Class Notes
Class Notes
Introduction
These are notes on complex geometry intended to accompany Math 545 at SBU in Fall,
2014. As explained in the course syllabus, the overall goal of the course is to present some
of the basic material on complex manifolds that are needed to work in complex or algebraic
geometry. We will try to emphasize concrete examples and calculations wherever possible.
The present notes generally speaking closely follow the presentations in the texts of
Griffiths-Harris [2] and Huybrechts [3] – and in later parts the notes of Schnell [4] – which
the reader can consult for more details. In particular, we borrow freely from these sources
without explicit attribution.
This section presents some preliminary facts concerning holomorphic functions of several
complex variables, following closely the presentations of Griffiths-Harris [2] and especially
Huybrechts [3].
the polydisk of (mult-)radius ε centered at c. Write T(ε; c) for the torus {z ∈ Cn | |zi − c| =
εi }.
As in the case of functions of one variable, there are various equivalent ways to charac-
terize holomorphic functions of several variables. For an open subset U ⊆ Cn , write C ∞ (U )
for the collection of all C ∞ complex-valued functions on U .
Theorem 1.1. Let U ⊂ Cn be an open set, and let f ∈ C ∞ (U ) be a smooth C-valued
function on U . Then the following three conditions are equivalent:
(ii). f is holomorphic separately in each of the variables zi , i.e. for each c = (c1 , . . . , cn ) ∈
U , the function
f (c1 , . . . , ci−1 , zi , ci+1 , . . . , cn )
is analytic in zi in a neighborhood of ci .
(iii). Given any point c = (c1 , . . . , cn ) in U , there is an open neighborhood V = V (c) of c
in U in which f (z) can be represented by a convergent power series:
∞
X
f (z) = aJ · (z − c)J .
|J|=0
Sketch of Proof of Theorem 1.1. The equivalences (i) ⇐⇒ (ii) and (iii) =⇒ (i) follow from
the one-variable theory. For the implication (ii) =⇒ (iii), one chooses 0 < ε 1 and uses
the Cauchy integral formula in each variable separately to write
Z Z
1 f (ξ1 , . . . , ξn )
f (z1 , . . . , zn ) = √ · ... · dξ1 · · · dξn
|ξn −cn |=ε (ξ1 − z1 ) · . . . · (ξn − zn )
(2π −1)n
|ξ1 −c1 |=ε
Z
1 f (ξ1 , . . . , ξn )
= √ · · dξ1 · · · dξn .
(2π −1)n T(ε;c) (ξ1 − z1 ) · . . . · (ξn − zn )
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 3
to replace the integrand by a power series, and then integrate term by term.
Using the expression of the am (w) as Cauchy-type integrals, one sees that the am (w) are
holomprphic in w. On the other hand, if |w2 | > 12 then φw (z1 ) is analytic on the whole unit
disk |z1 | < 1, and therefore am (w) = 0 for m < 0 when 21 < |w2 | < 1. It follows from the
Identity principle (Exercise 1.1 (ii)) implies that am (w) = 0 when m < 0 for every w with
|wi | < 1. But then (*) expresses f as a holomorphic function on all of ∆(ε). We refer to [3,
Proposition 1.1.4] for details.
4 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
We conclude this subsection with a Lemma that will be useful on several occasions.
Lemma 1.5. Working with variables (z, w1 , . . . , wn−1 ) on Cn , fix r > ε > 0 and εi > 0. Let
g(z, w1 , . . . , wn−1 ) be a function which is analytic on the cylindrical shell
S = {r − ε < |z| < r + ε} × {|wi | < εi }.
Then the function
Z
f (w1 , . . . , wn−1 ) =def g(ξ, w1 , . . . , wn−1 ) · dξ
{|z|=r}
The reader is asked to prove this as Exercise 1.5. (Or see [3, Lemma 1.1.3].)
In higher dimensions the situation is of course vastly more complicated and interest-
ing: the zeroes of an analytic function form (by definition) a hypersurface in its domain
of definition. The most one can say – which however is already something – is that such
a hypersurface is described locally by the vanishing of a monic polynomial in one variable
whose coeffients are holomorphic functions of the remaining ones. This is the content of the
Weierstrass preparation theorem.
z1
In other words, the zeroes of f in ∆ are given by those of its Weierstrass polynomial. See
Figure 1: the zeroes of the Weierstrass polynomial are shown in the indicated polydisk
around the origin.
Proof. We start with some preliminary remarks. Fix 0 < ε r, and consider for the moment
a function g(z) in a single complex variable z = z1 that is analytic in the disk {|z| < r + ε}.
We assume that g has no zeroes on the circle {|z| = r}. Denote by α1 , . . . , αd the zeroes of
g in ∆ = {|z| < r} (listed with multiplicity), and put
g 0 (ξ)
Z
1
σk = √ ξk · · dξ.
2π −1 |z|=r g(ξ)
A standard residue calculation shows that
d
X
σk = αik .
j=1
On the other hand, consider the polynomial p(z) = z d + a1 z d−1 + . . . + ad−1 z + ad defined by
p(z) = (z − α1 ) · . . . · (z − αd ).
6 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
By the theory of symmetric functions, the coefficients ai of p are given by (universal) poly-
nomials in the σk . On the other hand, we can write g = p · h, where – thanks to the Riemann
singularity theorem – h is an analytic function that is nowhere vanishing on ∆. The idea of
the proof of Theorem 1.6 is to repeat this discussion with g(z) replaced by gw (z1 ), and to
show that the resulting constructions vary holomorphically with w.
The last step in the argument just completed also yields the multivariable analogue of
the Riemann extension theorem.
h0 : U − Zeroes(f ) −→ C.
Assume that h0 is locally bounded near Zeroes(f ). Then h0 extends to a holomorphic function
h on all of U .
Theorem 1.8. (Weierstrass Division Theorem) In the setting of Theorem 1.6, suppose
that f is holomorphic in a neighborhood of the origin, and let p = p(z1 , w) be a Weierstrass
polynomial in z1 of degree d. Then in a suitable polydisk ∆ about the origin, one can uniquely
write
f = q · p + s,
where q is holomorphic on ∆ and s = s(z1 , w) is a polynomial in z1 of degree < d whose
coefficients are holomorphic functions of w.
Sketch of Proof. Keeping notation as in the proof of the Weierstrass preparation theorem,
define
Z
1 f (ξ, w)
q(z1 , w) = √ · · dξ.
2π −1 |z1 |=r p(ξ, w) · (ξ − z1 )
We assume that p(z1 , w) has no zeroes on the shell
S = {r − ε ≤ |z1 | ≤ r + ε} × {|wi | ≤ εi },
But since p is a Weierstrass polynomial of degree d in z1 , the second term in the integral is
a polynomial of degree ≤ d − 1 in z1 , as required. For the uniqueness statement, see [3, p.
15].
8 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Algebraic Properties of the Ring of Holomorphic Germs. We now apply the Weier-
strass theorems to study the algebraic properties of the ring of germs of holomorphic functions
in a neighborhood of the origin 0 ∈ Cn .
For local questions, it is useful to identify two functions if they agree in a neighborhood
of a fixed point, say 0 ∈ Cn : this leads to the notion of the germ of a holomorphic function.
More formally, consider pairs (U, f ) where U is a neighborhood of 0 ∈ Cn , and f is a
holomorphic function on U . Define an equivalence relation ∼ on such pairs by declaring that
(U1 , f1 ) ∼ (U2 , f2 ) if there is a neighborhood V ⊆ U1 ∩ U2 such that
f1 |V ≡ f2 |V.
The resulting equivalence classes are called germs of holomorphic functions about 0, and
form a ring
On = OCn ,0 . (1.4)
Via Taylor expansions, there is an isomorphism
On ∼= C{z1 , . . . , zn }
of On with the ring of convergent power series in z1 , . . . , zn . This is a local integral domain:
On has a unique maximal ideal, consisting of (germs of) functions vanishing at the origin.
Of course one can discuss in a similar manner germs of holomorphic functions at any fixed
point a ∈ Cn , or germs of smooth (or continuous, or ...) functions at a point.
Remark 1.9. As we shall see later, the notation OCn ,0 indicates the fact that On is the
stalk at 0 of the sheaf OCn of holomorphic functions on Cn .
Sketch of Proof. Induction on n, the case n = 1 being clear.1 Given f ∈ On vanishing at the
origin, we can assume after a linear change of coordinates that the hypothesis (1.2) of the
Weierstrass preparation theorem are satisfied. Thus we can write
f = p · u,
where p ∈ On−1 [z1 ] is (the germ of) a Weierstrass polynomial, and u ∈ On is a unit. Since
a polynomial ring over a UFD is itself a UFD, it follows by the induction hypothesis that
On−1 [z1 ] is a UFD. Thus p can be written essentially uniquely as a finite product of irreducible
elements of On−1 [z1 ], and it is elementary that any factor of a Weierstrass polynomial is
again a Weierstrass polynomial. The essential point is then to show that if p is irreducible
in On−1 [z1 ], then p remains irreducible considered as an element of On . To this end, suppose
1In fact, any f ∈ C{z} can be written uniquely in the form f = z m · u for some unit u ∈ Cz.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 9
Proof. By induction on n, we may suppose that On−1 is Noetherian, and then it follows
by the Hilbert basis theorem that the polynomial ring On−1 [z1 ] is also Noetherian. Taking
z1 , w2 , . . . , wn as the coordinates on Cn , we may view
On−1 [z1 ] ⊆ On
as a subring of On . Now let I ⊆ On be any non-trivial ideal. Note to begin with that –
possibly after a linear change of coodinates – I contains a Weierstrass polynomial p(z1 , w)
in z1 . In fact, if f ∈ I is any element satisfying (1.2), then f = p · h, for some Weierstrass
polynomial p and some unit h ∈ On , and hence p ∈ I. On the other hand, the ideal
I 0 = I ∩ On−1 [z1 ]
is finitely generated, and we choose generators p1 , . . . , pt ∈ I 0 . We assert that p, p1 , . . . , pt
generate I. In fact, consider any f ∈ I. By the Weierstrass division theorem, we can write
f = q · p + s,
where s ∈ On−1 [z1 ]. But evidently s ∈ I, so s is an On linear combination of p1 , . . . , pt , and
we are done.
We conclude this subsection by discussing ideals in On . These are related to (germs of)
analytic subvarieties of Cn . We start with a definition:
Definition 1.12. (Analytic subset or variety) Let U ⊆ Cn be an open set. An analytic
subset or subvariety 2 of U is a subset X ⊆ U with the property that for each point x ∈ X,
there is a neighborhood V 3 x of x in U such that
V ∩ X = Zeroes{f1 , . . . , fk }
for some holomorphic functions f1 , . . . , fk ∈ O(V ).
2In algebraic geometry, the term “variety” is sometimes reserved for an irreducible set.
10 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
It follows that there is a one-one correspondence between (germs of) analytic subvarieties
of Cn and radical ideals in OCn ,0 (Exercise 1.11). We will not discuss the proof of the
Nullstellensatz here. For a partial proof, see [3].
√
As a matter of notation, we take wj = uj + −1 · vj to be the coordinate functions on Cm ,
and we write √
w = f (z) = u(z) + −1 · v(z).
The first point is to discuss the derivatives of such a mapping. Suppose for the moment
that f is merely C ∞ . We have natural identifications
Cn = R2n , Cm = R2m ,
so we can view f as a mapping between domains in R2n and R2m . In particular, for each
a ∈ U we can form the derivative
DR f = (DR f )a : Ta R2n −→ Tf (a) R2m . (1.8)
This derivative is given by the usual real Jacobian matrix JR (f ) of f : with respect to the
standard real bases of the spaces in question, JR f is represented by the block matrix:
∂u ∂u
∂x ∂y
JR (f ) =
,
(1.9)
∂v ∂v
∂x ∂y
∂ui
where eg ∂u
∂x
denotes the matrix ∂xj
, and we have supressed evaluation of the derivatives
at a.
12 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Now we want to bring the complex structures back into the picture. Namely, there are
natural identifications
Ta Cn = Cn , Tf (a) Cm = Cm .
In particular,√the source and target of (1.8) are actually complex vector spaces, with multi-
plication by −1 given by
√ ∂ ∂ √ ∂ ∂
−1 · = , −1 · = . (*)
∂xj ∂yj ∂ui ∂vi
∂
We may provisionally take , ∂
∂xj ∂ui
as complex bases of the tangent spaces in question.3
The basic fact is that f is holomorphic if and only if its derivative is complex linear:
Proposition 1.17. The C ∞ -mapping f : U −→ V is holomorphic if and only if the derivative
(DR f )a : Ta Cn −→ Tf (a) Cm
is C-linear for every a ∈ U . In this case, the derivative is represented with respect to the
standard bases by the complex Jacobian matrix
∂fi
J(f )a =def (a)
∂zj
of f , where as above
∂fi √
∂fi 1 ∂fi
= − −1 · .
∂zj 2 ∂xj ∂yj
Proof. This is a restatement of the Cauchy-Riemann equations using (1.9). Omitting again
evaluation of the derivatives, JR (f ) is C-linear if and only if one has the matrix identities
∂u ∂v ∂u ∂v
= , = − ,
∂x ∂y ∂y ∂x
which are the Cauchy–Riemann equations. Moreover if these hold, then the C-linear mapping
determined by DR f is given by the m × n complex matrix
√
∂u ∂v ∂f
+ −1 · = .
∂x ∂x ∂z
As a consequence, we find:
For future reference, it will be useful to record the relation between the real and complex
derivatives in the case when f is merely C ∞ . To this end, it is convenient to complexify the
real tangent spaces in question, i.e. we consider DR f as defining a C-linear mapping
C2n = Ta R2n ⊗R C −→ Tf (a) R2m ⊗R C = C2m . (*)
We take the complex tangent vectors
∂ ∂ ∂ ∂
, , ,
∂zj ∂ z̄j ∂wi ∂ w̄i
as bases for the two sides in (*). Then the matrix representing (*) has the block form
∂f ∂f
∂z ∂ z̄
, (1.10)
∂ f¯ ∂ f¯
∂z ∂ z̄
where eg ∂f ∂fi
∂z
denotes the matrix ∂zj
, and we have again supressed mention of evaluation
at a. This yields convenient way of dealing with the chain rule.
The next point is that the inverse and implicit function theorems hold in the complex
setting.
Theorem 1.19. (Inverse and Implicit Function Theorems) Let U ⊆ Cn and V ⊆ Cm
be open sets, and let
f : U −→ V
be a holomorphic mapping. Assume that
Dfa : Ta U −→ Tf (a) V
is surjective at some point a ∈ U (and hence also in a neighborhood of a).
Sketch of Proof. Consider first the case n = m as in (i). Thanks to Proposition 1.18, it
follows from the usual smooth inverse function theorem that a local inverse g = f −1 exists as
a C ∞ mapping, and the issue is to show that g is holomorphic. Tothis end, we differentiate
the relation g ◦ f = Id. With notation as in (1.10) one has ∂(Id) ∂ z̄
= 0, and this yields the
matrix equation
¯
∂g ∂f ∂g ∂f ∂g
0 = · + · = · J(f ).
∂w ∂ z̄ ∂ w̄ ∂ z̄ ∂ w̄
Since J(f ) is non-singular, it follows that ∂∂gw̄ = 0, i.e. g is holomorphic. The proof of (ii)
is similar, and we refer to [3, p. 12] for details.
Exercise 1.4. (i). Prove that if f, g are holomorphic on U ⊆ Cn , then the product rule
holds for each of their z-derivatives, i.e.
∂(f g) ∂g ∂f
= f· +g· .
∂zi ∂zi ∂zi
(ii). Deduce that for a single variable z,
∂ m
(z ) = m · z m−1 .
∂z
(iii). In the setting of Theorem 1.1 (iii), show that the coefficient of z1j1 . . . znjn in the Taylor
series of f (z) is given by
1 ∂ j1 +···+jn f
aj1 ,...,jn = · j1 jn c .
j1 ! · · · jn ! ∂z1 · · · ∂zn
Exercise 1.5. Prove Lemma 1.5. (Cf. [3, Lemma 1.1.3].)
Exercise 1.6. Show that the Weierstrass polynomial associated to a given holomorphic
function f (z1 , w) is unique.
Exercise 1.7. Working in C2 with variables z, w, find the Weierstrass polynomial for
sin(w2 − z 3 ).
Exercise 1.8. Prove the Riemann Extension Theorem 1.7. Show that the conclusion holds
more generally if the hypersurface {f = 0} is replaced by any analytic subvariety X ⊆ U .
Exercise 1.9. Denote by An the ring of germs at 0 ∈ Cn of smooth functions. In contrast
to On , the ring An is not well-behaved algebraically. Specifically, prove that it is neither an
integral domain nor Noetherian.
Exercise 1.10. Prove the equality (1.5)
√
Exercise 1.11. Recall that an ideal J ⊆ OCn ,0 is said to be radical if J = J. Granting
the Nullstellensatz, prove that there is a one-to-one order-reversing correspondence between
germs of analytic subvarieties of Cn , and radical ideals in OCn ,0 .
Exercise 1.12. Prove Proposition 1.15 .
