Taylor@Notes On Several Complex Variables 1997 PDF
Taylor@Notes On Several Complex Variables 1997 PDF
J. L. Taylor
Department of Mathematics
University of Utah
July 27, 1994
Revised June 9, 1997
J. L. TAYLOR
TABLE OF CONTENTS
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
Holomorphic Functions
Local Theory
A Little Homological Algebra
Local Theory of Varieties
The Nullstellensatz
Dimension
Completion of Local Rings
Sheaves
Sheaf Cohomology
Coherent Algebraic Sheaves
Dolbeault Cohomology
Coherent Analytic Sheaves
Projective Varieties
Algebraic vs. Analytic Sheaves Serres Theorems
Stein Spaces
Frechet Sheaves Cartans Theorems
The Borel-Weil-Bott Theorem
p.
p.
p.
p.
p.
p.
p.
p.
p.
p.
p.
p.
p.
p.
p.
p.
p.
3
10
15
24
30
46
54
63
71
87
99
106
116
124
136
149
166
1. Holomorphic Functions
There are a number of possible ways to define what it means for a function defined on a
domain in Cn to be holomorphic. One could simply insist that the function be holomorphic
in each variable separately. Or one could insist the function be continuous (as a function
of several variables) in addition to being holomorphic in each variable separately. The a
priori strongest condition would be to insist that a holomorphic function have a convergent
expansion as a multi-variable power series in a neighborhood of each point of the domain.
The main object of this chapter is to show that these possible definitions are all equivalent.
In what follows, (a, r) will denote the polydisc of radius r = (r1 , r2 , ..., rn ) about
a = (a1 , a2 , ..., an ):
(a, r) = {z = (z1 , z2 , ..., zn ) : |zj aj | < rj
j = 1, 2, ..., n}
and (a,
r) will denote the corresponding closed polydisc. Note that (a, r) is just the
f (z) =
1
2i
n Z
|n an |=rn
|1 a1 |=r1
i1 ,...,in =0
is compact in U we may choose numbers ri > si such that the compact polydisc (a, r)
with polyradius r = (r1 , . . . , rn ) is also contained in U . We substitute into the integrand
of 1.1 the series expansion:
X
1
(z1 a1 )i1 ...(zn an )in
=
(1 z1 )...(n zn ) i ,...,i =0 (1 a1 )i1 +1 ...(n an )in +1
1
J. L. TAYLOR
This series converges uniformly in and z for |zi ai | < si < ri = |i ai |. Thus, the
series in the integrand of 1.1 can be integrated termwise and the resulting series in z is
i1 ,...,in =0
ci1 ...in =
1
2i
n Z
|n zn |=rn
|1 z1 |=r1
n
1
ci1 ...in =
1
2i
n Z
|n zn |=rn
|1 z1 |=r1
From the obvious bounds on the integrand of this integral it follows that
|ci1 ...in | M r1 i1 ...rn in
and the Cauchy inequalities follow from this.
In what follows, the notation |U | will be used to denote the volume of a set U Cn .
=
1.4 Theorem (Jensens inequality). If f is holomorphic in a neighborhood of
(a, r) then
Z
1
log |f (a)|
log |f (z)|dV (z)
||
Proof. We recall the single variable version of Jensens inequality:
1
log |g(a)|
2
and apply it to f considered as a function of only z1 with the other variables fixed at
a2 , ..., an . This yields:
1
log |f (a)|
2
We now apply Jensens inequality to the integrand of this expression, where f is considered
as a function of z2 with z1 fixed at aa +1 ei1 and the remaining variables fixed at a3 , ..., an
to obtain:
log |f (a)|
1
2
2 Z
2
0
log |f (a)|
1
2
n Z
...
0
Finally, we multiply both sides of this by 1 ...n and integrate with respect to d1 ...dn
over the set {i ri ; i = 1, ..., n} to obtain the inequality of the theorem.
In the next lemma we will write a point z Cn as z = (z 0 , zn ) with z 0 Cn1 . Similarly
we will write a polyradius r as r = (r0 , rn ) so that a polydisc (a, r) can be written as
(a0 , r0 ) (an , rn ).
1.5 Lemma (Hartogs lemma). Let f be holomorphic in (0, r) and let
f (z) =
fk (z 0 )znk
be the power series expansion of f in the variable zn , where the fk are holomorphic in
(0, r0 ). If there is a number c > rn such that this series converges in (0,
c) for each
0
0
0
z (0, r ) then it converges uniformly on any compact subset of (0, r )(0, c). Thus,
f extends to be holomorphic in this larger polydisc.
0 , s0 ) (0, r0 ) and
Proof.. Choose any point a0 (0, r0 ), choose a closed polydisc (a
choose some positive b < rn . Then we may choose an upper bound M > 1 for f on the
J. L. TAYLOR
0 , s0 ) (0,
polydisc (a
b). It follows from Cauchys inequalities that |fk (z 0 )| M bk for
0 , s0 ). Hence,
all z 0 (a
k 1 log |fk (z 0 )| k 1 log M log b log M log b = M0
0 , s0 ) and all k. By the convergence of the above series at (z 0 , c) we also
for all z 0 (a
0 , s0 ) and hence:
have that |fk (z 0 )|ck converges to zero for each z 0 (a
lim sup k 1 log |fk (z 0 )| log c
k
0 , s0 ).
for all z 0 (a
We have that the functions k 1 log |fk (z 0 )| are uniformly bounded and measurable in
(a0 ,s0 )
(a0 ,s0 )
0 0
|(a , s )| log c
By replacing c by a slightly smaller number if neccessary we may conclude that there is a
k0 such that
Z
k 1 log |fk (z 0 )|dV (z 0 ) |(a0 , s0 )| log c
(a0 ,s0 )
for all k k0 . Now by shrinking s0 to a slightly smaller multiradius t0 and choosing small
enough we can arrange that
(a0 , t0 ) (w0 , t0 + ) (a0 , s0 )
for all w0 (a0 , ), from which we conclude that
Z
k 1 log |fk (z 0 )|dV (z 0 ) |(a0 , t0 )| log c + M0 |(w0 , t0 + ) (w0 , t0 )|
(w0 ,t0 +)
Now if we choose a c0 slightly smaller than c we may choose small enough that the right
hand side of this inequality is less than |(w0 , t0 + )| log c0 . This yields
Z
k 1 log |fk (z 0 )|dV (z 0 ) |(w0 , t0 + )| log c0
(w0 ,t0 +)
for all k k0 and w0 (a0 , ). This implies that our series is uniformly convergent
in (a0 , ) (0, c0 ) and since a0 is arbitrary in (0, r0 ) and c0 is an arbitrary positive
number less than c, we have that our power series serves to extend f to a function which
is continuous on (0, r0 ) (0, c) and certainly holomorphic in each variable. It follows
from Osgoods lemma that this extension of f is, in fact, holomorphic.
0 , r0 )
previous lemma, we write z = (z , zn ) for points of Cn and write (a,
r) = (a
n , rn ).
(a
Consider the sets
n , rn /2) :
Xk = {zn (a
|f (z 0 , zn )| k
0 , r0 )}
z 0 (a
For each k this set is closed since f (z 0 , zn ) is continuous in zn for each fixed z 0 . On the
0 , r0 ) for each
other hand, since f (z 0 , zn ) is also S
continuous in z 0 and, hence bounded on (a
n , rn /2)
zn , we have that (a
k Xk . It follows from the Baire category theorem that
for some k the set Xk contains a neighborhood (bn , ) of some point bn (an , rn /2).
We now know that f is separately holomorphic and uniformly bounded in the polydisc
n , ). It follows from Cauchys inequalities in each variable separately that
(a0 , r0 ) (b
the first order complex partial derivatives are also uniformly bounded on compact subsets
of this polydisc. This, together with the Cauchy-Riemann equations implies that all first
order partial derivatives are bounded on compacta in this polydisc and this implies uniform
continuity on compacta by the mean value theorem. We conclude from Osgoods theorem
that f is holomorphic on (a0 , r0 ) (bn , ) and, in fact, has a power series expansion
about (a0 , bn ) which converges uniformly on compact subsets of this polydisc.
Now choose sn > rn /2 so that (bn , sn ) (an , rn ). Then f (z 0 , zn ) is holomorphic in
zn on (bn , sn ) for each z 0 (a0 , r0 ) and, hence, its power series expansion about (a0 , bn )
converges as a power series in zn on (bn , sn ) for each fixed point z 0 (a0 , r0 ). It follows
from Hartogs lemma that f is actually holomorphic on all of (a0 , r0 ) (bn , sn ). Since
a is in this set and a was an arbitrary element of U , the proof is complete.
We introduce the first order partial differential operators
=
i
=
+i
zj
2 xj
yj
zj
2 xj
yj
Clearly, the Cauchy-Riemann equations mean that a function is holomorphic in each
variable separately in a domain U if and only if its first order partial derivatives exist in U
and zj f = 0 in U for each j. We may write these equations in a more succinct form by
using differential forms. We choose as a basis for the complex one forms in Cn the forms
dzj = dxj + idyj ,
d
zj = dxj idyj ,
j = 1, ..., n
, where
then the differential df of a function f decomposes as f + f
f =
X f
dzj ,
z
j
j
=
f
X f
d
zj
j
j
J. L. TAYLOR
1.8 Theorem. Every closed bounded subset of H(U ) is compact, where A H(U ) is
bounded if {||f ||K ; f A} is bounded for each compact subset K of U .
Proof. Since H(U ) is a metric space we need only prove that every bounded sequence in
H(U ) has a convergent subsequence. Thus, let {fn } be a bounded sequence in H(U ). By
the previous theorem, we need only show that it has a subsequence that converges uniformly
on compacta to a continuous function - the limit will then automatically be holomorphic
as well. By the Ascoli-Arzela Theorem, a bounded sequence of continuous functions on a
compact set has a uniformly convergent subsequence if it is equicontinuous. It follows from
the Cauchy estimates that {fn } has uniformly bounded partial derivatives on each compact
set. This, together with the mean value theorem implies that {fn } is equicontinuous on
each compact set and, hence, has a uniformly convergent subsequence on each compact
set. We choose a sequence {Km } of compact subsets of U with the property that each
compact subset of U is contained in some Km . We then choose inductively a sequence
of subsequences {fnm,i }i of {fn } with the property that {fnm+1,i }i is a subsequence of
{fnm,i }i and {fnm,i }i is uniformly convergent on Km for each m. The diagonal of the
resulting array of functions converges uniformly on each Km and, hence, on each compact
subset of U . This completes the proof.
of
1.
2.
3.
4.
5.
6.
10
J. L. TAYLOR
2. Local theory
Let X be a topological space and x a point of X. If f and g are functions defined
in neighborgoods U and V of x and if f (y) = g(y) for all y in some third neighborhood
W U V of x, then we say that f and g are equivalent at x. The equivalence class
consisting of all functions equivalent to f at x is called the germ of f at x.
The set of germs of complex valued functions at x is clearly an algebra over the complex
field with the algebra operations defined in the obvious way. In fact, this algebra can be
described as the inductive limit lim F (U ) where F (U ) is the algebra of complex valued
functions on U and the limit is taken over the directed set consisting of neighborhoods of
x. The germs of continuous functions at x obviously form a subalgebra of the germs of all
complex valued functions at x and, in the case where X = Cn , the germs of holomorphic
functions at x form a subalgebra of the germs of C functions at x which, in turn, form a
subalgebra of the germs of continuous functions at X.
We shall denote the algebra of holomorphic functions in a neighborhood U Cn by
H(U ) and the algebra of germs of holomorphic functions at z Cn by Hz or by n Hz in case
it is important to stress the dimension n. We have that Hz = lim H(U ) where the limit is
2.2 Proposition. The algebra n Oz is just the ring of fractions of the algebra C[z1 , ..., zn ]
with respect to the multiplicative set consisting of polynomials which do not vanish at z.
By a local ring we will mean a ring with a unique maximal ideal. It is a trivial observation
that:
2.3 Proposition. The algebras n Oz and n Hz are local rings and in each case the maximal
ideal consists of the elements which vanish at z.
The algebras n Oz and n Hz are, in fact, Noetherian rings (every ideal is finitely generated). For n Oz this is a well known and elementary fact from commutative algebra. We
will give the proof here because the main ingredient (Hilberts basis theorem) will also be
needed in the proof that n Hz is Noetherian.
We will use the elementary fact that if M is a finitely generated module over a Noetherian ring A then every submodule and every quotient module of M is also finitely
generated.
11
2.4 Theorem (Hilbert basis theorem). If A is a Noetherian ring then the polynomial
ring A[x] is also Noetherian.
Proof. Let I be an ideal in A[x] and let J be the ideal of A consisting of all leading coeficients of elements of I. Since A is Noetherian, J has a finite set of generators {a1 , ..., an }.
For each i there is an fi I such that fi = ai xri + gi where gi has degree less than ri .
Let r = maxi ri and let f = axm + g (deg P
g < m) be an element
degree m r.
P of I of
mri
We may choose bi , ..., bn A such that a = i bi Ai . then f i bi fi x
belongs to
I and has degree less than m. By iterating this process we conclude that every element
of I may be written as the sum of an element of the ideal generated by {f1 , ..., fn } and
a polynomial of degree less than r belonging to I. However, the polynomials of degree
less than r form a finitely generated A module and, hence, the submodule consisting of its
intersection with I is also finitely generated. A generating set for this A module, together
with {f1 , ..., fn } provides a set of generators for I as an A[x]-module. This completes the
proof.
The above result and induction show that C[z1 , ..., zn ] is Noetherian. This implies that
Oz is Noetherian as follows: If I is an ideal of Oz and J = I C[z1 , ..., zn ] then J generates
I as an Oz -module but J is finitely generated as a C[z1 , ..., zn ]-module. It follows that I
is finitely generated as an Oz -module. In summary:
2.5 Theorem. The polynomial algebra C[z1 , ..., zn ] and the local algebra n Oz are Noetherian rings.
We now proceed to develop the tools needed to prove that n Hz is Noetherian.
A holomorhic function defined in a neighborhood of 0 is said to be regular of order k in
zn at 0 provided f (0, ..., 0, zn ) has a zero of order k at 0.
2.6 Theorem. If f is holomorphic in a neighborhood U of 0 in Cn and regular of order
k in zn at 0, then there is a polydisc (0, r0 ) (0, rn ) such that for each z 0 (0, r0 ),
as a function of zn , f (z 0 , zn ) has exactly k zeroes in (0, rn ), counting multipicity, and no
zeroes on the boundary of (0, rn ).
rn ) occur at zn = 0.
Proof. Choose rn small enough that the only zeroes of f (0, zn ) on (0,
Set
= inf{|f (0, zn )| : |zn | = rn }
and choose r0 small enough that
|f (z 0 , zn ) f (0, zn )| < whenever z 0 (0, r0 ),
|zn | = rn
Then Rouches theorem implies that for each z 0 (0, r0 ) the functions f (0, zn ) and
f (z 0 , zn ) of zn have the same number of zeroes in the disc (0, rn ). This completes the
proof.
A thin subset of Cn is a set which locally at each point of Cn is contained in the zero
set of a holomorphic function.
12
J. L. TAYLOR
|n an |=rn
f (z 0 , n )
dn
n zn
is holomorphic in all of (a, r) and agrees with f off T by the one variable removable
singularity theorem and the Cauchy integral theorem.
A Weierstrass polynomial of degree k in zn is a polynomial h
n1 H0 [zn ]
of the form
n1 H0 .
k
Y
j=1
Now the functions bj (z 0 ) need not even be continuous because of the arbitrary choices
made in labeling the zeroes of f . However, the functions aj (z 0 ) are, in fact, holomorphic.
To see this, note that these functions are the elementary symmetric functions of the bj s
and these, in turn, may be written as polynomials in the power sums sm where
0
sm (z ) =
k
X
j=1
0 m
bj (z )
1
=
2i
Z
m
||=rn
f (z 0 , ) d
f (z 0 , )
These functions and, consequently, the aj (z 0 ) are holomorphic in (0, r0 ). Note that the
bj s all vanish at z 0 = 0 and, thus, so do the aj s. We now have that h is a Weierstrass
polynomial.
13
The proof will be complete if we can show that u = f /h is holomorphic and nonvanishing in (0, r). For each fixed z 0 (0, r0 ) the function f (z 0 , zn )/h(z 0 , zn ) is holomorphic in zn in (0, rn ) since numerator and denominator have exactly the same zeroes
in this polydisc. Furthermore, h is bounded away from zero on (0, r0 ) (0, rn ). This
and the maximum modulus principal imply that f /h is bounded on (0, r). Since it is
holomorphic in this set except where h vanishs, the removable singularities theorem (2.7)
implies that it extends to be holomorphic and non-vanishing in the entire polydisc. The
uniqueness is clear from the construction. This completes the proof.
2.9 Theorem (Weierstrass division theorem). If h n1 H[zn ] is a Weierstrass polynomial of degree k and f n H, then f can be written uniquely in the form f = gh + q
where g n H and q n1 H[zn ] is a polynomial in zn of degree less than k. Furthermore,
if f is a polynomial in zn then so is g.
Proof. Choose representatives of f and h (still call them f and h) which are defined in a
r) which is chosen small enough that h(z 0 , zn ) has exactly
neighborhood of a polydisc (0,
Z
||=rn
f (z 0 , ) d
h(z 0 , ) zn
Z
||=rn
f (z 0 , ) h(z 0 , ) h(z 0 , zn )
d
h(z 0 , )
zn
h(z 0 , ) h(z 0 , zn )
zn
is a polynomial in zn of degree less than k (with coeficients which are functions of ) which
shows that q is a polynomial in zn of degree less than k.
To show that this representation is unique, suppose we have two representations
f = gh + q = g1 h + q1
with q and q1 both polynomials of degree less than k in zn . Then q q1 = h(g1 g)
is a polynomial of degree less than k in zn with at least k zeroes for each fixed value of
z 0 (0, r0 ). This is only possible if it is identically zero.
Now if f itself is a polynomial in zn then the usual division algorithm for polynomials
over a commutative ring gives a representation of f as above with g a polynomial in zn .
The uniqueness says that this must coincide with the representation given above. This
completes the proof.
14
J. L. TAYLOR
2.
3.
4.
5.
2. Problems
A unique factorization domain is an integral domain in which each element has a unique
(up to units) factorization as a finite product of irreducible factors. Prove that if A is
a unique factorization domain then so is A[x].
Prove that C[z1 , ..., zn ] and n O0 are unique factorization domains.
Prove that n H0 is a unique factorization domain.
Prove Nakayamas lemma: If M is a finitely generated module over a local ring A with
maximal ideal m and if mM = M then M = 0. Hint: Prove that if M has a generating
set with k elements with k > 0 then it also a generating set with k 1 elements.
Prove the implicit function theorem: If f is holomorphic in a neighborhood of a =
f
(a0 , an ), f (a) = 0, and
(a) 6= 0 then there is a polydisc (a, r) = (a0 , r0 )(an , rn )
zn
and a holomorphic map g : (a0 , r0 ) (a, r) such that g(a0 ) = a and, for each
z (a, r), f (z) = 0 if and only if z = g(z 0 ) for some z 0 (a0 , r0 ). Hint: Use the
Weierstrass preparation theorem with k = 1.
15
y
y
y
0
M N1
M N2
M N3
M A N1
M A N2
M A N3
However, it is not generally true that this functor preserves exactness at the left stage.
The tensor product relative to A is an important functor and the fact that it is right
exact but not exact cannot be ignored. This circumstance requires careful analysis and
the development of tools that allow one to deal effectively with the problems it poses.
A similar problem arises with the functor homA . Again, if M and N are modules over A
then the space of linear maps from M to N is denoted hom(M, N ) while the A module of Amodule homomorphisms from M to N is denoted homA (M, N ). It can be described as the
kernel of the map hom(M, N ) hom(AM, N ) defined by {am a(m)(am)}.
The functor hom(M, ) is covariant and exact while the functor hom(, N ) is contravariant
and exact. On the other hand, the functors homA (M, ) and homA (, N ) are not generally
exact. It is true that they are both left exact, as is easily seen from the definition and a
diagram chase like the one above. Here again, the fact that homA (, N ) is not exact leads
to the need to develop tools to deal with the problems that this poses.
The first step in this program is to understand exactly when the two functors in question
are exact.
16
J. L. TAYLOR
3.1 Definition. An A-module M for which M A () is an exact functor from the category
of A-modules to itself is called a flat A-module.
An A-module M for which homA (M, ) is an exact functor is called a projective Amodule.
An A-module for which homA (, M ) is an exact functor is called an injective A-module.
Of course, each of these functors already satisfies two of the three conditions for exactness. Thus, M is flat if and only if for every injection i : N1 N2 the induced morphism
M A N1 M A N2 is an injection . Similarly, M is projective if and only if for every surjection N2 N3 the induced morphism homA (M, N2 ) homA (M, N3 ) is a surjection ( ie.
that every morphism from M to N3 lifts to N2 ). Finally, an A-module M is injective if and
only if for every injection N1 N2 the induced morphism homA (N2 , M ) homA (N1 , M )
is surjective (ie. every morphism from N1 to M extends to N2 ).
3.2 Proposition. Let M be an A-module. Then there are natural isomorphisms
A A M M
Proof. The map a m am : A M M has kernel which contains the image of the
map
a b m ab m a bm : A A M A M
P
P
In fact the
ai mi = 0, then
ai mi is the image under the latter
P two are equal since, if
map of
1 ai mi . It follows from the definition of A that a m am : A M M
induces an isomorphism from A A M to M .
The case of hom(A, M ) is even easier. Each element m M determines a homomorphism a am : A M and every homomorphism from A to M arises in this way from
a unique element m. In fact, m is just the image of 1 under the homomorphism. This
completes the proof.
A trivial consequence of this proposition is that the functors M A () and homA (M, )
are exact in the case where M = A. This is also clearly true if M is a finite direct sum of
copies of A. It is only slighltly less trivial that this continues to hold if M is an arbitrary
direct sum of copies of A. A module M which is a direct sum of an (arbitrary) set of
copies of A is called a free A-module . Thus, the functors M A () and homA (M, ) are
exact when M is any free A- module. Finally, it is an easy consequence of the definition
of direct summand that if M is a direct summand of a free A-module then M A () is an
exact functor. The same arguments applied to the functor homA (M, ) show that it is also
exact when M is a direct summand of a free A-module. Thus, direct summands of free
modules are both flat and projective.
Since A is an algebra over a field, an A-module M is free if and only if it has the form
M = A X where X is a vector space over the same field. Here, the module action is
given by the action of A on the left factor in the tensor product; that is, a(b x) = ab x.
Note, every A-module M is a quotient of a projective, in fact a free, A-module. In fact
the morphism
: A M M, (a m) = am
17
. . . Fn (M ) n . . . 2 F1 (M ) 1 F0 (M ) M 0
where Fn (M ) is the free A-module (n+1 A) M , (a m) = am, and
n (a0 an m) =
n1
X
i=0
18
J. L. TAYLOR
and s1 : M F0 (M ) by
s1 (m) = 1 m
Note that though the maps n are A-module homomorphisms the maps sn are only linear
maps and not A-module homomorphisms. A direct calculation shows that
sn1 n + n+1 sn = 1,
n1
and
s1 + 1 s0 = 1
and this is exactly what is meant by the statement that {sn } is a contracting homotopy
for the above complex. It is immediate that a complex with a contracting homotopy is
exact. We shall write F (M ) for the complex
n+1
. . . Fn (M ) n . . . 2 F1 (M ) 1 F0 (M ) 0
and
F (M ) M 0
for the corresponding resolution of M . To make sense of this notation, just think of
as being a morphism between two complexes, where M is thought of as the complex
whose only non-zero term is the degree zero term which is the module M . Finally, note
that the each of the functors Fn is, by construction, an exact functor from A-modules to
free A-modules and, hence, F is an exact functor from A-modules to complexes of free
A-modules.
To summarize the above discussion, we have
3.5 Proposition. The Hochschild functor functor M F (M ) is an exact functor from
A-modules to complexes of free A-modules and for each M
F (M ) M 0
is a free resolution of M .
We have used the term complex several times in this discussion. Actually two types of
complexes occur and it is time to be more precise. A chain complex C of A-modules is a
sequence of modules and morphisms of the form
n+2
n+1
n1
. . . Cn+1 Cn n Cn1 . . .
If C is a chain complex then its nth homology is
Hn (C) = ker n /im n+1
A cochain complex is just a chain complex with the modules indexed in increasing order
instead of decreasing order. Also, it is traditional to index cochain complexes with superscripts. Thus, a cochain complex C is a sequence of modules and morphisms of the
form
n2
n1
n+1
. . . C n1 C n C n+1 . . .
If C is a cochain complex then its nth cohomology is
H n (C) = ker n /im n1
n
We can now define the functors torA
n and extA .
19
A quick look at the first two terms of the complexes M A F (N ) and homA (F (M ), N )
shows that:
0
3.7 Proposition. torA
0 (M, N ) = M A N and extA (M, N ) = homA (M, N ).
torA
1 (M3 , N ) M1 A N M2 A N M3 A N 0
Again, the same result holds with the roles of M and N reversed.
Proof. If we take the tensor product of the short exact sequence 0 M1 M2 M3 0
with the Hochschild complex F (N ) for N we obtain a short exact sequence of complexes
0 M1 A F (N ) M2 A F (N ) M3 A F (N ) 0
That this is exact follows from the fact that F (N ) is a complex of free A-modules. A simple
diagram chase proves the standard result that every short exact sequence of complexes
induces a long exact sequence of the corresponding homology. This completes the proof.
The same sort of arguments yield analogous results for ext:
3.9 Theorem. Let 0 M1 M2 M3 0 be a short exact sequence of A-modules
and let N be any A-module. Then there are natural long exact sequences
0 homA (M3 , N ) homA (M2 , N ) homA (M1 , N ) ext1A (M3 , N )
extnA (M3 , N ) extnA (M2 , N ) extnA (M1 , N ) extn+1
A (M3 , N )
and
0 homA (N, M1 ) homA (N, M2 ) homA (N, M3 ) ext1A (N, M1 ) . . .
extnA (N, M1 ) extnA (N, M2 ) extnA (N, M3 ) extn+1
A (N, M1 )
20
J. L. TAYLOR
3.10 Theorem. Let M be a module. Then M is flat if and only if tornA (M, N ) = 0 for all
n > 0 and all modules N . Of course, the same statement holds with M and N reversed.
The module M is projective if and only if extnA (M, N ) = 0 for all n > 0 and all modules
N.
The module M is injective if and only if extnA (N, M ) = 0 for all n > 0 and all modules
N.
Proof. That tornA (M, N ) = 0 for all N and all n > 0 if M is flat follows from the fact that
M A () preserves exactness and, in particular, preserves the exactness of the Hochschild
resolution. The reverse implication follows from the long exact sequence for tor. The
proofs of the results for ext are completely analogous.
The following results are trivial consequences of Theorem 3.10 and the existence of the
long exact sequences:
3.11 Theorem. If 0 K P M 0 is a short exact sequence of modules with P
projective and N is any module, then there are natural isomorphisms
A
(M, N ) ' tornA (K, N ) n > 0
torn+1
p1
and
torn+1 (Kq , N ) ' torn (Kq+1 , N ) n 0, q 0
By induction, this yields
torp (M, N ) = torp (K1 , N ) ' tor1 (Kp2 , N ) ' Hp (P N )
A similar argument works for ext. This leads to:
21
3.12 Theorem. Given any projective resolution P M , as above, and any module N
there are natural isomorphisms
tornA (M, N ) ' Hn (P A N ) and extnA (M, N ) ' H n (homA (P, N ))
The result for tor holds if P is just a flat resolution of M .
Finally, we return to the study of the algebras n O and n H with a result which shows
that they have particularly simple free resolutions. First, note that if A is Noetherian then
any finitely generated A-module has a resolution by free finitely generated modules, that
is, a resolution of the form
n+1
n1
gi fi = 0
with some gj a unit. Then the set of generators of Kn1 could be reduced by throwing
out the corresponding fj . We shall show that this implies that Kn = In Kn which, by
Nakayamas lemma, shows that Kn = 0, as desired.
The fact that Kn In Hs implies Kn = In Kn follows immediately from the case
k = j = n of the equality
Kp I j H k p = I j Kp
for 1 j p
which we shall prove by induction on j. We need only prove that Kp Ij Hkp Ij Kp since
the reverse containment is clear.
22
J. L. TAYLOR
23
3. Problems
1. Prove that if X is a vector space and hom(A, X) is given the obvious A-module structure,
then hom(A, X) is an injective A-module. Show that every A module is a submodule of
a module of this form. Then prove that a module is injective if and only if it is a direct
summand of a module of the form hom(A, X).
2. Fix C. Find a resolution of the form given in corollary 3.14 for the one dimensional
C[z]-module, C , on which each p C[z] acts as multiplication by the scalar p().
3. If V is any vector space and L end(V ) is any linear transformation, then we can make
V into a C[z]-module VL by letting p C[z] act on V as the linear transformation p(L).
A
Show that torA
1 (C , VL ) = ker ( L) and tor0 (C , VL ) = coker ( L). Thus, L
is invertible if and only if both of these tor groups vanish.
4. Prove a result analogous to the result of problem 3 but with tor replaced by ext.
5. Verify (if you have never done so before) that a short exact sequence of complexes yields
a long exact sequence of homology.
6. Show that each non-zero element of ext1A (M, N ) corresponds to a non-trivial extension
of M by N , that is, to a short exact sequence
0N QM 0
which does not split.
24
J. L. TAYLOR
V1 V2 = id V1 id V2 .
I1 I2 = loc I1 loc I2 .
V = loc id V .
I id loc I but they are not generally equal.
id(V1 V2 ) = (id V1 ) (id V2 ) (id V1 ) (id V2 ).
id(V1 V2 ) (id V1 ) + (id V2 ).
