.
Remark 2.41. In general, the rank of H 1 (X, O) ∼ = H 0 (X, Ω) (by Serre Duality)
is called the genus of a Riemann surface (or a complex algebraic curve) X.
Exercise 2.42. Let D = p1 + · · · + pn be a collection of n points in P1 . We say
that D is an effective divisor of degree n. Define the sheaf O(D) on P1 by
O(D)(U ) = {f ∈ M (U ) : f ∈ O(U \{p1 , . . . , pn }) with at worst a simple pole at each pi }.
Assume that the standard covering of P1 is acyclic respect to O(D). Use it to
calculate the cohomology groups H ∗ (P1 , O(D)).
As many other homology/cohomology theories, one can associate a long exact
sequence of cohomology to a short exact sequence. Suppose we have a short exact
sequence of sheaves
α β
0 −−−−→ E −−−−→ F −−−−→ G −−−−→ 0.
Then α and β induce maps
α : C k (U , E ) → C k (U , F ), β : C k (U , F ) → C k (U , G ).
Since the coboundary map δ is given by alternating sums of restrictions, α and
β commute with δ, hence they send a cocycle to cocycle and a coboundary to
coboundary. Consequently they induce maps for colomology
α∗ : H k (X, E ) → H k (X, F ), β∗ : H k (X, F ) → H k (X, G ).
Next we define the coboundary map
δ∗ : H k (X, G ) → H k+1 (X, E ).
For σ ∈ C k (U , G ) satisfying δσ = 0, after refining U (still denoted by U ) such
that there exists τ ∈ C k (U , F ) satisfying β(τ ) = σ, because β is surjective. Then
β(δτ ) = δ(β(τ )) = δσ = 0, hence after refining further there exists µ ∈ C k+1 (U , E )
satisfying α(µ) = δτ . Note that µ is a cocycle. It is because α(δµ) = δ(α(µ)) =
δδ(τ ) = 0 and α is injective, hence δµ = 0 and µ ∈ ker(δ). We thus take δ∗ σ :=
[µ] ∈ H k+1 (X, E ). One checks that this is independent of the choice of τ and µ.
44
x1 , . . . , xn as generators.
Example 2.53. Let X ⊂ Pn be a submanifold and Y ⊂ X a hypersurface, i.e.
Y is cut out (transversely) by a hypersurface F in Pn with X. We have the short
exact sequence
0 → IY /X → OX → OY → 0.
The ideal sheaf IY /X is an invertible sheaf. Indeed, for an open subset U ⊂ X,
IY /X (U ) can be expressed as (F |U ) · OX (U ), hence is locally free of rank 1. The
sheaf OY (extended to X by zero) is not locally free. For U ∩ Y = ∅, OY (U ) = 0
and for U ∩ Y 6= ∅, OY (U ) is non-zero. Later we will see how to construct a line
bundle corresponding to IY /X .
2.8. Divisors. Let X be a complex manifold. Suppose Y ⊂ X is an irreducible
subspace of codimension one satisfying that for every p ∈ Y there exists an open
neighborhood U ⊂ X of p such that U ∩ Y is cut out by a holomorphic function
f . Then we say that Y is an irreducible divisor of X. We call f a local defining
equation for Y near p. A divisor D on X is a formal linear combination of irreducible
divisors:
X n
D= ai Yi ,
i=i
47
P
Conversely, suppose D = ai Yi is a divisor on X with ai ∈ Z and Yi effective.
We can choose an open covering U = {Uα } such that Yi is locally defined by
giα ∈ O(Uα ). Consider Y
fα = (giα )ai ∈ M ∗ (Uα ).
i
Then we have
fα Y giα ai
= .
fβ i
giβ
Both giα and giβ cut out the same divisor Yi |Uα ∩Uβ in Uα ∩ Uβ , hence we conclude
that
giα fα
∈ O ∗ (Uα ∩ Uβ ), ∈ O ∗ (Uα ∩ Uβ ).
giβ fβ
Then {fα } defines a global section of M ∗ /O ∗ . Finally if D determines the zero
section of M ∗ /O ∗ (which is 1 since the group structure is multiplicative), it means
locally fα ∈ O ∗ (Uα ) (after refining the open covering). Then it does not have zeros
or poles, hence D|Uα = 0 for each Uα and D is globally zero. This shows the other
injection
Div(X) ,→ H 0 (X, M ∗ /O ∗ ).