Exercise 1.13. In contrast to the case of holomorphic mappings between domains in C, it
is not the case in general that a holomorphic map between domains in higher dimensions is
an open mapping. For example, show that
f : C2 −→ C2 , (z, w) 7→ (z, zw)
is not an open mapping.
Exercise 1.14. Let f : U −→ V be a holomorphic mapping between domains U, V ⊆ Cn .
Then
det JR (f ) = | det J(f )|2 .
In particular, a biholomorphic mapping is always orientation-preserving.
16 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
2. Complex manifolds
In this section we introduce our principle objects of interest, namely complex manifolds.
Definition and First Examples. As one might expect, a complex manifold is simply a
differentiable manifold that carries holomorphic coordinate charts.
Definition 2.1. (Holomorphic atlas, complex manifold) Let X be a smooth manifold.
Observe that since biholomorphic maps are orientation preserving (Exercise 1.14), a complex
manifold X is automatically orientable as a real manifold.
In the usual way, all of the notions studied locally in the previous section make sense on
such a manifold.
Definition 2.2. (Holomorphic function, mapping) Let X be a complex manifold.
These are independent of the choice of open set or the atlas thanks to the fact that the
functions appear in (2.1) and (2.2) are biholomorphic.
In the future, we will abbreviate this sort of definition by saying for example that
a function f on a complex manifold is holomorphic if it is given in local coordates by a
holomorphic function.
Definition 2.3. (Isomorphism) An isomorphism between complex manifolds is a holo-
morphic mapping
f : X −→ Y
that possesses a holomorphic inverse. In this case one says that X and Y are isomorphic or
biholomorphic.
Example 2.4. Any open subset U ⊆ X of a complex manifold is iteself a complex manifold.
In particular, any open subset U ⊆ Cn is a complex manifold of dimension n.
Example 2.5. (Riemann sphere) The standard Riemann sphere S = C ∪ {∞} is a
compact complex manifold (of dimension 1). Recall that in the finite part of the plane one
takes the usual coordinate function z as the local coordinate, while in a neighborhood of
the point at infinity one uses the coordinate w with w = z1 . In other words, the transition
function associated to this atlas is
1
h : C − {0} −→ C − {0} , z 7→ .
z
The Riemann sphere is also the one-dimensional instance of complex projective space Pn :
we will study this manifold in detail in the next subsection.
Example 2.6. (Analytic hypersurfaces) Let U ⊆ Cn+1 be an open set, let f : U −→ C
be a holomorphic function, and denote by
X = Zeroes(f ) ⊆ U
the zero-locus of f . Assume that for each point a ∈ X there is at least one index i such that
∂f
(a) 6= 0.
∂zi
Then X is a complex manifold of dimension n. (In fact, it follows from the Implicit Function
Theorem 1.19 (ii) that the n remaining coordinate functions on Cn+1 give local coordinates
in a neighborhood of a.) This construction leads to a vast number of important concrete
examples. Note that when n = 1 we get further examples of one-dimensional complex
manifolds.
Definition 2.7. (Riemann surface, complex curve) A Riemann surface is a complex
manifold of dimension one. These are also called smooth complex curves, especially when
they arise as in the previous two examples.
Example 2.8. Generalizing the construction of the previous example, suppose that V is a
complex manifold of dimension n + k, and that
X ⊆ V
18 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
One can also talk about possibly non-closed submanifolds, whose definition we leave to
the reader. However since we are mainly interested in closed ones, we adopt the
Convention 2.10. By a “submanifold” X ⊆ Y , we always mean unless otherwise stated a
closed submanifold.
Example 2.11. (Complex Lie Groups) A complex Lie group is a conplex manifold G,
which is simultaneously a group, in such a way that multiplication and inversion are given
by holomorphic mappings
mult : G × G −→ G , inv : G −→ G.
For example, there is a natural identification
2
Matn×n (C) = Cn
2
of the set of all n × n complex matrices with Cn : the coordinate functions are the entries
of the matrix. Then
GLn (C) =def {A ∈ Matn×n | det(A) 6= 0}
is an open subset of Matn×n (C), hence a complex manifold, and thanks to Cramer’s rule
multiplication and inversion are holomorphic maps. Thus GLn (C) is a complex Lie group.
The other standard linear groups – notably the special linear, orthognal and symplectic
groups SLn (C), On (C), Sp2n (C) – are closed submanifolds of GLn (C), and hence themselves
complex Lie groups (Exercise 2.6). On the other hand, observe that the unitary and special
unitary groups Un and SUn are not complex submanifolds of GLn (C). In fact, it turns out
that the only compact connexted complex Lie groups are complex tori, which we discuss
next.
Example 2.12. (Complex tori, I). Let
Λ ⊆ Cn
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 19
Remark 2.14. It is an interesting (but somewhat counter-intuitive) fact that if one starts
with a “random” lattice Λ ⊆ Cn , the resulting torus XΛ does not contain any proper analytic
submanifolds or subvarieties of positive dimension. Equivalently, there are no non-trivial Λ-
periodic sub varieties of Cn . On the other hand, if for instance
√
Λ = Zn + −1 · Zn ⊆ Cn ,
then XΛ is a product of one-dimensional complex tori, and so carries lots of submanifolds
(e.g. sub-products).
Example 2.15. (One-dimensional tori). We consider in more detail one-dimensional
complex tori. Denote by H the upper half plane in C. After a linear change of coordinates,
any lattice in C is spanned by 1 and some complex number τ ∈ H: write
Λτ = Z + Z · τ ⊆ C,
and let Xτ denote the corresponding one-dimensional torus. We will show that
aτ + b
Xτ ∼
= Xτ 0 ⇐⇒ τ 0 = (*)
cτ + d
for some
a b
∈ SL2 (Z).
c d
Granting this for the moment, it follows that isomorphism classes of one-dimensional complex
tori are parametrized by H/SL2 (Z) under the action defined by (*). This quotient space has
been much studied: for example, a fundamental domain is given (modulo some identifications
on the boundary) by
|τ | ≥ 1 , |Re(τ )| ≤ 12 .
In fact, it turns out (somewhat non-trivially) that H/SL2 (Z) can itself be given the structure
of a one-dimensional complex manifold with
H/SL2 (Z) ∼
= C.
In summary, isomorphism classes of one-dimensional complex tori are parametrized by the
complex plane: this is the first example of a moduli space, ie a complex manifold (or analytic
variety) that parametrizes isomorphism classes of complex manifolds of fixed diffeomorphism
type. As for the proof of (*), it follows from Example 2.13 that Xτ ∼ = Xτ 0 if and only if
there is a complex number µ ∈ C∗ such that multiplication by µ carries Λτ 0 onto Λτ . This
is equivalent to asking that
µ · τ 0 = aτ + b , µ · 1 = cτ + d
0 aτ +b a b
for some integers a, b, c, d ∈ Z. Hence τ = ,
and the resulting matrix must have
cτ +d c d
determinant = 1 since it is invertible and maps the upper half plane to iteself.
Example 2.16. (Hopf manifolds). Given a real number 0 < λ < 1 and an integer k ∈ Z,
define
σk : Cn − {0} −→ Cn − {0} , σk (z1 , . . . , zn ) = (λk z1 , . . . , λk zn ).
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 21
This defines a fixed point-free action of Z on Cn − {0}, and every point z ∈ Cn − {0} has
an open neighborhood disjoint from all of its images under the action. It follows that the
quotient
X = (Cn − {0}) /Z
carries in the natural way the structure of a complex manifold of dimension n, called the
Hopf manifold. It is diffeomorphic to the product S 2n−1 × S 1 . (See Exercise 2.9.) The
interest in this construction, as we will see later, is that it gives an example of a complex
manifold which cannot carry a Käbler metric.
Especially for algebraic geometers – who like to define spaces by means of defining equa-
tions – the construction in Example 2.6 and its generalizations are fundamental. But it
suffers from one serious drawback, namely that positive-dimensional closed submanifolds of
Cn are never compact. In fact, any submanifold of Cn carries lots of non-constant holo-
morphic functions (for instance the coordinate functions). On the other hand, one has the
elementary but basic
Proposition 2.17. Let X be a compact connected complex manifold. Then the only global
holomorphic functions on X are constants.
It is very important to have some ambient complex manifold in which one can “write
down” (using equations of some form) many compact submanifolds. We turn next to the
construction of projective space, which serves this function.
Projective Space. There are various equivalent definitions of projective space. As our base
construction we will take
Definition 2.18. Let C∗ act on Cn+1 − {0} by scalar multiplication, i.e.
λ · (a0 , . . . , an ) = (λ · a0 , . . . , λ · an ) for λ 6= 0. (2.3)
We define
Pn = Pn (C) =def Cn+1 − {0} /C∗
This specifies Pn as a set and a topological space. We will show momentarily how it carries
the structure of a complex manifold. Note that all the fibres of π are copies of C∗ sitting in
a one-dimensional complex subspace of Cn+1 .
22 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
A couple of preliminary remarks will be useful. First, note that any non-zero vector
0 6= a ∈ Cn+1 determines a unique one-dimensional complex subspace
C∼= C · a ⊆ Cn+1 .
Moreover, two non-zero vectors a, a0 ∈ Cn+1 determine the same subspace if and only if
a0 = λ · a for some λ ∈ C∗ .
Hence
Viewpoint 2.19. The projective space Pn parametrizes all one-dimensional vector sub-
spaces of Cn+1 .
This being the case, it is sometimes useful to take a coordinate-free approach. Specifically,
given any complex vector space V of dimension n + 1, we denote by P(V ) the n-dimensional
projective space of one-dimensional vector subspaces of V .5
Next, there is a very useful way to specify points in projective space by so-called homo-
geneous coordinates. Namely, we can view (2.3) as determining an equivalence relation on
all non-zero vectors 0 6= a ∈ Cn+1 , where
a0 ∼ a ⇐⇒ a0 = λ · a for λ ∈ C∗ ,
and then evidently n o
Pn = a ∈ Cn+1 − {0} / ∼ .
Given a non-zero vector a = (a0 , . . . , an ), denote by
[a] = [a0 , . . . , an ]
its equivalence class. Thus [a] determines a point π(a) ∈ Pn with homogeneous coordi-
nates [a0 , . . . , an ]. In other words, points in Pn are specified by homogeneous coordinates
[a0 , . . . , an ] with at least one non-zeroi entry, where
[a00 , . . . , a0n ] = [a0 , . . . , an ] ⇐⇒ a0i = λ · ai for some λ ∈ C∗ .
Remark 2.20. Note that one can replace C in these definitions and constructions by any
field k, leading to
Pn (k) = k n+1 − {0} /k ∗ .
In other words, Pn carries the structure of a complex manifold in such a way that first of all
π is holomorphic. Moreover we can find an open covering {Ui } of Pn by open sets with the
property that
π −1 (Ui ) ∼
= Ui × C∗ (2.5)
as complex manifolds.
It is perhaps not yet apparent that Pn is compact. To see this, let k k denote the
standard Euclidean norm on Cn , and consider the unit sphere
S 2n+1 = {k z k = 1} ⊆ Cn+1 .
Consider the restriction
h : S 2n+1 −→ Pn
to S 2n+1 of the basic quotient mapping π in (2.4). The compactness of Pn is a consequence
of the fact that h is surjective, which in turn follows from the fact that
a
π(a) = π
kak
for any non-zero vector a ∈ Cn+1 . To say the same thing a little differently, consider the unit
circle S 1 = {|z| = 1} ⊆ C of complex numbers of length 1. Then S 1 acts on S 2n+1 by scalar
multiplication, and h is the quotient map. The mapping h is called the Hopf vibration: the
reader should check that it expresses S 2n+1 as the total space of a locally trivial S 1 bundle
over Pn .
24 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Example 2.23. The case n = 1 of this discussion is already very interesting. Recalling that
P1 = S 2 is the Riemann sphere, the Hopf fibration in this case is a smooth mapping
h : S 3 −→ S 2
expressing the three-sphere as a circle bundle over the two-sphere. In particular, for any
point p ∈ S 2 , the fibre h−1 (p) is a circle in S 3 . The interesting fact is that for distinct
p, q ∈ S 2 , these circles are linked. (One can verify this pictorially by making a stereographic
projection from S 3 − {point} to R3 and working out explicitly the equations of two fibres.)
X0 ⊆ U0 = Cn
26 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
XY - Z2 = 0
Z=0
Y=0 X=0
Thanks to the compactness of Pr , any projective manifold is compact. It may happen (and
in fact always does) that a given complex manifold can be realized in different ways as
submanifolds of projective spaces of various dimensions.
Remark 2.29. (Chow’s Theorem). A remarkable theorem of Chow states that if X ⊆ Pr
is a complex submanifold, then in fact X is the algebraic set cut out by a collection of
homogeneous polynomials. The same statement remains true if one asks merely that X be
an analytic subvariety of Pr , ie locally cut out by holomprohic functions. Unfortunately we
won’t be able to prove this in Math 545, but it’s good to keep the fact in mind.
Proof. This follows from Example 2.6 (ie the implicit function theorem) together with Euler’s
Theorem: if F is homogeneous of degree d, then
∂F ∂F
+ ... + = d · F.
∂Z0 ∂Zn
We leave detais to the reader.
Intrinsically, given vector spaces V and W , one can describe the Segre embedding as the
map
P(V ) × P(W ) ,→ P(V ⊗ W ) , [v] × [w] 7→ [v ⊗ w].
(Recall that V ⊗ W consists of linear combinations of pure tensors v ⊗ w; the Segre variety
consists of the projectivization of the set of all elements of V ⊗ W that happen to be of the
form v ⊗ w for some v ∈ V and w ∈ W .)
is an easier “internal” way to work with these. Specifically, one says that a polynomial
P (Z, W ) = p(Z0 , . . . , Zr , W0 , . . . , Ws )
is bihomogeneous of bidegree (d, e) if P is homogeneous of degree d in the Zi and homoge-
neous of degree e in the Wj . For example
Z03 W0 W1 + Z12 Z2 W22
is bihomogeneous of degree (3, 2), but Z1 − W1 is not bihomogeneous. If P is bihomogeneous
of bidegree (d, e), then
P (λZ, µW ) = λd µe P (Z, W ),
so the zero-locus of a bihomogeneous polynomial is a well-defined subseteq of Pr × Ps . We
can then define an algebraic subset of Pr × Ps to be the common zero-locus of a collection
of bihomogeneous polynomials; the reader is asked to check in Exercise 2.17 that this is the
same thing as intersecting the Segre variety with an algebraic subset of Prs+r+s .
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 29
This is an algebraic subset: in fact, v = (v0 , . . . , vn ) lies on the line spanned by [a] =
[a0 , . . . , an ] if and only if the vectors v and a are linearly dependent, so B is cut out in
Pn × Cn+1 by the 2 × 2 minors of the matrix
Z0 . . . Zn
.
w0 . . . wn
It is very interesting to consider the two projections
p : Z −→ Pn , b : Z −→ Cn+1 .
For any point [a] ∈ Pn , the fibre p−1 ([a]) is the one-dimensional subspace C · a spanned by
a, and in fact Z is locally a product of Pn and C. On the other hand, consider the fibres
of b. If 0 6= v ∈ Cn+1 , then v lies on a unique line C · v ⊆ Cn+1 , and b−1 (v) consists of the
corresponding point of Pn ; in other words, b is an isomorphism away from 0. On the other
hand, b−1 (0) is all of Pn . The mapping b : Z −→ Cn+1 is called the blowing-up of Cn+1 at
the origin. Homotopically, Z is obtained from Cn+1 by removing a small disk about 0, with
boundary S 2n+1 , and gluing in a copy of Pn via the Hopf map h : S 2n+1 −→ Pn . See Figure
3.
Holomorphic Line and Vector Bundles. Given a projective manifold, we have seen
(Proposition 2.31) that one can use homogeneous polynomials to define holomorphic map-
pings to Ps . But suppose one starts with an abstract compact complex manifold X. How
then can one attempt to construct maps – and ideally embeddings – φ : X −→ Pr ? The
answer is that such morphisms are defined via sections of holomorphic line bundles; φ nec-
essarily has the form
φ(x) = [s0 (x), . . . , sr (x)]
where si (x) ∈ Γ(X, L) are sections of a holomorphic line bundle without no common zeroes.
In this subsection we give the basic definitions and examples of line and vector bundles.
First Viewpoint (Total Space). Intuitively, we can define a holomorphic vector bundle
on X to be a holomorphic family of vector spaces parametrized by X.
Definition 2.36. A holomorphic vector bundle on X of rank e is a complex manifold E,
together with a holomorphic mapping
p : E −→ X,
with the property that p realizes E a locally the product of X with an e-dimension vector
space in such a way that the transition functions are given by linear automorphisms.
The definition requires some elaboration. To begin with, we ask that there be given an
open covering {Ui } of X together with biholomorphic isomorphisms
∼
=
φi : p−1 (Ui ) −→ Ui × Ce
commuting with the projections to Ui :
∼
φi : p−1 (Ui )
= / Ui × Ce
p pr1
% {
Ui .
Now set
Uij = Ui ∩ Uj ,
and consider the “j-to-i” transition map
φij = φi ◦ φ−1 e e
j : Uij × C −→ Uij × C .