25
26
J. L. TAYLOR
with
coordinate
functions f1 , . . . , fm then the Jacobian of F is the m n matrix JF (z) =
fi
(z) .
zj
4.5 Theorem (Implicit mapping theorem). Let F be a holomorphic mapping as
above and suppose U and F () = 0. Suppose also that the last m columns of JF ()
form a non-singular m m matrix. Then there is a polydisc
(; r) = (0 ; r0 ) (00 ; r00 ) Cnm Cm
and a holomorphic map G : (0 ; r0 ) (00 ; r00 ) such that G(0 ) = 00 and F (z) = 0 for
z = (z 0 , z 00 ) (; r) if and only if G(z 0 ) = z 00 .
Proof. When m = 1 this is the implicit function theorem which is a simple corollary of the
Weierstrass preparation theorem in the case where the function is regular of degree one in
its last variable. We prove the general case by induction on m. Thus, we assume that the
result is true for m 1 and proceed to prove it for m.
Let JF () = (JF0 (), JF00 ()) be the separation of JF () into its first n m columns and
its last m columns. The hypothesis is that JF00 () is non-singular. By a linear change of
variables in the range space Cm we may assume that JF00 () is the m m identity matrix.
Then, since fm /zn = 1 at , it follows from the implicit function theorem that there is
a polydisc (; r) and a holomorphic mapping
h : (1 , . . . , n1 ; r1 , . . . , rn1 ) (n , rn )
such that h(1 , . . . , n1 ) = n and fm (z) = 0 for z (; r) exactly when zn =
h(z1 , . . . , zn1 ). Then we may define a holomorphic mapping
F 0 : (1 , . . . , n1 ; r1 , . . . , rn1 ) Cm1
0
to be
by defining its coordinate functions f10 , . . . , fm1
27
28
J. L. TAYLOR
29
polynomial of degree k plus an element of the k th power of the maximal ideal of m H , and
since each of F and maps the k th power of the maximal ideal of m H to the k th power
) belongs to the k th power
of the maximal ideal of V H , we conclude that F (f ) (f
of the maximal ideal of V H for every f m H and every positive integer k. However,
by Nakayamas lemma the intersection of all powers of the maximal ideal in a Noetherian
4. Problems
1. Give an example which shows that the implicit function theorem, the inverse mapping
theorem and the Weierstrass preparation theorem fail in the algebraic case.
2. Consider the polynomial on C2 defined by p(z, w) = z 2 w3 . Prove that p is irreducible
in both 2 H and 2 O and, hence, generates a prime ideal in each algebra.
3. Show that in either 2 H or 2 O the ideal generated by the polynomial p in problem 2 is
id V where V = {(z, w) C2 : p(z, w) = 0.
4. With V as in problem 2, show that there are irreducible elements f, g V H (V O) such
that f 2 = g 3 . Conclude that these algebras are not unique factorization domains.
5. With V as in problem 2, let M be the maximal ideal of V H. Show that M is generated
by two of its elements in such a way the the resulting morphism V H2 M has kernel
isomorphic to M as a V H -module. Conclude that the Hilbert syzygy theorem fails to
hold for V H.
30
J. L. TAYLOR
5. The Nullstellensatz
In rings A for which it is true, the Nullstellensatz says that
ideal is an ideal I whose radical I is prime and, in a Noetherian ring, each ideal I has
a primary decomposition I = m
j=1 Ij . Thus, if we assume the Nullstellensatz for prime
ideals, then we have
id loc I = id(
m
[
m
loc Ij ) = m
j=1 id loc Ij j=1 id loc
p
p
Ij = m
I
=
I
j
j=1
j=1
which is the Nullstellensatz for general ideals since we already have the reverse containment.
Our goal in this section is to prove the Nullstellensatz for the ring n H. However, we
first prove the much easier result that the Nullstellensatz holds for n O. It is easy to see
that the Nullstellensatz for C[z1 , . . . , zn ] implies the Nullstellensatz for n O (this is left as
an exercise) so we shall prove the Nullstellensatz for C[z1 , . . . , zn ]. We first need a couple
of lemmas from commutative algebra.
5.1 Lemma. Let A B C be rings with A Noetherian. Suppose that C is a finitely
generated A-algebra and C is integral over B. Then B is also a finitely generated A
algebra.
Proof. We first show that C is actually a finitely generated B-module. The fact that C is
finitely generated as an A-algebra means that it is also finitely generated as a B-algebra
and, hence, that every element of C is a polynomial in a finite set of generators c1 , . . . , ck
with coeficients in B. However, the fact that C is integral over B means that for each
c C there is an integer nc such that every polynomial in c is equal to one of degree
less than or equal to nc . Thus, the algebra generated over B by c1 is a finitely generated
B-module. An induction argument on the number of generators now shows that C is a
finitely generated B-module as claimed.
Now let x1 , . . . , xl generate C as an A-algebra and y1 , . . . , ym generate C as a B-module.
Then there exist elements bij and bijk in B such that
X
X
xi =
bij yj , yi yj =
bijk yk
j
Let B0 be the algebra generated over A by the bij and the bijk . We have A B0 B and
B0 is Noetherian since A is Noetherian and B0 is a quotient of a polynomial ring over A.
The above equations show that each element of C is a linear combination of the yi with
coeficients from B0 so that C is a finitely generated B0 -module. Since B0 is Noetherian and
B is a submodule of C it follows that B is also a finitely generated B0 -module. Since B0 is
finitely generated as an A-algebra it follows that B is finitely generated as an A-algebra.
This completes the proof.
31
implies that h1 k k[x1 , . . . , xj ] where k is some product of the gi0 s, which is impossible if
h is relatively prime to all the gi0 s in k[x1 , . . . , xj ]. The resulting contradiction shows that
{x1 , . . . , xj } must be empty. But in this case A/M is algebraic over k and, hence, equal
to k since k is algebraically closed.
5.3 Theorem (Nullstellensatz for polynomial algebras). If A = C[z1 , . . . , zn ] then
for each ideal I A we have
I = id loc I
Proof. We may assume I is prime. We know that I id loc I and so we need only prove
the reverse containment. Let f A be any element not in I and let B = A/I and C = Bf ,
the algebra of fractions over B with denominators that are powers of the image of f in
B. Now let M be a maximal ideal of C. Since C is finitely generated over C it follows
from the previous lemma that C/M = C. Then the images of z1 , . . . , zn in C/M are the
coordinates of a point Cn . It is clearly a point in loc I (since the maximal ideal it
determines contains I by construction) and a point at which f () 6= 0 (since f is invertible
in C). Thus, f is not in id loc I and the proof is complete.
The above proof depended heavily on the fact that quotients of C[z1 , . . . , zn ] are finitely
generated algebras over the ground field. We have no such finite generation conditions in
the holomorphic case and must use entirely different methods. The proof of the Nullstellensatz for n H depends on a fairly precise description of the locus of a prime ideal as the
germ of a certain kind of finite branched cover of a neighborhood in Cm . We now proceed
to develop this description.
In what follows we will be making fairly heavy use of field theory. Since this discussion
will not make much sense to someone who doesnt know a certain amount of field theory,
we present below a list of facts from this subject that we will use implicitly or explicitly.
This is presented as a study guide for those who need to brush up on the subject. Proofs
can be found in any algebra book with a good treatment of field theory (eg: Hungerford).
If K F are fields and x F then x is called algebraic over K if it is the root of a
polynomial with coeficients in K. The field F is called an algebraic field extension of K if
each of its elements is algebraic over K.
32
J. L. TAYLOR
Y
(xi xj )
i6=j
33
n1
xn
xn2
1
xn2
2
xn2
n
...
...
...
...
...
x1
x2
xn
A ring is said to be integrally closed if it is integrally closed in its quotient field. That
is, if whenever a monic polynomial with coeficients in the ring has a root in the quotient
field of the ring then that root actually lies in the ring.
F9 Theorem. A unique factorization domain is integrally closed.
The next result may be less well known than its predecessors and so we shall actually
prove it.
F10 Theorem. If A is a unique factorization domain with quotient field K of characteristic zero and if F = K(x) is the field extension generated by an element x, integral over
A, with minimal polynomial p of degree n, then every element of F which is integral over
xn1
x 1
A belongs to the A submodule of F generated by the elements
, . . . , , , where d is
d
d d
the discriminant of p.
Proof. For an element f (x) K(x), integral over A, we wish to find coeficients a0 , . . . , an1
such that
an1 xn1 + an2 xn2 + + a1 x + a0 = d f (x)
Let x1 = x and let x2 , . . . , xn be the other roots of p in a splitting field for p. We may then
write down a system of n equations, each of which is a copy of the one above but with x
replaced by xj in the j th equation. If we consider this as a system of equations in which
the unknowns are the elements a1 , . . . , an , Kramers rule give as solution
n1
x1
n1
x2
aj = d
n1
xn
xn2
1
xn2
2
n2
xn
...
...
...
...
...
x1
x2
xn
1
1 xn1
1
1 xn1
2
n1
1
xn
xn2
1
xn2
2
n2
xn
...
...
...
...
...
f (x1 )
f (x2 )
f (xn )
...
...
...
...
...
x1
x2
xn
where in the second determinant the f (xi ) replace the j th column of the first determinant.
Now, of course, the first determinant is the Vandermonde determinant which has square
equal to d by F8. Thus, aj is the product of the Vandermonde and the determinant
obtained from the Vandermonde by replacing its j th column with the column formed by
the f (xi ). Clearly this product is left fixed by any permutation of the roots x1 , . . . , xn
since this just amounts to applying the same permutation to the rows in both matrices.
Thus, the elements aj so determined are fixed by the Galois group of the splitting field of
p and belong to K by F5. However, since x and f (x) are integral over A so are all the xi
34
J. L. TAYLOR
and f (xi ), again by F5. It follows that the aj are also integral over A since they lie in the
ring generated by the xi and f (xi ). However, A is integrally closed in its quotient field K
by F9 and, hence, aj A for j = 1, . . . , n. This completes the proof.
We now return to our study of a prime ideal P n H.
Recall that a holomorphic function f in a neighborhood of 0 is called regular in zn if
f (0, . . . , 0, zn ) is not identically zero. In what follows, we consider j H for j n to be the
subring of n H consisting of functions that depend only on the first j variables.
5.4 Definition. An ideal I n H is called regular in the variables zm+1 , . . . , zn if m HI =
0 and for each j {m + 1, . . . , n} there is an element fj j H I which is regular in zj .
5.5 Lemma. For each non-zero ideal I n H there is a choice of a complex linear coordinate system for Cn and an m < n such that I is regular in the variables zm+1 , . . . , zn .
Proof. We can choose a non-zero fn I and then by a suitable linear change of coordinates
arrange that fn is regular in zn . Suppose we have chosen fj+1 , . . . , fn satisfying the
conditions of Definition 5.4. Then either j H I = 0, in which case we are through, or
there is a nonzero fj j H I. The function fj can be made regular in zj by a linear
change of coordinates that involves only the first j coordinates and, hence, does not effect
the regularity of the functions chosen previously. The Lemma follows by induction.
The notion of an ideal being regular in the variables zm+1 , . . . , zn seems to depend on
the ordering of these variables. However, the next lemma shows that it depends only on
the decomposition Cn = Cm Cnm and not on the choice of coordinate systems within
the two factors.
5.6 Lemma. An ideal I n H is regular in the variables zm+1 , . . . , zn if and only if m H
in n H
= n H/I and n H
is an integral algebraic extension of
is isomorphic to its image m H
j for
n H is obtained from m H by successive integral algebraic extensions by the elements z
j = m + 1, . . . , n. The theorem of transitivity of integral dependence now implies that
35
We can do even better in the case that our ideal, regular in zm+1 , . . . , zn , is a prime
= n H/P is an integral domain and has a field of quotients n M.
all roots in a splitting field of pj are also integral over m H by the transitivity of the Galois
group and, hence, the coeficients of pj , being elementary symmetric functions of the roots,
Since the latter algebra is a unique factorization domain and, hence,
are integral over m H.
and ,
integrally closed, we conclude that the coeficients of each pj are actually in m H
hence, may be considered elements of m H. Then each polynomial pj (zj ) is an element of
m H[zj ] n H which belongs to P since it vanishes mod P . The pj (zj ) are some of the
elements of P that we are seeking.
We use F10 to choose the remaining elements. For j = m + 2, . . . n the image zj of zj in
and
d I I 0 I
36
J. L. TAYLOR
Proof. The polynomial pn is a Weierstrass polynomial and so, for any f n H, the Weierstrass division theorem allows us to write f = pn gn + rn for some gn n H and some
rn n1 H[zn ] of degree less than the degree of pn . We now apply the Weierstrass division theorem to each coeficient of rn with the divisor being the Weierstrass polynomial
pn1 m H[zn1 ] n2 H[zn1 ]. If we gather together the terms this yields
f = pn gn + pn1 gn1 + rn1
with gj n H and rn1 n2 H[zn1 , zn ]. By repeating this argument as long as we have
pj s we eventually get
f = pn gn + + pm+1 gm+1 + rm+1
with gj n H and rm+1 m H[zm+1 , . . . , zn ]. Also note that the degree of rj is less than
the degree of pj for j = m + 1, . . . , n.
Now for each j = m + 1, . . . , n we have d zj = qj + sj and so, by the binomial theorem,
dk zjk = hjk qj + skj for each integer k and some hjk
P n H. If we apply this to each power
of each zj appearing in rm+1 and if we choose = (degree pj 1) then we conclude that
d rm+1 = qm+2 hm+2 + + qn hn + r0
for some hj n H and r0 m H[zm+1 ]. Another application of the Weierstrass division
theorem gives us r0 = pm+1 hm+1 + r where hm+1 n H and r m H[zm+1 ] with degree of
r less than the degree of pm+1 . Finally, this gives us
d f =
n
X
j=m+1
d gj pj +
n
X
hj qj + hm+1 pm+1 + r
j=m+2
37
38
J. L. TAYLOR
With : V W as in part (b) of the above proposition and q V , we know that there
are arbitrirly small neighborhoods of q on which is a finite branched holomorphic cover
of pure order. For small enough neighborhoods the order must stabilize at some positive
integer oq (). We call this integer the branching order of at q.
5.13 Lemma. Suppose pm+1 , . . . , pn are monic polynomials with coefficients in m HU for
some open set U Cm and with non-zero discriminants dm+1 , . . . , dn and let V = {z
Cn : pj (zj ) = 0, j = m + 1, . . . n}. Then the projection Cn Cm exhibits V as a finite
branched holomorphic cover of U .
Proof. We let D be the union of the zero sets of the discriminants dj . Then U0 = U D
is an open dense subset of U and we set V0 = 1 (U0 ) where : V U is the restriction
of the projection Cn Cm to V . We need to show that V0 is dense in V , that is proper
and finite to one on V and locally biholomorphic on V0 .
Now let K U be compact and set L = 1 (K). We claim that L is a bounded subset
of Cn . Clearly the j th components of points of L are bounded if j m since then they are
j th components of points of K. For j > m the j th components of points of L are bounded
because they satify the monic polynomial equation pj (zj ) = 0. If we divide this equation
n 1
by zj j , where nj is the order of pj , we may use the resulting equation to estimate |zj | in
terms of the coefficients of pj , on the set where |zj | 1. We conclude that |zj | is bounded
by the maximum of 1 and the sum of the suprema of the absolute values of the coefficients
of pj on K. Thus, the set of j th components of points of L is a bounded set for all j and,
therefore, L is bounded. The set L is also closed in Cn since it is just the set of points
in Cn which map to K under : Cn Cm and at which each pj vanishes. Hence, L
is compact. Thus, is a proper holomorphic map. The fact that each pj has a finite,
non-empty set of zeroes for each fixed value of (z1 , , zm ) shows that it is finite to one
and surjective.
Let a = (a0 , a00 ) be a point of V U Cnm with a0 the corresponding point of U .
Each polynomial pj is monic and, hence, pj (zj ) is regular of some order greater than zero
in the variable zj at a. Given > 0, Theorem 2.6 implies that we may choose a polydisc
0 , r0 ) (a
00 , r00 ) U Cnm , with r00 < , such that for each j and for
(a,
r) = (a
j
r0 ) the roots of pj (z 0 , zj ) all lie in the interior of (a
00 , r00 ) and the number of
z 0 (0,
j
0 , r0 ). This is a kind
these roots is constant, counting multiplicity, as a function of z 0 (a
of continuity of the roots result. It implies in particular that for each point of V there are
arbitrarily nearby points that lie over points of the open dense set U0 . In other words, V0
is dense in V .
Since U0 is exactly the set of z 0 U at which the roots of all the pj are distinct, the
inverse image V0 of this set under is the subset of U Cnm on which each pj vanishes
but its derivative with respect to zj does not vanish. Thus, the map F : Cn Cnm
defined by
F (z1 , , zn ) = (pm+1 (z1 , , zm , zm+1 ), , pn (z1 , , zm , zn ))
has Jacobian JF in which the last n m columns form a diagonal matrix with entries that
do not vanish on V0 . Thus, JF has rank n m in an open set containing V0 . It follows
from the implicit mapping theorem that, for each V0 , there is a neigborhood A of
39
with j 6= k
can be satisfied. By the choice of q, there must be at least one index i between 1 and q such
that no such equation holds. For this i, the functional fi separates the points z1 , , zr .
5.15 Lemma. Let : V W be a finite branched holomorphic cover where W is a
domain in Cm and let D be a proper subvariety of W , with the property that V0 =
V 1 (D) is dense in V and is locally biholomorphic on V0 . If V1 is a connected
component of V0 , then the closure V1 of V1 in V is a holomorphic subvariety of V .
Proof. The lemma is purely a local statement and so we may assume that W is a polydisc
in Cm and that V is a closed subvariety of some polydisc in Cn .
The set W0 = W D is connected (by Problem 5.2) and V0 has finitely many components, each of which is a finite unbranched holomorphic cover of W0 and one of which
is V1 . Let the points of V1 over a point w W0 be labeled 1 (w), . . . , k (w) and for a
function f holomorphic on V let p be the polynomial in the indeterminant x defined by
p(w, x) =
k
Y
(x f (j (w)))
j=1
40
J. L. TAYLOR
to do this globally. However, since the coefficients of the polynomial p are independent of
the labeling of the roots, they are well defined and holomorphic in all of W0 . In fact, if f is
holomorphic in all of V , it is locally bounded there, which implies that the coefficients of
p are as well. The generalized removable singularities theorem (Theorem 2.7) then implies
that they extend to be holomorphic in all of W . Now we have that p is a polynomial
with coefficients holomorphic in W and with the property that, whenever w W0 , the
roots of p(w, x) are exactly the values assumed by f on the set 1 (w) V1 . In particular,
p((z), z) vanishes on V1 . By continuity, this function also vanishes on V1 .
Since W0 is connected, is a cover of pure order r for some r. We apply the previous
lemma to obtain linear functionals f1 , , fq such that any set of r distinct points in
Cn can be separated by some one of the functionals fi . We then let p1 , , pq be the
polynomials constructed, as above, for the functions f1 , , fq . Each of the functions
pj ((z), fj (z)) vanishes on V1 and so we let V be the subvariety of V on which they all
vanish. We endeavor to prove that V1 = V . To this end, let a1 be a point of V and let
a1 , a2 , , ak be the distinct points of 1 ((a1 )). Note that k r and so there is an i for
which the numbers fi (a1 ), , fi (ak ) are all distinct. These are all roots of the polynomial
pi ((a1 ), x). Then for any w W0 , sufficiently near (a1 ), there must be a root of pi (w, x)
near fi (a1 ). But by the construction of pi this root must be a value assumed by fi on
1 (w) V1 . It follows that for any sequence {ws } W0 , converging to (a1 ) there is a
sequence {zs } V1 with (zs ) = ws and with fi (zs ) converging to fi (a1 ). Since is a
proper map, the sequence {zs } must have a limit point z V1 . Necessarily fi (z) = fi (a1 )
and, hence, z = a1 since fi separates the points a1 , , ak . Thus, a1 V1 and V1 = V .
This completes the proof.
The branch locus of a finite branched holomorphic cover is the set on which the branching order is at least 2.
5.16 Proposition. If : V W is a finite branched holomorphic cover over a domain
W in Cm and k is any positive integer, then the subset of V on which has branching
order at least k is a closed subvariety of V . In particular, the branch locus of is a closed
subvariety of V . Furthermore, the image under of the branch locus is a closed subvariety
of W .
Proof. With V0 and W0 as above and V1 = V0 we choose linear functionals f1 , , fq and
polynomials p1 , , pq as in the proof of the previous lemma. Then for any w W the
roots of the polynomial pj are exactly the values that fj assumes on 1 (w). A root is
(s)
a multiple root of pj of multiplicity k if and only if the polymomials pj vanish there for
s k. It follows that 1 (w) is a point of branching order at least k if and only if the
(s)
functions pj ((z), fj (z)) vanish at for s k and for all j. This proves the first part of
the Proposition. The second part follow from the fact that the set where the branching
order is at least two is the set where all the pj have multiple roots and this is the set
where the discriminants of all the polynomials pj vanish. The discriminants are functions
of w W only and the result follows.
5.17 Theorem. Let P n H be a prime ideal which is strictly regular in the variables
zm+1 , . . . , zn . Then the projection Cn = Cm Cnm Cm exhibits loc P as the germ of
41
a finite branched holomorphic cover of pure order r over Cm , where r is the order of the
defining polynomial for zm+1 .
Proof. We let the polynomials pm+1 , . . . , pn , qm+2 , . . . , qn be as in Lemma 5.9. Choose a
polydisc centered at 0 on which each of these germs has a representative and on which
each element of a generating set for P (containing the p0 s and q 0 s) also has a representative.
We let = 0 00 be the usual decomposition induced by representing Cn as Cm Cnm .
We replace each pj and each qk by its representative on . We let V be the locus of common
zeroes of the set of representatives on of our generating set for P . Then the germ of
V at 0 is loc P . We let W be the zero set in Cm+1 of pm+1 (zm+1 ). We also let D
be the zero set in of the discriminant d of pm+1 . Since d depends only on the first m
variables, we have D = D0 00 where D0 is the zero set in 0 of d. By Lemma 5.13 the
projection Cm+1 = Cm C Cm induces a holomorphic branched covering of pure order
r, 2 : W 0 . Also, if W0 = W (D W ) then W0 is the set on which this covering is
regular (locally biholomorphic).
Another finite branched holomorphic covering is obtained by applying Lemma 5.13 to
the set
V1 = {z : pm+1 (zm+1 ) = 0, . . . , pn (zn ) = 0}
and the map 1 : V1 0 induced by the canonical projection Cn = Cm Cnm Cm .
This is regular on the set V1 (E V1 ), where E = E 0 00 is the union of the zero sets of
the discriminants of pm+1 , , pn . We obtain the covering map that we are looking for
by restricting 1 to the subvariety V V1 . It remains to show that is a finite branched
holomorphic cover of pure order r.
It is clear that is a finite to one, proper, holomorphic map. The density of the
regular points and pure order r are not yet clear. We will exploit the fact that factors
as = 2 3 where 2 is the finite branched holomorphic cover introduced above and
3 : V W is induced by the canonical projection Cn Cm+1 . Note that 3 is also the
germ of a proper finite to one holomorphic map. We let V0 = V (E V ). We will show
that V0 is a dense open subset of V on which is locally biholomorphic. First note that,
by corollary 5.10, since D E, the germ of V0 at 0 is loc I 0 (E loc I 0 ). Thus, we may
choose small enough that (z 0 , zm+1 , . . . , zn ) is in V0 exactly when z 0
/ E 0 and
pm+1 (z 0 , zm+1 ) = 0
qj (z 0 , zm+1 , zj ) = d(z 0 )zj sj (z 0 , zm+1 ) = 0
and
for j = m + 2, . . . , n.
42
J. L. TAYLOR
43
id loc J which is J by the Nullstellensatz. This implies that zn J for some . In other
44
J. L. TAYLOR
that In1 id Vn1 I n1 and from this that loc In1 = Vn1 . Thus, In1 satisfies
condition (iii) as desired. Either m = n 1 or we may now conclude as above that there
is fm1 In1 = n1 H I which is regular in zn1 . Clearly we can repeat this procedure
n m times to achieve condition (i). This completes the proof.
We say that a germ : V W of a holomorphic map between two germs of varieties
is finite if 1 (0) = (0).
5.22 Theorem. A germ : V W of a holomorphic mapping between two germs of
holomorphic varieties is finite if and only if for each irreducible component Vi of V , the
image (Vi ) is the germ of a holomorphic subvariety of W and : Vi (Vi ) is a finite
branched holomorphic cover.
Proof. Suppose is finite. We may represent V and W by subvarieties of neighborhoods
in Cn and Cm (m n) in such a way that the projection Cn Cm induces , by Problem
8. Then 1 (0) = L V where L is the germ of {z Cn : z1 = = zm = 0}. Since
is finite we have L V = 0 . It then follows from the previous theorem that for each
irreducible component Vi of V , : Vi (Vi ) is the germ of a finite branched holomorphic
cover of the germ (Vi ) of a subvariety. The reverse implication is obvious and so the proof
is complete.
5. Problems
1. Assuming the Nullstellensatz for C[z1 , . . . , zn ], prove it for n O.
2. Prove that if U Cn is a connected open set and D U is a closed subvariety, then
U D is also connected.
3. Prove that if P n H is a prime ideal which is just regular in the variables zm+1 , . . . , zn ,
then the projection Cn = Cm Cnm Cm still exhibits loc P as a finite branched
holomorphic cover of pure order r over Cm for some r.
4. In a finite branched holomorphic cover : V W prove that each point where the
branching order is one is a regular point of the variety V - that is, a point which has a
neighborhood in V biholomorphic to a polydisc in Cm for some m.
5. Prove that if V is an irreducible germ of a variety at 0 Cn then V has a representative
in some neighborhood of 0 for which the set of regular points is connected.
6. Prove the Nullstellensatz for the local ring V H of the germ of a holomorphic variety.
45
46
J. L. TAYLOR
6. Dimension
We continue with our study of the local properties of regular and holomorphic functions
and algebraic and holomorphic varieties. In this section we will be concerned with a germ
V of an algebraic or holomorphic variety and with the local ring V O or V H. In each case
there are three notions of dimension of the local ring: a topological dimension, a geometric
dimension and a tangential dimension. We show that the first two agree and that they
agree with the third if and only if the variety is regular at the point in question. We shall
use this last result to prove that the singular locus of a variety is a subvariety. Here the
regular locus of a holomorphic variety is the set at which it is locally biholomorphic to
a polydisc in complex Euclidean space and the singular locus is the complement of the
regular locus. The regular and singular locus of an algebraic variety have not yet been
defined and, in fact, the usual definition is that the regular locus is the set where the
second and third of the three dimensions, referred to above, are equal.
We initially take as our definition of the dimension of a holomorphic variety the topological definition:
6.1 Definition. If V is a holomorphic variety then the dimension of V is the dimension
of the complex manifold that is the regular locus of V .
Note that if V then the dimensions of smaller and smaller neighborhoods of in V
eventually stabilize and so it makes sense to talk about the dimension of the germ of V at
. Also note that if V is not irreducible then its regular locus may decompose into germs
of manifolds of different dimensions. In this case, by dimension we mean the maximal
dimension that occurs. Thus, if V = V1 V2 Vk is the decomposition of a germ of
a variety into its irreducible components, then the dimension of V is the maximum of the
dimensions of the Vj .
6.2 Lemma. If V is the germ of a variety in Cn and I = id V is regular in the variables
zm+1 , . . . , zn then dim V = m.
Proof. If V is irreducible then this is an obvious corollary of Theorem 5.17. If V is not
irreducible, let V = V1 . . . Vr be a decomposition of V into irreducibles. By Theorem
5.21, the projection : Cn Cm determines a finite branched holomorphic cover of each
Vi onto a germ of a subvariety of Cm . The dimension of Vi is the same as the dimension of
(Vi ) and this is less than or equal to m for each i. To complete the proof, we must show
that one of these dimensions is equal to m. If not, then, for each i, (Vi ) is a germ of a
proper subvariety of Cm and, hence, there
is a non-zero element gi m H which vanishes
on (Vi ). Then g = g1 g2 . . . gr id V = I and so some power of g is a non-zero element
of I m H which violates the assumption that the ideal I is regular in zm+1 , . . . , zn . This
completes the proof.
6.3 Theorem. If V and W are germs of holomorphic varieties with V W then dim V
dim W and the two are equal exactly when V and W have a common irreducible component
of dimension dim W .
Proof. Since V W , each irreducible component of V is contained in some irreducible
component of W . Thus, the theorem can be reduced to the case where V and W are
47
48
J. L. TAYLOR
49
Proof. Note that it follows from the previous theorem that if W 0 W 00 are germs of
irreducible subvarieties of V and dim W 00 dim W 0 2 then there is another irreducible
subvariety W 000 which is properly contained in W 00 and properly contains W 0 . Indeed, there
must be a germ f V H which belongs to id W 0 but not to id W 00 . The zero locus in W 00
of such a function contains W 0 and is a subvariety of W 00 of pure dimension dim W 00 1.
Thus, some irreducible component of this variety properly contains W 0 and is properly
contained in W 00 and will do as our W 000 .
It is clear from the above paragraph that if 0 = Wd Wd1 W0 is a maximal chain of irreducible subvarieties of W , then successive varieties in the chain differ in
dimension by exactly one. It follows that d = dim W0 for such a chain. Clearly d will be
largest possible when W0 is an irreducible subvariety of V of largest dimension. That is,
when W0 is an irreducible component of V with the same dimension as V . This completes
the proof.
We now turn to the third notion of dimension tangential dimension.
6.10 Definition. If V is a germ at 0 of a algebraic or holomorphic variety then a tangent
vector to V is a derivation at 0 that is, linear map t : V H C (t : V O C) such that
t(f g) = f (0)t(g) + g(0)t(f ). The vector space of all tangent vectors is called the tangent
space to V and is denoted T (V ). Its dimension is the tangential dimension of V and is
denoted tdim V .