2.9. Line bundles. Recall that a line bundle L on X is a vector bundle of rank
1. Equivalently, it is a locally free sheaf of rank 1. Define the Picard group Pic(X)
parameterizing isomorphism classes of line bundles on X. The group law is given
by tensor product. We can interpret Pic(X) as a cohomology group.
Proposition 2.55. There is a one-to-one correspondence between the isomorphism
classes of line bundles on X and H 1 (X, O ∗ ), i.e.
Pic(X) ∼= H 1 (X, O ∗ ).
Proof. Take an open covering U = {Uα } of X with respect to the trivialization of
a line bundle L. The transition function
gαβ : (Uα ∩ Uβ ) × C → (Uα ∩ Uβ ) × C
can be regarded as a section of O ∗ (Uα ∩ Uβ ), satisfying
gαβ · gβα = 1,
gαβ · gβγ · gγα = 1.
Therefore, {gαβ } is a cocycle in C 1 (U , O ∗ ), hence represents a cohomology class in
H 1 (X, O ∗ ).
Suppose M is another line bundle with transition functions {hαβ }. If M and
L are isomorphic, then L ⊗ M ∗ is trivial, i.e. {gαβ /hαβ } are transition functions
of L ⊗ M ∗ , which has a nowhere vanishing section σ. Suppose on Uα we have
σα : Uα → C∗ as the restriction of σ. Then on Uα ∩ Uβ we have
gαβ
· σα = σβ .
hαβ
Therefore we conclude that
gαβ σβ
= ∈ δC 0 (U , O ∗ ).
hαβ σα
49
Remark 2.59. Replacing X by any open subset U , the proposition implies that
the sheaf O(D) can be regarded as gathering local holomorphic sections of the line
bundle [D]. If D ∼ D0 , i.e. D0 − D = (f ) for a global meromorphic function f ,
then for any g ∈ O(D0 )(U ), we have
for any open subset U , compatible with the sheaf restriction maps. In this sense,
the sheaf O(D) and the line bundle [D] have a one-to-one correspondence up to
isomorphism and linear equivalence, respectively, assuming that every line bundle
can be associated to a divisor.
Let |D| be the set of effective divisors that are linearly equivalent to D. We call
|D| the linear system associated to D.
i.e. an effective divisor in |D| and a holomorphic section of [D] (up to scalar)
determine each other.
1 1
P
Exercise 2.61. Let D = i pi be a divisor on P with ai ∈ Z and pi ∈ P . Define
aP
the degree of D by deg(D) = ai .
(1) Prove that D ∼ D0 if and only if deg(D) = deg(D0 ).
(2) Calculate the cohomology H ∗ (P1 , O(D)) in terms of deg(D).
52
Proof. Take a triangulation of Y such that every branch point is a vertex. Pull it
back as a triangulation of X. Note that it pulls back a face to d faces, an edge to
d edges and a vertex v to |π −1 (v)| vertices. Note that if
k
X
−1
π (v) = mi pi
i=1
Then L has a base point at p if and only if all holomorphic sections of L vanish at
p, i.e., H 0 (X, L ⊗ O(−p)) = H 0 (X, L). This proves (1).
For (2), a very ample line bundle is necessarily base point free by definition.
If p 6= q ∈ X have the same image under φL , it is equivalent to saying that the
subspace of sections vanishing at p is the same as the subspace of sections vanishing
at q, which is further equivalent to
h0 (X, L ⊗ O(−p)) = h0 (X, L ⊗ O(−p − q)) = h0 (X, L ⊗ O(−q)).
Moreover, φL induces an injection restricted to the tangent space Tp (X) if and
only if there exists a hyperplane such that it cuts out X locally a simple point at
p, namely, if and only if there is a section vanishing at p with multiplicity one, i.e.,
h0 (X, L ⊗ O(−2p)) < h0 (X, L ⊗ O(−p)).
But we have
h0 (X, L ⊗ O(−2p)) ≥ h0 (X, L ⊗ O(−p)) − 1.
Hence (2) follows from combining the two cases.
Remark 2.80. In (2), for p 6= q the condition geometrically means that the sections
of L separate any two points. When p = q, it says that the sections of L separate
tangent vectors at p.
Example 2.81. Let E be an elliptic curve, i.e., a Riemann surface of genus one.
Fix a point p ∈ E. Consider the morphism
τ : E → Pic0 (E)
by τ (q) = [q − p], where Pic0 (E) is the Picard group of isomorphism classes of line
bundles of degree zero. For q1 6= q2 , we have q1 − p 6∼ q2 − p, hence τ is injective.