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 31
Definition 2.37. A vector bundle of rank e = 1 is called a line bundle. Note that in this
case the gij (x) are simply nowhere-vanishing holomorphic functions on Uij , i.e.
gij ∈ O∗ (Ui ∩ Uj ).
Remark 2.38. These definitions make sense in any fixed geometric setting: one requires
that the local product isomorphisms φi and the transition matrices gij be morphisms in
the appropriate category. So for example one can discuss continuous real or complex vector
bundles on a topological space, or C ∞ vector bundles (real or complex) on a smooth manifold.
So for example, if M is a smooth manifold of dimension n, its tangent bundle T M is a real
vector bundle of rank n. (See Exercise 2.18.)
Example 2.39. The global product 1e = 1eX = X × Ce is called the trivial vector bundle
of rank e. As its name implies, this is not a very exciting example.
32 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Example 2.40. (OPn (−1)). As in Example 2.35, consider the incidence correspondence
Pn × Cn+1 ⊇ L = ([a], v) | v ∈ C · a .
Then the first projection p : L −→ Pn realizes L a line bundle on Pn , which is called (for
reasons that will become apparent later) OPn (−1).9 We work out the transition functions for
this bundle, assuming for ease of notation that n = 1. On P1 with homogeneous coordinates
[Z0 , Z1 ] we consider as usual the open sets
U0 = [1, ZZ10 ] , U1 = [ ZZ01 , 1]
In order to determine the transition function g10 , we compute the composition φ1 ◦ φ0−1 :
[1, ZZ01 ], t 7→ [Z0 , Z1 ], t · (1, ZZ01 ) = [Z0 , Z1 ], ( ZZ10 t) · ( ZZ01 , 1) 7→ [ ZZ01 , 1], ( ZZ10 t) .
In other words,
Z1
g10 [Z0 , Z1 ] = .
Z0
Similarly, with respect to the standard open covering Ui = {Zi 6= 0} of Pn , the transition
functions of OPn (−1) are given by
Zi
gij = . (2.9)
Zj
Example 2.41. (OPn (1)). In this example, we describe the hyperplane line bundle on Pn ,
which is also called OPn (1). Embed Pn into Pn+1 as the hyperplane Zn+1 = 0, and set
O = [0, . . . , 0, 1] ∈ Pn+1 . Linear projection from O defines a mapping
p : Pn+1 − {O} −→ Pn , [Z0 , . . . , Zn+1 ] 7→ [Z0 , . . . , Zn ].
Geometrically, p sends a point x ∈ Pn+1 − {O} to the unique point of intersection of the
line joining O and x with the hyperplane Pn ⊆ Pn+1 . (See Figure 5.) The reader should
check that p realizes Pn+1 − {O} as the total space of a line bundle on Pn , whose transition
functions with respect to the standard open covering {Ui } of Pn are given by
Zj
gij = . (2.10)
Zi
9This is sometimes also called the Hopf line bundle, or the tautological line bundle (since its fibre over a
point in Pn is “the line that that point is”).
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 33
Note that if p : E −→ X is a vector bundle, and x ∈ X is any point, then the fibre p−1 (x)
has the structrue of a vector space (defined up to isomophism): this comes from any of the
local trivializations of E and the fact that the comparison maps are linear isomorphisms.10
The general yoga is that any intrinsically defined notion or operation on vector spaces makes
sense for vector bundles. For example, if
p : E −→ X , q : F −→ X
are two vector bundles, then a homomorphism u from E to F is a holomorphic mapping
u : E −→ F,
commuting with p and q, such that u is linear on each fibre. (The reader should convince
him/herself that the linearity condition is well-defined.) Similarly, one can associate to
p : E −→ X the dual bundle E∗ −→ X whose fibre over a point x ∈ X is the vector space
of linear functionals on p−1 (x). (See Exercise 2.22.)
For our purposes, the most important feature of a bundle are its global sections:
Definition 2.42. Let p : E −→ X be a holomorphic vector bundle. A global section of E is
a holomorphic mapping
s : X −→ E such that p ◦ s = idX .
The condition on p ◦ s = idX guarantees that s(x) ∈ p−1 (x) for every x ∈ X. Thus a section
picks out a vector in each fibre of E over X. We denote by
Γ(X, E) or Γ(X, O(E))
the set of all global sections of E. Thanks to the vector space structure on the fibres of p,
we can add sections and multiply them by functions. Thus Γ(X, O(E)) is a module over the
ring of O(X) of holomorphic functions on X; in particular, Γ(X, E) is naturally a complex
vector space. The zero section 0 = 0E ∈ Γ(X, O(E)) is the section that picks out the zero
vector in each fibre. See Figure 6.
Example 2.43. A section of the trivial line bundle 1 is just a holomorphic function on X.
Example 2.44. Working in the C ∞ setting, a section of the tangent bundle of a smooth
manifold M is a vector field on M .
It is important to work out the meaning of sections in terms of transition data. Suppose
that p : E −→ X is described by data {Ui , gij }: in other words, {Ui } is an open cover of
X on which p trivializes, and gij ∈ GLe (O(Uij )) are the corresponding transition matrices.
Then the composition
∼
=
φi ◦ s : Ui −→ p−1 (Ui ) −→ Ui × Ce
is the graph of a vector-valued holomorphic function
fi : Ui −→ Ce :
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 35
The most important case of this principle occurs when e = rank E = 1, in which case
sections of suitable line bundles serve as substitutes for global holomorphic functions. This
is illustrated by the basic
Proposition 2.45. Global sections of OPn (1) are identified with linear forms on Pn , ie
Γ Pn , OPn (1) ∼
= linear forms in Z0 , . . . , Zn .
Proof. For simplicity we will treat the case n = 1. Recall that in general the transtion
functions of OPn (1) with respect to the standard open covering {Ui } of Pn are given by
gij = Zj
Zi
. In the case n = 1 this gives the one transition function
Z0
g10 = .
Z1
Now say s ∈ Γ(P1 , OP1 (1)). Then s|U0 and s|U1 are given by entire functions s0 ( ZZ01 ) and
s1 ( ZZ10 ) satisfying
s1 ZZ10 = ZZ01 · s0 ZZ10 . (*)
Writing
2
Z1
s 0 = a0 + a1 + a2 ZZ10 + . . .
Z0
2
s1 = b0 + b1 ZZ10 + b2 ZZ01 + . . . ,
(*) immediately implies that
a2 = a3 = . . . = 0 , b2 = b3 = . . . = 0,
and
a0 = b 1 , a 1 = b 0 .
Thus s may be identified with the linear form
s = a0 Z0 + a1 Z1 ,
s s
with s0 = Z0
and s1 = Z1
.
36 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Remark 2.46. (Finite dimensionality). Note that in particular Γ(Pn , OPn (1)) is a finite-
dimensional vector space. This is a special case of a non-trivial theorem to the effect that if
X is a compact complex manifold, then the space Γ(X, O(E)) of sections of any holomorphic
vector bundle is finite-dimensional.
Suppose that E −→ X and L −→ X are respectively a vector bundle and a line bundle
given by transition data {Ui , gij } and {Ui , hij }. Then we may form the vector bundle
E⊗L:
this is a bundle having the same rank as E, and transition matrices hij · gij .11 In particular
for k ≥ 1, the k-fold tensor product L⊗k is described by the transition functions (hij )k . For
projective space, the computation of Proposition 2.45 generalizes to yield:
Proposition 2.47. Set
⊗k
OPn (k) = OPn (1) .
Then
Γ Pn , OPn (k) ∼
= homogeneous polynomials of degree k in Z0 , . . . , Zn .
We can now explicate when data {Ui , gij } and {Ui , gij0 } determine isomorphic bundles.
By the discussion of the previous paragraph, this happens if and only if there exist invertible
matrices
νi ∈ GLe (O(Uij ))
of holomorphic functions such that
gij = νi−1 ◦ gij0 ◦ νj
on Uij . In particular, the bundle determined by data {Ui , gij } is isomorphic to the trivial
bundle if and only if
gij = νi−1 ◦ νj .
Given a compact complex manifold X, we can now begin to approach the question
whether X admits a holomorphic embedding into some projective space. In order to produce
such an embedding one needs to produce a holomorphic
line bundle L on X that in the first
place carries enough sections sα ∈ Γ X, O(L) so that (2.13) holds, and so that moreover
the resulting holomorphic mapping
φ : X −→ Pr
actually be an embedding. It is a remarkable fact that in many ways the most appealing
characterization of when this happens is differentio-geometric in nature. Specifically, one
starts by choosing a Hermitian metric on the fibres of L, which gives rise to a curvature form
Θ, which is a closed two-form representing the so-called first Chern class of L:12
[Θ] = c1 (L) ∈ H 2 (X, Z).
The first basic theorem – the Lefschetz (1, 1) Theorem (Theorem 6.29) – allows one under
suitable hypotheses to identifiy which cohomology classes arise in this fashion. The second
fundamental result – the Kodaira Embedding Theorem (Theorem 7.24) – states that if
Θ = Θ(L) satisfies a certain positivity hypothesis, then some large tensor power of L carries
enough holomorphic sections to define an embedding. Putting these two theorems together,
one arrives at a complete characterization of when a compact complex manifold admits a
projective embedding: it should carry a Kähler metric with rational periods (Exercise 7.7).
Our main goal in the rest of the class is to prove these results. This requires that we
study calculus on complex manifolds, to which we turn in the following sections.
Exercise 2.3. Let U ⊆ Cn be an open set, and f = f (z) ∈ O(U ) a non-zero holomorphic
function. Denote by X ⊆ U the complement of the zero-locus of f :
X = U − {f = 0} ⊆ U.
Show that X is isomorphic to a closed submanifold of U × C. (Hint: consider the function
f (z) · t − 1 ∈ O(U )[t].)
12We’re ignoring some normalizing constants here
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 39
(i). Prove that the condition of Example 2.6 is met, and hence that X is a complex
manifold of dimension 1.
(ii). By studying the projection
π : X −→ C , (z, w) 7→ z,
show that one can compactify X by adding two points “at infinity” to obtain a
compact Riemann surface X in such a way that π extends to a holomorphic map
π : X −→ S = C ∪ {∞}.
(Hint: Because p has even degree, the inverse image under π of the complement D∗
of a disk of very large radius consists of the disjoint union of two copies of D∗ . But
D∗ is a punctured neighborhood of the point at infinity, and therefore we need to
add two points to compactify X.)
Exercise 2.6. (Linear groups) Keeping the notation of Example 2.11, let
det : GLn (C) −→ C∗
be the holomorphic function that takes a matrix to its determinant. Prove that the complex
Jacobian of det is non-zero at any invertible matrix A. (Reduce to the case A = Id.)
Conclude that
SLn (C) =def {A | det A = 1}
GLn (C), and hence a complex Lie group. The complex orthogonal and symplectic groups
can be treated similarly.
Exercise 2.7. (Covering spaces). Let
f : Y −→ X
be a covering space in the sense of topology.
40 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
In this section we discuss the linear algebra underlying calculus on complex manifolds.
We follow closely the presentation of Huybrechts in [3, Chapter 1.2].
Note that giving such a J is equivalent to equipping V with the structure of a complex
vector space by taking √
−1 · v = J(v).
In particular, the existence of an almost complex structure on V implies that dimR V is
even, say
dimR V = 2n.
Remark 3.2. Suppose that V carries an almost complex structure J. Then one can find
vectors xi ∈ V such that the vectors
xi , yi =def J(xi )
form a basis for V (Exercise 3.1). With respect to this basis J is given by the block matrix
0 −1
1 0
0 −1
1 0
...
with the obvious expression for scalar multiplication by a complex number. Note that there
is an operation of complex conjugation on VC , given by the rule
√ √
v + −1 · w = v − −1 · w.
The real subspace V ⊆ VC consists of those vectors invariant under complex conjugation.
Note that JC : VC −→ VC satisfies JC2 = Id, and hence defines a complex structure on
JC . However this is not the complex structure defined √ via (3.1), which is the one we always
use. In fact, by definition JC acts
√ by multiplication by −1 on the subspace V 1,0 ⊆ VC , but
it acts by multiplication by − −1 on V 0,1 ⊆ VC . In fact:
Proposition 3.5. The composition
π 1,0
V ⊆ VC −→ V 1,0
gives rise to a C-linear isomorphism
√
(V, J) ∼
= (V 1,0 , · −1)
44 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
are C-bases of the two spaces in question, and in particular {zi , z i } is a complex basis for
VC .14
Note that
(V ∗ )1,0 ∼
= HomC−lin (V, J), C) , (V ∗ )0,1 ∼
= HomC−antilin (V, J), C).
If we choose a basis {xi , yi = J(yi )} for V as above, and denot by xi , y i the corresponding
dual basis of V ∗ , then
√ √
z i = xi + −1y i , z i = xi − −1y i (3.4)
are bases for (V ∗ )1,0 and (V ∗ )0,1 respectively, dual to the bases {zi } and {z i } appearing in
(3.2).
14In
effect we have seen these expressions before. If we think of V as the (real) tangent space to a complex
manifold, then (3.2) corresponds to the definitions
∂ 1 ∂ √ ∂ ∂ 1 ∂ √ ∂
= − −1 , = + −1 .
∂zi 2 ∂xi ∂yi ∂z i 2 ∂xi ∂yi
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 45
Next we turn to exterior products. Goven as before a real vector space V with dimR V =
2n, we consider the real and complex vector spaces
Λk V , Λk VC ,
of real and complex dimensions 2n
k
respectively. Now suppose that J : V −→ V is an
almost complex structure, giving rise to the decomposition
VC = V 1.0 ⊕ V 0,1 .
Then there is a canonical isomorphism:
Λk VC = Λk V 1,0 ⊕ V 0,1
M
Λp (V 1,0 ) ⊗ Λq (V 0,1 )
=
p+q=k
Explicitly, if
α1 , . . . , αp ∈ V 1,0 and β1 , . . . , βq ∈ V 0,1 ,
then
α1 ∧ . . . ∧ αp ∧ β1 ∧ . . . ∧ βq ∈ V p,q ,
and an arbitrary element of V p,q is a C-linear combination of terms of this form. As for the
decomposition (3.6), starting with µ1 , . . . , µk ∈ VC , start by writing
µi = µ1,0
i + µ0,1
i
as a sum of vectors of types (1, 0) and (0, 1). Then expand out the wedge product
µ1 ∧ . . . ∧ µk = (µ11,0 + µ0,1 1,0 0,1
1 ) ∧ . . . ∧ (µk + µk )
via linearity, and for each (p, q) with p + q = k, collect together those terms that are wedges
of p vectors of type (1, 0) and q vectors of type (0, 1). Note that
V p,q = V q,p .
In terms of the basis {z i , z i } appearing in (3.4), (V ∗ )pq has as basis the elements
z i1 ∧ . . . ∧ z ip ∧ z j1 ∧ . . . ∧ z jq
with i1 < . . . < ip , j1 < . . . < jq . Finally, consider the endomorphism
Λk JC : Λk VC −→ Λk VC
determined by J. Then the reader should verify that:
√
Λk JC acts on Λp,q V by multiplication by ( −1)p−q . (3.7)
46 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Metrics and the fundamental form. We now introduce metrics. Focusing still on a
finite-dimensional real vector space V , assume now that we are given a positive definite
inner product < , > on V .
Definition 3.6. An almost complex structure J on V is compatible with < , > if
< Jv, Jw > = < v, w > for all v, w ∈ V.
Note that a compatible complex structure J on (V, < , >) determines an orientation on
V : choosing vectors x1 , . . . , xn ∈ V such that
{x1 , Jx1 , . . . , xn , Jxn } (*)
is a basis of V , we require that the ordered basis (*) determine the positive orientation of V .
Example 3.7. It is instructive to consider the case dimR (V ) = 2. If we fix an orientation
on V , then a Euclidean structure < , > determines a unique compatible almost complex
structure. In fact, choose any non-zero vector v ∈ V , and define J by requiring that
< v, Jv >= 0 , ||v|| = ||Jv|| = 1 , J 2 (v) = −v
and that {v, Jv} be a positively oriented basis of V . Conversely, if J is an almost complex
structure on V , then up to scalar multiples there is a unique Euclidean inner product < , >
on V with respect to which J is compatible and {v, Jv} gives a positive orientation. See
also Exercise 3.3.
Fix a Euclidean space (V, < , >, J) with a compatible almost complex structure. These
data are combined in the crucial
Definition 3.8. The fundamental form associated to < , > and J is the two-form ω ∈ Λ2 V ∗
defined by
ω(v, w) = < Jv, w > = − < v, Jw > .
Proposition 3.9. The form ω is in fact alternating, and under the inclusion
Λ2 V ∗ ⊆ Λ2 VC∗ ,
ω has type (1, 1), ie:
ω ∈ Λ2 V ∗ ∩ Λ1,1 V ∗ .
We now discuss Hermitian forms arising from the data (V, J, < , >) of an almost complex
structure compatible with a Euclidean inner product. Define
√
h(v, w) = < v, w > − −1 · ω(v, w)
√ (3.8)
= < v, w > − −1· < Jv, w > .