50
J. L. TAYLOR
6.11 Theorem. The vector space T (V ) is naturally isomorphic to the dual of M/M 2
where M is the maximal ideal of V H (V O).
Proof. If t T (V ) then t(1) = 2t(1) and so t kills constants and is, thus, determined by its
restriction to M . However, if t is any linear functional on V H (V O) which kills constants
and f = 1 + f1 , g = 1 + g1 with f1 , g1 M then
t(f g) = t(f1 ) + t(g1 ) + t(f1 g1 ) = g(0)t(f ) + f (0)t(g) + t(f1 g1 )
from which we conclude that t is a tangent vector if and only if t vanishes on M 2 . Thus,
restriction to M defines an isomorphism between T (V ) and the dual of M/M 2 . This
completes the proof.
It is clear that if Cn is considered either a holomorphic or an algebraic variety then T (Cn )
is naturally isomorphic to Cn where a given point t = (t1 , . . . , tn ) Cn is associated to
P f
the derivation on n H or n O defined by t(f ) =
ti
(0).
zi
If F : V W is a germ of a holomorphic map between varieties, then F induces an
algebra homomorphism F : W H V H (F : W O V O) which, in turn, induces a linear
map dF : T (V ) T (W ) by dF (t)(f ) = t(F (f )).
6.12 Theorem. If V is a germ at 0 of a subvariety of Cn and F : V Cn is the
inclusion, then dF : T (V ) T (Cn ) is injective and its image is {t T (Cn ) : t(g) =
0 whenever g id V }.
Proof. We have that
dF (t)(g) = t(F (g)) = t(g F ) = t(g|V )
and so dF (t) = 0 if and only if t = 0. Furthermore, a derivation on n H of the form dF (t)
clearly vanishes on id V = {g n H : g|V = 0}. Conversely, if s is a derivation on n H which
vanishes on id V , then s determines a well defined linear functional t on V H = n H/ id V
such that t(g|V ) = s(g). Since s is a derivation, t clearly is as well. This completes the
proof for n H. The proof is the same for n O.
In the above theorem, suppose that id V is generated by g1 , . . . , gm . Then a derivation
in T (Cn ) vanishes on id V if and only if it vanishes at each gi . Thus, if we identify T (V )
with its image under dF then
T (V ) = {t Cn :
X gj
ti
(0) = 0
z
i
i
for j = 1, . . . , m}
51
52
J. L. TAYLOR
6.17 Theorem. If V is the germ of a holomorphic variety then dim V tdim V and V
is regular if and only if tdim V = dim V .
Proof. That dim V tdim V clearly follows from Theorem 6.14. A germ of a variety is
regular if and only if it is biholomorphic to the germ of a neighborhood of zero in Cn where
n = dim V . By Theorem 6.14 again, this is equivalent to tdim V = n.
6.18 Theorem. If V is a holomorphic subvariety of a domain in Cn then the singular
locus of V is a holomorphic subvariety of V .
Proof. We dont yet have all of the machinery necessary to prove this. but it fits naturally
into this circle of ideas. Thus, we will present a proof that assumes a result that will be
proved later (Theorem 12.8).
The result we are after is a local result and so we may assume V is a subvariety of
some polydisc and is the union of finitely many irrreducible components Vj . Then the
singular locus of V is the union of the singular loci of the Vj and the sets of intersection
Vi Vj for i 6= j. Thus, it is enough to prove the theorem in the case where V is irreducible.
However, for an irreducible subvariety V of a connected set, the dimension of the germ Vz
of the variety at z V is a constant m. On the other hand, tdim Vz is n rank JG (z)
where JG is the Jacobian of a holomorphic map G whose coordinate functions form a set
of generators of id Vz . By Theorem 12.8, we may choose small enough that there exists
a G such that the coordinate functions of G generate id V at every point z V . The set
where rank JG (z) < k is a subvariety for each k and so the set where tdim Vz > m = dim V
is a subvariety. By the previous theorem, this is the singular locus of V . This completes
the proof.
The dimension of the germ V at of an algebraic variety V is defined to be the Krull
dimension of the corresponding local ring V O . A point of an algebraic variety V is said
to be a regular point if tdim V = dim V . The singular locus of V is the set of singular
points.
Of course, to each algebraic variety there is associated a holomorphic variety V which
is the same point set but with a different topology and a different local ring V H associated
to a point. it makes sense to ask whether or not the germ af an algebraic variety at a
point has the same dimension (or tangential dimension), as the germ of the corresponding
holomorphic variety. It also makes sense to ask if the singular locus of an algebraic variety
is a proper subvariety and whether or not it agrees with the singular locus of V . The rest
of this section is devoted to showing that the answer to all these questions is yes. We will
need the following two lemmas from commutative algebra which we will not prove (see
Atiyah-Macdonald chapter 11):
6.19 Lemma. Let B A be integral domains with B integrally closed and A integral
over B. Then for each prime ideal M of A, M is maximal in A if and only if N = M B is
maximal in B and, in this case, the local rings AM and BN have the same Krull dimension.
6.20 Lemma. The Krull dimension of n O is n.
6.21 Theorem. If V is an irreducible algebraic subvariety of Cn then dim V = dim V
at any point V , where V is the holomorphic subvariety determined by V .
53
6. Problems
1. Let V be a germ of a subvariety of Cn . Prove that if id Vi is a principal ideal in n H for
every irreducible component Vi of V , then id V is also a principal ideal.
2. Let V be a subvariety of a domain in Cn . Prove that if the modulus of a holomorphic
function f on V has a local maximum at z V then f is constant on the irreducible
component of V containing z. Use this to prove that a compact subvariety of an open
set in Cn must be finite.
3. Prove that if V is a germ of a variety then tdim V is the minimal number of generators
for the maximal ideal of V H.
4. Prove that if V is a germ of a holomorphic variety and P V H is a prime ideal, then
depth(P ) + height(P ) = dim V , where depth(P ) is the maximal length of a strict chain
of primes with P at the bottom and height(P ) is the maximal length of a strict chain
of primes with P at the top.
5. Prove the first part of Lemma 6.19: If B A are integral domains with A integral over
B and if M A is a prime ideal and N = M B then M is maximal if and only if N
is maximal.
54
J. L. TAYLOR
55
7.3 Theorem. If A B C are algebras with C faithfully flat over A and C faithfully
flat over B then B is faithfully flat over A.
Proof. Suppose X Y is an injective morphism of A-modules and let N be the kernel of
B A X B A X. Then since C is B-flat we have an exact sequence
0 C B N C B (B A X) C B (B A Y )
but by the associativity of tensor product we have that C B (B A X) = C A X and
C B (B A Y ) = C A Y . But since C is A-flat we have that C A X C A Y is
injective and, hence, C A N = 0. But this implies that N = 0 since C is faithfully flat
over A. Thus, we have proved that B is A-flat.
Now suppose that X is an A-module and consider the maps X B A X C A X.
Since the composition is an injection due to the fact that C is faithfully flat over A it
follows that the first map is an injection as well and, hence, that B is faithfully flat over
A. This completes the proof.
Our strategy for proving that V H is faithfully flat over V O will be to inject both of them
into a third algebra the M -adic completion of V O with respect to its maximal ideal M
and to show that this algebra is faithfully flat over both V O and V H. Then the previous
theorem will give us the desired result. To this end, we need to study the completion A of
a local ring with respect to its maximal ideal.
To begin with we need a lemma about graded Noetherian rings. A graded ring is is a ring
A =
n=0 An which is the direct sum of subspaces An in such a way that An Am An+m
for all n, m. The elements of An are said to be homogeneous of degree n. A graded
Noetherian ring is a graded ring which is also Noetherian as a ring.
7.4 Lemma. Let A be a graded Noetherian ring. Then
(i) A0 is a Noetherian ring.
(ii) A is a finitely generated A0 -algebra.
Proof. (i) Put A+ =
n=1 An . Then A+ is an ideal in A and A0 = A/A+ .
(ii) A+ is finitely generated. Let x1 , x2 , . . . , xs be a set of generators of A+ . Without
loss of generality we may assume these generators are homogeneous since, otherwise, we
56
J. L. TAYLOR
Proof. Put A =
n=0 M . Then A has a natural structure of a graded ring. Let
(a1 , . . . , as ) be a set of elements in M with the property the the images of the ai in M/M 2
generate it. Then for each n, M n is generated as an A-module by the monomials of degree
n in the ai . Thus, we have a natural surjective morphism A[x1 , . . . , xs ]
A determined
by xi1 xin ai1 ain M n which implies that A is a graded Noetherian ring. Let
n
Y =
n=0 M Y . Then Y is a graded A -module. It is clearly generated by Y0 = Y as
In addition, put X =
n=0 (X M Y ) Y . Then
M p (X M n Y ) M p X M n+p Y X M n+p Y
implies that X is an A -submodule of Y . Since A is a Noetherian ring, X is finitely
n
0
generated. There exists m0 Z+ such that m
n=0 (X M Y ) generates X . Then for any
p Z+ ,
X M
p+m0
Y =
m0
X
M p+m0 s (X M s Y ) M p (X M m0 Y ) X M p+m0 Y.
s=0
57
When X = A we obtain a completion A for A itself. It is easy to see that A is also a ring
is a module over A.
In fact, X X
is a
and for each module X over A the completion X
sequence
0 Xp Yp Zp 0
where Yp = Y /M p Y and Zp = Z/M p Z. Now limits of inverse sequences preserve left exactness but do not always preserve right exactness. Right exactness is, however, preserved
in the case where the left hand sequence {Xp } is surjective in the sense that each map
Xp+1 Xp is surjective (Problem 7.1) as is true in our situation. It follows that
Y Z 0
0X
is exact, as required.
7.7 Theorem. If A is a Noetherian local ring and X is a finitely generated A-module,
is an isomorphism.
then A A X X
commutes with taking finite direct sums. Thus, since
Proof. It is clear that X X
y
y
y
0
58
J. L. TAYLOR
in which the bottom row is exact by Theorem 7.6, the top row is right exact and the
follows from another diagram chase that A A X X is injective. This completes the
proof.
7.8 Theorem. If A is a Noetherian local ring, then A is faithfully flat over A.
Proof. It is easy to see (Problem 7.2) that an A-module Y is flat if and only if whenever
X1 and X2 are finitely generated A-modules and X1 X2 is injective, then Y A X1
Y A X2 is also injective. Since Theorems 7.6 and 7.7 prove that A A () preserves
exactness of short exact sequences of finitely generated A-modules, we conclude that A is
flat as an A-module.
Now suppose that X is a finitely generated A-module. Then the kernel of the map
is E = n M n X. It follows from Artin-Rees applied to E X that M E = E.
X X
Then Nakayamas lemma implies that E = 0. Now by Theorem 7.7 we conclude that the
map X A A X is injective whenever X is finitely generated. But this clearly implies
that this map is injective in general and, hence, that A is faithfully flat over A.
7.9 Theorem. If A is a Noetherian local ring then
= AM
;
(i) the unique maximal ideal of A is M
n for all n Z+ ; and
(ii) M n = A M
M
n+1 is an isomorphism for all n.
(iii) A/M n A/
-adic topology.
(iv) A is complete in the M
Proof. Since M p (A/M ) = 0 for p Z+ we have that A/M is complete in the M -adic
topology. We apply the completion functor to the exact sequence
0 M A A/M 0
and use the fact that this functor is exact to conclude that we have an exact sequence
A A/M 0
0M
is a maximal ideal of A since A/M is a field. We also have that
This implies that M
= A A M = AM
, which implies that M
n = AM
n . Now since A is faithfully flat over
M
n A = M n . This
A, Lemma 7.2 implies that M n A A = M n and we conclude that M
proves (ii).
M
n+1 is surjective follows from the fact that a Cauchy sequence in
That A/M n A/
the M -adic topology is eventually constant modulo M n . That this map is injective follows
from (ii). This completes the proof of (iii).
-adic completion of A
Part (iv) follows immediately from (iii) which shows that the M
is A.
is the only maximal ideal of A.
To
To complete the proof of (i) we must show that M
do this, we need only show that 1 a is a unit in A for every a M . In fact, the inverse
-adic topology
of 1 a for a M is 1 + a + a2 + + an + . . . which converges in the M
of A.
59
7.10 Theorem. If A is a Noetherian local ring then A is also a Noetherian local ring.
Proof. We have that A is a local ring from the previous theorem. Thus, we need only
P
show that A is Noetherian. The graded ring G(A) =
M n /M n+1 associated to A is a
n=0
finitely generated algebra over the field A/M and is therefore Noetherian by the Hilbert
is also Noetherian since it is isomorphic to G(A) by
basis theorem. It follows that G(A)
If we give I the filtration {M
n I} then G(I)
Theorem 7.9(iii). Suppose I is an ideal of A.
and, as such, it is finitely generated. Let {
embedds as an ideal of G(A)
ai ; i = 1, . . . , n}
be a set of homogeneous generators of G(I), set ri = deg(
ai ) and let ai I M ri be a
representative of a
i for each i. Let J be the ideal in A generated by a1 , . . . , an .
We will prove that J = I. Clearly G(I) = G(J). Suppose u I. Since A is Hausdorff,
p M
p+1 . Then there exist v0i M
pri such that
therePexists p such that u M
p+1
u v0i ai I M
. By continuing this construction we obtain sequences {vji ; j
p+jri and
Z+ , i = 1, . . . , n} such that vji M
u
n X
s
X
p+s+1
vji ai I M
i=1 j=0
P
Since A is complete, the series j=0 vji converges to some vi A for each i and we have
Pn
u = i=1 vi ai . Thus, u J and the proof is complete.
We now return to the study of the algebras V O and V H. Note first that if A is n O
or n H and M is the maximal ideal of A then A/M p is just the quotient of the ring of
polynomials in n variables modulo the ideal consisting of polynomials all of whose terms
are of degree at least p. Thus, the following result is obvious from the definitions:
7.11 Lemma. The algebras n O and n H both have as completion the algebra of formal
power series C[[z1 , . . . , zn ]].
The following technical lemma due to Chevalley is the key to showing that V O and V H
also have the same completion.
7.12 Lemma. Let V be a germ of an algebraic subvariety of Cn . Then there are no
That is, V O
is reduced.
non-zero nilpotent elements of V O.
Proof. We first reduce to the case where V is irreducible. If V is not irreducible and
V = V1 Vk is its irreducible decomposition, then consider the map
V
O iVi O
iV O
O
i
60
J. L. TAYLOR
by Theorem 7.11 and, hence, is an integral domain. We denote its field of fractions by K
of V O is A A V O
and remark that this is an extension field of K. The completion V O
algebra L
to be the result of passing from L to A A L
by Problem 7.3 . We define a K
and then localizing relative to the multiplicative set consisting of the non-zero elements of
By Theorem 7.8, A is faithfully flat over A and so V O
= A A V O is embedded as a
A.
subalgebra of A A L. We claim that this algebra is, in turn, embedded as a subalgebra
That is, we must show that nothing is killed when we localize. This means we must
of L.
show that am = 0 for 0 6= a A and m A A L implies that m = 0. Since, {qi } forms
P
ci A. Then am = 0
a basis for L over K, we may write
m
=
ai qi /ci with ai A,
Q
implies that acm = 0 where c = ci A. This implies that
X
aai (c/ci ) qi = 0
and, hence, that aai (c/ci ) = 0 for each i. Since A is an integral domain, this implies that
is embedded as
ai = 0 for each i and, hence, that m = O. Thus, we have shown that V O
a subalgebra of L.
as K
K L, that is, as
It is clear from the construction that we may also describe L
This is clearly
the algebra obtained from L by extending its ground field from K to K.
61
p
= 0 we conclude that x = 0. This completes the proof.
that p=1 M
When confusion might otherwise result we will denote by V the germ of the holomorphic
variety associated to V . When it is clear which is meant we will simply write V for either
the algebraic or holomorphic variety.
7.13 Theorem. Let V be a germ of an algebraic variety in Cn and V be the corresponding
germ of a holomorphic variety. Then id V = n H id V .
Proof. Let I = id V n O.
If J = n HI then loc J =mV and so it follows from the
f in V O
Since f n H, it follows from Lemma 7.2(iii) that f J = n HI. We conclude that
idV = n H id V as claimed.
7.14 Theorem. If V is a germ of an algebraic variety then
=
H
O.
Proof. Let {p1 , . . . , pr } be a set of generators for id V n O. Then by the previous theorem
we have that it is also a set of generators for id V n H. Thus, we have a commutative
diagram
nO
nH
nO
nH
O 0
y
H 0
62
J. L. TAYLOR
7. Problems
1. Prove that an inverse limit of an inverse sequence of short exact sequences is exact
provided the left hand inverse sequence is surjective.
2. Prove that an A-module Y is flat if and only if whenever X1 and X2 are finitely generated
A-modules and X1 X2 is injective, then Y A X1 Y A X2 is also injective.
3. Prove that if A is a Noetherian local ring with maximal ideal M and B is a local ring
which is a finitely generated integral ring extension of A and if N is the maximal ideal
of B then the completion of B in the N -adic topology is the same as its completion as
an A-module with the M -adic topology.
4. Suppose A is an integrally closed integral domain and K is its field of fractions. Prove
that if a matrix with entries in K is integral over A then its trace lies in A.
5. Prove that the formal power series ring C[[z1 , . . . , zn ]] is a unique factorization domain.
Hint: Use induction on the number of variables, Gausss Theorem (A is a UFD implies
A[z] is a UFD), and an extension of the Weierstrass preparation theorem to formal
power series.
63
8. Sheaves
Sheaf theory provides the formal machinery for passing from local to global solutions for
a wide variety of problems as well as for classifying the obstruction to so doing when local
solutions do not give rise to global solutions. The following is a list of typical examples of
such local to global problems:
1. If X is a compact Hausdorff space and f is a continuous complex valued function on
X which never vanishes, then f locally has a continuous logarithm. Does it have a
logarithm globally? In other words, is there a continuous function g on X such that
f = exp g?
f
= g has a
2. If U is a domain in C and g is a C function on U then the equation
z
solution locally in a neighborhood of each point. Does it have a global solution on U ?
3. If U is a domain in Cn , V U is a holomorphic subvariety and f is holomorphic on V ,
then for each point V there is a holomorphic function defined in a neighborhood
U of in Cn whose restriction to U V agrees with that of f . Is there a holomorphic
function defined on all of U whose restriction to V is f ?
4. If U is a domain in Cn and V U a holomorphic subvariety then V is locally defined
as the set of common zeroes of some set of holomorphic functions. Is there a set of
holomorphic functions defined on all of U so that V is its set of common zeroes?
Generally these problems involve classes of functions continuous, holomorphic, C ,
etc. which make sense on any domain in the underlying space. The notion of sheaf
simply abstracts this idea:
8.1 Definition. Let X be a topological space. We consider the collection of open subsets
of X to be a category where the morphisms are the inclusions U V . Then a presheaf
on X is a contravariant functor from this category to the category of abelian groups. A
morphism between two sheaves on X is a morphism of functors.
Thus, a presheaf S on X assigns to each open set U X an abelian group S(U ) and to
each inclusion of open sets U V a group homomorphism U,V : S(V ) S(U ), called the
restriction map, in such a way that U,U = id for any open set U and U,W = U,V V,W
for any triple U V W .
A morphism : S T between two presheaves on X assigns a morphism U : S(U )
T (U ) to each open set U in a way which commutes with restriction. Unless the context
dictates otherwise, we shall usually drop the subscript from U and write simply .
An example of a presheaf is the assignment to each open subset U X of the algebra
of continuous functions C(U ). The restriction map U,V is just restriction of functions in
this case. The resulting presheaf C is called the presheaf of continuous functions. The
presheave, C , of C -funtions on a C -manifold and the presheaf, H, of holomorphic
functions on Cn are defined in the same way. If Cn is given the Zariski topology, then we
may define on it the presheaf, O, of regular functions. If X is any topological space and
G is a fixed abelian group, then we may define a presheaf called the constant presheaf by
assigning G to each non-empty open set and 0 to the empty set. The first four of the above
examples are actually presheaves of algebras not just of abelian groups. As we shall see,
64
J. L. TAYLOR
the existence of additional structure on the objects S(U ) for a presheaf S is the typical
situation, although the abelian group structure is all that is needed in much of the theory.
If S is a presheaf, then an element s S(U ) will be called a section of S over U . If
U V then the image of a section s S(V ) under the restriction map U,V : S(V ) S(U )
will often be denoted s|U and called the restriction of s to U .
A sheaf is a presheaf which is locally defined in a sense made precise in the following
definition:
8.2 Definition. If S is a presheaf on X then S is called a sheaf if the
S following conditions
are satisfied for each open subset U X and each open cover U = iI Ui of U :
(i) if s S(U ) is a section such that s|Ui = 0 for all i I, then s = 0;
(ii) if {si S(Ui )}iI is a collection of sections such that si |Ui Uj = sj |Ui Uj for all
i, j I, then there is a section s S(U ) such that s|Ui = si for all i I.
Note that since the empty cover is an open cover of the empty set, it follows from 8.2(i)
that S() = 0 if S is a sheaf.
The presheaves of continuous, C , holomorphic and regular functions described earlier
are obviously sheaves. However, the presheaf which assigns a fixed group G to each nonempty open set is not a sheaf unless the underlying space has the property that every open
set is connected. There is, however, a closely related sheaf: the sheaf of locally constant
functions with values in G. In fact, for every presheaf there is an associated sheaf, as we
shall show in Theorem 8.6.
If S is a presheaf on X then the stalk, Sx , of S at x X is the group lim{S(U ) : x U }.
65
such that (sUx ) = t|Ux . Then for x, y U , (sUx |Ux Uy sUy |Ux Uy ) = 0. Since Ux Uy
is injective, this implies that sUx |Ux Uy = sUy |Ux Uy and, by 8.2(ii), that there exists a
section s S(U ) such that s|Ux = sUx for each x U . It follows that (s)x = tx for each
x U and, hence that (s) = t. Thus, U is surjective and the proof is complete.
S
If S is a presheaf on X then we may construct a topological space S =
Sx by
xX
66
J. L. TAYLOR
Note that condition (i) above implies that the images of local sections of S form a
neighborhood base for the topology of S and, in view of this, (iii) is equivalent to the
requirement that for each open set U X, the sum of two sections of S over U is again
a section. With the above definition, a morphism : S T of sheaves over X is a
continuous map which commutes with projection and is a group homomorphism Sx Tx
between stalks. The equivalence between the category of sheaves in our previous sense and
the category of sheaves in this sense is given by the constructions S S and S S of
Theorems 8.5 and 8.6. We will normally stick with our original definition, but on occasion
it will be useful to use the fact that it has the above alternate formulation.
One of the things the alternate definition of sheaf allows us to do easily is define the
notion of a section of a sheaf : S X over a subset Y X which is not necessarily
open. A section : Y S is a continuous map such that = id. The group of all
sections over Y will be denoted by (Y, S). We shall have more to say about this later
(Theorem 8.12). Of course, for an open set U X, (U ; S) = S(U ).
An additive category is a category such that for each pair A, B hom(A, B) has an
abelian group structure satisfying a distributive law relative to composition, direct sums
are defined and there is a zero object. If : A B is a morphism in an additive category,
then a kernel ker for is an object with a morphism : ker A such that = 0
and such that any morphism : C A with = 0 factors through . Similarly, a
cokernel coker for is an object with a morphism : B coker such that = 0
and such that any morphism : B D with = 0 factors through . In general, a
morphism need not have a kernel or a cokernel. When they do exist they are unique up
to isomorphism.
An abelian category is an additive category such that every morphism : A B has
both a kernel and a cokernel and the natural map
coim = coker(ker A) im = ker(B coker )
is an isomorphism.
A morphism : S T in the category of presheaves has both a kernel and a cokernel
and these are the obvious presheaves: U ker U and U coker U . It is easy to see
that the presheaves on a given space X form an abelian category. However, we are not
really interested in this category. We are interested in the category of sheaves.
Suppose : S T is a morphism of sheaves. It is easy to see that the kernel of as a
morphism of presheaves (the presheaf U ker U ) is, in fact, a sheaf and is the category
theoretic kernel of as a sheaf morphism (Problem 8.2). However, the analogous statement
is not true in general for the cokernel (Problem 8.3). In other words, U coker U need
not be a sheaf. However, we have the following:
8.8 Theorem. Suppose : S T is a morphism of sheaves. Then the sheaf of germs of
the presheaf U coker U is a cokernel for .
Proof. If C is the presheaf U coker U and C is its sheaf of germs, then the composition
of the presheaf morphisms T C and C C is a sheaf morphism such that = 0.
Any sheaf morphism : T D such that = 0 must factor through T C as a
presheaf morphism but since it is a sheaf morphism it must actually factor through by
Problem 8.4. This completes the proof.
67
We shall denote the sheaf of germs of the presheaf U coker U by coker . This might
seem ambiguous since one might use the same notation for U coker U itself since it is
the cokernel of as a presheaf. However, we will never do this, since our focus will be on
the category of sheaves.
Note that if U is an open set, then the space of sections (coker )(U ) is not the obvious
candidate, coker U , since U coker U is not a sheaf; however, it is true that we get
the obvious thing at the level of stalks: that is, (coker )x = coker x . This follows from
the fact that a presheaf and its sheaf of germs have the same stalks (Theorem 8.6(b)). It
follows that im = ker(T coker ) is characterized by
(im )(U ) = {t T (U ) : tx im x
for all
x U}
Proof. Let 0
A
B
C be an exact sequence of sheaves on X. This means
that 0
Ax
Bx
Cx is exact for each x X. Let U X be open. Then
U : A(U ) B(U ) is injective by Theorem 8.4. Now suppose b B(U ) is in the kernel
of U . For each x U there is a neighborhood Vx of x and a section ax A(Vx ) such
that (ax ) = b|Vx . Then (ax )|Vx Vy = (ay )|Vx Vy for each pair x, y U and it follows
from the injectivity of Vx Vy that ax |Vx Vy = ay |Vx Vy . From the definition of a sheaf it
now follows that there exists a A(U ) such that a|Vx = ax for each x and from this that
(a) = b. This completes the proof.
We shall show that the category of sheaves has enough injectives to construct injective
resolutions for each sheaf. From this it follows that we have naturally defined right derived
functors for every left exact functor in particular, for the functors (U ; ). The resulting
functors are those of sheaf cohomology and are the subject of the next section.
We end this section with a discussion of the operations on sheaves that are induced by
a continuous map f : Y X.
68
J. L. TAYLOR
U }.
Part (b) of the above definition is a case where it would be more instructive to use the
alternate definition of sheaf given in 8.7. In fact if : S X is a sheaf in this sense, then
f 1 S is just the pullback of : S X via f : Y X. This is the topological space
f 1 S = {(s, y) S Y : (s) = f (y)}
with projection f 1 S Y given by projection on the second coordinate. This description
of f 1 S makes it clear that it is a sheaf on Y which has stalk at y Y equal to the stalk
of S at f (y). Form this it follows easily that f 1 is an exact functor from sheaves on X
to sheaves on Y .
The direct image functor f from sheaves on Y to sheaves on X is not exact in general.
To see this, consider the map f : Y pt; in this case, f T is the sheaf which assigns to pt
the group (Y ; T ) and we know that is not always exact (Problem 8.6). On the other
hand, an argument like the one in Theorem 8.9 shows that f is alway left exact. Thus, it
is another functor for which we expect to be able to construct right derived functors.
A special case of the inverse image functor is the restriction functor. Here, Y X is a
subset and i : Y X is the inclusion. For a sheaf S on X, i1 S is denoted S|Y and is
called the restriction of S to Y . Using the alternate definition of sheaf, the restriction of
: S X to a subset Y is the space SY = 1 (Y ) with topology and projection SY Y
inherited from S and . The group of global sections of the restricted sheaf, (Y, S|Y ), is
the same as the group (Y ; S) of continuous sections of S over Y . It follows from Theorem
8.9 that
8.11 Corollary. For any subset Y X and any sheaf S on X, (Y ; ) is a left exact
functor.
Thus (Y ; ) is also a functor which we expect to have right derived functors if we can
show that there are enough injective sheaves.
If S is a sheaf on X, Y is a subspace of X and U is an open set containing Y , then
restriction defines a morphism S(U ) = (U ; S) (Y ; S). Thus, restriction defines
a morphism : lim{(U ; S) : Y U } (Y ; S). This is often, but not always an
isomorphism:
8.12 Theorem. The morphism
: lim{(U ; S) : Y U } (Y ; S)
69
injective.
Now suppose s (Y ; S). Then there is an open cover {Ui }iI of X and a family
of sections {si (Ui ; S)}iI such that s|Vi Y = si |Vi Y for each i. If X is Hausdorff
and Y compact or if X is paracompact and Y closed then we may assume that the cover
{Ui } is a locally finite cover of X and, furthermore, that
S there is locally finite collection
of open sets {Vi }iI with Vi Ui for each i and Y iI Vi . Now for each x X set
I(x) = {i I : x Vi } and
W = {x
Vi : six = sjx
for all
i, j I(x)}
Then I(x) is a finite set. For a given x let U be a neighborhood of x which meets Vi for i
in only a finite set, J I. Then the set
Wx = U
Vi
iII(x)
is an open set containing x. In fact, Wx is the set of y U such that I(y) I(x).
Now if x W choose Nx to be a neighborhood of x contained in iI(x) Ui Wx on which
si |Nx = sj |Nx for i, j I(x). Then the conditions for membership in W will also be satisfied
for any point y Nx . In other words, W is an open subset of X. The fact that the si
fit together to define a section s on Y means that Y W . By construction, the si fit
together on W to define a section in (W ; S) which restricts to s on Y .