For any line bundle L of degree zero, by Riemann-Roch h0 (L ⊗ O(p)) ≥ 1, hence
L ⊗ O(p) has a section vanishing at a single point q. It implies that L = [q − p],
hence τ is surjective. Therefore, we thus conclude that τ is an isomorphism. This
defines a group law on E with respect to p, i.e., q + r = s, where s ∈ E is the unique
point satisfying that
(q − p) + (r − p) ∼ s − p.
Now consider the linear system |3p| on E. Since
h0 (E, O(3p)) = 3, h0 (E, O(2p)) = 2, h0 (E, O(p)) = 1,
O(3p) is very ample. It induces an embedding of E into P2 as a plane cubic curve.
A line cuts out a divisor of degree three in E, say, q + r + s (not necessarily distinct)
if and only if
q + r + s ∼ 3p.
Note that the tangent line L of E at p is a flex line, i.e., the tangency multiplicity
(L · E)p = 3. Such p is called a flex point.
Exercise 2.82. Show that there are in total nine flex points in a smooth plane
cubic curve.
Let V ⊂ |L| be a linear subspace. We say that V is a linear series of L. The
linear system |L| is also called a complete linear series. The above definitions and
properties go through similarly for the induced map φV .
59
Exercise 2.83. Write down a linear series of |O(3)| on P1 such that it maps P1
into P2 as a singular plane cubic curve. How many different types of such singular
plane cubics can you describe?
2.15. Canonical maps. Let K be the canonical line bundle on a Riemann surface
X. If X is P1 , deg(K) = −2 and K is not effective. If X is an elliptic curve,
then K ∼ = O and the induced map φK is onto a point. From now on, assume
that the genus of X satisfies g ≥ 2. We say that X is hyperelliptic if it admits a
degree 2 branched cover of P1 . Two points p, q ∈ X are called conjugate if they
have the same image in P1 . A ramification point of the double cover is called a
Weierstrass point of X, i.e., it is self conjugate. By Riemann-Hurwitz, a genus
g ≥ 2 hyperelliptic curve possesses 2g + 2 Weierstrass points.
Lemma 2.84. If X is a hyperelliptic curve of genus ≥ 2, then X admits a unique
double cover of P1 .
Proof. Otherwise suppose h0 (X, O(p + q)) = 2 and h0 (X, O(p + r)) = 2 for q 6=
r. Let L = OX (p + q + r). If h0 (X, L) = 3, since h0 (X, L(−x − y)) ≤ 1 and
h0 (X, L(−x)) ≤ 2 by degree reason, φL would map X as a plane cubic of genus
one, leading to a contradiction. Then we conclude that
H 0 (X, O(p + q)) = H 0 (X, O(p + r)) = H 0 (X, O(p + q + r)),
which implies that both q, r are base points of |p + q + r| and h0 (X, O(p)) = 2,
X∼= P1 , leading to a contradiction.
Proposition 2.85. Let X be a curve of genus g ≥ 2. Then the canonical line
bundle K is base point free. The induced map
φK : X → Pg−1
is an embedding if and only if X is not hyperelliptic. If X is hyperelliptic, φK is a
double cover of a rational normal curve in Pg−1 .
Proof. First, let us show that K is base point free. For any point p ∈ X, by
Riemann-Roch we have
h0 (X, K ⊗ O(−p)) − h0 (X, O(p)) = 1 − g + (2g − 3),
h0 (X, K ⊗ O(−p)) = g − 1 = h0 (X, K) − 1.
Hence K satisfies the criterion of base point freeness.
Next, K fails to separate p, q (not necessarily distinct) if and only if
h0 (X, K ⊗ O(−p − q)) = h0 (X, K ⊗ O(−p)) = g − 1,
which is equivalent to, by Riemann-Roch again, that
h0 (X, O(p + q)) = 2.
In other words, the linear system |p + q| induces a double cover X → P1 .
Finally, if X is hyperelliptic of genus ≥ 2, it admits a unique double cover of P1 .
By the above analysis, two points p, q have the same image under the canonical
map if and only if h0 (X, O(p + q)) = 2, i.e., p, q are conjugate. Then the canonical
map is a double cover of a rational curve of degree deg(K)/2 = g − 1 in Pg−1 , i.e.,
a rational normal curve. A hyperplane section of φK (X) pulls back to X a divisor
g−1
X
(pi + qi ),
i=1
60
References
[GH] Phillip Griffiths and Joe Harris, Principles of Algebraic Geometry.
[S] Wilhelm Schlag, A concise course in complex analysis and Riemann surfaces.
[T] Joseph Taylor, Complex Variables.