This is a Hermitian form on V with respect to the complex structure given by J.15 For
another viewpoint, denote by < , >C−lin the C-linear extension of < , > to VC . Then we can
extend < , > to a Hermitian inner product < , >herm on VC by setting
< µ, ν >herm = < µ, ν >C−lin .
√
Then under the isomorphism (V, J) ∼
= (V 1,0 , · −1) given by Proposition 3.5, one has
1
2
· h( , ) = < , >herm |V 1,0 . (3.9)
15We take Hermitian forms to be C-linear in the first argument, and conjugate linear in the second.
48 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
The Hodge ∗-operator. Before proceeding, we pause to recall some facts about the Hodge
∗-operator. Let W be a real vector space of dimension d equipped with a positive definite
inner product g =< , >. Giving < , > is the same thing as giving a symmetric linear mapping
T : W −→ W ∗ , T (v)(w) = < v, w >
with the property that T (v)(v) > 0 for all 0 6= v ∈ W . On the other hand, T determines by
fuctoriality a symmetric mapping
Λk T : Λk W −→ Λk W ∗
satisfying the analogous positivity property. In other words, an inner product < , > on W
induces one on each exterior product Λk W , which we continue to denote by < , >. Concretely,
if
e1 , . . . , ed ∈ W (*)
is an orthonormal basis for W , then < , > is defined on Λk W by taking as an orthonormal
basis the multi-vectors
eI = ei1 ∧ . . . ∧ eik for I = {i1 < . . . < ik }.
We also fix the volume form
vol = e1 ∧ . . . ∧ ed ∈ Λd W,
arising from the orthonormal basis (*).
The basic properties of this operator are summarized in the following Proposition, whose
proof we leave to the reader.
Proposition 3.10. Fix as above an orthonormal basis e1 , . . . , ed ∈ W .
(i). If I = {i1 < . . . < ik } and J = {j1 < . . . < jd−k } are disjoint multi-indices such that
I ∪ J = [1, d], then
∗ eI = σI,J · eJ ,
where eI = ei1 ∧ . . . ∧ eik , with eJ defined similarly, and σI,J is the sign of the
permutation (I, J).
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 49
(iii). If α ∈ Λk W , then
∗ ∗ α = (−1)k(d−k) · α.
Now let (V, < , >, J) be as above a real Euclidean space of dimension 2n with a compat-
ible almost complex structure. We apply the previous discussion to W = V ∗ , which inherits
an inner product < , > and orientation from V and J. In particular, one then gets an inner
product < , > on Λ∗ V ∗ , and the Hodge ∗-operator
∗ : Λ∗ V ∗ −→ Λ∗ V ∗ (3.16)
is defined. Moreover, the inner product < , > on V ∗ extends to a positive definite Hermitian
form < , >herm on VC∗ , which then determines a positive definite Hermitian inner product
< , >herm on Λ∗ VC∗ . Thanks to the fact (Exercise 3.5) that the subspaces
V ∗1,0 , V ∗0,1 ⊆ VC∗
are orthogonal with respect to < , >herm , one finds that
Λk V ∗ = ⊕p+q=k Λp,q VC∗
is an orthogonal decomposition with respect to < , >herm . Finally, the Hodge ∗-operator
(3.16) extends C-linearly to
∗ : Λk VC∗ −→ Λ2n−k VC∗ ,
and satisfies
α ∧ ∗β = < α, β >herm ·vol. (3.17)
p0 ,q 0
Observe that if β ∈ Λp,q V ∗ and α ∈ Λ V ∗ , then
< α, β >herm = 0 unless p0 = q , q 0 = p.
It follows that
∗ Λp,q VC∗ ⊆ Λn−q,n−p VC∗ .
Lefschetz operator and decomposition. Consider once again a triple (V, < , >, J) con-
sisting of a real Euclidean space of dimension 2n and a compatible almost complex structure.
Denote by
ω ∈ Λ2 V ∗
the corresponding fundamental form.
Definition 3.13. The Lefschetz operator L associated to ω is the operator given by wedge
product with ω:
L : Λk V ∗ −→ Λk+2 V ∗ , α 7→ ω ∧ α.
The dual or adjoint Lefschetz operator is the operator
Λ : Λk+2 V ∗ −→ Λk V ∗
determined by requiring that
< Λα, β > = < α, Lβ > for all β ∈ Λk V ∗ .
We sometimes view L and its adjoint Λ as operators on the full exterior algebra of V :
L : Λ∗ V ∗ −→ Λ∗ V ∗ , Λ : Λ∗ V ∗ −→ Λ∗ V ∗ .
Note that it follows from Proposition 3.10 that the Hodge ∗-operator is invertible. We
can then use it to express Λ in terms of L:
Proposition 3.14. One has
Λ = ∗−1 ◦ L ◦ ∗.
Proof. Since
< Λα, β > = < α, Lβ > = < Lβ, α >,
the issue is to show that if α ∈ Λk+2 V ∗ and β ∈ Λk V ∗ , then
< Lβ, α > ·vol = < β, ∗−1 L ∗ α > ·vol.
But
< β, ∗−1 L ∗ α > ·vol = β ∧ L(∗α)
= β ∧ ω ∧ ∗α
= ω ∧ β ∧ ∗α
= < Lβ, α > ·vol,
as required.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 51
We now come to the Lefschetz decomposition. The first point is to define primitive
forms:
Definition 3.15. (Primitive vectors). A k-vector α ∈ Λk V ∗ is primitive if
Λ(α) = 0.
We denote by P k ⊆ Λk V ∗ the subspace of primitive elements.
We will see shortly that α ∈ Λk V ∗ is primitive if and only if k ≤ n and Ln−k+1 (α) = 0. One
uses the same definition to define the subspace of complex primitive forms PCk ⊆ Λk VC∗ .
The statement in (i) is the Lefschetz decomposition. The assertion in (ii) will later on lead
to the Hard Lefschetx theorem. The next few pages of this subsection will be devoted to the
proof of the Theorem.
52 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
We next review the structure of representations of sl2 : this will lead to the proof of
Theorem 3.16. Recall that sl2 is the Lie algebra of all 2 × 2 matrices of trace 0; we will work
with real matrices and representations, but nothing changes if one complexifies. Put
0 1 1 0 0 0
X = , B = , Y = .
0 0 0 −1 1 0
These satisfy the commutation relations
[B, X] = 2X , [B, Y ] = −2Y , [X, Y ] = B.
Lemma 3.21. Let W be an irreducible finite-dimensional representation of sl2 . Then there
exists a non-negative integer m such that W is the direct sum
W = W−m ⊕ W−m+2 ⊕ . . . ⊕ Wm−2 ⊕ Wm ,
where Wj ⊆ W is a one-dimensional eigenspace for B, with eigenvalue j. Moreover, Y maps
Wj onto Wj−2 and X maps Wj onto Wj+2 . In particular, any vector u ∈ W can be written
uniquely in the form
u = u0 + Xu1 + X 2 u2 + . . . + X m um
where Y uj = 0 for each j.
54 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
So it suffices to show that the B-eigenvalue of w is −m. To this end, suppose that Bw = λw.
Then
Y Xw = XY w − Bw = −λw,
and
Y X 2 w = XY Xw − BXw
= −λXw − (λ + 2)Xw.
In general
Y X kw = − λ − (λ + 2) − . . . − λ + (2(k − 1) (X k−1 w)
= (−kλ − k 2 + k)X k−1 w.
Since X m+1 w = 0 but X m w 6= 0, this implies that
−(m + 1)λ − (m + 1)2 + (m + 1) = 0,
ie λ = −m.
2
Remark 3.22. Let C denote the standard two-dimensional representation of sl2 . Then the
representation W considered in the previous lemma arises as Symm C2 .
Corollary 3.23. Let U be a finite dimensional representation of sl2 , and let
P = {u ∈ U | Y u = 0} ⊆ U
be the subspace of primitive vectors. Assume that B acts on P by multiplication by −m.
Then m ≥ 0, and
U = P ⊕ X(P ) ⊕ X 2 (P ) . . . ⊕ X m (P ).
Moreover X m+1 (P ) = 0, while
X m (u) 6= 0 ∀ 0 6= u ∈ P.
Proof. In fact, U is a direct sum of irreducible representations, and then the corollary follows
from the previous lemma and its proof.
16It will emerge shortly that B actually has real eigenvalues.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 55
Proof of Theorem 3.16. Proposition 3.20 asserts that Λ∗ V ∗ is a representation of sl2 , with
L ↔ X , H ↔ B , Λ ↔ Y.
Furthermore, by definition H acts as multiplication by k−n on P k V ∗ ⊆ Λk V ∗ . The existence
of the Lefschetz decomposition then follows from the previous corollary as does statement
(iii) of the Theorem. Moreover the corollary shows that the mapping
Ln−k : P k V ∗ −→ Λn−k V ∗
is injective. It then follows from (i) that
Ln−k : Λk V ∗ −→ Λ2n−k V ∗
is injective, and since both sides have the same dimension Ln−k must be an isomorphism.
Note that the operators Ln−k−j and ∗Lj both map Λk V ∗ to Λ2n−k−j V ∗ , so it is natural
to ask for a comparison of these. To this end, write
J = Λ∗ J ∗ : Λ∗ V ∗ −→ Λ∗ V ∗
for the endomorphism of Λ∗ V ∗ determined by the almost complex structure J.
Proposition 3.24. Let α ∈ P k be a primitive k-form. Then
k(k+1) j!
∗Lj (α) = (−1) 2 · · Ln−k−j J(α).
(n − k − j)!
The operators L, Λ, H pass by linearity to the complexification Λ∗ VC∗ , and all the results
of this subsection remain valid in this setting. Moreover since these operators act bihomo-
geneously with respect to the (p, q) grading on Λ∗ VC∗ , the Lefschetz decomposition respects
this bigrading. Specifically, let
P p,q = PCk ∩ Λp,q VC∗ .
Then PCk = ⊕p+q=k P p,q , and
M
Λp,q VC∗ = Lj P p−i,q−i .
i≥0
Example 3.26. Taking k = 2, note that Λ2,0 V ∗ and Λ2,0 V ∗ are primitive for reasons of
type. The Lefschetz decomposition in this case takes the form
Λ2 VC∗ = P 2,0 ⊕ P 1,1 ⊕ C · ω ⊕ P 0,2 .
We conclude by studying the sign of an important quadractic form defined with the help
of ω. Define
Q : Λk V ∗ × Λk V ∗ −→ R
by requiring that
k(k+1)
(−1) 2 α ∧ β ∧ ω n−k = Q(α, β) · vol.
We extend this by linearity to a symmetric bilinear form
Q : Λk VC∗ × Λk VC∗ −→ C,
called the Hodge-Riemann pairing. The basic fact is:
0 0
Theorem 3.27. (Hodge-Riemann bilinear relations). If α ∈ Λp,q V ∗ and β ∈ Λp ,q V ∗ ,
then
Q(α, β) = 0 unless p = q 0 , q = p0 .
If p + q ≤ n and α ∈ P p,q is a primitive (p, q)-form, then
√
( −1)(p−q) · Q(α, α) = n − (p + q) !· < α, α >herm .
(*)
√
In particular ( −1)(p−q) Q(α, α) > 0 if α 6= 0.
Proof. Only (*) needs proof, for which we apply Proposition 3.24 with j = 0 and k = p + q
to the primitive form (α) ∈ P q,p . One finds that
k(k+1) 1 √
∗α = (−1) 2 · · ( −1)q−p · α ∧ ω n−k ,
(n − k)!
so that
k(k+1) √
α ∧ ω n−k = (−1) 2 · (n − k)! · ( −1)p−q · ∗α.
Recalling that
< α, α >herm ·vol = α ∧ ∗α,
(*) follows.
Example 3.28. Take k = 2, and consider the form
Q : Λ1,1 V ∗ × Λ1,1 V ∗ −→ R , (α, β) 7→ α ∧ β ∧ ω n−2
(where we identity Λn V ∗ with R via vol). Recall that
Λ1,1 V ∗ = P 1,1 ⊕ (R · ω).
Theorem 3.27 shows that Q is negative definite on P 1,1 , but positive-definite on the one-
dimensional subspace spanned by ω. So on Λ1,1 V ∗ Q has one positive eigenvalue, and the
rest negative. This will lead to the Hodge Index theorem in the global setting.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 57
We now apply the linear algebra of the previous section to the tangent and cotangent
spaces of a complex manifold. We start with the local picture, where we can make explicit
calculations.
58 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
As in the pointwise situation of the previous section, this leads to a C-linear direct sum
M
Λk TC∗ U = Λp,q T ∗ U,
p+q=k
p,q ∗
with a basis for Λ T being given by forms of the sort
dzI ∧ dz J , |I| = p , |J| = q.
This in turn leads to a decomposition of the spaces of smooth differential forms on U . Write
AkC (U ) for the space of smooth complex-valued differential k-forms on U , i.e. C ∞ global
sections of Λk TC∗ U . Then we have
M
AkC (U ) = Ap,q (U ), (4.1)
p+q=k
the fI,J being smooth functions on U . We say that such a form is of type (p, q), and we
denote by
π p,q : AkC (U ) −→ Ap,q (U )
the projections.
So far all of these constructions are completely algebraic in nature, ie they depend only
on the point-wise situation studied in the previous section. Calculus enters the picture with
the exterior derivative. Given a smooth C-valued function f ∈ A0C (U ) on U , we define
∂f ∈ A1,0 (U ) , ∂f ∈ A0,1 (U )
to be respectively the (1, 0) and (0, 1) components of df ∈ A1 (U ), i.e. ∂f = π 1,0 (df ) and
∂f = π 0,1 (df ). Thus df = ∂f + ∂f , and in terms of local coordinates:
n
X ∂f
∂f = · dzi
i=1
∂z i
n
X ∂f
∂f = · dz i .
i=1
∂z i
Therefore the derivative of an arbitrary (p, q)-form is the sum of a (p + 1, q)-form and a
(p, q + 1)-form. It follows that the full exterior derivative
d : AkC (U ) −→ Ak+1
C (U )
(i). d = ∂ + ∂.
2
(ii). ∂ 2 = ∂ = 0 and ∂∂ + ∂∂ = 0.
(iii). These operators satisfy the Leibnitz rule: if α ∈ Ap,q (U ), then
∂(α ∧ β) = ∂(α) ∧ β + (−1)p+q α ∧ ∂(β)
∂(α ∧ β) = ∂(α) ∧ β + (−1)p+q α ∧ ∂(β).
∂-Poincaré Lemma. In this subsection we will establish the analogue for ∂ of the classical
Poincaré lemma that a closed form is locally exact. Specifically, we aim to prove (most of)
the following:
Theorem 4.2. Let ∆ ⊆ Cn be a polydisk, and let
α ∈ Ap,q (∆)
be a ∂-closed smooth (p, q)-form on δ. If q ≥ 1, then there exists a smooth (p, q − 1)-form β
on ∆ such that
α = ∂β.
It turns out that most of the work occurs in the case n = 1. We start by recording a
C ∞ version of the Cauchy integral formula.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 61
The following proposition essentially proves the ∂-Poincaré lemma in one variable for
the form f (z)dz (which is automatically ∂-closed).
Proposition 4.4. Let ∆ ⊂ C be a disk, and let f ∈ C ∞ (∆) be a smooth function on an
open neighborhood of ∆. Then the function
Z
1 f (w)
h(z) = √ dw ∧ dw
2π −1 ∆ w − z
is defined and C ∞ in ∆, and satisfies
∂h
= f.
∂z
We now turn to the proof of (a slightly weakened form of) Theorem 4.2. To begin with,
we reduce to proving the theorem in the case p = 0. In fact, suppose that
X
α = fI,J dzI ∧ dz J
I,J
P
(#I = p, #J = q) is a form of type (p, q). Then we can write α = I dzI ∧ αI , where
X
αI = fI,J dz J
J
is a (0, q)-form. Then ∂α = 0 if and only if ∂αI = 0 for every I, and if each αI is ∂-exact
then so it α. So it suffices to prove the theorem for (0, q)-forms.
Now let ∆ = ∆(ε) ⊆ Cn be a bounded polydisk, and that for q > 0 we are given a
∂-closed (0, q)-form α ∈ A0,q (U ) defined and smooth on an open neighborhood U ⊇ ∆. We
will show that
α = ∂β for some β ∈ A0,q−1 (∆).18
P
Write then α = fI dz I , and let 1 ≤ k ≤ n be the least index such that no dz i appears
non-trivially in this expression for i > k. Then we can write
α = α1 ∧ dz k + α2 ,
where α2 is a (0, q)-form involving only the variables z1 , . . . , zk−1 . Setting
∂
∂i = · dz i ,
∂z i
the assumption ∂α = 0 implies that
∂i α1 = ∂i α2 = 0
for i > k, and hence the coefficients fI of α are holomorphic in the variables zk+1 , . . . , zn .
the integral being taken over a suitable disk B ⊂ C. By Proposition 4.4, one has
∂hI
= fI
∂z k
on B. Moreover hI is holomorphic in zk+1 , . . . , zn , and smooth in the other variables. This
being said, put X
γ = hI dz I−{k} .
I3k
Local geometry of Kähler forms. We now bring metrics into the picture.