The functors indroduced in the following definition are also left exact functors on sheaves
(Problem 8.7):
8.13 Definition. If S is a sheaf on X and s (X; S) is a section, then the support of
s is the (necessarily closed) set K = {x X : sx 6= 0}. If Y X is any subset of X
then Y (X; S) is the group of sections s (X; S) with support contained in Y . If is a
family of subsets of X which is closed under finite unions, then (X; S) is the group of
sections s (X; S) with support contained in some member of .
A family as above is called a family of supports. A common situation in which
(X; S) is useful is when X is locally compact and the family of supports is is the
family of compact subsets of X.
8. Problems
1. Prove Theorem 8.5.
2. Prove that if is a morphism of sheaves then the presheaf U ker U is a sheaf and is
a kernel for in the category of sheaves.
3. Give an example of a morphism of presheaves such that the presheaf U coker U
is not a sheaf.
4. Prove that if S is a presheaf then each presheaf morphism S T , where T is a sheaf,
Prove that
factors uniquely through the morphism S S of S to its sheaf of germs S.
70
J. L. TAYLOR
this means that S S is a left adjoint functor for the forgetful functor which regards
a sheaf as just a presheaf.
5. Prove that the category of sheaves is an abelian category.
6. Give an example to show that the functor (X; ) is not necessarily exact.
7. Prove that the functors Y (X, ) and (X, ) of Definition 8.11 are left exact.
71
9. Sheaf Cohomology
Recall that in any abelian category an object A is called injective if the functor hom(, A)
is exact and is called projective if the functor hom(A, ) is exact. In this section we shall be
concerned with injective objects. The following summarizes the most elementary properties
of injectives (Problem 9.1):
9.1 Theorem. In any abelian category:
(1) An object A is injective if and only if for every monomorphism i : B C, each
morphism : B A extends to a morphism : C A such that = i.
(2) Every monomorphism i : A B, with A injective, splits (has a left inverse).
(3) If 0 A B C 0 is an exact sequence and A is injective, then B is injective
if and only if C is injective.
Our abelian category is said to have enough injectives if for every object A there is an
injective object I and a monomorphism A I. This, and the fact that every morphism
has a cokernel, allows one to construct injective resolutions
0 A I0 I1 In
for each object A. There is another way of thinking about such resolutions which yields
both economy of notation and additional insight and so is worth introducing. We will let
I denote the complex
0 I0 I1 In
and identify A with the complex
0A0
where A appears in degree zero. Then an injective resolution of A may be thought of as a
morphism of complexes
i:AI
where I is a complex of injective objects (zero in negative degrees) and i induces an
isomorphism on cohomology (both complexes have vanishing cohomology in all degrees
except zero where the cohomology is A). A morphism of complexes which induces an
isomorphism of cohomology is called a quasi-isomorphism. Thus, an injective resolution of
an object A is a quasi-isomorphism A I where I is a complex of injectives which vanishes
in negative degrees. Actually, insisting that I vanish in negative degrees is equivalent, for
the purposes of this theory, to insisting that it be bounded on the left that is, vanish for
sufficiently high negative degrees.
A morphism : X Y of complexes is homotopic to zero if there are morphisms
n
n
hn : X n Y n1 such that hn+1 X
+ Yn1 hn = n for each n. Here, X
: X n X n+1
and Yn1 : Y n1 Y n are the differentials in the complexes X and Y . Two morphisms
, : X Y of complexes are said to be homotopic if their difference, is homotopic
to zero. A key result is the following (Problem 9.2):
72
J. L. TAYLOR
y
y
B J
is commutative. Furthermore, any two maps I J with this property are homotopic.
When Theorem 9.2 is applied to two different injective resolutions of A and the identity
morphism from A to A, it implies the following:
9.3 Corollary. If i : A I and j : A J are two injective resolutions of A then there
are morphisms : I J and : J I such that
I
I
x
J A J
is a commutative diagram and both and are homotopic to the identity.
If X = {X n , n } is a complex, then its cohomology is the graded group {H n (X)} where
H n (X) = ker n / im n1 . A morphism : X Y of complexes induces a morphism
H() : H(X) H(Y ) of cohomology. Two morphisms which are homotopic induce the
same morphism of cohomology. Also, if two morphisms and are homotopic and F is
a functor into another abelian category, then F () and F () are also homotopic. These
facts and Theorem 9.2 imply the following:
9.4 Corollary. If : A B is a morphism, i : A I and j : B J are injective
resolutions of A and B and F is a functor from our category to another abelian category,
then the morphisms
: I J of Theorem 9.2 induce a single, well defined morphism
: H(F (I)) H(F (J)). This is an isomorphism if is an isomorphism.
In any reasonable category with enough injectives there is a way of assigning to each
object A a particular injective resolution I(A). In a small category this can be done using
the axiom of choice. We shall show how to do this in the case of the category of sheaves.
In many cases this can be done in such a fashion that A I(A) is a functor ideally an
exact functor. This is nice when it can be done (and it can be for sheaves) but it is not
necessary for the developement of the theory. In any case, assuming that we are working
in a category with enough injectives and with some way of assigning an injective resolution
to an object, we may construct the higher derived functors of a left exact functor F as
follows: the n-th derived functor of A is
Rn F (A) = H n (F (I(A)))
the n-th cohomology of the complex obtained by applying the functor F to the complex of
injectives I(A). Of course, by Corollary 9.4, different choices of resolutions I(A) will yield
73
isomorphic objects Rn F (A) but, in order that the Rn F be functors, it is important that
a specific way of making the assignment A I(A) be available. That Rn F is a functor
then follows from Corollary 9.4.
The fact that F is left exact means that if A I 0 I 1 . . . is an injective resolution,
then on applying F exactness is preserved at the first two terms; that is,
0 F (A) F (I 0 ) F (I 1 )
is exact. From this it follows that:
9.5 Theorem. If F is a left exact functor from an abelian category with enough injectives
to an abelian category, then there is an isomorphism of functors F R0 F .
If (as is the case for the category of sheaves) A I(A) is an exact functor then an
exact sequence of objects
0ABC0
yields an exact sequence of complexes
0 I(A) I(B) I(C) 0
because the objects in these complexes are injective, this short exact sequence splits and,
hence, remains exact when we apply a left exact functor F . This yields a short exact
sequence of complexes
0 F (I(A)) F (I(B)) F (I(C)) 0
which, in turn, yields a long exact sequence of cohomology
0 F (A) F (B) F (C) R1 F (A) R1 F (B) R1 F (C) R2 F (A)
However, to get this result it is not neccessary to assume that A I(A) is exact or is even
a functor. In fact, given injective resolutions A I(A) and C I(C) one may construct
an injective resolution B J along with morphisms I(A) J and J I(C) such that
0 I(A) J I(C) 0
is an exact sequence of complexes. To do this, we set J n = I n (A) I n (C) and define
morphisms j : B J 0 and n : J n J n+1 as follows: Using the fact that I 0 (A) is
injective, we extend iA : A I 0 (A) to a morphism j1 : B I 0 (A). We let j2 : B I 0 (C)
be the composition of B C with iC : C I 0 (C). Then j = j1 j2 . Clearly j is a
monomorphism of B into J 0 and the diagram
0
iA y
jy
iC y
0 I 0 (A) J 0 I 0 (C) 0
74
J. L. TAYLOR
is commutative. We repeat this argument with iA : A I 0 (A) and iC : C I 0 (C) replaced by coker iA I 1 (A) and coker iC : I 1 (C), respectively, and obtain a commutative
diagram
0 A B C 0
jy
iA y
iC y
0 I 0 (A) J 0 I 0 (C) 0
0
0
0
A
y
y
Cy
0 I 1 (A) J 1 I 1 (C) 0
Continuing in this way, we construct an injective resolution B J of B and morphisms
of complexes I(A) J and J I(C) for which the diagram
0
iA y
jy
iC y
0 I(A) J I(C) 0
is commutative with exact rows. On applying any left exact functor F we conclude, without
the assumption that I() is exact or even a functor, that
9.6 Theorem. If F is a left exact functor from an abelian category with enough injectives
to an abelian category, and if
0ABC0
is a short exact sequence in the first category, then there is a long exact sequence
0 F (A) F (B) F (C) R1 F (A) R1 F (B) R1 F (C) R2 F (A)
In this result, the morphisms Rn F (A) Rn F (B) and Rn F (B) Rn F (C) are just
those induced by A B and B C as in Corollary 9.4; i. e. they are the images of these
morphisms under the functor Rn F . The connecting morphisms Rn F (C) Rn+1 F (A) a
priori depend on the choices made in the construction of J. In fact, they do not depend on
these choices. They are well defined and depend functorially on the short exact sequence
0 A B C 0. To prove this requires some diagram chasing which we choose not
to do here. It is done in any number of books on homological algebra.
Of course, one does not really compute the objects Rn F (A) using injective resolutions.
In practice, one uses the long exact sequence to reduce the computation of Rn F (A) for
complicated objects A to that for simpler objects or one uses Theorem 9.8 below which
often allows one to compute Rn F (A) using much simpler resolutions of A.
An object C is said to be acyclic for the functor F if Rn F (C) = 0 for n > 0.
9.7 Theorem. Let F be as in Theorem 9.6. Then injective objects are F -acyclic.
Proof. If I is injective, then 0 I I 0 is an injective resolution of I. It follows from
Corollary 9.4 that Rn F (I) may be computed by applying F to this resolution and taking
cohomology. Thus, Rn F (I) = 0 for n > 0 and I is acyclic.
75
p0
where K 0 = A. Using the fact that Rn F (J p ) = 0 for n > 0, we conclude from these long
exact sequences that
Rq F (K p ) ' Rq1 F (K p+1 ),
and
p 0, q > 1
76
J. L. TAYLOR
77
H p (X; MU ) = Hp (U ; M)
78
J. L. TAYLOR
The definition of fine sheaf needs some comment. The statement that i is supported
in Ui means
P that there is a closed subset Ki of Ui such that x = 0 for all x X Ki .
The sum i i makes sense as a morphism from M to M because the open cover {Ui }
is locally finite. This and the fact that i is supported in Ui implies
P that, in a sufficiently
small neighborhood of each point, only finitely many terms of i i are non zero and,
hence, the sum makes sense in such a neighborhood; but then these local morphisms fit
together to define a morphism of sheaves globally.
In what follows, the term -acyclic will mean acyclic for the functor (X; ).
9.14 Theorem. If X is any topological space, then in the category of sheaves of Rmodules on X
(i) if the first two terms of a short exact sequence are flabby then so is the third;
(ii) injective flabby -acyclic;
Proof. For U open and M a sheaf of R-modules on X, let RU be the restriction of R to
U followed by extension by zero to all of X. Then we have an inclusion RU R. Thus,
if M is injective then each morphism from RU to M extends to a morphism from R to
M. But a morphism RU M is just a section of M over U and a morphism R M is
just a section of M over X (Problem 9.5). Thus, sections of M over U extend to sections
of M over X. Thus, injective flabby.
Now suppose that A is a flabby sheaf and
0 A B C 0
is an exact sequence of sheaves of R-modules. We wish to prove that
79
some section of B over U . But, since B is flabby, the section b is the restriction to U of a
global section b0 of B. Then (b0 ) provides an extension of c to a global section of C.
Now to prove that a flabby sheaf A is -acyclic we embed it in an injective I and use
the long exact sequence of cohomology for the short exact sequence
0 A I C 0
where is the inclusion and C is its cokernel. Since I is -acyclic we have that
H p (X; A) ' H p1 (X; C)
p>1
and, by what we proved in the first part of the argument, H 1 (X; A) = 0. Also, C is flabby
because A and I are flabby. Thus, H 2 (X; A) ' H 1 (X; C) = 0. By iterating this argument
we conclude that H p (X; A) = 0 for all p > 0. Thus, we have proved that flabby acyclic.
There is a similar result with a similar proof for soft sheaves. It is slightly more complicated and requires that the space be paracompact.
9.15 Theorem. If X is paracompact, then for sheaves of R-modules on X
(i) if the first two terms of a short exact sequence are soft then so is the third;
(ii) flabby soft -acyclic;
(iii) fine soft -acyclic.
Proof. Let U X be open and let M be a sheaf of R-modules on X. Since X is paracompact, By Theorem 8.12, every section of a sheaf on a closed set Y extends to an open set
containing Y and, hence, extends to all of X if the sheaf is flabby. Thus, flabby soft.
Now suppose that A is a soft sheaf and
0 A B C 0
is an exact sequence of sheaves of R-modules. We wish to prove that
80
J. L. TAYLOR
0 A I C 0
where is the inclusion and C is its cokernel. Since I is acyclic we have that
H p (X; A) ' H p1 (X; C)
p>1
and, by what we proved in the first part of the argument, H 1 (X; A) = 0. Also, C is soft
because A and I are soft. Thus, H 2 (X; A) ' H 1 (X; C) = 0. By iterating this argument
we conclude that H p (X; A) = 0 for all p > 0. Thus, we have proved that soft acyclic.
It remains to prove that fine soft. Thus, let Y X be closed and let M be a
fine sheaf on X. If s (Y ; M) then for each x Y the germ of s at x is represented
by a section defined in a neighborhood of x which agrees with s when restricted to that
neighborhood intersected with Y . Thus, we may choose an open cover {Ui } of X and
elements si (Ui ; M) such that s|Ui Y = si |Ui Y for each i - one of these open sets
will be the complement of Y and will have the zero section assigned to it and the others
will be neighborhoods of points of Y . Because X is paracompact, we may assume that
{Ui } is locally finite. Now because M is fine, we may
P choose for each i an endomorphism
i : M M supported in Ui in such a way that i i = id. For each i we interpret i si
to be a section
on all of X by extending it to be zero on the complement of Ui . We then
P
set s0 = i i si (X, M). That this sum makes sense follows from the local finiteness
of the open cover which means that in a neighborhood of any point we are summing only
finitely many non-zero terms. We also have that s0x = sx at each point x Y so that s0 is
an extension of s to all of X. It follows that M is soft. Thus, we have proved that fine
soft and completed the proof of the Theorem.
We conclude this chapter with some examples of acyclic resolutions which show that
certain classical cohomology theories are just examples of sheaf cohomology.
By C 0 we shall mean the sheaf of continuous functions on X. If X is the appropriate
kind of differentiable manifold, we shall denote by C p and C the sheaves of functions
with continuous partial derivatives up to order p and functions with continuous partial
derivatives of all orders, respectively.
9.16 Theorem. Each of C 0 , C p and C is a fine sheaf if X is a paracompact space and,
in the case of C p and C , X has the appropriate differentiable manifold structure.
Proof. Paracompact implies normal which implies that Urysohns Lemma holds. Urysohns
Lemma can be used to constuct continuous partitions of unity subordinate to any locally
finite open cover. A continuous function with support inside a given open set defines, by
81
d1
dp1
dp
0 C E 0 E 1 E p
where dp is exterior differentiation and C here stands for the constant sheaf with stalks
C. The Poincare Lemma says that if X is any open ball in Rn then the corresponding
sequence of sections is exact (this is proved by constructing an explicit homotopy between
the identity and zero using integration along lines from the center of the ball). Since a C
manifold looks locally like a ball in Rn it follows that the de Rham complex is exact as
a complex of sheaves. Thus, it defines a resolution C E of the constant sheaf C by a
complex E of fine sheaves. On passing to global sections of E, we obtain the classical de
Rham complex E(X) of differential forms on X. The cohomology of this complex is called
the de Rham cohomology (with coefficients in C) of X. By theorem 9.8 we have:
9.17 Theorem. There is a natural isomorphism between the de Rham cohomology of a
C manifold X and the sheaf cohomology H(X; C) for the constant sheaf C on X.
each open
on U , for the cover U, is the direct
Q set U X, the module of Cech p-cochains
product I p+1 (U U, S) and is denoted C p (U, S)(U ). In other words, a p-cochain for
U on U is a function f which assigns to each I p+1 an element f () (U U ; S). If
V U then restriction clearly defines a morphism C p (U , S)(U ) C p (U, S)(V ) and, thus,
U C p (U, S)(U ) is a presheaf. In fact, it is clearly a sheaf. We denote it by C p (U, S).
The module of global sections of this sheaf is C p (U, S)(X). This is the classical space of
Cech
cochains for S and the cover U and will be denoted C p (U, S). We next define a
coboundary mapping
p : C p (U, S) C p+1 (U, S)
by
p f () =
p+1
X
(1)j f (j )|U U
j=0
where f C p (U , S)(U ) and j I p+1 is obtained from I p+2 by deleting its j th entry.
82
J. L. TAYLOR
p1
0 C 0 (U, S) C 1 (U, S) C p (U , S)
is a complex of sheaves and sheaf morphisms.
Proof. For I p+3 let j,k denote the result of deleting both the j th and the k th entries
from . Then
(j )k = j,k if 0 k < j p + 2
and
(j )k = j,k+1
if 0 j < k p + 1
p+2
X
(1)j
j=0
"p+1
X
#
(1)k f ((j )k )|Uj U |U U
k=0
k<j
X
(1)j+k f (j,k+1 )|U U
kj
=0
due to the fact that, in the middle line above, the second term is equal to the negative of
the first, which is evident from the change of variables j k, k + 1 j applied to the
second term and the observation that j,k = k,j .
The restriction maps S(U ) S(Ui U ) define a sheaf morphism : S C 0 (U , S)
whose composition with 0 is 0. In fact, much more is true:
Theorem 9.19. If U is an open cover of X and S a sheaf on X then the complex
0
p1
0 S C 0 (U , S) C 1 (U, S) C p (U, S)
is an exact sequence of sheaves.
Proof. Fix an x X. Let U be any neighborhood of x which is contained in a member of
U, say U Uk . Now suppose f C p (U, S)(U ) and p f = 0. Define g C p1 (U, S)(U ) by
g() = f ((k, ))
where we set (k, ) = (k, i0 , . . . , ip1 ) I p+1 for = (i0 , . . . , ip1 ) I p . Note that
g C p1 (U, S)(U ) due to the fact that U Uk which implies that Uk U U = U U .
Then for I p+1 ,
p
0 = f ((k, )) = f ()|Uk U U
p
X
j=0
= f ()
p1
g()
83
again due to the fact that Uk U U = U U . This shows that, locally, the kernel of p
is the image of p1 for p > 0. The same argument also works for p = 0 with 1 replaced
by . This proves the exactness of the above sequence.
the Cech
complex and we shall call the global Cech
complex for the sheaf S and the open
cover U.
1
and the direct image of a flabby sheaf is flabby by Problem 9.8. It follows that (i ) i S
is flabby for each and, hence, that C p (U, S) is flabby for each p. Thus, the complex in
Theorem 9.19 is an exact sequence of flabby sheaves. Then Theorem 9.14 implies that the
complex obtained from it by applying (X, ) is also exact. Part (ii) follows.
Let S be a sheaf on X, U an open cover of X and S C(U, S) the corresponding
Cech
resolution as in Theorem 9.19. Let S I be an injective resolution of S. Then the
injectivity of the terms in I can be used, as in Theorem 9.2, to inductively construct a
morphism of complexes C(U, S) I for which the diagram
S C(U , S)
y
S
I
is commutative. This morphism is unique up to homotopy and, hence, after we apply
p (U, S) H p (X, S).
(X, ) it determines a well defined morphism H
The open cover U is called a Leray covering for the sheaf S if for each multi-index
the sheaf S is acyclic on U . For example, it follows from the Poincare lemma that the
deRham cohomology of any convex open set in Rn vanishes in positive degrees. From this
and Theorem 9.17 we conclude that the constant sheaf C is acyclic on any convex open
set in Rn . This implies that any open cover of Rn by open convex sets is a Leray cover for
the constant sheaf C.
84
J. L. TAYLOR
9.22 Theorem. If U is a Leray cover for the sheaf S then the natural morphism
p (U, S) H p (X, S)
H
is an isomorphism.
Proof. Choose an embedding S F of S in a flabby sheaf (an injective, for example).
Let G be the cokernel and consider the short exact sequence:
0 S F G 0
Then from the long exact sequence and the fact that S is acyclic on U we conclude that
0 S(U ) F(U ) G(U ) 0
is exact for each multi-index and from this that we have an exact sequence
0 C(U, S) C(U, F) C(U, G) 0
of global Cech
complexes. This, in turn, gives us a long exact sequence of Cech
cohomology,
which, along with the long exact sequence for sheaf cohomology, the induced morphisms
from Cech
to sheaf cohomology discussed above and the fact that F is flabby, gives us the
following commutative diagram with exact rows:
0 (U , S) H
0 (U , F) H
0 (U , G) H
0 (U, S) 0
0 H
y
y
y
y
0 (X, S) (X, F) (X, G) H 1 (X, S) 0
The first three vertical arrows are isomorphisms and, hence, so is the fourth. This establishes our result in the case p = 1. We prove the general case using the following
commutative diagram, which is also part of the the diagram which comes from the long
exact sequences associated to the above short exact sequence:
p (U, G) H
p+1 (U, S) 0
0 H
y
y
0 H p (X, G) H p+1 (X, S) 0
for p > 0. This has exact rows and so if the first vertical morphism is an isomophism so
is the second. Consider the class of all sheaves T such that the open cover U is a Leray
cover for T . From the long exact sequence for sheaf cohomology it follows that if the
first two sheaves in a short exact sequence belong to this class then so does the third. By
hypothesis, S belongs to this class and F does since it is flabby. It follows that G belongs
as well. Thus, if we have proved the theorem for all members of this class and for all
85
degrees less thatn or equal to p then the above diagram shows that it is true for degree
p + 1 as well. By induction, the proof is complete.
Cech cohomology for a sheaf without reference to a cover is obtained by passing to a limit
over covers. More precisely, if V is an open cover which is a refinement of the open cover U
then the restriction maps define a morphism of complexes of sheaves C(U, S) C(V, S). By
passing to the limit over the directed set of open covers of X we obtain a complex of sheaves
is called the limit global Cech complex. Its cohomology is the Cech cohomology of S on
p (X, S)}. Clearly the morphisms H
p (U, S) H p (X, S) induce a
X and is denoted {H
p (X, S) H p (X, S).
morphism H
9.23 Theorem. If X is paracompact then the natural morphism
p (X, S) H p (X, S)
H
is an isomorphism.
Proof. . We shall prove that S C(S) is a resolution of S by soft sheaves. Since soft
sheaves are acyclic, the result will then follow from Theorem 9.8.
We have already remarked that S C(S) is a quasi-isomorphism in other words,
that the exactness of the sequences in Theorem 9.19 is preserved on passing to the direct
limit since direct limits generally preserve exactness. It remains to prove that each
C p (S) is a soft sheaf. Thus, let Y X be closed and suppose f is a section of C p (S)
over Y . Since X is paracompact and Y is closed, we have that f may be represented by a
section in a neighborhood U of Y . This section may, in turn, be represented by an element
f 0 C p (U, S)(U ) for some open cover U of X. We may choose a locally finite refinement
V of U with the property that each set in V is either contained in U or is contained in
X Y . We define a section g 0 of C p (V, S) on X by setting g 0 equal to the image of f 0
under the refinement map on those multi-indices for which all the correspondings sets in
V are contained in U and setting it equal to zero otherwise. On passing to the image g of
the classical Cech cohomology with coeficients in the group G from the theory of algebraic
topology.
We end this section with an example which shows how to solve one of the local to
global problems posed in chapter 8. This is the problem of finding a global logarithm for a
non-vanishing continuous function on X. We assume X is paracompact. Let C denote the
sheaf of continuous functions on X with addition as group operation and C 1 the sheaf
of invertible (non-vanishing) continuous functions with multiplication as group operation.
Then, due to the fact that a non-vanishing continuous function has a logarithm locally in
a neighborhood of each point, the sequence of sheaves:
exp
0 2iZ C C 1 0
86
J. L. TAYLOR
is exact. Since C is a fine sheaf and hence acyclic, we conclude from the long exact sequence
of cohomology for this short exact sequence that
1 (X, Z)
C 1 (X)/ exp(C(X)) ' H 1 (X, Z) ' H
where, of course, Z stands for the constant sheaf Z and we use the fact that 2iZ ' Z.
Thus, every non-vanishing continuous function on X has a global logarithm if and only if
9. Problems
1.
2.
3.
4.
5.
6.
7.
8.
9.
H p (X; MU ) = Hp (U ; M)
87
z V}
Then the map z (z, f (z)1 ) : Vf W is a morphism of ringed spaces (Problem 10.4)
with inverse (z, zn+1 ) z : W Vf . Thus, Vf is affine since it is isomorphic to W .
An algebraic prevariety is an ringed space which has a finite cover by open sets which
are affine varieties. Note that by Theorem 10.4 each open subset of an affine variety V
is, itself, a finite union of open sets which are affine varieties (sets of the form Vf ). Thus,
88
J. L. TAYLOR
89
P
I and, hence, in I. Thus, the equation
pj hj = 1 has a solution. When we restrict the
hj s to V , the last k m restrict to zero. Thus,P
the restriction to V of p1 , , pm gives
the required solution g1 , , gm to the equation
gi fi = 1. This shows that v Mv is
a bijection from points of V to maximal ideals of A. The topology of V is the relative
topology inherited from Cn . This is the topology in which closed sets are common zero
sets of families of polynomials. Thus, the closed subsets of V are common zero sets of
families of polynomials restricted to V . This is the Zariski topology induced on V by A.
Thus, we have proved part (i).
One direction of (ii) is trivial: Restriction to Vf defines a natural map Af V O(Vf ).
Suppose g/f k Af (g A) determines the zero section of V O over Vf . This means that
g = 0 on Vf which implies that f g = 0 in A which implies that g/f k = 0 in Af . Thus,
Af V O(Vf ) is injective. The surjectivity will be proved after we prove (iii).
We now prove (iii). Certainly the restriction to V of a polynomial on Cn is a regular
function on V and so A O(V ). Thus, we must show that the restriction map is surjective.
Suppose now that h O(V ). We can cover V with a collection {Ui } of open sets such that
f |Ui has the form pi /qi , O(Ui ) where pi and qi are elements of A and qi does not vanish
90
J. L. TAYLOR
91
by
g/f k m gm/f k
Its inverse is the morphism determined by
m/f k 1/f k m : Mf M (Vf )
It is clear that both morphisms are well defined and they are inverses of one another.
10.9 Theorem. Let V be an affine variety. Then
(i) for each open set U V of the form U = Vf , the functor M M (U ) =
V O(U ) O(V ) M is exact; that is, V O(U ) is a flat O(V )-module;
is exact.
(ii) the localization functor M M
Proof. Using the description of M (U ) as V O(U )O(V ) M makes it clear that this functor is
right exact. Thus, we need only prove that if N M is a submodule then N (U ) M (U ) is
injective. To prove this we use the fact, proved in the previous theorem, that M (U ) = Mf
and N (U ) = Nf if U = Vf . Thus, let n/f k represent an element of Nf and suppose that
it determines the zero element of Mf . This means that f p n = 0 for some p. But if this
equation holds in M it also holds in N and, hence, n/f represents the zero element of N
as well. This proves (i); however, (ii) is an immediate consequence of (i) and the fact that
sets of the form Vf form a basis for the topology of V
10.10 Theorem. If V is an affine variety and M is an O(V )-module then
) is
(i) on an open set of the form Vf the natural morphism Mf = M (Vf ) (Vf ; M
an isomorphism;
) is an isomorphism.
(ii) in particular, M (V ; M
92
J. L. TAYLOR
Proof. Statements (i) and (ii) are actually equivalent since, if V is affine, so is each Vf .
Thus, we will just prove (ii).
This is almost the same as the proof of Theorem 10.7 but there are a few differences.
over V . This means that we may
Suppose m M (V ) determines the zero section of M
(Ui ) for each i. Without
cover V with finitely many open sets Ui such that m|Ui = 0 in M
loss of generality we may assume that these sets are of the form Ui = Vqi where qi O(V ).
Then m|Ui = 0 means that for each i there is an integer ni such that qini m = 0. Since the
sets Ui cover V , the collection {qi } has no common zero on V . Since V is affine, Theorem
P ni
(V ). However, this
10.7 (i) implies there exists a set {gi }, such that
gi qi = 1 in M
ni
)
implies that m = 0 since m is killed by qi for every i. We conclude that M (V ) (V ; M
is injective.
). We can cover V with a finite collection {Ui = Vq } of
Now suppose that s (V ; M
i
) of an element mi /q ni , M
(Ui ),
basic open sets such that s|Ui is the image in (Ui ; M
i
where mi M . In fact by relabeling each qini as qi we may assume that s|Ui is the
) of an element of the form mi /qi . Since these elements fit together
image in (Ui ; M
to form a section over V , we may assume (after refining the cover if necessary) that
(mi /qi )|Ui Uj (mj /qj )|Ui Uj = 0. However, Ui Uj = Vqi qj and so this equality means
that there is a positive integer n so that
(qi qj )n (qj mi qi mj ) = 0
in M for each pair i, j. If we simply relabel qin mi by mi and qin+1 by qi , then the fractions
mi /qi dont change but the above equality becomes simply
qj mi qi mj = 0
Since
V is affine and the sets Ui cover V , we may find elements gi
P
gi qi = 1 on V . Then, the equation
X
X
qj (
gi mi ) = (
gi qi )mj = mj
P
holds in M . It says that mj = qj m in M where m =
gi mi . Then m is an element of M
which restricts to mj /qj on Uj for each j. In other words, s is the image of m under the
).
map M (V ; M
10.11 Definition. A sheaf M of X O-modules on an algebraic variety X is called a quasicoherent sheaf if each point of X is contained in an affine neighborhood V such that M|V
for some O(V )-module M . A quasi-coherent sheaf is called coherent if
is isomorphic to M
for each point of X this can be achieved with a module M which is finitely generated over
O(V ).
Note that the structure sheaf X O of an algebraic variety X is a coherent sheaf, since
on any affine open set V it is the localization to V of the ring O(X). It follows that direct
sums of copies of the structure sheaf are quasi-coherent and finite direct sums are coherent.
93
94
J. L. TAYLOR
Vfi , such that for each i, S|Vfi is the localization of a finitely generated module necessarily
isomorphic to Mfi . In other words, there is a finite set fi O(V ) with no common zeroes
on V and with the property that Mfi is finitely generated for each i. This implies that M
is finitely generated (Problem 10.5). This completes the proof.
is an equivalence of
10.14 Corollary. If V is an affine variety, then the functor M M
categories from the category of O(V )-modules to the category of quasi-coherent sheaves of
V O-modules on V . The functor (V ; ) is its inverse functor.
and S (V ; S)
Proof. By Theorems 10.10 and 10.13, the composition of M M
in either order is naturally isomorphic to the identity. Thus, each is an equivalence of
categories and they are inverses of one another.
Our next major result is that quasi-coherent sheaves on an affine variety are -acyclic.
This will follow easily from Theorem 10.16. But first we need to prove the following
technical lemma:
10.15 Lemma. If A is a Noetherian ring, I an injective A-module and K an ideal of
A, then the submodule J I defined by J = {x I : K n x = 0 for some n} is also
injective.
Proof. To prove that J is injective, it suffices to prove that if N M are A-modules with M
finitely generated, then every morphism N J extends to a morphism M J (Problem
10.6). Thus, suppose that : N J is such a morphism. Then since (N ) J and N
is also finitely generated, we may choose a fixed n such that (K n N ) = K n (N ) = 0. By
Krulls Theorem (Problem 10.7), there is an integer m such that K m M N K n N . Thus,
factors through N N/(K m M N ). Since I is injective, the map N/(K m M N )
J I induced by extends to a morphism : M/K m M I. However, the image of lies
in J since it is, necessarily, killed by K n . Then the composition of with M M/K n M
is the required extension of . This completes the proof.
10.16 Theorem. Let V be an affine variety. If I is an injective module over the ring
O(V ), then I is a flabby sheaf on V .
Proof. We first show that for any f O(V ) the natural map I If is surjective. To this
end, let x/f n be an element of If , with x I and n a non-negative integer. We consider
the morphism f n+1 g f gx : f n+1 O(V ) I. This is well defined since if f n+1 g is zero
then so is f g and, hence, f gx. Since I is injective, this morphism extends to a morphism
: O(V ) I such that (f n+1 g) = f gx. Then f n+1 y = f x if y = (1). However, this
implies that x/f n is the image of y under the localization map I If . Thus, I If is
(U ; I)
is surjective in the case
surjective. This proves that the restriction map (V ; I)
where U V is an open subset of the form Vf .
(U ; I)
is surjective if U is an
To complete the proof we must show that (V ; I)
If
arbitrary open subset of V . Let Y be the subvariety of V which is the support of I.
Y U = then we are through since the only section of I over U is then zero. Suppose
Y U 6= . Then there is an open set of the form Vf U such that Y Vf 6= . If
then by the first paragraph, the restriction of s to Vf is also the restriction
s (U ; I),
has its support in
to Vf of a global section t. Then s = t|U + r where r (U ; I)
95
X O(V
) is
z Y V }
is finitely generated. We claim that IY (Vf ) = If for any basic open set Vf V with
f X O(V ). In fact, an element of IY (Vf ) is a function on Vf of the form g/f n , with
g X O(V ), which vanishes on Y Vf . Then f g X O(V ) vanishes on Y V and, hence,
belongs to I. Then, on Vf we have
g/f n = gf /f n+1 If .
This proves that, on V , the ideal sheaf I is the localization I of the finitely generated ideal
I. Since X may be covered by such sets V , we have proved that the ideal sheaf is coherent.
Obviously, Corollary 10.14 implies that on an affine variety a coherent sheaf of ideals I
is the localization of an ideal, (V ; I), of O(V ) and, conversely, an ideal I O(V ) is the
96
J. L. TAYLOR
10.20 Theorem. Let V be an algebraic variety. Then the following statements are equivalent:
(i) V is affine;
(ii) H p (V ; S) = 0 for p > 0 and for all quasi-coherent sheaves S on V ;
(iii) H 1 (V ; S) = 0 for all coherent S sheaves on V .
Proof. That (i) implies (ii) is Theorem 10.17 and the implication (ii) implies (iii) is trivial;
thus, to complete the proof we must prove that (iii) implies (i).
Assume (iii) holds. The strategy is to prove that V is the space of maximal ideals of the
algebra O(V ) with the Zariski tolology and with its structure sheaf given by localization
of O(V ). Then we will prove that O(V ) is finitely generated and, hence, is the algebra
O(W ) for some affine variety. Necessarily, then, V is isomorphic to the affine variety W
by Theorem 10.7.
For x V let U be an affine open set containing x and set Y = V U . Consider the
exact sequence of sheaves
0 IY {x} IY C{x} 0
where IY {x} and IY are the ideal sheaves in V O of the subvarieties Y {x} and Y ,
respectively and C{x} is the skyscraper sheaf which is C at x and zero at all other points.
Since the ideal sheaf IY {x} is coherent, H 1 (V ; IY {x} ) = 0 and the long exact sequence
of cohomology implies that
0 (V ; IY {x} ) (V ; IY ) (V ; C{x} ) 0
is also exact. This just means that there is a function f O(V ) which vanishes on Y
and does not vanish at x. In other words, every affine neighborhood U of x contains a
neighborhood of x of the form Vf where f is a global section in O(V ). This implies, in
particular, that the functions in O(V ) separate points in V . It also implies that the open
sets Vf form a basis for the topology of V and, hence, that V has the Zariski topology
determined by the algebra O(V ) that is, the topology in which the closed sets are the
common zero sets of finite sets of functions from O(V ). It also implies that we may cover
V with a finite collection of affine open sets {Vfi }ni=1 with fi O(V ) for each i.
We claim that every maximal ideal of the algebra O(V ) has the form id({x}) = {f
O(V ) : f (x) = 0} for some point x of V . In fact if M is a maximal ideal which does
not have this form then there is no point of V at which all functions in M vanish. This
means the the collection on open sets {Vf : f M } covers V and this, in turn, implies
that some finite subcollection covers V by Problem 10.3. In other words, there is a finite
set {fi } M with no common zero on V . Given any such set {fi } consider the sheaf
morphism
n V O V O
P
defined by (g1 , . . . , gn ) i gi fi . This map is surjective, since at each x in V some fi
is non-vanishing and, hence, invertible in some neighborhood. The kernel of this map is
coherent (Problem 10.9) and, thus, has vanishing first cohomology. It follows from the long
97
10. Problems
1. Prove that the product V W of two prevarieties and the projection morphisms V :
V W V and V : V W V have the following universal property: For each
prevariety Q and pair of morphisms f : Q V and f : Q W there is a morphism
h : Q V W such that f = V h and g = W h.
98
J. L. TAYLOR
2. Prove that an algebraic prevariety V is an algebraic variety if and only if the diagonal
is closed in V V .
3. Prove that every open cover of an algebraic variety has a finite refinement.
4. Prove that if V is a subvariety of an open set in Cn , W is a subvariety of an open set
in Cm and f : V W is algebraic (has coordinate functions which are regular) then f
is a morphism of ringed spaces.
5. Prove that if V is an affine variety, {fi } a finite set of elements of O(V ) which have no
common zero on V and M is an O(V )-module such that Mfi is finitely generated for
each i, then M is finitely generated.
6. Use a trascendental induction argument to prove that I is injective if and only if for
every singly generated module M and every submodule N M , any morphism N I
has an extension to M .
7. Prove Krulls Theorem: If A is a Noetherian ring, K an ideal of A, M a finitely generated
A-module and N M a submodule, then for each positive integer n there is a positive
integer m such that K m M N K n N .
8. A sheaf supported on a single point is called a skyscraper sheaf. Prove that every
skyscraper sheaf is flabby.
9. Prove that the kernel, image and cokernel of a morphism between quasicoherent (coherent) sheaves on an algebraic variety is also quasi-coherent (coherent).
99
X i1 ...ip
i
xi
It follows from the Poincare Lemma that E = {E P , dp } is a complex which provides a fine
resolution C E of the constant sheaf C.
If U is an open set in Cn , then we consider it as an open subset of the real vector
space R2n with basis {x1 , . . . , xn , y1 , . . . , yn }. However, instead of the usual bases for the
complexified tangent and cotangent spaces we use the basis consisting of
dzi = dxi + i dyi ,
d
zi = dxi i dyi ,
i = 1, . . . , n
100
J. L. TAYLOR
= 1/2
i
,
= 1/2
+i
,
zi
xi
yi
zi
xi
yi
i = 1, . . . , n
for the complexified tangent space. Note that these are dual bases to one-another. In terms
of the above basis for the cotangent space, a differential form in E r (U ) may be written as
X
ji ...jp k1 ...kq dzj1 dzjp d
zk1 d
zkq
p+q=r
with coeficients ji ...jp k1 ...kq C (U ). If we let E p,q (U ) denote the differential forms in
E p+q (U ) which are of degree p in the dzi and of degree q in the d
zi , then we have a direct
sum decomposition:
X
E p,q (U ).
E r (U ) =
p+q=r
p,q
Forms in the space E (U ) are said to have bidegree (p, q) and total degree p + q.
When restricted to E p,q , exterior differentiation defines a map
d : E p,q E p+1,q + E p,q+1 .
If we define and to be the operators which, on forms of bidegree (p, q), act as d followed
by projection on E p+1,q and E p,q+1 , respectively, then
: E p,q E p+1,q , : E p,q E p,q+1
and
d = + .
d( ) = d + (1)r d,
it follows that
( ) = + (1)r ,
) =
+ (1)r
X f
d
zi
zi
i
101
11.1 Definition. For a domain U in Cn the pth Dolbeault complex is the complex
f ()
d
f () d d
d f ()
=
d d =
z
z
z
Thus, if r is the boundary of the disc D(z, r) and if r is chosen small enough that this
disc is contained in U , then Stokes Theorem implies that
ZZ
Z
Z
f () d d
d
d
=
f
()
f
()
z
z
z
r
Ur
where Ur = U D(z, r). Now ( z)1 is integrable on any bounded region of the plane,
as is easily seen by integrating its absolute value using polar coordinates centered at z.
Thus,
ZZ
ZZ
f () d d
f () d d
lim
=
.
r0
z
z
Ur
Also,
d
lim
= lim
f ()
r0
z r0
r
The result follows.
Z
0
102
J. L. TAYLOR
is defined for all z C and defines a function g C (C). We calulate the derivative g/ z
of g using the change of variables + z:
ZZ
1
d d
g(z) =
f ( + z)
z
2i z
ZZ
ZZ
1
f ( + z) d d
1
f ( + z) d d
=
=
2i
z
2i
ZZ
f () d d
1
=
= f (z)
2i
z
where the last line follows from reversing the change of variables and using the generalized
Cauchy integral theorem on U (recall that f vanishes on the complement of a compact
subset of U and so the line integral in Theorem 10.2 vanishes). Thus, g/ z = f on all of
C. Of course, we modified f on the complement of V and so this equation holds for our
original f only on V , but this is what was to be shown.
If K Cn is compact, we denote by E p,q (K) the space (K; E p,q ) of C forms of
bidegree (p, q) defined in a neighborhood of K. The Dolbeault cohomology Hp.q (K) for K
0 =
zk
103
d
If j > k then no cancellation can occur between terms of
zk involving d
zj and
terms of involving d
zj . It follows that such terms individually vanish and, hence, the
coeficients of and are holomorphic in the variables zk+1 , , zn . Now it follows from
which
Theorem 11.3 that if f is a coeficient of then f = g/ zk for some g C ()
is also holomorphic in the variables zk+1 , , zn . That the solution g given by Theorem
10.3 is actually C in all the variables and not just in k and that it is holomorphic in
the variables zk+1 , , zn follows from the fact that these things are true of f and the
solution g is given in terms of f by an explicit integral formula which commutes with the
differential operators in question. By replacing each coeficient f of by the corresponding
g as above, we obtain a (p, q 1)-form with the property that
= + d
zk
=
where is a form involving only the conjugate differentials d
z1 , , d
zk1 . Then
=
for
( ) = 0. Thus, by the induction hypothesis, we conclude that
p,q1
p,q1
some E
(), from which it follows that = with = + E
(). This
completes the proof.
be a polydisc. It
Note that it was not really important in the above argument that
was important that it be a Cartesian product that is, a set of the form K1 K2 Kn
for some collection of compact sets Ki C. This is due to the fact that the solution was
obtained by applying Theorem 11.3 in each variable separately while treating the other
variables as parameters.
The following theorem concerns an open polydisc = (, r). Note that we allow
some or all of the radii ri to be infinite, Thus, Cn itself is an open polydisc.
11.5 Theorem. Let be an open polydisc. Then Hp,q () = 0 for q > 0 and for all p.
j j+1
Proof.
of open polydiscs with compact closure such that
S Let j be a sequence
p,q
p,q1
() such
and j j = . If E () and = 0 then we will construct E
j = on j and j+1 |j = j . Clearly this will give the desired result, since we can
then define a solution E p,q1 () by |j = j . Suppose the sequence {j } has been
constructed with the above properties for j < k. Then we use Dolbeaults Lemma to find
= in a neighborhood of
k ) such that
k . We then have
E p,q1 (
k1 ) = 0
(
k1 and, since q > 1, we may apply Dolbeaults Lemma again to
in a neighborhood of
p,q2
= k1 in a neighborhood of
k1 . By multiplying
find E
(k1 ) such that
k1 and has compact support in
by a C function which is one in a neighborhood of
k1 , we may assume that is actually
an appropriate slightly larger neighborhood of
p,q1
p,q2
in E
(k ). Then k = E
(k ) gives the required next function in our
sequence since
k = (
)
=
104
J. L. TAYLOR
105
11.8 Corollary. If X is a complex manifold of dimension n, then for each p the sheaf
cohomology H q (X; Hp ) vanishes for q > n.
11. Problems
1. Give a coordinate free definition of the space E p,q .
2. Prove that if is an open polydisc and f H() then a function g H() belongs
to the ideal generated by f in H() if and only if its germ at belongs to the ideal
generated by the germ of f at in H for each f 1 ({0}).
3. Prove Hartogs extension theorem: If K is a compact subset of an open set U Cn ,
U K is connected and n > 1, then each function f which is holomorphic in U K
extends to be holomorphic in U . Hint: Let K 0 be chosen so that K 0 is a compact set
containing K in its interior, K 0 U and U K 0 is connected. Then multiply f by a C
function which is zero in a neighborhood of K and one in a neighborhood of U K 0 .
The resulting function then extends to a C function g in U which agrees with f on
the connected set U K 0 . Now use Corollary 11.7 and Problem 1.5 to show that you
can find a C function h on U which vanishes on U K 0 and is such that g h is
holomorphic in all of U .
4. Prove that H q (U, Hp ) = 0 for q > 0 and for any open subset U C. Hint: use Theorem
11.3 and an approximation argument like that used in the second half of the proof of
Theorem 11.5.
5. Prove that H q (P 1 (C), Hp ) = C p,q , where P 1 (C) is the Riemann sphere. Hint: Cover
the sphere with two open disks which overlap in an annulus. By the previous problem,
106
J. L. TAYLOR
Om V On S|V 0
12.3 Okas Theorem. Let U Cn be an open set. Then the kernel of any
morphism U Hm U Hk is locally finitely generated.
107
U H-module
Hp
Hm
Hk1
with the top row exact. We also have, by assumption, that the kernel of : V Hp V H is
locally finitely generated. This means that, after shrinking V if necessary, we can find q > 0
and a morphism : V Hq V Hp which maps onto ker . But ker = ker ker =
(ker ) = im . Thus, the kernel of is also locally finitely generated as was to be
shown.
Thus, we have reduced the proof to showing, in dimension n, that the kernel of a
morphism of the form : U Hm U H is locally finitely generated, under the assumption
that the theorem holds for all m and k in dimensions less than n. The strategy of the proof
is to use the Weierstrass theorems to reduce the problem to an analogous one involving
polynomials of a fixed degree in zn and then to apply the induction assumption to the
coeficients of these polynomials.
Given a point x U we must show that ker is finitely generated in some neighborhood
of x. Without loss of generality, we
Pmay assume that the point x is the origin. The map
has the form (g1 , . . . , gm ) =
fi gi for an m-tuple of functions fi n H(U ). By
appropriate choice of coordinates, we may assume that the germ at 0 of each fi is regular
of some degree at 0 and, hence, by the Weierstrass preparation theorem, has the form ui pi
where pi is a Weierstrass polynomial and ui is a unit. We may replace these germs by
their representatives in some neighborhood V of 0 and by choosing V small enough we may
assume the ui are non vanishing in V . Then the map (g1 , . . . , gm ) (u1 g1 , . . . , um gm )
is an automorphism of V Hm which maps the kernel of to the kernel of the morphism
determined by the m-tuple (p1 , . . . , pm ). Thus, without loss of generality,Pwe may replace
the fi s with the pi s and assume that has the form (g1 , . . . , gm ) =
pi gi . Let d be
the maximum of the degrees of the pi s. We may assume that V has the form V 0 V 00 for
open sets V 0 Cn1 and V 00 C. Let Kd denote the sheaf on V 0 defined as follows
Kd (W ) ker : V Hm (W V 00 ) V H(W V 00 )
108
J. L. TAYLOR
is the subspace consisting of m-tuples whose entries are polynomials of degree less than or
equal to d in the variable zn with coeficients in n1 H(W ). This is a sheaf of V 0 H-modules.
We shall show that it is locally finitely generated on V 0 . Now for each neighborhood
W V 0 , the space of m-tuples (q1 , . . . , qm ) where each qi is a polynomial in zn of degree
at most d, with coeficients which are in n1 H(W ) is a free module of rank m(d + 1) over
n1 H(W ). The map, determines a n1 H(W )-module morphism of this free module into
the free n1 H(W )-module of rank 2d + 1 consisting of polynomials of degree at most 2d
in zn with coeficients in n1 H(W ). Furthermore, Kd (W ) is the kernel of this morphism.
In other words, we may regard as determining a morphism of sheaves of V 0 H-modules,
m(d+1)
V 0 H2d+1 and our sheaf Kd is its kernel. By the induction hypothesis, Kd is
V 0H
locally finitely generated as a sheaf of V 0 H-modules. To complete the proof, we will show
that Kd generates ker as a sheaf of V H-modules.
Let be a point of W V 00 . We must show that the stalk of Kd at generates the
stalk of ker at . By performing a translation, we may assume that the point is the
origin. In the process, however, we lose the fact that the polynomials pi are Weierstrass
polynomials at 0. They are, however, still polynomials in zn and, by the Weierstrass
preparation theorem, we may factor p1 as p1 = p01 p001 where the germ at 0 of p01 is a
Weierstrass polynomial and the germ at 0 of p001 is a unit. We set
d01 = deg p01 ,
and note that
If h = (h1 , . . . , hm ) n H0m we then use the Weierstrass division theorem on each component of h to write h = p01 h00 + r0 , where r0 , h00 n H0m and r0 has components which are
polynomials in zn of degree less than d01 . If we set h0 = (p001 )1 h then
h = p1 h0 + r0
Also,
0
p1 h = q +
m
X
h0j ej
j=2
where
q=(
pi h0i , 0, . . . , 0)
and ej = (pj , 0, . . . , p1 , 0. . . . , 0)
with p1 occuring in the j th position of ej . Note that each ej belongs to (Kd )0 . Thus, if
we set r = r0 + q, we have
m
X
h=r+
h0j ej
j=2
109
which implies that p01 p001 r1 = p1 r1 is a polynomial in zn of degree less than d + d01 since this
is true of each term on the right above. However, since p01 is a Weierstrass polynomial, the
Weierstrass division theorem implies that p001 r1 must be a polynomial of degree at most d.
Then the entries of p001 r all have degree at most d and we conclude that p001 r (Kd )0 . Since
p001 is a unit we conclude that r and, thus, h belongs to the submodule of n H0 generated by
Kd . Thus, we have shown that at every point of V the stalk of Kd generates the stalk of the
kernel of . Since, Kd is itself locally finitely generated over V 0 O, the proof is complete.
12.4 Corollary. If U is an open set in Cn and M is a locally finitely generated sheaf of
submodules of U Hk , then M is coherent. In particular, if : U Hm U Hk is a morphism
of analytic sheaves , then ker and im are coherent analytic sheaves on U .
Proof. Since M is locally finitely generated, for each point of U there is a neighborhood
V on which M is the image of a morphism : V Hm V Hk of sheaves of V H-molules.
By Okas Theorem we know that ker is locally finitely generated. Thus, by shrinking V
if necessary, we may assume ker is the image of a morphism : V Hp V Hm . Then, on
V , M is the cokernel of . Thus, M is coherent.
Since the image of a morphism : U Hm U Hk of analytic sheaves is finitely generated
by definition and the kernel is locally finitely generated by Okas Theorem, they are both
coherent by the result of the previous paragraph.
12.5 Corollary. If U is an open set in Cn and M and N are coherent sheaves of submodules of U Hm , then so is M N .
Proof. Every point of U has a neighborhood V on which there are morphisms : V Hp
m
with image M|V and : V Hq V Hm with image N |V . Consider the map :
VH
p+q
V Hm defined by writing V Hp+q as V Hp V Hq and setting (f g) = (f )(g).
VH
The kernel of is coherent by Okas Theorem and, hence, is locally finitely generated.
Furthermore, on V , M N is the image of the kernel of under . Thus, after shrinking
V if neccessary, we may choose a finite set of generators for ker . Then the image of this
set under will generate M N on V . Thus, M N is locally finitely generated. In view
of the previous corollary, this proves that it is coherent.
If U Cn is an open set and I and J are sheaves of ideals of U H then I : J will denote
the sheaf which assigns to the open set V U the ideal I(V ) : J (V ) = {f n H(V ) :
f J (V ) I(V )}.
12.6 Corollary. If U is an open set in Cn and I and J are coherent ideal sheaves, then
so is I : J .
Proof. First, suppose that J is generated by a single element h n H(U ) and I is generated
by the elements g1 , . . . , gk n H(U ). Then consider the map : U Hk+1 U H defined by
(f0 , . . . , fk ) = hf0 g1 f1 gk fk
Then I : J is the image of the kernel of under the projection (f0 , . . . , fk ) f0 :
k+1
U H. The kernel of is locally finitely generated by Okas Theorem and, thus,
UH
so is I : J . It is then coherent by Corollary 12.4.
110
J. L. TAYLOR
For the general case, for any point of U we may choose a neighborhood V in which J
is finitely generated, say by h1 , . . . , hm . Also, on V , I : J is just the intersection of the
sheaves I : Ji , where Ji is the sheaf of ideals on V generated by hi . Thus, it follows from
the previous corollary that I : J is coherent.
We will prove that the ideal sheaf IY of a subvariety Y U of an open subset of Cn is
coherent. The key to the proof is the following lemma:
12.7 Lemma. Let U be an open set in Cn containing the origin. Let f1 , . . . , fp be a set of
holomorphic functions on U with Y as its set of common zeroes and I as the ideal sheaf it
generates. Suppose that I0 is a prime ideal of n H0 , and I = id Y at all points Y Z
where Z is a holomorphic subvariety of Y with Z0 6= Y0 . Then there is a polydisc U
centered at 0 such that I = id Y at all points .
Proof. Since the
functions f1 , . . . , fp determine Y , it follows from the Nullstellensatz that
I0 id Y
0 I0 . However, by assumption I0 is prime and, hence, we have that I0 =
id Y0 = I0 . Let d be a function holomorphic in a polydisc centered at 0 such that
its germ d0 at 0 belongs to id Z0 but not to id Y0 . We may assume (by shrinking if
neccessary) that d vanishes on Z but does not vanish identically on Y . By Corollary
12.6, the sheaf I : n Hd on is locally finitely generated and, hence, we may assume it is
finitely generated by shrinking . Let g1 , . . . , gq n H() be a set of generators for this
sheaf. Then, dgj I() and, in particular, its germ at 0 belongs to I0 . However, this is a
prime ideal and it does not contain the germ of d. Hence, we must have that the germ of
gj at 0 belongs to I0 for every j. That is, the germs at 0 of each gj belong to the ideal at
0 generated by the fi . If this is true at 0, it is true in a neighborhood of 0 and, hence, we
may as well choose small enough that it is true at every point of . This implies that
I : n H0 d = I
on
111
Corollary 5.18 the subvariety Y is actually a complex manifold outside a proper subvariety
Z. Furthermore, it can be seen from the proof of Lemma 5.13 and Theorem 5.17 that
the functions pm+1 , qm+2 , . . . , qn can be taken as the last n m coordinate functions of a
coordinate system near any point of Y Z a coordinate system in which Y is expressed
as the set where these last n m coodinates vanish. It follows that the functions f1 , . . . , fp
generate the germ of id Y at such points. It then follows from the previous lemma that
these functions generate (IY ) = id Y at all points of some polydisc containing 0. Thus,
IY is coherent.
If the germ Y0 is not irreducible, then we write Y = Y1 Yq in some neighborhood
of 0, where the varieties Yj have irreducible germs at 0. Then IYj is finitely generated in
some neighborhood of 0 for each i by the previous paragraph. We may choose to be a
neighborhood in which this is true for all j. Then each ideal sheaf IYj is coherent on
and, hence, so is the intersection IYj IYq by Corollary 12.5. But this intersection is
just IY and, thus, the proof is complete.
We are now in a position to prove the strong form of Okas Theorem:
12.9 Okas Theorem on Varieties. Let : X Hm X Hk be a morphism of analytic
sheaves on an analytic variety X. Then ker is coherent.
Proof. This is a local result and so we may assume that X is a subvariety of an open
set U contained in Cn . The morphism is determined by a k m-matrix with entries
which are holomorphic functions on X, and, hence, extend locally to be holomorphic in
neighborhoods in U . Again, since we are proving a local result, we may as well assume that
the entries of this matrix extend to be holomorphic in U . Then extends to a morphism
: U Hm U Hk of analytic sheaves. Furthermore, by the previous result, we know that
the ideal sheaf IX is locally finitely generated and, hence, we may as well assume that it is
finitely generated on U . Then we may represent it as the image of a U H-module morphism
: U Hp U H. Let k : U Hkp U Hk denote the morphism that is just the direct sum
of k copies of . Then consider the morphism
: U Hm U Hkp U Hk
where
) k (g)
(f, g) = (f
112
J. L. TAYLOR
113
m
m
S
UH
UH
x
x
x
UH
UH
UH
UH
UH
UH
The top row of this diagram is exact and the vertical maps are all surjective. It evident
from the diagram that the kernel of is 1 (im ) = im . Thus, by definition, coker is
coherent.
It remains to prove that ker is coherent. To this end, we fix x X and choose a
neighborhood U of x small enough that we may find surjective morphisms of analytic
114
J. L. TAYLOR
sheaves : U Hm M and : U Hn N such that ker and ker are finitely generated
and, thus, coherent. We then construct the following commutative diagram after shrinking
U appropriately:
0 ker |U M|U N |U
x
x
UH
UH
UH
UH
UH
Here we use Lemma 12.12 to construct the map . Its image is finitely generated and, hence,
coherent. Then ker im is coherent and, hence, is the image of a morphism as above.
Then is obtained as another application of Lemma 12.12. The map is a morphism
of analytic sheaves which maps onto the kernel of and it exists, for small enough U , by
Corollary 12.10(ii). Now a simple diagram chase shows that (+) : U Hp U Hq M|U
has ker |U as its image. This completes the proof that ker is locally finitely generated
and, hence, coherent.
The above theorem implies that the category of coherent analytic sheaves is an abelian
category.
Finally, we have the following result:
12.16 Theorem. If
0 K M N 0
is an exact sequence of analytic sheaves and if any two of the three are coherent, then the
third is also coherent.
Proof. Two of the three cases have already been proved in the preceding theorem. It
remains to prove that if K and N are coherent then M is coherent. To this end, let x be
a point of X and U a neighborhood of x for which we may find surjective morphisms of
analytic sheaves : U Hk K|U and : U Hn N |U with finitely generated kernels. We
may then use Lemma 12.12 to construct the following commutative diagram with exact
rows:
0 K|U M|U N |U 0
x
x
x
0 U Hk U Hk+n U Hn 0
Here the morphisms on the bottom row are just the canonical injection and projection
associated with writing U Hk+n as U Hk U Hn . The morphism is constructed, for
sufficiently small U , by lifting to a morphism 0 : U Hn M|U with 0 = using
Lemma 12.12, then writing U Hk+n as U Hk U Hn and defining by
(f g) = (f ) + 0 (g).
115
Now the fact that and are surjective implies that is also surjective. Also, the kernels
of the vertical maps form a short exact sequence
0 ker |U ker |U ker |U 0
in which the first and third terms are locally finitely generated and, hence, coherent. Thus,
we may repeat the above argument for this sequence and conclude that, for sufficiently
small U , ker |U is also finitely generated. It follows that M is coherent and the proof is
complete.
12. Problems
1. Prove Corollary 12.14.
2. Prove that if : X Hk X Hm is a morphism of analytic sheaves then x : X Hxk X Hxm
is surjective if and only if the matrix of holomorphic functions on X defining has rank
m at x.
3. Use the result of the preceding problem to prove that if M is a coherent analytic sheave
on a holomorphic variety X, then Support(M) = {x X : Mx 6= 0} is a holomorphic
subvariety of X.
4. Use the result of the preceding problem to prove that if : M N is a morphism of analytic sheaves between two coherent analytic sheaves, then {x X : x is not injective}
and {x X : x is not surjective} are holomorphic subvarieties of X.
116
J. L. TAYLOR
zi1 zi+1
zn
z0
,...,
,
,...
i (z0 , . . . , zn ) =
zi
zi
zi
zi
The map i is clearly well defined and, in fact:
117
Thus, the image of Ui Z(p) under 0 is a subvariety of Cn that is, a closed set. It
follows that the image of every closed subset of Ui is closed in Cn .