Consider a Riemannian metric g =< , > on U , i.e a positive definite bilinear form gx =<
, >x on real tangent space Tx U varying smoothly with x. We assume that g is compatible
with the (almost) complex structure on U , i.e. we assume that
< Jv, Jw > = < v, w >
for all local vector fields v, w on U . Then by our discussion of Hermitian linear algebra, g
determines a real (1, 1)-form ω in U given by
ω(v, w) = < Jv , w >,
and √
h = g − −1 · ω
is a positive-definite Hermitian metric on U . As in the pointwise setting, one calls ω the
fundamental form associated to the metric. If we define
∂ ∂
hij (z) = h , ,
∂xi ∂xj z
then hij is a positive-definite matrix of smooth functions, and
√
−1 X
ω = · hij (z) dzi ∧ dz j .
2 i,j
Since conversely ω determines g and h, one sometimes abusively refers to the fundamental
form ω as the compatible metric.
Such metrics will play a central role in what follows. For now we want to prove an important
local characterization of Kähler metrics.
We say that the Hermitian metric h on U (or the underlying Riemannian metric g) os-
culates to the identity to order 2 at a point p ∈ U if in suitable local holomorphic coordinates
centered at p one write
h = Id + O(|z|2 ), (4.2)
where the term on the right indicates a matrix of functions vanishing to order ≥ 2 at 0. This
is a useful condition, because it allows one to check identities among first order differential
operators by working with the standard metric.
The basic result is that the Kähler condition is equivalent to the being able to choose
coordinates so that g osculates to the identity.
Theorem 4.6. Let g be a compatible metric on U 3 p, with fundamental form ω. Then
dω = 0 if and only if one can choose coordinates centered at p with respect to which g
osculates the identity to order 2.
Sketch of proof. Assuming that dω = 0 we will show that after a coordinate change we can
arrange for (4.2) to hold. To this end, write
√
−1 X
ω = · hij (z)dzi ∧ dz j ,
2 i,j
where
X X
hij (z) = δij + aijk zk + a0ijk z k + O(|z|2 ),
k k
Then a computation using the relations in (*) shows that in these new coordinates,
√
−1 X
ω = · dwj ∧ dwj + O(|w|2 ),
2
as required. (See [3, p. 49] for details.)
Almost complex structures. Let X be a real manifold of even (real) dimension 2n. An
almost complex structure on X is a bundle endomorphism
J : T X −→ T X with J 2 = Id.
Thus J gives an almost complex structure Jx on each of the (real) tangent spaces Tx X. As
in §3, an almost complex structure gives rise to a decomposition into type
T ∗ X = T ∗1,0 ⊕ T ∗0,1 X,
and similarly for k-forms. Moreover we can define operators
∂ : Ap,q (X) −→ Ap+1,q (X) , ∂ : Ap,q (X) −→ Ap,q (X)
by composing the deRham d with projection onto the spaces of (p + 1, q) and (q, p + 1)-forms
respectively. Note that on functions d acts as ∂ + ∂, but there is no reason that this has to
be true on higher forms, since it could happen eg that there is a (1, 0) form α such that dα
has a non-trivial (2, 0) component.
Evidently every complex manifold has a canonical such almost complex structure, so
it is natural to for conditions under which an almost complex structure is integrable, i.e
arises from a complex structure. This is the content of a famous theorem of Newlander and
Nierenberg.
Theorem 4.7. (Newlander–Nierenberg). Let (X, J) be an almost complex manifold.
Then X carries a complex structure giving rise to J if and only if any of the following
equivalent conditions are satisfied:
The proof that these statements are equivalent is elemantary (see [3, p. 107, 108]). The
serious assertion is that these imply the existence of a complex structure on X giving rise to
J.
66 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
In this section, we give a brief sketch of the theory of sheaves and their cohomology, and
prove Dolbeaut’s theorem.
Statement of Dolbeaut’s theorem. Let X be a complex manifold. Writing Ap,q (X) for
the space of global smooth (p, q)-forms on X, we have the ∂-operator
∂ : Ap,q (X) −→ Ap,q+1 (X)
2
satisfying ∂ = 0. Therefore we can make the
Definition 5.1. The Dolbeaut cohomology groups of X are defined as the vector space of
∂-closed (p, q)-forms modulo ∂-exact (p, q)-forms:
p,q ker ∂ : Ap,q (X) −→ Ap,q+1 (X)
H∂ (X) = .
im ∂ : Ap,q−1 (X) −→ Ap,q (X)
One can see these groups as holomorphic analogues of the DeRham cohomology groups of
a smooth manifold. So it is natural to ask what invariants of X they compute. Dolbeaut’s
theorem asserts that they are isomorphic to certain basic sheaf-theoretic invariants of X:
Theorem 5.2. (Dolbeaut’s theorem, I). There are canonical isomorphisms:
H∂p,q (X) = H q (X, ΩpX ).
In particular,
H∂0,q (X) = H q (X, OX ).
The groups on the right are the cohomology of X with coefficients in the sheaf of holomorphic
p-forms and the structure sheaf of X, respectively. Our goal in the rest of the section is to
explain the meaning of these groups, and to sketch the proof of the theorem. (Along the
way we’ll give the sheaf-theoretic proof of De Rham’s theorem, which follows formally from
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 67
the d-Poincaré lemma in the same way that the Dolbeaut theorem will follow from the ∂-
Poincaré lemma.) However before turning to sheaf theory, we want to indicate a “twisted”
analogue of Dolbeaut cohomolgy.
It is a wonderful fact that given any holomorphic vector bundle E, the ∂ operator on
forms extends canonically to an operator
∂ = ∂ E : Ap,q (X, E) −→ Ap,q+1 (X, E). (5.1)
The critical case to understand is that when p = 0:
∂ E : A0 (E) −→ A0,1 (E). (*)
In other words, we need to explain how to take the ∂-derivative of a smooth section of E.
On a trivializing open set Uj , we represent s ∈ A0 (E) by a vector sj of smooth sections,
and these data patch together via the rule si = gij sj on Uij . We locally define ∂s to be the
vector of (0, 1)-forms obtained by differentiating each of the components of sj , i.e.
∂ E (s) =locally ∂(sj ) ∈ A0,1 (Uj , E).
We need to check that the vectors of (0, 1)-forms so obtained transform in the required
manner. But ∂gij = 0 thanks the fact that the gij are matrices of holomorphic functions,
and hence differentiating the relation si = gij sj one finds that
∂si = ∂gij · sj + gij · ∂sj
= gij · ∂sj ,
as required. The reader can then check
Lemma 5.3. The operator ∂ E : A0 (E) −→ A0,1 (E) defined in (*) extends uniquely to an
operator
∂ = ∂ E : Ap,q (X, E) −→ Ap,q+1 (X, E)
satisfying the Leibnitz rule
∂ E (f · s) = (∂f ∧ s) + f · ∂ E (s),
2
and ∂ E = 0.
68 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Sheaves. One can think of the theory of sheaves as a tool for managing the questions
involved in trying to pass from local to global constructions.
More succinctly, one can say that a presheaf is a contravariant functor from the category
of open subsets of X to the category of abelian groups. One defines presheaves of rings, or
vector spaces, or ..., similarly. One typically calls the elements of F (U ) sections of F over
U.
Example 5.6. (Holomorphic functions and sections). Let X be a complex manifold.
The presheaf OX of holomorphic functions on X associates to an open set U the holomorphic
functions on U :
OX (U ) = {holomorphic functions on U },
with ρU,V being given by restriction. Similarly, given a holomorphic vector bundle E on X,
we define OX (E) by the rule
OX (E)(U ) = Γ(U, E) = {holomorphic sections of E over U },
again with the ρ being ordinary restrictions. Thus OX is a presheaf of rings (and OX (E) is
a sheaf of modules over OX in a sense that we will define shortly).
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 69
Example 5.7. (Ideal of analytic subvariety). Let X be a complex manifold, and let
V ⊆ X be an analytic subvariety. Then one defines the ideal presheaf IV of V by taking
IV (U ) = {f ∈ OX (U ) | f |V ≡ 0}.
This is a sub-presheaf of OX in the sense that IV (U ) ⊆ OX (U ) for all U ⊆ X.
Example 5.8. (Smooth forms). Let M be a smooth real manifold. Then the presheaf
AkM of smooth k-forms on X is given by
AkM (U ) = Ak (U ),
with the evident restrictions. So for exampe A0M is the presheaf of smooth functions on M .
If X is a complex manifold, one defines presheaves Ap,q
X in the analogous fashion.
Example 5.9. (Constant presheaf ). Let X be a topological space. The constant presheaf
Zconst on X is the presheaf that assigns to U ⊆ X all constant functions U −→ Z, ie
Zconst (U ) = Z,
with all non-trivial restrictions being the identity.
In all the examples so far, the restriction homomorphisms ρ have been actual restrictions
of functions, sections or forms. Although this need not be true in general, it is customary to
lighten the notation by writing
ρU,V (s) = s|V
for s ∈ F (U ) and V ⊆ U .
In other words, the condition asks that the si should patch together to give a section on all
of U .
Example 5.11. The presheaves in Examples 5.6, 5.7 and 5.8 are sheaves.
Example 5.12. The presheaf Zconst from Example 5.9 is not in general a sheaf. For example,
take X = R, and consider the disconnected subset U = R − {0} with the open cover
U = (−∞, 0) ∪ (0, ∞).
70 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Example 5.16. Let X be a complex manifold, and fix an index p ≥ 0. Then the ∂-operator
gives rise to a sheaf homomorphism
∂ : Ap,q p,q+1
X −→ AX
of sheaves of vector spaces.19 In fact, these fit together into a complex
. . . −→ Ap,q−1 −→ Ap,q
X −→ A
p,q+1
−→ . . . (5.3)
of sheaves of vector spaces. The proof of the Dolbeaut theorem will revolve around the
analysis of this complex.
There are many constructions that lead naturally to a presheaf F , and it is then impor-
tant to know that there is a canonical way to pass to a sheaf that is “closest” to F . This is
the sheafification of F , or the sheaf associated to F .
Proposition 5.19. Let F be a presheaf on a topological space X. Then there is a sheaf F + ,
together with a morphism
θ : F −→ F +
characterized by the property that any homomorphism φ : F −→ G from F to a sheaf G
factors uniquely through θ. Moreover, the canonical homomorphism
θx : Fx −→ Fx+
is an isomorphism for all x ∈ X.
19Note that since ∂ is not linear over the ring of smooth functions, 5.3 is not a homomorphism of A0X -
modules.
72 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
We refer eg to [] for the proof. However we can quicky explain the construction of F + .
Start by forming the disjoint union
a
U = Fx
x∈X
Note in particular that the surjectivity of φ does not imply the surjectivity of φU :
F (U ) −→ G(U ) for every open set U ⊆ X. For example, let X = C, and consider the
exponential map
∗
exp : OX −→ OX .
As noted in the previous example, this is surjective as a morphism of sheaves. However if
∗
U = C − {0}, then the function z ∈ OX (U ) is not in the image of exp (one cannot define a
single-valued branch of log(z) on C − {0}).
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 73
Having discussed quotients and kernels, one can then define exact sequences of sheaves:
0 −→ G −→ F −→ Q −→ 0,
which implies that G is a sub sheaf of F , and that Q = F/G. In general, a sequence of
sheaves on X is exact if and only if the corresponding sequences of stalks are exact at every
point.
are exact, where for each multi-index I = {i0 < . . . < ip }, UI denotes the intersection of the
corresponding elements of the covering.
We claim now that s lifts to a section t ∈ Γ(X, E) if and only if we can find ri ∈ F (Ui ) such
that
rij = ri − rj | Uij .. (*)
In fact, supposing that (*) holds, define
t0i = ti − ri .
Then (t0i − t0j ) = 0 on Uij , so the t0i patch to give a global section t ∈ Γ(X, E) that maps to s.
We now turn to the formal definitions. Let X be a topological space, let A = {Ui } be
an open covering of X, and let F be a sheaf on X.
Definition 5.27. The group of Cech p-cochains is
Y
C p (A, F ) = F (Ui0 i1 ...ip ).
i0 ,i1 ,...,ip
Thus an element γ ∈ C p (A, F ) consists in giving for each p + 1-fold intersection of the Ui a
section
γ(i0 , . . . , ip ) ∈ F (Ui0 i1 ...ip
of F over the corresponding open set.
Note that we do not require that the Ui appearing in the definition be distinct.
In the usual way, one checks that δ 2 = 0, and then we can take cohomology:
Definition 5.28. The Cech cohomology of F with respect to the covering A the cohomology
of the Cech complex just constructed:
Ȟ p (A, F ) = H p C ∗ (A, F ) .
Of course one would like to remove the dependence on the covering. Although we will
gloss over this point, for this one passes to refinements. In brief, suppose that A0 ≺ A is a
refinement of A. One can then construct chain maps
C p (A, F ) −→ C p (A0 , F )
which are uniquely defined up to homotopy. This gives rise to canonically defined homomor-
phisms
Ȟ p (A, F ) −→ Ȟ p (A0 , F ),
and one then passes to a direct limit:
Definition 5.31. The Cech cohomology of X with coefficients in a sheaf F is detined to the
the direct limit
Ȟ p (X, F ) = lim Ȟ p (A, F ),
−→
where the limit is taken over all open covers.
It turns out that for most of the sheaves F that arise in day to day complex or algebraic
geometry – specifically “coherent analytic” or “coherent algebraic” sheaves – there is a single
cover A with the property that all the cohomology groups Ȟ p (X, F ) can be computed as
H p (A, F ). So in practice one usually doesn’t need to pass to the limit.
∗
Example 5.32. Given a complex manifold X, denote by OX the sheaf of nowhere-vanishing
holomorphic functions on X: this is a sheaf of abelian groups under multiplication. Then
n o
∗
Ȟ 1 (X, OX ) = isomorphism classes of holomorphic line bundles on X .
∗
In fact, Ȟ 1 (A, OX ) classifies isomorphism classes of line bundles that trivialize on the open
∗
cover A: the corresponding transition functions gij ∈ OX (Uij ) give the cocycle determined
by a line bundle.
The most important property of the theory is that a short exact sequence
0 −→ F −→ E −→ G −→ 0 (*)
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 77
of sheaves on X should give rise to a long exact sequence on cohomology extending (5.5).
This is not universally true for Cech cohomology, but it does work on paracompact Hausdorff
spaces.
Theorem 5.33. Let X be a paracompact Hausdorff space, and suppose given a short exact
sequence (*) of sheaves on X. Then (*) induces a long exact sequence
0 −→ Ȟ 0 (X, F ) −→ Ȟ 0 (X, E) −→ Ȟ 0 (X, G) −→ Ȟ 1 (X, F ) −→ Ȟ 1 (X, E) −→ Ȟ 1 (X, E)Ȟ 2 (X, F ) −→ . . .
of cohomology groups.
We will not try to prove this in general, but the result is easy to derive assuming that
we are in the situation of the simplfying ssumption 5.26. In fact, it follows from (5.4) and
the definitions that the short exact exact sequence (*) gives rise to a short exact sequence
of complexes
0 −→ C • (A, F ) −→ C • (A, E) −→ C • (A, G) −→ 0.
This gives rise to a long exact sequence of the cohomology groups of A, and since we assume
that 5.26 holds for arbitrarily fine covers, we get the conclusion of 5.33 by passing to a limit.
Example 5.34. Consider the exponential sequence
∗
0 −→ ZX −→ OX −→ OX −→ 0
on a compact connected complex manifold X. Since
∗
H 0 (X, OX ) = C , H 0 (X, OX ) = C∗
(the only holomorphic functions on X are constant), the corresponding long exact sequence
on cohomology effectivelly starts with H 1 and takes the form:
∗
0 −→ H 1 (X, Z) −→ H 1 (X, OX ) −→ H 1 (X, OX ) −→ H 2 (X, Z) −→ H 2 (X, OX ) −→ . . . .
∗
Recalling that H 1 (X, OX ) classifies isomorphism classes of holomorphic line bundles on X,
this will later give us a rather precise description of all such. In particular this sequence will
lead to the proof of the Lefschetz (1, 1)-theorem.
Proof of the Dolbeaut theorem. We’re now ready to indicate the proof of the Dolbeaut
theorem.
We denote by Ap,q = Ap,qX the sheaf of smooth (p, q) forms on X. Thus the space A (X)
p,q
of global (p, q)-forms on X is just the group of global sections of this sheaf. Fixing an integer
p ≥ 0, we get as in Example 5.16 a complex of sheaves Ap,• :
0 −→ Ap,0 −→ Ap,1 −→ . . . −→ Ap,q −→ . . . (5.6)
6. Kähler Manifolds
In this section, much of the preceeding material comes together in the Kähler package,
which shows that compact complex manifolds that carry a Kähler metric have a number of
truly miraculous prorperties.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 79
The Kähler condition. Let X be a complex manifold of (complex) dimension n, and let
J be the corresponding almost complex structure. We suppose that X carries a Riemannian
metric g =< , > that is compatible with J, so that the fundamental two-form
ω ∈ A1,1 (X) , ω(v, w) = < Jv , w >
is defined. As in Definition 4.5, we say that (X, J, ω) is a Kähler manifold – and that g (or
ω) is a Kähler metric if dω = 0.20
Example 6.1. Any Riemann surface (complex manifold of dimension = 1) is Kähler.
Example 6.2. The standard constant metric g on Cn , with fundamental form
√
−1 X
ω = dzi ∧ dz i ,
2
is Kähler. Since the metric is invariant under translations, it follows that any complex torus
X = Cn /Λ is Kähler.
Example 6.3. (Submanifolds) If (X, g) is a Kähler manifold, then any complex subman-
ifold Y ⊆ X inherits a natural Kähler structure.
Example 6.4. Let (X, g, ω) be a compact Kahler manifold of dimension n. Since ω is closed,
it defines a class
[ω] ∈ H 2 (X, C).
Moreover since ω n is (a multiple of) the volume form determined by g, one has
Z
ω n > 0,
X
Fubini-Study metric. What gives the Kähler condition its real importance is that projective
space carries a natural SU(n+1)-invariant Kähler metric, the so-called Fubini–Study metric.
Therefore any complex submanifold of projective space also carries a Kahler metric. We
construct the Fubini-Study metric by building an SU(n + 1)-invariant Hermitian metric
HFS on Pn . The Fubini–Study form ωFS will then arise as the negative imaginary part
ωFS = −Im HFS of this Fubini–Study metric.21
20One sometimes says that a manifold in Kähler if it admits a Kähler metric, but for now we’ll generally
suppose given a particular Kähler metric.
21The following paragraphs are lifted from Section 1.2.C of my positivity book.
80 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
More explicitly, write Wx ⊆ V for the Hx0 -orthogonal complement to C · x ⊆ V , and let
πx : V −→ Wx be orthogonal projection:
h(v, x)
πx (v) = v − · x.
h(x, x)
Then Wx is identified with Tρ(x) Pn and πx with dρx , and
Hρ(x) dρx v, dρx w FS = Hx0 (πx v, πx w)
We next verify that ωFS is indeed a Kähler form, i.e. that ωFS is closed. Following
Mumford, one can use the SU(n + 1) invariance to give a quick proof of this. In fact, given
p ∈ Pn choose an element γ ∈ SU(n + 1) such that γ(p) = p while dγp = −Id. Then for any
three tangent vectors u, v, w ∈ Tp Pn one has
dωFS (u, v, w) = γ ∗ (dωFS )(u, v, w) = dωFS (−u, −v − w),
and hence dωFS = 0.
22It
may be useful to consider here an n-dimensional vector space W = Cn , with its standard Hermitian
product h(u, v) = t u · v. Then as w varies over W the expressions
0
ηw (u, v) = −Im h(u, v) , ηw (u, v) = −Im h(u, w)h(w, v)
define (1,
√
1)-forms η and η 0 on W , which
√
in terms of standard linear coordinates w1 , . . . , wn on W are given
by η = 2 · dwα ∧ dwα and η 0 = 2−1 ·
−1 P P P
wα dwα ∧ wα dwα .
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 81
Another approach to the Fubini–Study metric involves the Hopf map. Keeping the
notation of the previous example, consider the unit sphere
Cn+1 ⊇ S 2n+1 = S
with respect to the standard inner product h , i, with
p : S −→ Pn
the Hopf mapping. Denote by ωstd the standard Kähler form on Cn+1 , i.e.
X
ωstd = dxα ∧ dyα ,
where zα = xα + iyα are the usual complex coordinates on Cn+1 . Then ωFS is characterized
as the unique symplectic form on Pn having the property that
p∗ ωFS = ωstd | S.
(This follows from the construction in the previous example.)
Remark 6.6. (Normalization) There are various different normalizing conventions in the
literature. As shown in [3, p. 119] (where Huybrechts makes a different choice), one has
Z
ωF S = π.
P1
Noting that the Fubini-Study metric on Pn restricts to the Fubini–Study metric on any
linear subspace, it follows that π1 · ωFS represents a generator of
H 2 (Pn , Z) = Z.
The Kähler Identitites. We next study various pointwise and differential operators on a
complex manifold, and the relations that hold among them in the Kähler setting.
We now bring differentiation into the picture. Specifically, we introduce three operators:
d∗ : Ak (X) −→ Ak−1 (X)
∂ ∗ : Ap,q (X) −→ Ap−1,q (X) (6.1)
∗
∂ : Ap,q (X) −→ Ap,q−1 (X)
defined as the compositions
d∗ = − ∗ d ∗
∂∗ = − ∗ ∂ ∗ (6.2)
∗
∂ = −∗ ∂∗
∗
Thus d∗ = ∂ ∗ + ∂ .
One can interpret these operators as (formal) adjoints to d, ∂, ∂ with respect to natural
inner products on the spaces of forms in question. Specifically, given a compact complex
manifold (X, g) as above define a Hermitian inner product on the spaces Ak (X), Ap,q (X) by
the formula Z
(α, β)X = α ∧ ∗β. (6.3)
X
If as in §3 we denote by < , >herm the Hermitian extension of g, then it follows from equation
(3.17) that Z
(α, β)X = < α, β >herm dvol. (6.4)
X
∗
The adjoint property of ∂ , ∂ ∗ , d∗ are given by
Proposition 6.7. For every α ∈ Ap,q−1 (X) and β ∈ Ap,q (X) one has
∗
(∂α, β)X = (α, ∂ β)X .
and analogously for ∂ ∗ and d∗ .
On the other hand, since ∂η = ∂η, and since ∗2 = (−1)k on k-forms, one has
− ∗ ∗∂ ∗ β = (−1)p+q ∂ ∗ β.
Therefore Z Z
∗ p+q
(α, ∂ β)X = α ∧ (− ∗ ∗∂ ∗ β) = (−1) α ∧ ∂ ∗ β,
X X
as required.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 83
We next study the commutation relations among these operators when X is Kähler.
The first remark is
Proposition 6.8. Assume that X is Kähler. Then
[∂, L] = [∂, L] = 0 (6.5)
∗
[∂ , Λ] = [∂ ∗ , Λ] = 0. (6.6)
Proof. Note that ∂ω and ∂ω are respectively the (2, 1) and (1, 2) components of dω, and
hence dω = 0 if and only if ∂ω = ∂ω = 0. Therefore
∂(α ∧ ω) = ∂α ∧ ω , ∂(α ∧ ω) = ∂α ∧ ω,
which gives (6.5). For (6.6) one first uses that Λ = ∗−1 L∗ (Proposition 3.14) and the fact
that ∗2 = (−1)k on k-forms to prove that
∗
[∂ , Λ] = − ∗ [∂, L]∗,
and the assertion follows.
The deeper identity involves the commutation relations between Λ and ∂ and ∂.
Theorem 6.9. Assume that (X, g) is Kähler. Then
√ √ ∗
[Λ, ∂] = −( −1) ∂ ∗ , [Λ, ∂] = ( −1) ∂ (6.7)
∗ √ √
[L, ∂ ] = = −( −1) ∂ , [L, ∂ ∗ ] = ( −1) ∂ (6.8)
Turning to a sketch of the proof of the theorem, note that the assertion is an identity
between first-order differential operators. Therefore thanks to Theorem 4.6 it suffices to prove
it in a neighborhood of the origin in Cn with the standard (constant) metric. While one could
work on Cn with compactly supported forms, it is convient to keep dealing with compact
manifolds. Therefore, as do Griffiths–Harris, we will work in the following paragraphs with
a complex torus T = Cn /Λ with the flat Kähler metric coming from Cn . I will closely follow
Schnell’s presentation in [4].
In the sequel, we write dzI , dz J etc for the forms on T induced by the indicated trans-
lation invariant forms on Cn . We assume that the flat metric g is normalized so that
volg (T ) = 1, and we denote by h the corresponding Hermetian metric on T . Note that
h(dzI ∧ dz J , dzI ∧ dz J ) = 2|I|+|J| ,
and hence
(dzI ∧ dz J , dzI ∧ dz J )T = 2|I|+|J| .
Thus √ √
−1 X −1 X
L(α) = dzi ∧ dz i ∧ α = ei ei (α).
2 2
One next defines the adjoint
e∗i : Ap,q (T ) −→ Ap−1,q (T )
by the condition that e∗i be adjoint with respect to <, >herm of the point-wise operator given
by the same formula as ei , so that
(ei α, β)T = (α, e∗i β)T .
A combinatorial argument now proves the:
Lemma 6.10. Fix an index 1 ≤ i ≤ n.
(i). If i 6∈ J then
e∗i (dzJ ∧ dz K ) = 0 , e∗i dzi ∧ dzJ ∧ dz K
= 2dzJ ∧ dz K .
(ii).
(
2 · Id if i = j
ej e∗i + e∗i ej =
0 otherwise
Laplacians and harmonic forms. Let (X, g) be a compact complex manifold with a
compatible metric. Associated to each of the operators d, ∂, ∂ ∗ there is a Laplace operator
whose kernel consists (by definition) of harmonic forms. The Hodge theorem will assert
that any cohomolgy class has a unique harmonic representative. In general there need be
no particular connection between d-, ∂ and ∂ ∗ -harmonic forms, but in the case of Kähler
manifolds the three Laplacians essentially coincide. As we shall see, fact has many remarkable
consequences.
We start by defining the three Laplace operators in question. As in the previous para-
graph, let (X, g) be a compact complex manifold with a compatable metric.
Definition 6.12. We define the Laplace operators associated to d, ∂, ∂ to be the operators
∆d : Ak (X) −→ Ak (X)
∆∂ : Ap,q (X) −→ Ap,q (X) (6.12)
∆∂ : Ap,q (X) −→ Ap,q (X)
given by:
∆d = dd∗ + d∗ d
∆∂ = ∂ ∂ ∗ + ∂ ∗ ∂ (6.13)
∗ ∗
∆∂ = ∂ ∂ + ∂ ∂.
We emphasize that these spaces depend on the chosen metric on X. It will turn out that
they are finite dimensional. Note that in the case of d, one only needs (X, g) to be a compact
oriented Riemannian manifold.
Example 6.14. To explain the terminolgy, consider the Laplace operator ∆d acting on
functions f ∈ A0c (Rn ) where Rn is given the usual flat Euclidean metric.23 Then d∗ f = 0
since f is a zero-form, hence with Euclidean coordinates x1 , . . . , xn one has
X
∗ ∂f
∆d (f ) = d d f = − ∗ d ∗ dxi
∂xi
X
∂f
= −∗d ± dx1 ∧ . . . ∧ dxi ∧ . . . ∧ dxn
c
∂xi
X 2
∂ f
= −∗ dx1 ∧ . . . ∧ dxn
∂x2i
X 2
∂ f
=− .
∂x2i
So up to a sign, ∆d (f ) is just the classical Laplacian of f . One can compute ∆∂ (f ) and
∆∂ (f ) similarly (see [2, p. 83]).
Proposition 6.15. Let X be a compact complex manifold. Then a form α ∈ Ak (X) is
d-harmonic if and only if
dα = d∗ α = 0,
with the analogous condition for a form α ∈ Ap,q (X) to be ∂- or ∂-harmonic.
Proof. We use Proposition 6.7 (or more precisely its analogue for d∗ ). In fact, that result
shows that
(∆d α , α)X = (d d∗ α, α)X + (d∗ dα, α)X
= (d∗ α, d∗ α)X + (dα, dα)X .
Since both terms on the right are in any event non-negative, the assertion follows.
23Assume for simplicity that n is even to avoid sign worries involving the definition of d∗ given above.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 87
The crucial point for us is that on a Kähler manifold, the three different Laplacians that
we have defined essentially coincide.
Theorem 6.16. Let (X, g) be a compact Kähler manifold. Then
1
∆∂ = ∆∂ = ∆ ,
2 d
∗
and ∆d commutes with the operators ∗, ∂, ∂, ∂ ∗ , ∂ , L and Λ, and it preserves the decomposi-
tion into type.
Proof. We first show that ∆∂ = ∆∂ . Using equation (6.7) in Theorem 6.9 and recalling that
∂ and ∂ anticommute we find:
∆∂ = ∂ ∗ ∂ + ∂ ∂ ∗
√
= ( −1) [Λ, ∂] ∂ + ∂ [Λ, ∂]
√
= ( −1) Λ∂∂ − ∂Λ∂ + ∂Λ∂ − ∂ ∂Λ
√
= ( −1) Λ∂ ∂ − ∂[Λ, ∂] + ∂ ∂Λ + [∂, Λ]∂ + Λ∂ ∂ − ∂ ∂Λ
√ √ ∗
√ ∗
= ( −1) Λ∂ ∂ − ( −1)∂ ∂ − ∂ ∂Λ + −( −1)∂ ∂ + Λ∂ ∂ − ∂ ∂Λ
√ √ ∗ √ ∗
= ( −1) −( −1)∂ ∂ − ( −1)∂ ∂
= ∆∂ .
= ∆∂ + ∆ ∂ .
The proofs that ∆d commutes with all the holomorphic and anti-holomorphic operators
are similar, and are left to the reader. That ∆d commutes with the projection operators
π p,q : Ak (X) −→ Ap,q (X) follows from the facts that ∆d = 2∆∂ and that ∆∂ preserves the
type decomposition since it is the compostion of operators that do.
88 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
These identities already have an interesting corollary. Recall that a holomorphic p-form
on a complex manifold X is a (p, 0)-form η ∈ Ap,0 (X) such that ∂η = 0. Equivalently, η has
a local expression of the form X
η =locally fI dzI
where the fI are holomorphic. In complex dimension 1 any holomorphic (1, 0)-form is closed
simply because d = ∂ in (1, 0)-forms. However in higher dimensions there is no local reason
that a holomorphic form should be closed: eg η = z dw is not closed on C2 . However in the
global Kähler setting a holomorphic form is automatically closed:
Corollary 6.17. If X is a compact Kähler manifold, then any holomorphic form η on X is
closed.
We will see later (Exercise 6.4) that if η 6= 0, then moreover η cannot be exact.
We close this subsection with a proposition that provides yet another characterization
of harmonic forms. It will be used in the next subsection to give an intuitive heuristic for
the Hodge decomposition theorem.
Proposition 6.18. With X as above, consider the Hermitian inner product (6.3), and let
α ∈ Ak (X)
be a d-closed form on X. Then α is d-harmonic if and only if α has minimial length in its
deRham cohomology class, i.e.
α ∈ Hdk (X) ⇐⇒ ||α + dη||2 > ||α||2
for every η ∈ Ak−1 (X) with dη 6= 0. The analogous statements hold for ∂ and ∂ cohology
and ∆∂ and ∆∂ -harmonic forms.
Proof of Proposition 6.18. First assume that α is harmonic, so that dα = d∗ α = 0. Then for
any η ∈ Ak−1 (X) one has
|| α + dη ||2 = ||α||2 + (α, dη)X + +(dη, α) + X ||dη||2
= ||α||2 + (d∗ α, η)X + (η, d∗ α)X + ||dη||2
= ||α||2 + ||dη||2 ,
so α has minimal norm. Conversely, if α has minimal norm, then taking η = t d∗ α in the
above we find that
||α + t dd∗ α|| = ||α||2 + 2t||d∗ α||2 + t2 ||dd∗ α||2 .
This has a minimum at t = 0 if and only if ||d∗ α|| = 0, ie iff α is harmonic.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 89
The Hodge Theorem. We now turn to the Hodge theorem and its consequences.
Let (X, g) be a compact connected complex manifold with a compatible complex struc-
ture. The Hodge theorem – which we will momentarily state precisely – asserts that there
is a unique harmonic representative in each d−, ∂− or ∂− cohoomology class. The proof
of the theorem is analytic in nature – it’s ultimately a consequence of regularity theorems
for elliptic PDE – and we do not attempt to give anything like a real explanation. However
there is an enjoyable heuristic argument (which most of us learned from Griffiths-Harris)
suggesting why the statement is not unreasonable.
The starting point is Proposition 6.18, which asserts that a form is harmonic (with
respect to any of the three Laplacians) if and only it has minimal length in its (d-, ∂- or ∂-)
cohomology class. Focusing for concreteness on the d-Laplacian ∆d and DeRham cohomolgy,
imagine – which unfortunately is not the case – that Ak (X) were a Hilbert space with respect
to the inner product (, )X and that
Im d : Ak−1 (X) −→ Ak (X) ⊆ Ak (X)
were a closed subspace. Then we could find a unique element of minimal lengh in the affine
subspace
[α] = α + Im(d),
which according to the Proposition would be the unique harmonic form in this cohomology
class. This provides at least some intuitive reason to imagine that something like the Hodge
theorem could be true.
Here is the formal statement of the Hodge decomposition theorem. To clarify where this
is and isn’t used, for the moment we do not impose the Kähler condition.
Theorem 6.19. (Hodge Theorem). Let (X, g) be a compact complex manifold with a
compatible metric. Then the three spaces
Hdk (X, g) ⊆ Ak (X) and H∂p,q (X, g) , H∂p,q (X, g) ⊆ Ap,q (X)
of harmonic forms are finite dimensional. Morover:
The orthogonality in the theorem is of course with respect to the Hermitian inner produce
(6.3). We will say a few words in Remark ?? about the general PDE facts underlying the
theorem.