The closed subsets of Cn are finite intersections of zero sets of polynomials. Let q be
a polynomial in z1 , . . . , zn of degree k. Then we may define a homogeneous polynomial of
degree k in z0 , . . . , zn by
p(z0 , z1 , . . . , zn ) = z0k q(z1 /z0 , . . . , zn /z0 )
and p is zero exactly at points which 0 maps to zeros of q. That is, Z(p) U0 is the
inverse image under 0 of the zero set of q. It follows that the inverse image under 0 of
any closed set in Cn is a closed set in U0 .
To finish the proof, we must show that 0 induces an isomorphism between the structure
sheaves of U0 and Cn . This amounts to showing that, for each open set V U0 , a complex
valued function on V has the form f for a homogeneous regular function f on 1 (V )
if and only if it has the form g 0 for a regular function g on 0 (V ). In other words,
for each open set W = 0 (V ), we must show that g g 0 is an isomorphism from
the regular functions on W to the homogeneous regular functions on 1 (W ). Since 0
is algebraic and homogeneous it is clear that g g 0 is a ring homomorphism from
O(W ) to regular homogeneous functions on 1 (W ). To see that it is an isomorphism, we
simply note that its inverse is given by f f where f(z1 , . . . , zn ) = f (1, z1 , . . . , zn ). This
completes the proof.
13.6 Theorem. The ringed space P n is an algebraic variety.
Proof. We have that, {Ui } is a cover of P n by open subsets which are isomorphic as ringed
spaces to Cn . Thus, P n is an algebraic prevariety It is also clear that, given any two points
p and q of P n , we may choose our coordinate system in Cn+1 in such a way that one of
the Ui contains both p and q. In other words, given any two points, there is an affine open
subset containing both. By Theorem 10.6 this implies that P n is actually an algebraic
variety.
13.7 Definition. A projective variety is an algebraic variety which is isomorphic to a
closed subvariety of P n for some n.
To any algebraic variety we may associate in a canonical way a holomorphic variety. An
algebraic variety is locally isomorphic, as a ringed space, to an algebraic subvariety of Cn .
There is a canonical way to associate to an algebraic subvariety V of Cn an holomorphic
subvariety V h we simply give V the Euclidean topology instead of the Zariski topology
118
J. L. TAYLOR
and let its structure sheaf be the sheaf of holomorphic functions rather than the sheaf
of regular functions. This means that for any algebraic variety we have a canonical way
of changing the topology and ringed space structure on any affine open subset in such
a way as to make it a holomorphic variety. This is canonical because any two ways of
representing an affine open subset as subvarieties of complex Euclidean space are related by
a biregular map between the two subvarieties which will necessarily be a biholomorphic
map between the associated holomorphic subvarieties. In particular, this implies that
holomorphic space structures defined in this way on affine subsets of an algebraic variety
will agree on intersections and, thus, define a global structure of a holomorphic variety.
One thing that does need to be checked is that the resulting topological space is Hausdorff
(Problem 13.2).
From the above discussion, we conclude that projective space P n may also be considered
as a holomophic variety in fact, as a complex manifold, since the maps i : Ui Cn give
local biholomorphic maps onto Cn . We initially defined regular functions on U P n as
just regular functions on 1 (U ) which are homogeneous of degree zero. It turns out that
the analogous statement is true of holomorphic functions:
13.8 Definition. For each integer k we define a sheaf H(k) on P n with the Euclidean
topology as follows: If U P n is open in the Euclidean topology, then H(k)(U ) is the
space of functions in H( 1 (U )) which are homogeneous of degree k.
13.9 Theorem. The sheaf H(0) is a sheaf of rings canonically isomorphic to the structure
sheaf H of P n and the sheaves H(k) are sheaves of H-modules.
Proof. Clearly H(0) is a sheaf of rings and each H(k) is a sheaf of modules over H(0).
Thus, we need only show that H(0) is canonically isomorphic to H. The isomorphism is
obviously the one which sends f H(U ) to f . It is easy to see, using the definition of
the holomorphic structure on P n , that : Cn 0 P n is holomorphic, so that f f
is an homomorphism (obviously injective) of H to H(0). It only remains to show that this
is surjective. If g is a homogeneous holomorphic function on an open set 1 (U ) then
f ((z)) = g(z) certainly defines a function f on U . The only question is whether or not it
is holomorphic. It suffices to prove this in the case where U Ui for some i and, without
loss of generality, we may assume that i = 0. Then to prove that f is holomorphic, we must
1
n
show that f 1
0 is holomorphic on 0 (U ) C . But f 0 (z1 , . . . , zn ) = g(1, z1 , . . . , zn )
and this is certainly holomorphic.
We now turn to the study of the sheaves O(k) and H(k) introduced above. The sheaves
O(k) are sheaves of O = O(0) modules. If V is an open subset of Ui then it is easy to
see that f O(V ) if and only if zik f O(k)(V ). Thus, f zik f defines an O|Ui -module
isomorphism from O|Ui to O(k)|Ui . In other words, O(k) is locally free of rank one as
a sheaf of O-modules. Exactly the same thing is obviously true of H(k) as a sheaf of H
modules. A sheaf of modules with this property is called an invertible sheaf since such
a sheaf always has an inverse under tensor product relative to the structure sheaf. In
this case, the inverse of O(k)(H(k)) is obviously O(k)(H(k)) in view of the following
theorem:
13.10 Theorem. If j and k are integers then multiplication defines an isomorphism
O(j) O O(k) O(j + k).
119
Lerays theorem implies that we may compute the cohomology of H(k) or T0 using Cech
cohomology for the cover {Ui }.
120
J. L. TAYLOR
each permutation and each multi-index . This is equivalent to the definition we used
earlier the proof of this may be found in nearly any standard text in which Cech
Theory
is developed (or do Problem 13.1).
in which am0 ...mn is allowed to be non-zero only for terms such that either mi 0 or
mi < 0 and i is one of the ij appearing in . For such an expansion to converge on U ,
the coeficients am0 ...mn must decay faster than r|m0 |+...|mn | for every positive number r.
In other words, we may identify T (U ) with the space of functions m am : Z n+1 C
which decay at infinity faster than any geometric series and which are supported on
{m = (m0 , . . . , mn ) : mi 0 if
i
/ {i0 , . . . , ip }}
121
Also, the same statements are true with O(k) replaced by H(k).
for i
/ {i0 , . . . , in }}.
the cochains with support in L. Thus, PL also defines a projection operator on Cech
cohomology as well. In particular, the cohomology of O(q) can be obtained from
P that of S
by applying the projection PLk determined by the set Lk = {(m0 , . . . , mn ) :
mi = k}.
Also, it makes sense to talk about an element of cohomology of S being supported in a set
L that is, H p (P n , S) is supported in L if PL = . With this in mind, the strategy
of the proof will be to prove that, for p > 0, every element of H p (P n , S) is supported in
the set K = {(m0 , . . . , mn ) : mi < 0 i}.
We know that H p (Uj , S) = 0 for p > 0, since Uj is affine and S is quasi-coherent (it
is an infinite direct sum of coherent sheaves). In terms of our Fourier coeficient picture,
this means that, for a fixed j and p > 0, any p-cocycle a = {a
m0 ,...,mn } for S and {Ui } is,
when restricted to Uj , the coboundary of some p 1-cochain b = {bm0 ,...,mn } for S and
{Uj Ui }i . Now generally b will have some non-zero coeficients with negative index mj
and a which does not include j among its entries. This, of course, just means that b
is not generally a cochain for {Ui }. However, if Wj = {m Z n+1 : mj 0} then PWj b
is a cochain for {Ui } which is mapped by the coboundary onto PWj a. This implies that
PWj kills elements of H p (P n , S) for p > 0 and, hence, that H p (P n , S) is supported on the
complement of Wj . Since this is true for every j, we conclude that H p (P n , S) is supported
on K as claimed. This implies that H p (P n , S) = 0 for 0 < p < n since every p-cochain
for p < n is supported on the complement of K. We already know that H p (P n , S) = 0 for
p < 0 and for p > n. This completes the proof of (c) in the algebraic case, since we know
that we can obtain H p (O(k)) as a direct summand of H p (P n , S).
For an n-cochain, the only multi-index we need consider is = (0, 1, . . . , n), since all
others without repeated entries are permutations of this one. For this index the set
W is all of Z n+1 . Thus, an n-cochain for S and {Ui } is just a single coeficient function
a = am0 ,...,mn with no restriction other than that only finitely many coeficients are nonzero. Now we know from the previous paragraph that H n (P n , S) is supported on K.
However, every n- cochain is a cocycle and the space of (n 1) cochains is supported in
the complement of K. Thus, it follows that H n (P n , S) is actually isomorphic to the space
of coeficient functions supported in K. This proves
P (a) since an index (m0 , . . . , mn ) which
mi = n 1 if and only if each mi is
belongs to K, that is mi < 0 i, can satisfy
n1
n
1. Thus, H
(P , O(k)) consists of functions with a single non-vanishing coeficient
corresponding to the index (1, . . . , 1). This proves part (a).
122
J. L. TAYLOR
We have from the previous theorem that H 0 (P n , S) may be regarded as the space of
finitely non-zero coeficient functions which are supported in K+ = {(m0 , . . . , mn ) : mi
0 i} and from the previous paragraph, that H n (P n , S) may be regarded as the space of
finitely non-zeroP
coeficient functions which are supported on K. Clearly reflection through
the hyperplane
mi = (n 1)/2 is a bijection between K+ and K. This pairs basis
elements for H 0 (P n , S) and H n (P n , S). Clearly it pairs functions supported on Lk with
functions supported on Lnk1 and, thus, it pairs basis elements for H 0 (P n , O(k)) and
H n (P n , O(n k 1k)) as require in the theorem. This completes the proof of part (a)
in the algebraic case.
The argument is almost the same in the analytic case, where O(k) is replaced by H(k).
The difference is that we must work with T0 rather than S. However, T0 is just the direct
sum of the H(k) and, hence, also has vanishing cohomology on each Ui . The rest of the
argument is the same if one observes that the projection operators which play such a
prominent role preserve the faster than exponential decay at infinity that defines coeficient
functions in T (U ).
Since we use the same open cover by Zariski open sets to compute both, there is a map
H p (P n , O(k)) H p (P n , H(k)) defined by inclusion. Clearly, the above theorem has as a
consequence:
13.14 Corollary. The map H p (P n , O(k)) H p (P n , H(k)) is an isomorphism for every
p and every k.
We complete this section with some results which show how to make strong use of the
sheaves O(k).
13.15 Definition. If F is a coherent algebraic sheaf on P n then we set F(k) = F O O(k).
The sheaf F(k) is said to be obtained from F through twisting it by O(k).
Note that k F(k) is a sheaf of modules over the sheaf of rings S = k O(k) where the
product hf of h O(k) and f F (m) is h f F(k + m).
Note also, that, since O(k) is locally free of rank one, F(k) is locally isomorphic to
F, but it is not generally globally isomorphic to F because of the twist introduced by
tensoring with O(k).
13.16 Theorem. If F is a coherent algebraic sheaf on P n then for some k 0, the sheaf
F(k) is generated by finitely many of its global sections.
Proof. For each i the open set Ui is affine and, hence, the coherent sheaf F|Ui is the image
under localization of (Ui , F) which is finitely generated. Let {fij }j be a finite generating
set of sections for (Ui , F). Now zi is a global section of O(1) which vanishes exactly on
the complement of Ui . For some integer m, which may be chosen large enough to work
for all i, j, the product zim fij extends to be a global section of F(m) (problem 13.4). The
resulting collection of global sections {zim fij }ij clearly generate F(m) on each Ui and,
hence, on all of P n .
13.17 Theorem. If F is a coherent algebraic sheaf on P n , then F is the quotient of a
sheaf which is a finite direct sum of sheaves of the form O(k)
123
Proof. By the previous theorem, there exists a non-negative integer m such that F(m)
is generated by finitely many global sections. If there are r of these sections then they
determine a surjection Or F(m). On tensoring this with O(m), we get a surjection
O(m)r F as required.
13.18. If F is a coherent algebraic sheaf on P n then
(a) H p (P n , F) is finite dimensional for each p and vanishes for p > n;
(b) there exists an integer m0 such that H p (P n , F(m)) = 0 for all p > 0 and all
m > m0 .
Proof. That H p (P n , F) vanishes for p > n is just the fact that we can use alternating
Cech
cochains for the cover {Ui } to compute it. Note that this implies that the m0 in part
(b) can be chosen independent of p if it can be chosen depending on p since there are only
finitely many ps to worry about. With these things in mind, the proof is by reduction on
p. We assume both statements are true for p + 1 n + 1 and then prove they are also true
for p. We express F as a quotient of a finite direct sum i O(ki ) so that we have a short
exact sequence
0 K i O(ki ) F 0
Then part(a) for p follows from the long exact sequence for coholomogy associated to
this short exact sequence, the assumption that the theorem is true for p + 1 and Theorem
13.13. Part(b) follows in a similar fashion from Theorem 13.13 and the long exact sequence
associated to the short exact sequence
0 K(m) i O(m + ki ) F(m) 0.
obtained by twisting by the sheaf O(m). This completes the proof.
13. Problems
124
J. L. TAYLOR
The second equality above follows easily from the definitions of direct limit and tensor
product (Problem 14.1).
The next theorem is a direct application of our work on faithful flatness in Chapter 7.
125
126
J. L. TAYLOR
14.5 Theorem. If i : Y X is an embedding of a holomorphic variety Y in an holomorphic variety X and M is a coherent analytic sheaf on Y , then i M is a coherent analytic
sheaf on X. The analogous statement is true for an embedding of algebraic varieties and
a coherent algebraic sheaf.
Proof. . We prove this in the holomorphic case. The proof in the algebraic case is even
easier.
If M is a coherent analytic sheaf on Y then for each y Y there is a neighborhood
V Y of y and an exact sequence
Hk
Hm M|V 0
If V = U Y for an open set U X, then we may rewrite this in terms of the image of
this sequence under the exact functor i restricted to U :
kp
X Hmp |U
X H |U
m
k
y
y
XH
k
j y
|U
XH
jmy
|U
y
y
0
0
with exact rows and columns. From the diagram it is evident that j m : X Hm |U
i M|U is surjective with kernel equal to the sum of the images of m : X Hmp |U X Hm |U
and k : X Hkp |U X Hm |U . Thus, i M is coherent.
This is a very important fact. It will allow us to prove a great many theorems about
coherent sheaves on projective varieties once we know they are true for coherent sheaves
on projective space. To make use of it in our study of the functor M Mh we need the
following:
14.6 Theorem. If i : Y X is an embedding of algebraic varieties and M is a coherent
algebraic sheaf on Y , then (i M)h = i (Mh ).
Proof. There is an inclusion M0 Mh of sheaves of Y O0 -modules which yields a morphism
i (M0 ) i (Mh ) of sheaves of X O0 -moldules. Clearly, i (M0 ) = (i M)0 and so we have
127
X Ox0 Mx Y Hx Y Ox0 Mx
Ox0 ) X Ox0 Mx
by using the fact that X Hx X Ox0 Y Ox0 ' Y Hx , which is deduced from tensoring X Hx
with the exact sequence 0 IY X O Y O 0 and using flatness and Theorem
14.4(c). The theorem then follows from the associativity of tensor product and the fact
that Y Ox0 X Ox0 Mx ' Mx .
We are now in a position to prove the three main theorems of Serres GAGA paper.
Note that if X is an algebraic variety and M is a sheaf on X then M and M0 have the
same global sections and, in fact, the same cohomology. The latter is due to the fact that a
flabby resolution of M0 will also be a flabby resolution of M when restricted to the Zariski
open sets. Thus, H p (X, M) = H p (X h , M0 ). Furthermore, the map m 1 m : M0
Mh induces a morphism H p (X h , M0 ) H p (X h , Mh ). Combining these facts gives us a
morphism H p (X, M) H p (X h , Mh ). The following is the first of Serres three GAGA
theorems:
14.7 Theorem. If X is a projective algebraic variety and M is a coherent algebraic sheaf
on X, then the natural map H p (X, M) H p (X h , Mh ) is an isomorphism for every p.
Proof. Since X is projective, we may assume that it is embedded as a subvariety of P n . If
i : X P n is the embedding, then i M is a coherent algebraic sheaf on P n by Theorem
14.5. Furthermore, the cohomology groups of M and i M are the same (Problem 14.2).
Likewise, i Mh is a coherent analytic sheaf on P n by Theorem 14.5. and the cohomology
groups of Mh and i Mh are the same. By Theorem 14.6, (i M)h = i (Mh ). It follows
that the theorem is true in general if it is true for coherent algebraic sheaves on P n . Thus,
we may assume that X = P n .
By Theorem 13.17, M is the quotient of a sheaf F which is a finite direct sum of sheaves
of the form O(k). If K is the kernel of F M, then we have a short exact sequence of
coherent algebraic sheaves
0 K F M 0
128
J. L. TAYLOR
If we apply ( )h to this sequence and use Theorems 14.3 and 14.4 then we have an exact
sequence
0 Kh F h Mh 0
of coherent analytic sheaves with F h a finite direct sum of sheaves of the form H(k).
If we apply the morphism of cohomology induced by ( )h to the long exact sequences
of cohomology corresponding to these short exact sequences, we obtain a commutiative
diagram
H p (K) H p (F) H p (M) H p+1 (K) H p+1 (F)
1 y
2 y
3 y
4 y
5 y
H p (Kh ) H p (F h ) H p (Mh ) H p+1 (Kh ) H p+1 (F h )
where, to save space, we have suppressed the P n in each cohomology group and written,
for example, H p (M) rather than H p (P n , M). In this diagram, 2 and 5 are always
isomorphisms by Corollary 13.14 and the fact that F is a direct sum of sheaves O(k).
Suppose that H p+1 (M) H p+1 (Mh ) is a isomorphism for every coherent algebraic sheaf
M. Then 4 is an isomrphism. This implies that 3 is surjective for every M and, hence,
that 1 is surjective. This, along with the fact that 2 and 4 are isomorphisms, implies that
3 is also injective and, thus, is an isomorphim. Thus, the proof will be complete if we can
show that there is a p0 so that H p (M) H p (Mh ) is an isomorphism for all M for p > p0 .
and
T 0 (E) =
129
Then : T T 0 is a transformation of functors and we are to show that it is an isomorphism on finitely generated modules. Clearly is an isomorphism if E = A, since T (A)
and T 0 (A) are both equal to F A B in this case. Similarly is an isomorphism if E = An
for some n. Note that T and T 0 are left exact since hom is left exact in its first variable
and B is faithfully flat over A. Since A is Noetherian, for each finitely generated A-module
E we can construct an exact sequence of the form
An Am E 0
On applying T and T 0 , we obtain a diagram
0 T (E) T (Am ) T (An )
y
y
y
0 T 0 (E) T 0 (Am ) T 0 (An )
with exact rows and with the last two vertical maps isomorphisms. It follows that the first
vertical map is also an isomorphism and the proof is complete.
14.9 Theorem. If M and N are two coherent algebraic sheaves on a projective algebraic
variety X, then every morphism of analytic sheaves Mh N h is induced by a morphism
M N of algebraic sheaves.
Proof. Let A denote the sheaf hom(M, N ). This is the sheaf which assigns to an open
U X the O(U )-module consisting of morphisms M|U N |U in the category of sheaves
of U O-modules. Similarly, let B = hom(Mh , N h ) be the analogous sheaf for the sheaves
of H modules Mh and N h . The functor ( )h clearly defines a sheaf morphism A0 B
and, since B is an H-module, this induces a morphism Ah = H A0 B, where, in
this argument, will mean tensor product relative to O. We claim that this morphism
Ah B is an isomorphism. As usual, it suffices to check this for the stalks at each point
of X. The fact that M is coherent and, hence, locally finitely generated implies that each
Ox -module homomorphism from Mx to Nx extends to a morphism from M|U to N |U for
some neighborhood U of x and that a morphism from M|U to N |U which vanishes at x
also vanishes in a neighborhood of x. These two statements, taken together, and their
analogues for coherent analytic sheaves mean that
Ax = hom(Mx , Nx )
Mxh = Mx Hx
and
Nxh = Nx Hx
Thus, our claim will be established if we can show that the natural homomorphism
hom(Mx , Nx ) Hx hom(Mx Hx , Nx Hx )
130
J. L. TAYLOR
is bijective. But this follows from the previous lemma since Ox is Noetherian, Mx is finitely
generated and Hx is faithfully flat over Ox by Corollary 7.15.
To finish the proof, we consider the morphisms
H 0 (X, A) H 0 (X h , Ah ) H 0 (X, B).
The first of these morphisms is an isomorphism by Theorem 14.7 provided we can show that
A is coherent. This is done in Problem 14.3. The second morphism is an isomorphism by
the claim proved in the previous paragraph. Thus, the composition H 0 (X, A) H 0 (X, B)
is an isomorphism. This completes the proof of the theorem since a global section of B is
a morphism Mh N h while a global section of A is a morphism M N .
The geometric fiber of a coherent analytic sheaf S at a point x is the Hx module
Sx /Mx Sx , where Mx is the maximal ideal of Hx . The geometric fiber of a coherent algebraic
sheaf is defined analogously.
14.10 Lemma. If S is a coherent analytic sheaf on an holomorphic variety X, x X and
F H 0 (X, S) is a set of sections which generates the geometric fiber of S at x, then F
generates S|U for some neighborhood U of x. The analogous statement is true for coherent
algebraic sheaves.
Proof. It follows from Nakayamas Lemma that if F generates Sx /Mx Sx then it generates
Sx (Problem 14.5). However, by Problem 12.4 the set of y at which F fails to generate Sy
is a closed subvariety. Hence, there is a neighborhood U of x such that F generates S|U .
The third and most difficult of Serres GAGA theorems is the following:
14.11 Theorem. If X is a projective algebraic variety and M is a coherent analytic sheaf
on X h , then there is a coherent algebraic sheaf N on X such that N h ' M. Furthermore,
N is unique up to isomorphism.
Proof. The uniqueness is an immediate consequence of the preceding theorem.
Claim 1. The theorem is true if it is true for X = P n for all n.
If i : X P n is an embedding of X as a subvariety of P n , then i M is a coherent
analytic sheaf on P n . Suppose there is a coherent algebraic sheaf S on P n with S h ' i M.
Then we claim that S is i N for a coherent algebraic sheaf N on X. In fact, if I is the
ideal sheaf of X and f Ix for some x P n , then multiplication by f determines
an endomorphism : S|U S|U for some nieghborhood U of x. The corresponding
morphism h : S h |U S h |U is still multiplication by f and is zero in a neighborhood of
x because S h ' i M and M is a sheaf of X H-modules and the H-module action on i M
factors through the quotient map H X H. But if h vanishes in a neighborhood then
so does by Theorem 14.3. Thus, we have proved that IS = 0. This implies that S is
supported on X and its restriction to X is a sheaf N of X O-modules. It is easy to see
that N is a coherent algebraic sheaf on X (Problem 14.4) and, obviously, i N = S. Now
Theorem 14.6 implies that i N h ' (i N )h = S h ' i M. On restricting to X, this implies
that N h ' M. Thus, the theorem is true for any projective variety if it is true for P n .
We have reduced the proof to the case where X = P n . We will now prove it in this
case by induction on n. It is trivial when n = 0 since P 0 is a point and coherent algebraic
131
and analytic sheaves are just finite dimensional vector spaces. Thus, we will assume that
n > 0 and that the theorem is true in dimensions less than n.
Note that for a coherent analytic sheaf M on P n we may twist by H(k) to construct a
coherent analytic sheaf M(k) = MH H(k) for each k just as we did for coherent algebraic
sheaves. Note also that N (k)h ' (N h )(k) due to the fact that O(k)h = H(k).
Claim 2. Let E be a hyperplane in P n and A a coherent analytic sheaf on E. Then,
under our induction assumption, H p (E, A(k)) = 0 for large enough k.
We have that E is the subvariety of P n defined by the zero set of a linear functional
on Cn+1 (a homogeneous polynomial of degree one). Then E is a copy of P n1 and so,
by assumption, the theorem is true for X = E. That H p (E, A(k)) = 0 for a coherent
analytic sheaf A on E and large enough k then follows from Theorems 14.7 and 13.18 and
our induction assumption which implies that A = B h for some coherent algebraic sheaf B.
The key to the proof of the Theorem is the next claim:
Claim 3. Under our induction assumption, for every coherent analytic sheaf M on X
there is an integer kM such that for every k > kM the sheaf M(k) is generated over H by
its space of global sections H 0 (P n , M(k)).
Note that if M(k)x is generated by H 0 (P n , M(k)) then M(k + p)x is generated by
H (P n , M(k+p)) for all p > 0. This is due to the fact that M(k+p)x = M(k)x Hx H(p)x
and H(p)x is generated by its global sections if p 0 (since H(p) is the sheaf of holomorphic
sections of a line bundle, one only needs to have one global section which is non-vanishing
at x in order to have the global sections generate H(p)x and the existence of such a section
follows from Theorem 13.12). In view of these remarks, the compactness of P n , and Lemma
14.10, to prove Claim 3 it suffices to prove that for each x P n there is a k for which the
geometric fiber of the module M(k)x is generated by H 0 (P n , M(k)).
Now let E be a hyperplane in P n and let IE be its ideal sheaf in H. We may as well
assume that E is the hyperplane on which the coordinate function z0 vanishes. Consider
the exact sequence
0
0 IE H
EH
EH
132
J. L. TAYLOR
If we set Lk = ker(M(k) B(k)), then this sequence breaks up into two short exact
sequences of coherent analytic sheaves
0 C(k) M(k 1) Lk 0
and
0 Lk M(k) B(k) 0.
Now we can apply the long exact sequences of cohomology to these two short exact sequences. The relevant parts for us are
H 1 (P n , M(k 1)) H 1 (P n , Lk ) H 2 (P n , C(k))
and
H 1 (P n , Lk ) H 1 (P n , M(k)) H 1 (P n , B(k))
Now B = E H H M and C = torH
1 (E H, M) are coherent sheaves of H-modules but the
action of H factors through E H and, hence, they are actually coherent analytic sheaves
on E (Problem 14.4). It follows that B(k) and C(k) are also coherent analytic sheaves on
E. By our induction assumption this means that they are images under the functor ( )h
of coherent algebraic sheaves on E. It then follows from Theorem 14.7 and Therem 13.18
that B(k) and C(k) have vanishing pth cohomology for p > 0 and k sufficiently large. Thus,
the above sequences imply that for large k we have surjective maps
H 1 (P n , M(k 1)) H 1 (P n , Lk ) and
H 1 (P n , Lk ) H 1 (P n , M(k))
133
The latter module is just the quotient module M(k)x /(IE )x M(k)x . Since H 0 (P n , M(k))
generates this quotient module of M(k)x , it generates the geometric fiber of M(k)x . This
completes the proof of Claim 3 and puts us in a position to complete the proof of the
Theorem.
By Claim 3 we know that if M is a coherent analytic sheaf then there is an integer k
such that M(k) is generated by H 0 (P n , M(k)). Since H 0 (P n , M(k)) is finite dimensional
by Theorem 16.18, there is a surjection Hp M(k) for some p. If we twist this morphism
by the sheaf H(k), we obtain a surjection Hp (k) M. Now by applying the same
analysis to the kernel of this map, which is also a coherent analytic sheaf, we obtain an
exact sequence of the form
Hq (j) Hp (k) M 0
Now Hq (j) = Oq (j)h and Hp (k) = Op (k)h and so, by Theorem 14.9, the morphism
is induced by a morphism of coherent algebraic sheaves : Oq (j) Op (k). If N is
the cokernel of , then the exact functor ( )h applied to the exact sequence
Oq (j) Op (k) N 0
yields an exact sequence
Hq (j) Hp (k) N h 0.
But this implies that M ' N h which completes the induction and the proof of the theorem.
The results of Theorems 14.7, 14.9 and 14.11 (Serres Theorems 1, 2, and 3) can be
summarized as follows:
14.12 Theorem. If X is a projective algebraic variety, then the functor M Mh is a
cohomology preserving equivalence of categories from the category of coherent algebraic
sheaves on X to the category of coherent analytic sheaves on X h .
The above results have a wide variety of applications. We state some of these below, but
prove only a couple of them. The proofs of the others require knowledge of results from
algebraic geometry which we have not developed here. For a more complete discussion of
applications we refer the reader to Serres paper.
At this point, we will drop the use of the X h notation except in situations in which it
is needed to avoid confusion. If X is an algebraic variety then will generally also use X
to denote the corresponding holomorphic variety that is, we will think of an algebraic
variety as having two ringed space structures one algebraic and one holomorphic. We
will call a holomorphic variety X algebraic if it is the holomorphic variety associated to
some algebraic variety.
Our first application is the following theorem of Chow:
14.13 Corollary. If X is a projective variety, then every holomorphic subvariety of X is
algebraic.
Proof. Let Y be an analytic subvariety of X and consider the ideal sheaf IY . Then the
quotient H/IY is a coherent analytic sheaf on X which is isomorphic to iY H, where
134
J. L. TAYLOR
i : Y X is the inclusion. The support of the sheaf H/IY is clearly the subvariety
Y . Now by Theorem 14.12, there is a coherent algebraic sheaf N with the property that
N h ' H/IY . The support of N is the same pointset as the support of N h due to the fact
that Hx is faithfully flat over Ox for each x. Thus, the support of N is Y . However, the
support of a coherent algebraic sheaf is an algebraic subvariety (see Problem 12.4 and note
that the proof works equally well in the algebraic case). This completes the proof.
One can combine this with another result of Chow (on representing a general algebraic
variety as the image under a regular map of a dense open subset of a projective variety)
to obtain (cf. Serre):
14.13 Corollary. If X is an algebraic variety, then every compact holomorphic subvariety
of X h is algebraic.
With this and a little additional work, one can prove (cf. Serre):
14.14 Corollary. Every holomorphic map from a compact algebraic variety to an algebraic variety is regular.
This has the obvious consequence that:
14.15 Corollary. A compact holomorphic variety has at most one structure of an algebraic variety (up to isomorphism).
The category of algebraic vector bundles on an algebraic variety may be identified with
the category of locally free finite rank sheaves of O modules a vector bundle is identified
with its sheaf of sections. In the same way, the category of holomorphic vector bundles may
be identified with the category of locally free finite rank sheaves of H-modules. Clearly,
the equivalence of categories M Mh of Theorem 14.12 has the property that M is free
of finite rank if and only if Mh is free of finite rank. The corresponding functor on vector
bundles is just the functor which assigns to an algebraic vector bundle : E X over an
algebraic variety the holomorphic bundle h : E h X h obtained by putting the canonical
analytic structure on both total space and base. Thus, we have proved:
14.16 Corollary. If X is a projective algebraic variety then the category of algebraic
vector bundles on X is equivalent to the category of holomorphic vector bundles on X h
under the natural correspondence.
Serres paper contains a more general result of the above type which concerns bundles
with structure groups other than Gln (C). It also contains, as an entirely different kind of
application of the GAGA theorems, a proof of the following conjecture of A. Weil:
14.17 Corollary. If V is a projective non-singular variety defined over an algebraic number field K. Then the complex projective variety determined by an embedding of K in C
has Betti numbers which are independent of the embedding that is chosen.
135
14. Problems
1. Prove that if M and N are sheaves of modules over a sheaf of rings R, then
(M R N )x = lim
{M(U ) R(U ) N (U ) : x U } = Mx Rx Nx
2.
3.
4.
5.
136
J. L. TAYLOR
0 (, Lk ) (, Hp ) (, Lk1 ) 0
is exact. The theorem follows from descending induction using this fact, beginning on the
left at k = m.
In a corollary to Hilberts Syzygy Theorem (Corollary 3.14) we proved that every finitely
generated module over the local ring n H has a terminating free finite rank resolution a
terminating syzygy. Thus, given a coherent sheaf S on an open set in Cn , a sequence like
the one in the above lemma can always be constructed at each point. Using the results on
coherence of Chapter 12, we are able to construct such a sequence in a neighborhood of
any point:
137
15.2 Lemma. If S is a coherent analytic sheaf defined on an open set U in Cn , then for
each point x U there is a neighborhood W of x in U and an exact sequence:
n
138
J. L. TAYLOR
and so it is clear that Bx = y if x = lim xn . It also follows from our estimate on the norms
of the un that ||x|| 2K0 . This establishes the Lemma with K = 2K0 .
If f is a function from an open subset U of a Banach space X to a Banach space Y ,
then the derivative (if it exists) of f at x U is a bounded linear map f 0 (x) : X Y with
the property that
lim ||u||1 ||f (x + u) f (x) f 0 (x)u|| = 0.
u0
f (x + u) f (x) f (x)u =
[f 0 (x + tu) f 0 (x)]u dt
Thus, by shrinking U if necessary and using the continuity of f 0 again, we may assume
that
||f (x + u) f (x) f 0 (x)u|| < (2K)1 ||u||
for x, x + u U .
The remainder of the proof is just an application of Newtons method. We choose > 0
so that ||u|| < 2 implies that u U . We will show that for ||y f (0)|| < K 1 we can
solve the equation f (x) = y. We proceed as in Newtons method, using x0 = 0 as our
initial guess. We then choose x1 X so that f 0 (0)x1 = y f (0) and ||x1 || K||y f (0)||.
Note that ||x1 || < . We then inductively choose xn so that xn = xn1 + un where
f 0 (xn1 )un = y f (xn1 ).
and
||un || K||y f (xn1 )||.
Then we have
||y f (xn )|| = ||y f (xn1 ) (f (xn ) f (xn1 ))||
= ||f 0 (xn1 )(xn xn1 ) (f (xn ) f (xn1 ))|| < (2K)1 ||un ||
So that
Now ||u1 || < implies that ||u2 || < 21 and, in general, that ||un || < 2n+1 . This
implies that the sequence {xn } is contained in U and converges to an element x U . Our
estimate above on ||y f (xn )|| shows that f (x) = y. This completes the proof.
139
Roughly speaking, the above theorem says that a certain non-linear problem has a
solution if the linearized version is solvable. In our application of this result, the solvability
of the linearized problem is given by the next lemma. In this lemma and in what follows
we will use the following geometric situation. By an open (compact) box in Cn we will
mean an open (compact) set U which is the cartesian product of intervals one from each
of the 2n real and imaginary coordinate axes. An aligned pair of open (compact) boxes
will be a pair (U1 , U2 ) which is the cartesian product of two ordered sets of intervals which
are identical except in one (real or imaginary) coordinate and in that coordinate the two
intervals are overlapping. The coordinate in which the defining intervals are allowed to be
different will be called the exceptional coordinate. It is clear from the definition that if
(U1 , U2 ) is an aligned pair of boxes, then U1 U2 and U1 U2 are also boxes and they are
obtained from U1 and U2 by taking intersection or union of the two defining intervals in
the exceptional coordinate and leaving the defining intervals in all other coordinates the
same.
15.5 Lemma. Let (U1 , U2 ) be an aligned pair of open boxes in Cn . Then each bounded
holomorphic function f on U1 U2 is the difference f1 f2 of a bounded holomorphic
function f1 on U1 and a bounded holomorphic function f2 on U2 .
Proof. Without loss of generality we may assume that the exceptional coordinate for
(U1 , U2 ) is x1 , so that the pair (U1 , U2 ) has the form
U1 = I1 iK W
U2 = I2 iK W
where I1 and I2 are overlapping open intervals on the line, K is an open interval on the
line and W is an open box in Cn1 .
We choose a bounded C function on R that is one in a neighborhood of I1 I2
and is zero in a neighborhood of I2 I1 . We then consider to be a function defined on
Cn which is constant in all variables except x1 . Then (1 )f extends by zero to be a
bounded C function g1 in U1 while f extends by zero to be a bounded C function g2
in U2 . Furthermore, on U1 U2 , g1 g2 = f . In other words, we have solved our problem
in the class of bounded functions which are C in the variable z1 and holomorphic in
the remaining variables. Now we need to modify this solution to arrive at one which is
holomorphic in z1 as well.
The fact that g1 g2 = f is holomorphic on U1 U2 implies that
g1
g2
=
z1
z1
on U1 U2 and that implies that they define a bounded C function g on U1 U2 which
gi
is
on Ui . Let V be an open set with compact closure in (I1 I2 ) K and let 1 be a
z1
C function of z1 which is 1 on V and which has compact support in (I1 I2 ) K. Set
2 = 1 1 . We then proceed as in the proof of Dolbeaults Lemma. We set D = U1 U2
and
Z
1
i (1 )g(1 , z2 , . . . , zn ) d1 d1
hi (z) =
2i D
1 z1
140
J. L. TAYLOR
an ;
n=0
X
an
(c) there is a map a exp(a) =
from A to A1 which is a homeomorphism
n!
n=0
from some neighborhood of 0 in A to the neighborhood {b : ||1 b|| < 1} in A1 ;
X
(1 b)n
(d) on {b A : ||1 b|| < 1 the map b log b =
is an inverse for exp;
n
n=1
(e) the subgroup of A1 generated by the image of exp is open and is equal to the
connected component of the identity in A1 .
In what follows, Gln (C) will denote the group of invertible nn complex matrices. This
is the group of invertible elements of the algebra Mn (C) of all n n complex matrices.
The latter is a Banach algebra under the standard matrix norm
||a|| = sup{||ax|| : x Cn , ||x|| 1}
141
and, hence, with the topology determined by this norm, Gln (C) is a topological group. If U
is a domain in Cn , we will also be concerned with the Banach algebra Hb (U, Mn ) of bounded
Mn -valued functions on U . Here the norm is given by ||f || = sup{||f (x)|| : x U }. The
invertible group of this Banach algebra is the group of bounded holomorphic functions
with values in Gln (C).
15.7 Theorem. If U Cn is the Cartesian product of simply connected open subsets of
C and K is a compact subset of U , then each holomorphic mapping f : U Gln (C) may
be uniformly approximated on K by holomorphic mappings from Cn to Gln (C).
Proof. By the Riemann mapping theorem we may, without loss of generality, assume that
U is a polydisc centered at the origin. We let V be an open polydisc centered at the
origin, containing K and with compact closure in U . Then, on V , f and f 1 are bounded
holomorphic functions with values in Gln (C) that is, f is an element of the invertible
group Hb (U, Gln (C)) of the Banach algebra Hb (V, Mn (C)). Then a curve t ft , t [0, 1]
in Hb (V, Gln (C)) joining f to the constant matrix f (0) may be constructed by setting
ft (z) = f (tz) for each t [0, 1]. Since Gln (C) itself is connected (Problem 15.2), this
proves that Hb (V, Gln (C)) is connected. By Theorem 15.6(e), this implies that f is a
product of elements in the range of the exponential function. Hence, on V ,
f = exp(g1 ) exp(g2 ) . . . exp(gk ) with
g1 , g2 , . . . gn Hb (V, Mn (C)).
Now each gi may be regarded as a matrix with entries which are bounded holomorphic
functions on V . By truncating the power series of each entry of each gi we may approximate
each gi by matrices hi with polynomial entries as closely as we like in the uniform topology
on K. Then
f = exp(h1 ) exp(h2 ) . . . exp(hk )
will be a holomorphic Gln (C)-valued function on all of Cn . Clearly, the hi can be chosen
so that f approximates f arbitrarily closely on K.
The next lemma is the key to the vanishing theorem we are seeking:
15.8 Cartans Lemma. Let (K1 , K2 ) be an aligned pair of compact boxes in Cn . Then
each holomorphic Gln (C)-valued function f defined in a neighborhood of K1 K2 may be
factored as f = f21 f1 where fi is a holomorphic Gln (C)-valued function in a neighborhood
of Ki for i = 1, 2.
Proof. We may construct an aligned pair (U1 , U2 ) of open boxes such that Ki Ui and
f and f 1 are holomorphic and bounded in U1 U2 . Let Ai be the Banach algebra
Hb (Ui , Mn (C)) for i = 1, 2 and let B be the Banach algebra Hb (U1 U2 , Mn (C)) and
consider the non-linear map : A1 A2 B defined by
(g1 , g2 ) = log((exp g2 )1 exp g1 ) on
U1 U2
Now it is easy to see that a function from a Banach algebra to itself which is defined by a
convergent power series is infinitely differentiable. It is also easy to see that the analogues of
the chain rule and the product rule hold for functions between Banach algebras. It follows
142
J. L. TAYLOR
that is infinitely differentiable in a neighborhood of (0, 0). We have that (0, 0) = 0 and
a simple calculation (Problem 15.3) shows that
0 (0, 0)(h1 , h2 ) = h1 h2
By Lemma 15.5, 0 (0, 0) is surjective. By Theorem 15.4, the image of contains a neighborhood of 0 in B. After composing with exp, we conclude that the map
(g1 , g2 ) (exp g2 )1 exp g1
has image which contains a neighborhood of the identity in B. This means that the
theorem is true for f sufficiently close to the identity in B. However, by the previous
theorem, we may approximate f arbitrarily closely on U1 U2 by a Gln (C)-valued function
h which is holomorphic on all of Cn . Then, for an appropriate choice of h we will have
f h1 sufficiently close to the identity in B that we may write f h1 = g21 g1 with gi a
holomorphic Gln (C)-valued function on Ui . Then the desired solution is f = f21 f1 with
f1 = g1 h and f2 = g2 .
The procedure used in the next theorem is sometimes known as amalgamation of syzygies.
15.9 Theorem. Let (K, L) be an aligned pair of compact boxes in Cn . Let S be a
coherent analytic sheaf defined in a neighborhood of K L and suppose that there are
exact sequences of analytic sheaves
2
1
0
0 Hpm m . . .
Hp1
Hp0
S 0
0
0 Hp0
S 0
0 Hq0 S 0
With the first defined over a neighborhood of K and the second over a neighborhood of
L. On K L, the composition = 01 0 : Hp0 Hq0 is an isomorphism over a
143
neighborhood of K L. This implies that p0 and q0 are the same integer k and that
is determined by a holomorphic Glk (C)-valued function in a neighborhood of K L. By
Cartans Lemma, this function may be factored as = 1 where () is a holomorphic
Glk (C)-valued function in a neighborhood of K (L). Then 0 1 = 0 1 in a
neighborhood of K L and so these two morphisms fit together to define an isomorphism
0
0 Hk S 0
over a neighborhood of K L, as required.
For the induction step, we assume that the theorem is true of all pairs of sequences,
as above, of length less than m and we suppose we are given a pair of length m. By
applying the Riemann mapping theorem in each variable we see that K L has arbitrarily
small neighborhoods U which are biholomorphically equivalent to open polydiscs. On a
sufficiently small such neighborhood, Lemma 15.1 implies that the sequences
2
1
0
. . .
(U, Hp1 )
(U, Hp0 )
(U, S) 0
0
Hp0 (U )
(U, S) 0
Hq0 (U ) (U, S) 0
Now is a matrix with entries which are holomorphic functions on U and, as such, it defines
a morphism of analytic sheaves : Hp0 Hq0 over U such that 0 = 0 . A similar
argument shows that we may construct a morphism of analytic sheaves : Hq0 Hp0
over U with 0 = 0 . We now modify each of the sequences so that the free modules
that appear in degree 0 will be identical. Thus, the first sequence is modufied by taking
its direct sum with the exact sequence
id
0 Hq0 Hq0 0
to obtain
1
0
Hp1 Hq0
Hp0 Hq0
S 0
where
1 = 1 id and
0 (f g) = 0 (f ). Similarly, we modify the second sequence by
taking its direct sum (on the other side) with the exact sequence
id
0 Hp0 Hp0 0
144
J. L. TAYLOR
to obtain
1
1
, =
=
1
1
A calculation shows that 0 =
0,
0 = 0 and = 1 . Thus, is an isomorphism
and we have the following commutative diagram over U
1
0
Hp1 Hq0
Hp0 Hq0
S 0
2
1
0 Hpm m . . .
Hp1
K 0
0 Hrm . . . Hr1 K 0
in a neighborhood of K L. Combining this with the morphism 0 gives us the required
sequence for S over a neighborhood of K L. This completes the proof of the theorem.
145
b = K;
(ii) a compact subset K of X is said to be holomorphically convex in X if K
b is compact for every compact subset
(iii) X is said to be holomorphically convex if K
K X.
b is always a closed subset of X and so it will be compact if it is contained
Note that K
in a compact subset of X. Note also that if X is a holomorphic variety, U an open subset
of X and K a compact subset of U , then it makes sense to talk about the holomorphically
convex hull of K in X and in U . These are not necessarily the same. The open set
146
J. L. TAYLOR
i = 1, . . . k} U
Thus, the functions {fi }ki=1 are the coordinate functions of a holomorphic map 0 : X Ck
which maps W1 into (0, 1). In fact, 0 is a proper holomorphic map of W1 into (0, 1)
since the inverse image in W1 of any compact subset of (0, 1) will be closed not only in
W1 , but also in the compact set U .
Now, by (ii) of Definition 15.13, for each point x of W 1 there is a finite set of global
sections of H which vanish at x and generate the maximal ideal of Hx . Without loss of
generality, we may assume that these functions all have modulus less than 1 at each point
of W1 . By Theorems 6.15 and 6.16, the map X Cm with these functions as coordinate
functions is a biholomorphic map of some neighborhood of x onto a closed subvariety of
some neighborhood of f (x). Since W 1 is compact, finitely many such neighborhoods will
cover W 1 . By adjoining all the corresponding functions to the set {fi }ki=1 we obtain a set
{fi }qi=1 of functions which are the coordinate functions of a holomorphic map 1 : X Cq
147
which is proper on W1 and is also has the property that in a neighborhood of each point
x W 1 it is a biholomorphic map onto a closed subvariety of a neighborhood of f (x).
In particular, 1 is locally one to one. This means that the diagonal in W 1 W 1 has
a neighborhood V in which (x) (y) is non-vanishing except on the diagonal itself.
Another compactness argument and (iii) of Definition 15.13 show that we can find another
finite set of global sections of H (again with modulus less than one at points of W1 )
such that whenever (x, y) W 1 W 1 V there is some function f in this set such that
f (x) 6= f (y). By adjoining this finite set to the set {fi }qi=1 , we obtain a set {fi }m
i=1 of
functions which are the coordinate functions of a holomorphic map : X Cm which is
proper, injective and locally biholomorphic on W1 .
A proper, injective, continuous map from one locally compact space into another has
closed image and is a homeomorphism onto its image (Problem 15.7). Thus, the image Y
of on W1 is a closed subset of (0, 1) and is a holomorphic homeomorphism onto Y .
Furthermore, for each x W1 , maps some neighborhood of x biholomorphically onto an
open set in Y which is a closed holomorphic subvariety of an open set in (0, 1). It follows
that Y is a closed subvariety of (0, 1) and maps W1 biholomorphically onto Y . Thus,
W1 is an Oka-Weil subdomain of X.
For each polyradius r which has all coordinates less than or equal to 1 we set Wr =
1 ((0, r)). Then each Wr is obviously also an Oka-Weil subdomain of X. If r < s means
that each coordinate of r is less than the corresponding coordinate of s, then W r Ws
whenever r < s. Fix an r < 1 such that K Wr . If a coherent sheaf S is defined in
a neighborhood of W r , then we may choose an s > r such that Ws is contained in that
neighborhood. Then S may be considered a coherent analytic sheaf defined on Y (0, s)
and may be extended by zero to a coherent analytic sheaf on all of (0, s). It follows from
Corollary 15.11 that such a sheaf will be acyclic on (0, r) and, hence, on Y (0, r).
Thus, if we let W = Wr , then W is an Oka-Weil subdomain of X which contains K, has
compact closure contained in U and has the property that every coherent analytic sheaf
defined in a neighborhood of W is acyclic on W . This completes the proof.
15.16 Corollary. If X is a Stein Space then X is the union of a sequence {Wn } of OkaWeil subdomains such that W n Wn+1 for each n and each coherent analytic sheaf on
W n is acyclic on Wn .
Proof. Since X is countable at infinity, it is the union of an increasing sequence {Kn }
of compact sets. Suppose we have managed to find Oka-Weil subdomains Wj with the
required acyclic property for coherent sheaves and such that Kj Wj and W j1 Wj
bn is also
for j n. Then Cn = W n1 Kn is compact. Since X is a Stein space, C
compact. By Theorem 15.15, Cn is contained in an Oka-Weil subdomain Wn with the
required acyclic property. Thus, the Corollary is true by induction.
15.17 Corollary. Every holomorphic variety X has a neighborhood basis W with the
property that given any x X and any open set U containing x, there is a W W, with
x W such that W has compact closure W U and every coherent analytic sheaf defined
on a neighborhood of W is acyclic on W .
Proof. Every holomorphic variety has a neighborhood base consisting of closed subvarieties
of open polydiscs. Since a closed subvariety of an open polydisc is a Stein space, an
148
J. L. TAYLOR
15. Problems
1.
2.
3.
4.
149
16. Fr
echet Sheaves Cartans Theorems
A topological vector space is a vector space with a topology under which the vector
space operations are continuous. A topological vector space is locally convex if it has a
neighborhood base at zero consisting of convex balanced sets, where a set is balanced if
it is closed under multiplication by scalars of modulus one. A topological vector space is
locally convex if and only if its topology is given by a family { } of seminorms, where
a semi-norm on a vector space X is a function from X to the non-negative reals which
satisfies
(x + y) (x) + (y) and (x) = ||(x)
for x, y X and C. The topology determined by a family { } of seminorms is the
topology in which a basis of neighborhoods at zero is given by the collection of all sets of
the form
{x X : (x) < }.
A continuous linear functional on a topological vector space is a linear complex valued
function which is continuous. The space of all continuous linear functionals is called the
dual of X and is denoted X . There are a number of topologies that can be put on X
which make it a locally convex topological vector space. They have different properties
and are used in different circumstances.
A Frechet space is a locally convex topological vector space F which is complete and
which has its topology defined by a sequence {n } of semi-norms. Without loss of generality, the sequence {n } may be chosen to be increasing. Equivalently, a Frechet space is
a topological vector space which is the inverse limit of a sequence of Banach spaces and
bounded linear maps. Equivalently, a Frechet space is a complete locally convex topological vector space with a topology defined by a translation invariant metric. A bounded set
in a Frechet space is a set B with the property that each of the defining semi-norms n is
bounded on B equivalently, B is bounded if for every neighborhood U of 0, there is a
positive number k such that B kU . A Frechet space F is called a Montel space if every
closed bounded subset of F is compact. A separated quotient of a topological vector space
is a quotient by a closed subspace.
We will assume knowledge of the following elementary facts concerning locally convex
topological vector spaces. The proofs can be found in any text on functional analysis or
topological vector space theory.
TVS 1. Closed subspaces, separated quotients and countable direct products of Frechet
spaces are Frechet spaces.
TVS 2. Closed subspaces, separated quotients, and countable direct products of Montel
spaces are Montel spaces.
TVS 3 (Open Mapping Theorem). A surjective continuous linear map from a Frechet
space to a Frechet space is an open map.
TVS 4 (Closed Graph Theorem). A linear map from a Frechet space to a Frechet
space is continuous if and only if its graph is closed.
150
J. L. TAYLOR
TVS 5 (Hahn-Banach Theorem). The strong form says that a real linear functional
on a subspace of a real vector space, dominated by a convex functional, extends to the
whole space with preservation of the dominance. We will need two corollaries of this:
(a) every continuous linear functional on a subspace of a locally convex topological
vector space extends to a continuous linear functional on the whole space;
(b) if B is a closed convex balanced set in a locally convex topological vector space X
and x0 X is a point not in B, then there exists a continuous linear functional f
on X such that |f (x)| 1 for all x B but |f (x0 )| > 1.
TVS 6. Every locally compact topological vector space is finite dimensional.
If U is a domain in Cn then H(U ) is a Frechet space in the topology of uniform convergence on compact sets. In fact, by Theorem 1.8, H(U ) is a Montel space. The same thing
is true for holomorphic functions on a variety:
16.1 Theorem. If U is an open subset of a holomorphic variety X, then H(U ) is a Montel
space in the topology of uniform convergence on compact subsets of U .
Proof. We may express U as the union of an increasing sequence {Kn } of compact subsets,
then the sequence {n } of seminorms defined by
n (f ) = sup{|f (x)| : x Kn }
determine the topology of H(U ). It is not obvious that H(U ) is even complete in this
topology. However, by Corollary 15.18, X has a neighborhood base consisting of sets W
for which W may be imbedded as a closed subvariety of an open polydisc and the
restriction map H() H(W ) is surjective. Since the kernel of this map consists of the
functions which vanish on W , it is closed in H(). Hence, H(W ) is a separated quotient
of a Montel space and is, therefore, Montel. Now for a general open set U , let {Wi } be
a countable open cover of U by sets with the aboveQ
property. Then the map f {f |Wi }
embedds H(U ) as a subspace of the Montel space i H(Wi ). In fact, it is embedded as
the closed subspace {{fi } : (fi )|Wi Wj = (fj )|Wi Wj i, j}. Thus, H(U ) is a Montel space.
Given a coherent analytic sheaf S on a holomorphic variety X, we will define a canonical
Frechet space topology on its space S(U ) of sections over an open set U . The strategy for
doing this is to do it first for small neighborhoods of the kind given by Corollary 15.17
and then to argue that this is enough to determine a Frechet space structure on spaces of
sections over arbitrary open sets.
If x X then there is a neighborhood W of x on which S is the cokernel of a morphism
of analytic sheaves : Hm Hk . By Corollary 15.17, by shrinking W if necessary, we
may assume that W has the property that ker is acyclic on W . From the long exact
sequence of cohomology, we conclude that : H(W )m H(W )k has S(W ) as cokernel,
so that S(W ) is a quotient of the Frechet space Hk (W ). If : H(W )m H(W )k has
closed image then S(W ) is a quotient of Hk (W ) by a closed subspace and, hence, is also
a Frechet space. Unfortunately, it is not trivial to prove that has closed image. This
requires an argument using the Weierstrass theorems which is reminiscent of the proof of
Okas Theorem. First a lemma:
151
0
K
Hm
L
0.
From the long exact sequence of cohomology and the vanishing theorem of the previous
chapter (Corollary 15.11), we conclude that the corresponding sequence of sections over
is also exact. Thus, : Hm () L() is surjective. Since the closure of the image
of : Hm () Hk () is contained in L(), we conclude that the image and its closure
coincide.
The proof of part (a) will be by induction on n. Part (a) is trivially true in the case
n = 0 and so we assume that n > 0 and that the theorem is true whevever U is an open
subset of Cn1 . Under this assumption, we next prove part (a) when U is an open subset
U of Cn and k = 1.
Thus, let M be an ideal in Hz . We may as well assume that z = 0. After a linear change
of coordinates, if necessary, we may assume that there is a polydisc centered at 0 and
a function h H() which is regular in zn and has germ h0 belonging to M . Then the
Weierstrass preparation theorem implies that h0 is a unit times a Weierstrass polynomial.
Hence, by shrinking if necessary, we may assume that h itself is a Weierstrass polynomial
say of degree k. Then, by the Weierstrass division theorem, every germ f H0 has a
unique representation as f = gh + q where g is the germ of a function holomorphic in a
neighborhood of 0 and q is a polynomial in zn of degree less than k with coeficients which
are germs at 0 of holomorphic functions in the variables z1 , . . . , zn1 . Recall from the proof
of the Weierstrass division theorem, that g is given by an integral formula
Z
1
f (z 0 , ) d
g(z) =
2i ||=rn h(z 0 , )( z)
where z = (z 0 , zn ) (0, r) and r = (r0 , rn ) is chosen so that h(z) has no zeroes on the
part of (0, r) where |zn | = rn . This clearly means that if is chosen small enough, then
g may be chosen so as to have a representative in H() for all f H() and, furthermore,
that this representative will depend linearly and continuously on f . In other words, for
small enough , there is a continuous linear map : H() H() such that f (f )h
is a polynomial of degree less than k in n for all f H().
Now let N be the set of elements of M which are polynomials in zn of degree less than
k. Clearly, N is a module over the ring n1 H0 of germs of holomorphic functions in the
152
J. L. TAYLOR
Mk = {fk : (f1 , . . . , fk ) M }
Then Mk is isomorphic to M/M0 under the projection map which sends a k-tuple of
functions to its kth element. We choose a polydisc centered at 0 and a finite set of
elements of Hk () whose germs at 0 form a set of generators for M . This set of elements
then determines a morphism of analytic sheaves over
: Hm Hk
with the image of 0 equal to M . Over we define : Hm H to be followed by the
projection of Hk on its last component. Then the image of 0 is Mk . Now, by shrinking
if necessary, we may assume that and are defined in an open polydisc containing
the closure of and that has compact closure. Under our induction assumption, we
know part (a) holds for k = 1. Hence, Part (b) holds for k = 1. This means that
: Hm () H() has closed image. By the open mapping theorem for Frechet spaces,
this implies that is an open mapping onto its image. Now suppose that {fj } is a
sequence in Hk (U ) with (fj )|0 M for every j and suppose that this sequence converges
to f Hk (U ). If gj and g are the last components of fj and f , then gj g and,
since is an open map, there exists a convergent sequence hj h in Hm () such that
gj = (hj ) (Problem 16.1). Then fj (hj ) is a convergent sequence in Hk1 () consisting
of elements whose germs at 0 belong to M0 . Thus, by our induction assumption on k, its
limit f (h) also has germ at 0 belonging to M0 . However, (h)0 M and M0 M .
Hence, f0 M . This completes the induction on k and also the induction on n. Hence,
part(a) and part(b) are both proved in general.
16.3 Theorem. Let X be a holomorphic variety and S a coherent subsheaf of Hk for
some k. Then S(X) is a closed subspace of the Frechet space Hk (X).
Proof. We first note that part(a) of Theorem 16.2 extends to varieties. That is, if fj f
in X Hk (X) and if x X, then Corollary 15.18 and the open mapping theorem imply that
there is a neighborhood W of x which may be identified with a closed subvariety of an open
polydisc Cn and there is a convergent sequence gj Hk () such that gj |W = fj |W
(Problem 16.1). Thus, if the functions fj all belong to some submodule M X Hx , then
the gj will belong to the inverse image N n Hxk of this submodule under the restriction
map n Hxk X Hxk . It follows from Theorem 16.2 that if g is the limit of the sequence
153
{gj }, then g has germ at x belonging to N and, hence, its restriction f to W has germ at
x belonging to M . Thus, part(a) of Theorem 16.2 holds with U replaced by an arbitrary
holomorphic variety X.
Now Theorem 16.3 is an immediate consequence, since a global section of S is just a
global section of Hk which has germ at x belonging to Sx for each x X.
A Frechet sheaf on a space X is just a sheaf of Frechet spaces on X. Of course, the
restriction maps are required to be morphisms in the category of Frechet spaces that is,
continuous linear maps.
16.4 Theorem. If S is a coherent analytic sheaf on an analytic space X, then there is a
unique structure of a Frechet sheaf on S with the property that if U is any open set and
Hk S is a surjective morphism of analytic sheaves from a free finite rank H module to
S defined over U , then Hk (U ) S(U ) is continuous. Furthermore, S is a Montel sheaf
with this structure.
Proof. By Corollary 15.17, we may choose a neighbohood base U for the topology of X
consisting of sets U which have compact closure and have the property that coherent
sheaves defined in a neighborhood of U are acyclic on U . Furthermore, we may choose U
so that for each U U , the sheaf S is the cokernel of a morphism between free finite rank
sheaves on a neighborhood of U .
Fix U U . Then there is an exact sequence of analytic sheaves
Hm Hk S 0
defined in a neighborhood of U . The fact that every coherent sheaf defined in a neighborhood of U is acyclic on U implies that the induced sequence on sections over U :
Hm (U ) Hk (U ) S(U ) 0
is also exact. Furthermore, by Theorem 16.3, the image of : Hm (U ) Hk (U ) is closed.