90 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Corollary 6.20. In the situation of the Theorem, there are canonical isomorphisms
∼
=
Hdk (X, g) −→ HdR
k
(X)
∼
=
H∂p,q (X, g) −→ H∂p,q (X)
∼
=
H∂p,q (X, g) −→ H∂p,q (X).
Lemma 6.21. One has
ker d : Ak (X) −→ Ak+1 (X) = Hdk (X, g) ⊕ d Ak−1 (X),
(6.15)
with analogous statements for ker ∂ and ker ∂.
Proof. Evidently the RHS of (6.15) contains the LHS. For the reverse inclusion, suppose
that dα = 0. By the Hodge decomposition one can write α = dα1 + h + d∗ α2 , where h is
harmonic. It suffices to show that α2 = 0. Now since dα = 0, we have dd∗ α2 = 0, but then
||α22 || = (d∗ α2 , d∗ α2 ) = (dd∗ α2 , α2 ) = 0,
as required.
Proof of Corollary 6.20. We prove the first statement. Since a harmonic form is d-closed,
there is a natural map
Hdk (X, g) −→ HdR
k
(X) , α 7→ [α]. (*)
It follows from the orthogonality of the Hodge decomposition that a non-zero harmonic form
cannot be exact, hence (*) is injective. Moreover, equation (6.15) implies that any closed
form is cohomologous to a (unique) harmonic form, hence (*) is surjective.
Assume now that (X, g) is compact Kähler. Then the three Laplacians in play all
coincide, so we expect the Hodge decomposition to have particularly striking consequences.
This is very much the case. In fact, by the deRham theorem we have
H k (X, C) = Hdk (X)
and we write H p,q (X) = H∂p,q (X) for the ∂- cohomogy of X, so that
H p,q (X) = H q (X, ΩpX )
thanks to the Dolbeaut theorem.
Proof. Fix a Kähler metric g on X, and consider the resulting spaces of harmonic forms. In
the first place, by Corollary 6.20 we have isomorphisms
∼
=
H∂p,q (X, g) −→ H p,q (X),
∼
=
Hdk (X, g) −→ H k (X, C).
Furthermore, since ∆d commutes with taking (p, q)-components (Theorem 6.16), we have a
decomposition
Hdk (X, g) = ⊕p+q=k Hdp,q (X, g).
On the other hand, thanks to Theorem 6.16 it is the same to be harmonic for d or for ∂, and
therefore Hdp,q (X, g) = H∂p,q (X, g). Putting this together, we get isomorphisms
H k (X, C) ∼
= Hdk (X, g) = ⊕ H∂ (X, g) ∼
p,q
= ⊕ H p,q (X).
It remains to show that the resulting isomorphism between the outer terms is independent
of the choice of Kahler metric. Fixing a second Kähler metric g1 , this amount to showing
that if we have harmonic forms
α ∈ Hp,q (X, g) , α1 ∈ Hp,q (X, g1 )
representing the same cohomology class in H p,q (X) then α and α1 represent the same class
in H p+q (X, C). The hypothesis means that
α − α1 = ∂β,
and by the directness of the decomposition in Theorem 6.19 (ii) for ∂, it follows that α − α1
is orthogonal to H∂p,q (X, g) = Hdp,q (X, g). Since d(α − α1 ) = 0, it follows from Proposition
6.21 that α − α1 = dγ, as required. Finally if α ∈ Hp,q (X), then
α ∈ Hq,p (X) , ∗α ∈ Hn−q,n−p (X)
(Exercise 6.5), which gives the isomorphisms (6.17).
We conclude this section with a useful result, called the ∂∂- Lemma.
Proposition 6.23. Let X be a compact Kähler manifold, and let
α ∈ Ap,q (X)
be a d-closed (p, q)-form. Then the following are equivalent:
Proof. Since α is a form of pure type (p, q), it is d closed if and only if it is both ∂- and
∂-closed. Therefore it follows from Theorem 6.19 that each of the other conditions implies
(v), and clearly (iv) implies (i) – (iii). So it suffices to show that (v) ⇒ (iv). Since ∂α = 0
and α is perpendicular to H∂p,q (X), Lemma 6.21 for ∂ implies that α = ∂β for some β. Now
use the Hodge decomposition for ∂ to write
∗
β = ∂γ1 + γ2 + ∂ γ3 ,
where γ2 is harmonic. Therefore
∗
α = ∂β = ∂∂γ1 + ∂∂ γ3
∗
= −∂∂γ1 − ∂ ∂γ3 .
∗
On the other hand, ∂α = 0, so ∂∂ ∂γ3 = 0, which implies that
∗ ∗
(∂ ∂ ∂γ3 , ∂γ3 )X = ||∂ ∂γ3 ||2 = 0,
and hence
∗ ∗
−∂∂ γ3 = ∂ ∂γ3 = 0.
Thus α = ∂∂γ1 , as required.
Let (X, g) be a compact Kähler manifold, with Kähler form ω. Denote as usual by L and
Λ the Lefschetz operator and it’s dual. Since these commute with the Laplacian ∆ = ∆d ,
they give rise to homomorphisms
L : Hp,q (X) −→ Hp+1,q+1 (X) , Hp,q (X) −→ Hp−1,q−1 (X),
and hence to
L : H p,q (X) −→ H p+1,q+1 (X) , Λ : H p,q (X) −→ H p−1,q−1 (X).
Note that L is given by cup product with [omega] ∈ H 1,1 (X), and hence depends only on
this class. One can show that the same is true of Λ.
Note that the two groups appearing here are Poincaré dual, so the deep fact is not that there
exists an isomorphism between the groups, but rather that (n − k)-fold iteration of L does
the job.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 93
Moreover
P k (X) = ker Ln−k+1 : H k (X, C) −→ H 2n−k+2 (X, C) ,
Proof. By the Hodge decomposition theorem, primitive classes are (uniquely) represented by
primitive harmonic forms:
P k (X) =def ker Λ : Hk (X) −→ Hk−2 (X) ,
(*)
and similarly for P p,q . So it is equivalent to prove the Lefschetz decomposition for harmonic
forms. For this one can either reduce to Theorem 3.16, or one can note that L and Λ a
generate an SL2 action on primitive forms, and argue as in §3. We leave details to the
reader.
Remark 6.26. Note that the condition (*) defining primitive harmonic forms operates
pointwise. Therefore if α ∈ P k is a primitive harmonic form, then α(x) ∈ Λk T ∗ X is a
primitive vector for every x ∈ X (although possibly α(x) = 0 at some x ∈ X.
Theorem 6.27. Let X be a compact Kähler manifold of dimension n, with Kähler class ω.
Let 0 6= α ∈ P p,q (X). Then
√
Z
(p+q)(p+q−1)
p−q
( −1) (−1) 2 α ∧ α ∧ ω n−(p+1) > 0.
X
Proof. It is enough to prove the statement when α is a primitive harmonic form. Fix a point
x ∈ X at which α(x) 6= 0. Then thanks to Remark 6.26 the assertion follows from the
pointwise statement Theorem 3.27 established in §3.
We now turn to the remarkable Lefschetz (1, 1)-theorem, which describes the holomor-
phic line bundles on a compact Kähler manifold. Recall first that isomorphism classes of
smooth complex line bundles on X are classified by H 2 (X, Z): this arises from the isomor-
phism
∼
=
H 1 (X, A∗ ) −→ H 2 (X, Z) , L 7→ c1 (L) (*)
∗
deduced from the exponential sequence 0 −→ Z −→ A −→ A −→ 0 where A is the sheaf
of smooth C-valued functions on X. Here c1 (L) ∈ H 2 (X, Z) is the first Chern class of a
line bundle L: for now one can take (*) as the definition of c1 (L), but we will discuss other
manifestations shortly.
where ι : H 2 (X, Z) −→ H 2 (X, C) is the natural map.25 In other words, if we ignore torsion
and pretend that H 2 (X, Z) embeds in H 2 (X, C), then one has the more picuresque relation
H 1,1 (X, Z) = H 2 (X, Z) ∩ H 1,1 (X).
Theorem 6.29. (Lefschetz (1,1) theorem) A class γ ∈ H 2 (X, Z) first Chern class of a
holomorphic line bundle if and only γ ∈ H 1,1 (X, Z).
25Thus ι embeds H 2 (X, Z) modulo its torsion subgroup into
H 2 (X, C) = H 2 (X, Z) ⊗Z C.
We note that some authors define H 1,1 (X, Z) to be the intersection
im H 2 (X, Z) −→ H 2 (X, C) ∩ H 1,1 (X).
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 95
∗
Proof. We consider the exponential seqence 0 −→ Z −→ OX −→ OX −→ 0, which gives rise
to an exact sequence
∗ c
H 1 (X, OX 1
) −→ H 2 (X, Z) −→ H 2 (X, O). (6.18)
Our question is to understand the image of the map
∗
Pic(X) = H 1 (X, OX ) −→ H 2 (X, Z),
so the Theorem is equivalent to the assertion that
H 1,1 (X, Z) = ker H 2 (X, Z) −→ H 2 (X, O) . (*)
We assert:
Claim 6.30. Under the Dolbeaut isomorphism
H 2 (X, OX ) = H 0,2 (X),
the map H 2 (X, Z) −→ H 2 (X, OX ) coming from the exponential sequence associates to a
class γ ∈ H 2 (X, Z) the (0, 2)-component of its image in H 2 (X, C). (Exercise 6.9.)
Granting this, one immediately gets the inclusion ⊆ in (*), and the other inclusion follows
from the observation that since the image of H 2 (X, Z) −→ H 2 (X, C) consists of real classes,
2,0
ι(γ)0,2 = 0 ⇐⇒ ι(γ)0,2 = ι(γ) = ι(γ)2,0 = 0.
To complete the picture, we say a word about the Picard torus of a compact Kähler
manifold. Define
Pic0 (X) = ker Pic(X) −→ H 2 (X, Z)
to be the group of topologically trivial holomorphic line bundles on X. This fits into the
exact sequence
0 −→ H 1 (X, Z) −→ H 1 (X, O) −→ Pic0 (X) −→ 0.
Now it follows from the Dolbeaut isomorphism that H 1 (X, OX ) = H 0,1 (X), and as above
the map H 1 (X, Z) −→ H 0,1 (X) occuring here is the composition
H 1 (X, Z) −→ H 1 (X, C) −→ H 0,1 (X)
coming from the Hodge decomposition of H 1 . Now note that
rank H 1 (X, Z) = b1 (X) = 2 · h01 (X) = 2 dimC H 1 (X, OX ).
This suggests that one might hope that H 1 (X, Z) sits as a lattice inside H 1 (X, OX ). In fact,
this is the case (Exercise 6.10). Therefore
Pic0 (X) = H 1 (X, OX )/H 1 (X, Z)
has in the natural way the structure of a complex torus, called the Picard torus of X.
Thus we have the beautiful picture that the set of isomorphism classes of holomorphic line
bundles having a given topological type is parametrized by a torus of complex dimension
h0,1 (X) = 21 b1 (X).
96 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
(a). Prove that this is indeed an open convex cone. (Cf. [3, p. 130].)
(b). Describe the Kähler cone explitly for X = P1 × P1 .
We remark that in all but the simplest cases it can be difficult to compute this cone explicitly,
especially if one wants to know the subspace spanned by the integral classes. For example,
if X is a compact Riemann surface of genus g > 2, then K ∩ H 2 (X, Z) can depend on the
intrinsic geometry of X, and in fact it is not known what K ∩ H 2 (X, Z) is for “general” X.
See [PAG, §1.5.B] for a survey of the algebro-geometric analogue of this question.
Exercise 6.8. Let X be a compact Kähler maifold.
(b). Prove that π1 (X) cannot be isomorphic to the free group on three generators.
(c). Prove that π1 (X) cannot be isomorphic to the free group on two generators. [Hint:
Otherwise there would be a two-sheeted covering space of X whose π1 is free on three
generators, and such a covering is again a compact Kähler manifold.]
Exercise 6.9. Prove Claim 6.30. (See [3, Lemma 3.3.1].)
Exercise 6.10. Using the facts that
H 1 (X, C) = H 1,0 (X) ⊕ H 0,1 (X) , H 0,1 = H 1,0 ,
show that the image of H 1 (X, Z) −→ H 0,1 (X) is a lattice. (See [3, Cor. 3.3.6].)
Exercise 6.11. (The case of complex tori) Let Λ ⊆ V be a lattice in an n-dimensional
complex vector space V . We consider the complex torus X = V /Λ with the flat (constant)
metric on V . As usual, we may consider forms on V that are invariant under translation by
Λ to be forms on X.
(a). Prove that the harmonic forms on X are just the constant ones of the sort
X
aI,J dzI ∧ dz J , (*)
show that Z
1,1
η ∈ H (X, Z) ⇐⇒ η ∈ Z ∀i, j. (*)
γi ∧γj
(c). Show by explicit computation that there exist Λ ⊂ V that determine tori X for which
H 1,1 (X, Z) = 0. [Hint: By a suitable choice of coordinates one can suppose that
λn+1 = e1 , . . . , λ2n = en ,
where ei are the standard basis vectors. Then compute the integrals appearing in
(*) in terms of the coordinates of λ1 , . . . , λn , and check that for sufficiently general
choices of these λi there are no solutions to (*).]
98 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
In this final section we study the notion of a positive line bundle on a Kähler manifold,
and we prove the Kodiara vanishing and embedding theorems.
Positive Line Bundles. We start with the notion of positivity for a (1, 1)-form on a com-
plex vector space.
Definition 7.1. Let (V, J) be a real vector space with an almost complex structure. A
(1, 1)-form η ∈ Λ1,1 V ∗ is positive if
η(v, Jv) > 0 for all 0 6= v ∈ V.
Equivalently, with notation as in §3, η is positive if one can write
√
−1 X
η = hij z i ∧ z j
2
where hij is a positive definite Hermitian metric.
Yet another way to phrase the definition is to require that if W ⊆ V is any J-stable subspace
of real dimension 2 – so that (W, J|W ) is a two-dimensional almost complex vector space –
then η|W = c volW for some c > 0.
Similarly:
Definition 7.2. Let X be a complex manifold, with J the corresponding almost complex
structure, and let η be a (1, 1)-form on X. We say that η is positive if
ηx ∈ Λ1,1 Tx∗ X
is a positive form for every x ∈ X.
The plan now is to introduce now a positivity notion for a holomorphic line bundle L
by asking in effect that c1 (L) be representable by a positive (1, 1)-form.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 99
Suppose that (L, h) is a hermitian line bundle, and consider a holomorphic section
s ∈ Γ(U, L) of L on an open subset U ⊂ X. Then we take the fibre-wise norm square of s
to get a non-negative function
|s|2 = |s|2h = h(s, s)
on U . If L|U = U × C is trivial, then conversely a hermitian metric on L|U is specified by
giving a positive smooth function
h : U −→ R>0 (7.1)
which we declare to be the (square) length of the section x −→ 1 ∈ C. In general, if L is
described by transition data (Ui , gij ) then a metric is given by smooth functions
hi : Ui −→ R>0
such that
hi = |gij |2 hj on Uij . (7.2)
A simple argument with partitions of unity shows that any holomorphic (or, for that matter,
smooth) line bundle admits (many different) hermitian metrics. (Exercise 7.1.)
Let (L, h) be a hermitian line bundle described by local data (Ui , hi ) as in (7.2). We
claim that the (1, 1)-forms
ti =def −∂∂ log hi
patch together to give a globally defined form
Θ(L, h) ∈ A1,1 (X),
which is called the curvature form of (L, h). In fact, on Ui ∩ Uj one has
log hi = log |gij |2 + log hj ,
and
∂∂ log |gij |2 = ∂∂ gij g ij
A basic fact is that Θ(L, h) essentially represents the first Chern class of L:
Proposition 7.6. One has
√
−1
c1 (L) = 2π
Θ(L, h) ∈ H 2 (X, C).
The prototypical example is the hyperplane line bundle OPn (1) on projective space:
Proposition 7.9. The hyperplane line bundle on Pn carries a positive hermitian metric.
Proof. We will show that OPn (1) carries a Hermitian metric h whose curvature form is a
multiple of the Fubini–Study form ωFS . In fact, the standard Hermitian product hv, wi = t v·w
on V = Cn+1 gives rise to a Hermitian metric on the trivial bundle VPn on Pn = P(V ).
Then OPn (−1) inherits a metric as a sub-bundle of the trivial bundle Pn × V , which in turn
determines a metric h on OPn (1). Very explicitly, write [x] ∈ Pn for the point corresponding
to a vector x ∈ V − {0} and consider a section
s ∈ V ∗ = H 0 Pn , OPn (1) .
An important fact is that on a Kähler manifold, the positivity of a line bundle L depends
only on the cohomology class of c1 (L).
Proposition 7.12. Assume that X is a compact Kähler manifold. Consider a hermitian
line bundle such that
√
−1
[ 2π Θ(L, h0 )] = [ω] ∈ H 2 (X, C).
where ω is a positive (1, 1) form. Then L carries a hermitian metric h such that
√
−1
2π
Θ(L, h) = ω.
In particular, L is positive.
But by assumption the left hand side of (*) is d-exact, hence by the ∂∂ lemma (Proposition
6.23) it is ∂∂-exact, as required.
The first statement is the Kodaira vanishing theorem, while the second is usually called
Nakano vanishing.
Harmonic theory for hermitian line bundles. The first ingredient going into Theorem
7.13 is an extension of the Hodge decomposition theorem for hermitian line bundles. The
analysis underlying this result (which in any event we ignore) is nothing beyond that which
goes into the classical statement, once things have been set up properly. This set-up is what
we now discusss. Most of this material generalizes with no change to the case of hermitian
vector bundles of higher rank, but for simplicity we stick with line bundles.
Consider then a hermitian holomorphic line bundle (L, h) on a compact complex man-
ifold (X, g) of dimension n with a compatible Riemannian metric. The metric h allows us
first of all to define a Hodge ∗-operator on L-valued forms. Namely, one can view h as
giving a C-anti-linear isomorphism h : L −→ L∗ , and then one can define a C-anti-linear
isomorphism
∗L : Λp,q T ∗ X ⊗ L −→ Λn−p,n−q T ∗ X ⊗ L∗
by the rule
∗L α ⊗ s = ∗α ⊗ h(s) = ∗α ⊗ h(s) .
The metrics h and g also give rise to a global hermitian product on the spaces Ap,q (L) of L
valued forms via Z
α ⊗ s, β ⊗ t L = h(s, t) · (α ∧ ∗β), (7.3)
X
extending the definition in (6.3). In particular, if φ ∈ Ap,q (X, L), then
Z
φ ∧ ∗L φ = (φ, φ)L .
X
We now bring ∂ into the picture. Recall from (5.1) that because L is holomorphic there
is a canonically defined operator
∂ L : Ap,q (L) −→ Ap,q+1 (L).
We then define the adjoint
∗
∂ L : Ap,q (L) −→ Ap,q−1 (L)
by the formula
∗
∂ L = −∗L∗ ◦ ∂ L∗ ◦ ∗L .
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 103
(See Exercise 7.4). As in Proposition 6.7, this is adjoint to ∂ L with respect to the inner
product (7.3):
Lemma 7.14. For any α ∈ Ap,q (L) and β ∈ Ap,q+1 (L) one has
∗
α , ∂ Lβ L = ∂ Lα , β L.
And just as in §6, one concludes that every class in H∂p,q (X, L) = H q (X, ΩPX ⊗ L) has a
unique harmonic representative:
Corollary 7.17. The natural map
Hp,q (X, L) −→ H∂p,q (X, L)
is an isomorphism.
Before going on, we pause to note an important consequence of this picture, namely the
famous Serre duality theorem. Let L be a holomorphic line bundle on a compact complex
manifold X of dimension n. Specifically, there is a natural pairing
Z
p,q n−p,n−q ∗ n,n
H (X, L) × H (X, L ) −→ H (X) = C , (α, β) 7→ α ∧ β. (7.4)
X
∗
Here we view α ∧ β as an (n, n)-form with values in L ⊗ L = 1X .
Theorem 7.18. (Serre duality). The map (7.4) is a perfect pairing. In particular,
H p,q (X, L) ∼
= H n−p,n−q (X, L∗ )∗ .
104 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Sketch of Proof. We will give the local construction of ∇L . Note first that the difference
of any two connections on L is a 1-form on X. That said, choose a local trivialization
LU ∼= U × C of L over some open set U ⊆ X. Then on U sections of L are identified with
smooth functions, and ∇0,1 1,0
L = ∂ L acts simply as ∂. On the other hand, ∇L differs from ∂
by wedge product with a (1, 0)-form. In other words, one can write locally on U :
∇L = ∂ + θ + ∂
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 105
where θ ∈ A1,0 (U ). It remains to see what is the condition on θ imposed by the requirement
that ∇ be hermitian. For this, let e ∈ Γ(U, L) be the section determined by the constant
vector 1 ∈ C under the local trivialization of L. Then
h(e, e) = h
is the smooth function on U determining the metric as in equation (7.1), and ∂e = ∂e = 0.
Hence (7.5) yields
dh = h · θ + h · θ.
Since dh = ∂h + ∂h and since θ has type (1, 0), we find:
1
θ = · ∂h = ∂ log(h). (7.6)
h
This proves the existence and uniqueness of ∇L .
Proof of Lemma 7.20. For simplicity, we check the stated identity when the two sides act
on smooth sections of L. Working locally, we may identify such a section with a smooth
function s. Then, using the local description of ∂L derived in the proof of Proposition 7.19,
we have
∂L ∂ L (s) = ∂(∂s) + θ ∧ ∂(s)
where θ = ∂(log h) is the 1-form appearing in (7.6). On the other hand,
∂ L ∂L (s) = ∂ L ∂s + θ ∧ s
= ∂∂(s) + ∂θ ∧ s − θ ∧ ∂s.
Therefore
∂L ∂ L + ∂ L ∂L (s) = ∂∂ log(h)) · s,
as required.
106 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
There are three more operators that come into the picture. First, wedging with the (1, 1)
form ω determined by the metric g gives in the natural way an extenstion of the Lefschetz
operator
L : Ap,q (X, L) −→ Ap+1,q+1 (X, L) , α ⊗ s 7→ ω ∧ α ⊗ s.26
As before, we have the adjoint operator Λ:
Note that these are point-wise operators, which coincide with the previously studied versions
when we take a local trivialization of L. One can check moreover that Λ is the global adjoint
of L with respect to the hermitian inner product ( , )L introduced above. Finally we denote
by ∂L∗ : Ap,q (L) −→ Ap−1,q (L) the adjoint of ∂L with respect to ( , )L . Then the Kähler
identity (6.7) generalizes to
This is sometimes called the Nakano identity. It fairly easily reduces to (6.7): see for instance
[3, Lemma 5.2.3].
Proof of Theorem 7.13. Since L is positive, it carries a hermitian metric h such that
√
−1
ω =def Θ(L, h)
2π
is a closed positive (1, 1)-form. We may then choose a Kähler metric on X with ω as the
corresponding Kähler form. With these choices, (7.8) becomes
√
∂L ∂ L + ∂ L ∂L = −2π −1 L,
26We apologize for the two conflicting meanings of the symbol L, but there doesn’t seem to be any easy
way to avoid this, and hopefully no confusion will result.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 107
Thanks to Corollary 7.17, the issue is to show that α = 0 if p + q > n. For this, we compute:
√ ∗ ∗
−1
Λα , (∂L∗ ∂ L + ∂ L ∂L∗ )α L
Λα , Λα L = 2π (by (7.10))
√ ∗ ∗
−1
Λα , ∂ L ∂L∗ α L
= 2π (since ∂ L α = 0)
√
−1
= 2π
∂ L Λα , ∂L∗ α)L
√
−1
− [Λ, ∂ L ]α, ∂L∗ α L
= 2π
(since ∂ L α = 0)
√
−1
√ ∗ ∗
= 2π
−1∂ L α, ∂ L α L
(by Proposition 7.22)
1
∂L∗ α, ∂L∗ α L .
= − 2π
But by comparing the signs of the first and last expressions we find that (Λα, Λα)L = 0, and
hence Λα = 0. But this means that α is locally represented by a primitive form. On the
other hand, by Theorem 3.16 (iii), there are no non-vanishing primitive forms of degree > n.
Hence α = 0, and we are done.
Corollary 7.23. Let L be a positive line bundle on a compact Kähler manifold X. Then
given any line bundle P on X, there exists a constant m0 = m0 (L, P ) such that if m ≥ m0
then
H q (X, L⊗m ⊗ P ) = 0
for q > 0.
This is a special case of the Serre vanishing theorem. The reader is asked to prove the
statement in Exercise 7.6.
The Kodaira embedding theorem. We now come to our final theorem, which ties to-
gether much of what we’ve studied up to now.
Theorem 7.24. (Kodaira Embedding Theorem). Let X be a compact Kähler manifold,
and let L be a positive holomorphic line bundle on X. Then for m 0 there exists an
embedding
φ = φm : X −→ Prm
(for some large r = rm ) such that
L⊗m = φ∗ OPr (1).
In other words, a compact Kähler manifold admits a projective embedding if and only if it
carries a positive line bundle.
The plan is to show first of all that L⊗m is globally generated for m 0, so that L⊗m
defines a holomorphic mapping φ : X −→ Pr . Then we will show that – possibly after
increasing m – the resulting mapping is actually an embedding. We will concentrate on
proving the first point, the second being similar.
108 NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014
Fix a point x ∈ X. We want to show that if m 0 then L⊗m has a holomorphic section
that doesn’t vanish at x. To this end, denote by mx ⊆ OX the ideal sheaf of holomorphic
functions vanishing at x, and consider the exact sequence
0 −→ mx −→ OX −→ O{x} −→ 0,
so that O{x} is a one-dimensional sky-scraper sheaf supported at x. Tensoring by L⊗m we
arrive at
0 −→ L⊗m ⊗ mx −→ L⊗m −→ L⊗m ⊗ O{x} −→ 0,
and the long exact cohomology sequence gives
ev
H 0 (X, L⊗m ) −→
x
C −→ H 1 (X, L⊗m ⊗ mx ). (7.11)
Here C = H 0 (L⊗m ⊗ O{x} ) and the map evx is evaluation at x. Now suppose one knew that
H 1 (X, L⊗m ⊗ mx ) = 0 for m 0. (7.12)
Then evaluation the evaluation map evx would be surjective, and we could conclude that
L⊗m has a global section that is non-vanishing at x.27 The question being reduced to a
vanishing statement involving a positive line bundle, one can hope to apply Theorem 7.13.
Unfortunately it doesn’t apply directly because – when n = dim X ≥ 2 – the ideal sheaf
mx ⊆ OX is not a line bundle. However we can circumvent this problem by blowing up the
point x ∈ X, which reduces one to a question about locally free sheaves.
Blowing up a point. Recall that in Example 2.35 we discussed the construction of the blow-
up of a point in Cn . Since this construction is local about a neighborhood of 0 ∈ Cn , we
will be able carry it over to any complex manifold X to produce a new manifold X 0 in which
the ideal sheaf of a point is in effect replaced by a line bundle.
We start by reviewing and analyzing the construction in Cn . Viewing Pp−1 as the space
of lines through the origin in Cn , consider
Z = ([a], v) | v ∈ C · a ⊆ Pn−1 × Cn .
Write
b : Z −→ Cn , f : Z −→ Pn−1
for the two projections. Thus f realizes Z as the total space of the line bundle OPn−1 (−1)
over Pn−1 . In particular, the fibre of b over 0 ∈ Cn is a copy of Pn−1 which we is called the
exceptional divisor of Z:
E = Pn−1 = b−1 (0) ⊆ Z.
Note that E is naturally identified as the projective space PT0 Cn of lines in the (holomorphic)
tangent space of Cn at 0. Furthermore, note that if ∆ is any neighborhood of 0 ∈ Cn , and
if ∆∗ = ∆ − {0}, then b restricts to an isomorphism
∼
=
b−1 (∆∗ ) −→ (∆∗ ). (7.13)
We set
∆0 = Bl0 (∆) =def b−1 (∆) :
27A similar argument with mx replaced by the ideal mx,y of two points x, y ∈ X would show that the
morphism φ is one-to-one, and replacing mx by m2x yields that dφ is injective.
NOTES ON COMPLEX GEOMETRY FOR MATH 545, FALL 2014 109
It is useful for computations to observe that if Ui ⊆ Pn−1 are the standard open subsets
of Pn−1 , then there are natural isomorphisms f −1 (Ui ) ∼ = Cn under which b is locally identified
n n
with the maps b : C −→ C given by
(w0 , w1 , . . . , wn−1 ) 7→ wi w0 , wi w1 , . . . , wi , . . . , wi wn−1 (7.14)
(Exercise 7.8). In these local coordinates on Z, the exceptional divisor E is (locally) defined
in Z by the equation wi = 0, and f is given by (w0 , . . . , wi , . . . wn−1 ) 7→ [w0 , . . . , 1, . . . , wn−1 ].
Denote by OZ (−E) ⊆ OZ the ideal sheaf of E, i.e. the sheaf of holomorphic functions
vanishing on E. Then, using the local description just given, one can check (Exercise 7.8):
Lemma 7.25. The ideal sheaf OZ (−E) is a line bundle on Z, and
OZ (−E) ∼ = f ∗ OPn−1 (1). (7.15)
(See Exercise 7.8.) This will allow us shortly to put a metric on OZ (−E). The reader should
also check by taking the Jacobian determinant of (7.14) that
b∗ ΩnCn = ΩnZ ⊗ OZ (−(n − 1)E);
very concretely, the pull-back via b of the form dz1 ∧ . . . ∧ dzn vanishes to order (n − 1) along
E.
The next point is to study the positivity of the relevant line bundles.
Lemma 7.27. Let L be a positive line bundle on X, and fix an integer a > 0 and an arbitrary
line bundle P on X. Then for m 0 the line bundle
µ∗ (L⊗m ⊗ P ) ⊗ OX 0 (−aE))
is positive on X 0 .
Proof. Fix a positive metric hL on L and an arbitrary metric hP on P . Then h⊗m L ⊗ hP pulls
back to a metric hm on µ∗ (L⊗m ⊗ P ). By taking m 0 we can arrange that it is positve
away from E, but it won’t be positive along E itself since Θ(hm ) vanishes on vectors tangent
to E. On the other hand, via the local description (7.15) of OX 0 (−E), we can find (using a
partition of unity) a metric ha on OX 0 (−aE) that in a neighborhood of E agrees with a power
of the pullback f ∗ h⊗a
FS of the Fubini-study metric on OPn−1 (1) (but we don’t know anything
about ha away from E). We claim that if m 0 then hm ⊗ ha is everywhere positive. This
will follow from Exercise 7.3 (a) away from E, so it remains to check the positivity along
E. For this, note that by construction Θ(ha ) is positive on tangent directions to E. On the
other hand, Θ(hm ) is positive on directions normal to E since (as we may assume) it’s the
pull-back of a positive (1, 1)-form on X. Therefore hm ⊗ ha is positive everywhere on X 0 for
m 0.
Corollary 7.28. Let L be a positive line bundle on X. If m 0, then
H 1 X 0 , µ∗ (L⊗m ) ⊗ OX 0 (−E) = 0.
Proof. We apply the previous Lemma with P = (ΩnX )∗ and a = n = dim X. Then by (7.16),
one has:
µ∗ L⊗m ⊗ OX 0 (−E) = ΩnX 0 ⊗ µ∗ L⊗m ⊗ (ΩnX )∗ ⊗ OX 0 (−nE) ,
We now indicate the proof of the Kodaira embedding theorem. Given a point x ∈ X,
we focus on showing that L⊗m has a section that doesn’t vanish at x for m 0: this implies
that the mapping φm appearing in Theorem 7.24 is well-defined, and as indicated above the
argument that it is one-to-one with injective derivative is similar. For simplicity we assume
n ≥ 2 (the case of Riemann surfaces being elementary given what we already know).
Then the restriction of s0 to X 0 − E can be viewed as a section s0 ∈ Γ(X − {x}, L⊗m ) of L⊗m
on the complement of x. But by Hartog’s theorem, s0 extends to a section s ∈ Γ(X, L⊗m )
which pulls back to s0 .
evx
H 0 X, L⊗m ) / H 0 X, L⊗m ⊗ O{x}
The vertical map on the left is an isomorphism, and since µ∗ L is trivial along E, so is the
vertical map on the right. Thus we can identify the top map with evaluation of sections at
x ∈ X. On the other hand, it follows from Corollary 7.28 – and this is the serious point –
that the top vertical map is surjective if m 0. Therefore evx is surjective when m 0,
and so we have produced a section
sx ∈ Γ X, L⊗m with sx (x) 6= 0.
(a). Let ω be a positive (1, 1)-form on X, and let η be an arbitrary (1, 1)-form. Show that
if m 0, then mω + η is positive.
(b). Let L be a positive holomorphic line bundle on X, and let P be an arbitrary holo-
morphic line bundle on X. Show that if m 0, then L⊗m ⊗ P is again a positive
line bundle.
Exercise 7.4. Show that if L = X × C is the trivial line bundle with the constant metric,
∗ ∗
then ∂ L = ∂ .
Exercise 7.5. Prove that if α ∈ Hp,q (X, L), then
∗L α ∈ Hn−p,n−q (X, L∗ ).
Exercise 7.6. Prove Corollary 7.23. (Hint: Use Exercise 7.3 to write L⊗m0 ⊗ P = ΩnX ⊗ N
where N is a positive line bundle.)
Exercise 7.7. Prove that a compact Kähler manifold X admits a projective embedding if
and only if it carries a Kähler form ω with rational periods, ie with the property that
Z
ω ∈ Q
γ
for all γ ∈ H2 (X, Z). (Hint: The condition is equivalent to the condition that ω ∈ H 2 (X, Q),
and then after replacing ω by multiple one can assume that ω ∈ H 2 (X, Z). Now apply the
Lefschetz (1, 1)-theorem.) Such a manifold is sometimes called a Hodge manifold.
Exercise 7.8. Prove the local description of blowing up given in equation (7.14). Use this
to verify Lemma 7.25.
References