Hence, S(U ), as a separated quotient of Hk (U ), inherits a Frechet space topology, in fact,
a Montel space topology. Now suppose that V U is another set in our basis U and
suppose for this set we have an exact sequence
Hp Hq S 0
defined in a neighborhood of V . Then we may construct the following commutative diagram
with exact rows:
Hm (U ) Hk (U ) S(U ) 0
r
y
V,U y
y
Hp (V ) Hq (V ) S(V ) 0
Where rV,U is the restriction map and and are constructed by using the familiar
lifting argument (the projectivity of free modules). The maps and are module homomorphisms and, hence, are given by matrices of holomorphic functions on V through
154
J. L. TAYLOR
vector-matrix multiplication. This clearly implies that they are continuous. If S(U ) has
the topology it inherits from being a quotient of Hk (U ), then Hk (U ) S(U ) is an open
map. Also, the map Hk (V ) S(V ) is continuous if S(V ) is given the quotient topology.
These facts combine to force the map rV,U to be continuous. We draw two conclusions
from this:
(1) The topology on S(U ) is independent of the way in which it is expressed as a
quotient of a free finite rank H(U )-module; and
(2) if V U are two sets in our basis, then the restriction map rV,U : S(U ) S(V )
is continuous.
It is now clear how to define the topology on S(U ) for a general open set U . We cover
U
Q by a countable collection {Wi } of sets from our basis. Then f {f |Wi } : S(U )
i S(Wi ) is an
Qinjective continuous linear map of S(U ) onto a closed subspace
Q of the
Montel space i S(Wi ). The image is closed because it is the subspace of i S(Wi )
consisting of {gi } such that gi = gj on Wi Wj for all i, j. Since a closed subspace of a
Montel space is Montel, this serves to put a Frechet space structure on S(U ) under which
it is actually a Montel space. Note that, by construction, this topology has the property
that for each x U there is a basic neighborhood W U containing x such that the
restriction map S(U ) S(W ) is continuous. Now suppose that V U is another open
set and S(V ) is given a Frechet space topology with this same property. Then we claim
that the restriction map rV,U : S(U ) S(V ) has closed graph and, hence, is continuous.
To see this, let {(fn , rV,U (fn ))} be a sequence in the graph which converges to the point
(f, g). Then fn f and rV,U (fn ) g. Now for each point x of V we can choose a basic
neighborhood W of x such that the restrictions S(U ) S(W ) and S(V ) S(W ) are
both continuous. This clearly implies that rV,U (f )|W = g|W . But since this is true for a
neighborhood of each point of V , we conclude that rV,U (f ) = g and, hence, that the graph
of rV,U is closed as required. We draw two conclusions from this:
(1) The Frechet space topology on S(U ) is uniquely defined by the property that for
each basic open set W U the restriction map S(U ) S(W ) is continuous; and
(2) if V U are two open sets, then the restriction map rV,U : S(U ) S(V ) is
continuous.
Thus, we have proved the existence of a Frechet sheaf structure on S under which it is
a Montel sheaf. Furthermore, it is clear from the construction that a morphism T S
is continuous if and only if it is continuous locally, that is, if and only if T (W ) S(W )
is continuous for a neighborhood W of each point of X. This, and the construction of
the topology on basic open sets shows that the topology has the property that if Hk S
is a surjective morphism over an open set U , then Hk (U ) S(U ) is continuous. The
uniqueness is obvious from the construction.
In view of the above theorem, we may, henceforth, assume that every coherent analytic
sheaf comes equipped with a structure of a Montel sheaf. A morphism : S T between
two Frechet sheaves is said to be continuous if : S(U ) T (U ) is continuous for each
open set U . When is a morphism of coherent analytic sheaves continuous? Always!
16.5 Theorem. A morphism : S T of sheaves of H-modules between two coherent
analytic sheaves is automatically continuous.
155
Proof. As was pointed out in the proof of the previous theorem, a morphism between
coherent analytic sheaves is continuous if and only if it is continuous locally. If x is a
point of X then we may choose a basic neighborhood W for x of the kind used in the
proof of the previous theorem. Then there are surjective morphisms : Hk S and
: Hm T defined over U which are also surjective on sections over U . Then the usual
lifting argument gives us the in the following commutative diagram
Hk (W ) S(W ) 0
y
y
Hm (W ) T (W ) 0
It follows that is continuous because and are continuous and is open. Thus, is
locally continuous and, hence, continuous.
We are now in a position to prove the main theorem in the subject Cartans Theorem
B. The first step is an approximation theorem:
16.5 Theorem. If X is a Stein space, S a coherent analytic sheaf on X and W X is an
Oka-Weil subdomain, then the space of restrictions to W of global sections of S is dense
in S(W ).
Proof. We first prove that this is true in the case where S is the structure sheaf H. Since
W is an Oka-Weil subdomain, there is a holomorphic map : X Cn such that maps
W biholomorphically onto a closed subvariety of the unit polydisc (0, 1) centered at 0.
If f is a holomorphic function on W and K is a compact subset of W then it follows from
Corollary 15.18 that f has the form g in a neighborhood of K, where g is a holomorphic
function on an open polydisc with closure contained in (0, 1). If {hj } is a sequence of
polynomials converging to g in the topology of uniform convergence on compact subsets
of , then fj = hj defines a sequence of holomorphic functions on X which converge
uniformly to f on K.
In order to prove the theorem for a general coherent analytic sheaf S, we choose a
sequence {Wn } of Oka-Weil subdomains with W = W0 and W n Wn+1 for each n.
We claim that the image of S(Wn ) under restriction is dense in S(Wm ) if m < n. In
fact, we can find a surjective morphism Hk S in a neighborhood of W n . Then both
Hk (Wn ) S(Wn ) and Hk (Wm ) S(Wm ) are surjective. Thus, an element f of S(Wm )
can be lifted to an element g of Hk (Wm ) and this can be expressed as the limit of a
sequence of restrictions of elements hj Hk (Wm ) by the result of the above paragraph.
The image of the sequence {hj } in S(Wn ) will then have the property that its restriction
to S(Wm ) converges to f .
To finish the proof, we choose a translation invariant metric n defining the topology
n
P
of S(Wn ) for each n. Since the metric n may be replaced by the metric
i without
i=0
changing the topology it defines, we may assume without loss of generality that the sequence of metrics is increasing in the sense that m (f ) n (f ) if m < n and f S(Wm ).
Then if > 0 and f S(W ) = S(W0 ), we choose g1 S(W1 ) such that 0 (f g1 ) < /2.
156
J. L. TAYLOR
0 S F 0 F 1 F p
be a flabby resolution of S. Suppose p > 1 and f F p (X) with f = 0. Then we will
prove by induction that there is a sequence {gn } with gn F p1 (Wn ), gn = gn1 on
Wn1 and gn = fn on Wn . Clearly, if we can show this then part(a) of the Lemma will
be established. since such a sequence determines a global section g F p1 (X) such that
g = gn on Wn and, consequently, g = f on all of X.
Suppose we have managed to construct the sequence {gn } for all n m. Because S
is acyclic on each Wn , there exists a section gm+1 F p1 (Wm+1 ) such that gm+1 = f .
However, gm+1 gm may not be zero on Wm . Whatever it is, it is in the kernel of and,
hence, there exists hm+1 F p2 (Wm ) such that hm+1 = gm+1 gm on Wm . Since F p2
is flabby we may assume that hm+1 is actually a section defined on all of X. We then set
157
gm+1 = gm+1 hm+1 on Wm+1 . Clearly, this serves to extend our sequence {Wn } to
n = m + 1 and completes the induction. This completes the proof of part(a). Note that
this proof does not work when p = 1 since, in this case, there is no F p2 .
We now proceed with the proof of part(b). By Lerays Theorem we may compute
1
space of 0-Cech
cochains on Wj . As in the proof of Theorem 16.5, we may assume that the
sequence {j } is increasing in the sense that j (g|Wj ) k (g) if j < k and g is a 0-cochain
on Wk . Suppose f is a 1-cocycle for the cover {Wn }. We inductively construct a sequence
{gj } where gj is a 0-cochain on Wj such that gj = f on Wj and j (gj gj+1 |Wj ) < 2j
for each j. Suppose such a sequence {gj } has been constructed for all indices j < k. We
use the fact that H 1 (Wk1 , S) = 0 to find a 0-cochain t on Wk such that t = f on Wk .
Then,
(t gk1 ) = 0
in Wk1 . This means that t gk1 is the 0-cochain on Wk1 determined by a section r
S(Wk1 ). Using the density hypotheses, we may choose a global section that approximates
this section as closely as we desire. Thus, there is a global 0-cochain s such that s = 0
and k1 (t gk1 s) < 2k+1 . Then gk = t s has the properties that gk = f on Wk
and k1 (gk gk1 ) < 2k+1 . Thus, by induction, we may construct the sequence {gj } as
claimed. Now on a given Wk consider the sequence {gj )|Wk }
j=k . This is a Cauchy sequence
j
in the metric k since k (gj+1 gj ) < j (gj+1 gj ) < 2 for j k. Furthermore, the
terms of this sequence differ from the first term by cocycles (since the difference is killed
by . Thus, the sequence may be regarded as a fixed cochain plus a uniformly convergent
sequence of cocycles. It follows that this sequence actually converges in the topology of
0-cochains on Wk and the limit hk satisfies hk = f |Wk . Furthermore, hk+1 = hk on Wk
and, hence, the hk determine a 0-cochain h on X. Clearly, h = f . Thus, every 1-cochain
is a coboundary and the proof is complete.
Cartans Theorem B has a host of corollaries. We list some of these in the next few
pages. The first five follow immediately using sheaf theory techniques which are, by now,
familiar to us and so we leave their proofs as excercises:
Corollary 16.9. If X is a Stein space, then every surjective morphism S T of coherent
analytic sheaves induces a surjective morphism S(X) T (X) on global sections.
Corollary 16.10. If Y is a closed subvariety of a Stein space X then every holomorphic
function on Y is the restriction of a holomorphic function on X.
Corollary 16.11. If X is a Stein space, an if {fi } is a finite set of holomorphic functions on
X which does not vanish simultaneously
at any point of X then there is a set of holomorphic
P
functions {gi } on X such that
gi fi = 1. In other words, each finitely generated ideal of
H(X) is contained in a maximal ideal of the form Mx = {f H(X) : f (x) = 0} for some
x X.
158
J. L. TAYLOR
0 Z H H 0
we conclude that:
16.14 Corollary. If X is a Stein space then
(a) H 1 (X, Z) ' H (X)/ exp(H(X));
(b) H 2 (X, Z) ' H 1 (X, H );
Using Corollary 16.14 as in problem 9.9, one can now show that
16.15 Corollary. If X is a Stein space then the group of isomorphism classes of holomorphic line bundles is isomorphic to the group of isomorphism classes of continuous line
bundles which is isomorphic to H 2 (X, Z).
Now suppose X is a connected complex manifold. The rings H(U ) for U a connected
open subset of X are all integral domains. Let M denote the sheaf on X generated by the
presheaf which assigns to each connected open set U the quotient field of H(U ). Then M
is called the sheaf of meromorphic functions on U . We let M be the sheaf of non-zero
elements of M under multiplication. Clearly, H M . The quotient sheaf is denoted D
and called the sheaf of divisors on X.
The long exact sequence of sections associated to the short exact sequence
0 H M D 0
combined with Corollary 16.14 yields an exact sequence
(X, M ) (X, D) H 1 (X, H ) H 1 (X, M )
An element of (X, D) is called a divisor and its image in H 1 (X, H ) = H 2 (X, Z) is
called its Chern class. Now by Corollary 16.15 an element of H 1 (X, H ) corresponds to a
holomorphic line bundle L on X and L corresponds to the zero element of H 1 (X, H ) if
and only if it is the trivial line bundle that is, if and only if it has a section that is nowhere
vanishing. However, if L just has a section that is not identically zero, then it is easy to
see that the class in H 1 (X, H ) corresponding to L is sent to 0 by the map H 1 (X, H )
H 1 (X, M ) (problem 16.7). However, the holomorphic sections of a holommorphic line
bundle form a coherent analytic sheaf. By Cartans Theorem A, every coherent sheaf on
a Stein space is generated by its global sections. We conclude that every holomorphic line
bundle on X has a section which is not identically zero. Hence, H 1 (X, H ) H 1 (X, M )
is the zero map, and we conclude from the above exact sequence and Corollary 16.14 that:
159
y
y
Hk (U ) S(U ) 0
There is an neighborhood of zero in Hk (W ) whose image in Hk (U ) has compact closure
by the result of the previous paragraph. The image of this neighborhood in S(W ) is a
neighborhood of zero in S(W ) by the open mapping theorem and its image in S(U ) will
have compact closure by the commutativity of the diagram. Thus, S(W ) S(U ) is a
compact map. However, S(X) S(U ) is the composition of this map with the restriction
map S(X) S(W ). Since, the composition of a continuous linear map with a compact
map is clearly compact, the proof is complete.
We now prove the Cartan- Serre theorem assuming Schwartzs Theorem. We will end
the Chapter with a proof of Schwartzs Theorem.
16.18 Cartan-Serre Theorem. If X is a compact holomorphic variety and S is a coherent analytic sheaf, then H p (X, S) is finite dimensional for all p.
Proof. We choose a finite open cover W = {Wi } of X consisting of sets which are Stein
spaces. Finite intersections of sets in this cover are also Stein spaces (Problem 15.4). We
160
J. L. TAYLOR
then choose another such cover U = {Uj } which is a refinement of the first cover and,
in fact, has the property that for each j there is an integer (j) such that U j W(j) .
Then for each multi-index = (j0 , . . . , jp ) the multi-index () = ((j0 ), . . . , (jp )) has the
property that U W() so that the restriction map S(W() ) S(U ) is a compact
map. It follows that the refinement morphism : C p (W) C p (U) from the space of
Cech
p-cochains for W to the space of Cech
p-cochains for U is a compact map between
Frechet spaces. Since both covers W and U are Leray covers for S the map induces an
isomorphism of cohomology. Hence, if Z p (W) and Z p (U) are the spaces of Cech
p-cocycles
for W and U, then the map
f g p1 (f ) + (g) : C p1 (W) Z p (W) Z p (W)
is surjective. Since is compact, it follows from Schwartzs theorem (Theorem 16.20)
that p1 has closed image and finite dimensional cokernel. Hence, H p (X, S) is finite
dimensional.
It remains to prove Schwartzs theorem. We first prove a dual version of Schwartzs
theorem from which the theorem itself will follow:
16.19 Theorem. Let X and Y be locally convex topological vector spaces and let A :
X Y be a continuous linear map which has closed image and is a topological isomorphism
onto its image, Let C : X Y be a compact continuous linear map. Then B = A + C
has finite dimensional kernel K, closed image I and the induced map B : X/K I is a
topological isomorphism.
Proof. A and C agree on the kernel K of B. Thus, A|K is a topological isomorphism of
K onto a subspace of Y , but it is also a compact map. This implies that the image of K
in Y under A has a neighborhood of zero with compact closure. However, a topological
vector space which is locally compact is necessarily finite dimensional. It follows that K
is finite dimensional.
Now there is a closed subspace L X which is complementary to K in X. This follows
from the Hahn-Banach theorem. In fact, if {xi } is a basis for the vector space K, then the
Hahn-Banach theorem implies that we can find for each i a continuous linear functional
fi on X with fi (xj ) = ij . The intersection of the kernels of the fi will then be a closed
complement for K. If L is such a complement, then A|L and C|L are continuous linear maps
of L into Y , the first a topological isomorphism onto its image and the second a compact
map. Thus, we are back in our original situation except now we have that A|L + C|L is
injective. Thus, to complete the proof it is enough to prove the theorem in the case where
A + B is injective.
We need to prove that B = A+C has closed image and is a topological isomorphism onto
its image. Thus, let x be a net in X and suppose that B(x ) converges to y Y . Since
C is compact, there is a continuous seminorm on X such that U = {x X : (x) < 1}
is a neighborhood which C maps to a set with compact closure in Y . Suppose {(x )} is
bounded, say by M . Then C(x ) is contained in the set (M +1)C(U ), which has compact
closure in Y . Thus, C(x ) has a cluster point z Y and A(x ) = B(x ) C(x ) has
y z as a cluster point. Since the image of A is closed, we have that y z = Ax for a
161
unique x X. Since A is a topological isomorphism onto its image, we conclude that the
original net {x } has x as a cluster point and Bx = y since B is continuous. Thus, in the
case where {(x )} is bounded, we have that {x } has a cluster point x and B(x) = y.
Now suppose that {(x )} is unbounded, then, after possibly modifying {x } to eliminate terms where (x ) = 0, we may consider the net {x0 } defined by
x0 =
x
(x )
This net has the property that B(x0 ) converges to zero and {(x0 )} is bounded, in fact
(x0 ) = 1 for every . Thus, we are back in the previous case except that y has been
replaced by zero. We conclude that {x0 } has a cluster point x and B(x) = 0. However,
this is only possible if x = 0, since we are assuming that B is injective, and zero cannot
be a cluster point of a net of elements x0 with (x0 ) = 1. Thus, the first case was the
only one possible and we conclude that every net {x } X for which B(x ) converges in
Y has a cluster point in X. This is enough to imply that the image of B is closed and the
inverse of B on that image is continuous. This completes the proof.
It turns out that the previous theorem is the dual of the one we want. In order to prove
Schwartzs theorem we must first define a topology on the dual of a Frechet space and
prove an important theorem of Mackey and Arens concerning duality. There are many
ways to topologize the dual of a locally convex topological vector space. The useful ways
are of the following type:
A saturated family of bounded subsets of a topological vector space X is a family which
is closed under subsets, multiplication by scalars, finite unions and closed convex balanced
hulls. Such a family is said to cover X if X is the union of the sets in the family.
16.20 Definition. If X is a locally convex topological vector space and is a saturated
family of bounded subsets of X which covers X, then X will denote the space of continuous
linear functionals on X with the topology of uniform convergence on sets in .
Clearly X is a locally convex topological vector space. A family of seminorms defining
the topology is the family of all seminorms of the form K where K is a set in and
K (f ) = sup{|f (x)| : x K}. The family of sets we want to use is the family c of all sets
with compact closure. This is not always a saturated family since it is not always true
that the closed convex, balanced hull of a compact set is compact. However, for Frechet
spaces we have:
16.21 Theorem. If X is a Frechet space, then the closed convex, balanced hull of every
compact subset of X is also compact.
Proof. If D is the closed unit disc in C and K X is compact, then D K is the image of
the compact set D K under the scalar multiplication map and is, hence, compact. The
closed convex hull of D K will be a closed convex balanced set containing K and, hence,
to prove the Theorem it suffices to prove that the closed convex hull of a compact set in a
Frechet space is compact.
Recall that a subset of a complete metric space is compact if and only if it is closed and
totally bounded. A subset S of a metric space is totally bounded if for each > 0 there is
a finite set F in S so that each point of S is within of some point of F .
162
J. L. TAYLOR
si yi : S X
P
where S is the simplex {(s1 , , sn ) (R+ )n :
si = 1}. Thus, L is compact. Furthermore, every element of the convex hull of K is within /2 of a point of L. Since L is
compact, it is also totally bounded and we may find a finite set of points {xj } such that
every point of L is within /2 of some xj . It follows that every point of the convex hull
of K is within of some xj . Thus, we have proved that the convex hull of K is totally
bounded. The closure of a totally bounded set is clearly totally bounded as well. Thus, the
closed convex hull of a compact subset of a Frechet space is totally bounded and, hence,
compact.
The above theorem implies that the family c of sets with compact closure in a Frechet
space is a saturated family. The topology on X which it determines is the topology of
uniform convergence on compact subsets of X. The space X with this topology will be
denoted Xc .
Our final preliminary result before proving Schwartzs Theorem is the following theorem
of Mackey-Arens. Note that, for any saturated family covering X, each element x X
determines a continuous linear functional on X by f f (x) : X C. Thus, X embedds
in the second dual (X ) . In the case where the sets in are precompact, every continuous
linear functional on X has this form.
16.22 Mackey-Arens Theorem. If X is a locally convex topological vector space and
is a saturated family of precompact subsets of X which covers X, then every continuous
linear functional on X is determined by an element of X. Thus, X = (X ) .
Proof. If is a continuous linear functional on X then there is a neighborhood V of zero
in X such that |(f )| < 1 for all f V . We may assume that the neighborhood V has
the form
V = {f Xc : sup |f (x)| < 1}
xK
for some K . Without loss of generality we may assume K is compact convex and
bounded since is saturated and consists of precompact sets. We regard X as embedded
in (X ) and we give the later space the weak-* topology that is, the topology of pointwise
convergence as functions on X . This may also be described as the topology of uniform
convergence on the family of sets which are convex balanced hulls of finite sets. On the
image of X in (Xc ) the topology is weaker than the original topology on X and, hence,
K is also compact in this topology. If we knew the result of the theorem in the case of the
topology, then we could conclude that every continuous linear functional on (Xc ) with
the topology is given by an element of Xc . Let us assume this for the moment. If is
not an element of K, then it follows from the Hahn-Banach theorem (TVS 5(b)) that there
exists a continuous linear functional f on (Xc ) such that |f ()| > 1 and |f (x)| < r < 1
163
for all x K. However, under our assumption about the topology, we must have that
f Xc and it then follows that f V . This is a contradiction, since |f ()| = |(f )| < 1
if f V . We conclude that does belong to K and, in particular, belongs to X.
It remains to prove the theorem in the case where is the family generated by finite
sets. In this case, let and V and K be as above. Since, = we have that K is the
convex balanced hull of a finite set {xi }ni=1 . Then the set of f X such that f (xi ) = 0
for i = 1, . . . , n is just k k 1 V and so it follows that vanishes on this set. It then follows
from elementary linear algebra that must be a linear combination of the xi and, hence,
must belong to X.
16.23 Theorem (Schwartz). Let X and Y be Frechet spaces and suppose that A : X
Y is a surjective continuous linear map and C : X Y is a compact continuous linear
map. Then B = A + C has closed image and finite dimensional cokernel.
Proof. We consider the duals Xc and Yc of X and Y in the topology of uniform convergence
on compact subsets. Then the continuous linear map A : X Y has a dual A : Yc Xc
defined by
A (f )(x) = f (A(x))
Since A is continuous, it maps compact sets to compact sets, from which it follows that
A is continuous. Similarly, C is continuous.
Claim 1. The linear map C : Yc Xc is compact.
Since C is compact, there is a zero neighborhood U in X such that C(U ) has compact
closure K in Y . Then the zero neighborhood VK = {f Yc : supxK |f (x)| < 1} has
the property that C (VK ) is a family of continuous functions which is uniformly bounded
by one in modulus on U . If L is any compact subset of X then C (VK ) is uniformly
bounded on L, since L kU for some k > 0, and C (VK ) is equicontinuous on L, since
|f (x)f (x0 )| < if xx0 U and f C (VK ). It follows from the Ascoli-Arzela Theorem
that C (VK ) has compact closure in the space of all functions on X in the topology of
uniform convergence on compact subsets of X. However, the space of continuous linear
functionals is clearly closed in this topology since X is a metric space. Thus, C (VK ) has
compact closure in Xc and C is a compact operator.
Claim 2. The linear map A : Yc Xc has closed image and is a topological isomorphism
onto its image.
Let {f } be a net in Yc such that A (f ) converges in Xc to g. Let K be a compact
subset of Y . The fact that A is an open map implies that K is the image under A of a
compact subset L of X (Problem 16.8). We have that A (f ) = f A converges to g
uniformly on L and this implies that f converges uniformly on K. Since this is true for
every compact set K Y , the net {f ) converges in the topology of Yc to an element f .
Clearly A (f ) = g. Hence, the image of A is closed and, on the image, the inverse map
is continuous. This establishes Claim 2.
We now have the hypotheses of Theorem 16.19 satisfied for the pair of operators A and
C . We conclude that B = A + C has finite dimensional kernel Z and closed image and
induces a toplogical isomorphism from Yc /Z to its image in Xc . Now the map Yc /Z Xc
164
J. L. TAYLOR
induced by B is the dual of the continuous linear map B 0 obtained by composing B with
the map Y Z defined by y {f f (y)} f Z, y Y . Thus, to finish the proof
we need to show that B 0 is surjective. In other words, the proof will be complete if we can
prove:
Claim 3. If B : X Y is a bounded linear map between Frechet spaces and B : Yc
Xc is injective, has closed image and is a topological isomorphism onto its image, then B
is surjective.
The map B induces a continuous linear map X/ker B Y which has as dual the
map B as a map from Yc to {f Xc : f (x) = 0 x ker B}. This map has closed
image and so if it is not surjective, then the Hahn-Banach Theorem implies that there is a
continuous linear functional F on X which vanishes identically on B (Yc ) and does not
vanish identically on {f X : f (x) = 0 x ker B}. However, by the Mackey-Arens
Theorem (Theorem 16.22), the functional F has to have the form F (f ) = f (x) for some
x X. In other words, there is an x X such that B (g)(x) = g(B(x)) for all f Yc
but g(x) 6= 0 for some f Xc which vanishes on ker B. This is impossible, since such an
x would necessarily be in ker B. Thus, B is surjective as a map from Yc to the dual of
X/ ker B. It follows that we may assume without loss of generality that B is injective and
B is surjective.
Thus, suppose that {xn } is a sequence in x such that {B(xn )} converges to y Y .
Then the set S = {B(xn )} {y} is compact. Hence, the set
U = {f Y : |f (y)| < 1
n}
x K} B (U ) = {f Xc : |f (xn )| < 1
x K}
or
{f X : |f (x)| < 1
x 1 K} B (U ) = {f Xc : |f (xn )| < 1
x K}
From the Hahn-Banach theorem (TVS 5(b)), it now follows that the sequence {xn } lies
in the compact convex balanced hull of 1 K and, hence, lies in a compact subset of X.
Then it has a cluster point x and, clearly, B(x) = y. Thus, B is surjective. This completes
the proof of claim 3 and the proof of the Theorem.
16. Problems
1. Prove that a bounded linear map : X Y between two Frechet spaces is an open
map if and only if every convergent sequence in Y lifts under to a convergent sequence
in X.
2. Prove Corollary 16.9.
3.
4.
5.
6.
7.
165
166
J. L. TAYLOR
h}.
Since h is maximal abelian, the space g0 is h itself. The non-zero elements h for
which g is non-empty are called roots. We denote the set of all roots by . Thus,
X
g=h
g .
167
if
6= .
Since, h = g0 , this implies that h is othogonal to each g for a root. This in turn implies
that the Killing form remains non-singular when restricted to h. Thus, the Killing form
induces an isomorphism between h and its dual h . That is, to each h there is a unique
h such that () = h, i. Then h, i = h , i defines a non-singular bilinear form
on h which we shall also call the Killing form. If we let h0 be the real subspace of h
spanned by the set of roots, then one can prove that h0 is a real form of h i. e. is a
real subspace with the property that h = h0 ih0 . Furthermore, the restriction of the
Killing form to h0 is positive definite. Again, for proofs we refer the reader to any standard
treatment of Lie algebra theory.
A positive root system for a Cartan h is a subset + such that: (i) it is closed
under addition in the sense that if , + , then + + provided + is a root;
and (ii) for every root , exactly one of , belongs to + . Such a system may
be constructed by choosing a real hyperplane through 0 in h which does not meet and
then letting + be the set of all roots on one side of this hyperplane. If + is a positive
root system, then so is its complement in and this is usually denoted . A system of
positive roots determines two Borel subalgebras b+ and b by
X
X
b+ = h
g , b = h
g ,
+
168
J. L. TAYLOR
17.3 Theorem. If B is a Borel subgroup of a complex semisimple Lie group G, then G/B
is a compact complex manifold.
Thus, the set X of Borel subalgebras of g may be given the structure of a compact
complex manifold through its identification with G/B. We call this the flag manifold
of G. Note also that G acts on X through left multiplication if X is represented as
the coset space G/B or, equivalently, through Ad if X is represented as the set of Borel
subalgebras of g. This action is, of course, transitive and is holomorphic in the sense that
g x gx : G X X is a holomorphic map.
We are interested in studying the finite dimensional representations of g. Let (, V ) be
such a representation. By Lie theory, exponentiates to a holomorphic representation of
the Lie group G on V which we also denote by . If h is a Cartan subalgebra of g, and
H is the connected Lie subgroup of G corresponding to h, then H is a connected complex
abelian Lie group; hence, it must be a complex torus that is, a product of copies of
the punctured plane C . The corresponding product of unit circles is a maximal compact
subgroup T of H. The complex vector space V may be given a T -invariant inner product
(by integrating any inner product over the compact group T ). Then (, V ) is a finite
dimensional unitary representation of T . It follows that V may be written as a direct sum
of subspaces on which the representation of T is irreducible and each of these subspaces
consists of eigenvectors for the operators (t) for t T . It follows that they are also
eigenvectors for the operators () for h. In other words,
17.4 Theorem. The restriction of the representation to h is completely reducible.
From this it follows that V is a direct sum of subspaces V where h and
V = {v V : ()v = ()v
h}
The elements h for which V 6= 0 are called the weights of the representation .
Since the representation is finite dimensional, there can be only finitely many of them.
How does the rest of the Lie algebra g act on V ? We get an idea by using the root space
decomposition. Thus if and g , then a calculation shows that ()V V+ .
Now, if the representation is irreducible, then for any weight the weight space V must
generate V . It follows that one must be able to obtain all the weights from a given one by
successively adding roots.
17.5 Definition. For a finite dimensional representation (, V ), a Cartan h, a positive
root system + and weights and , we say < if is a sum of positive roots. A
weight is called a highest weight if it is maximal relative to this relation. If a weight has
the property that, for each root , either + or is not a weight for (, V ), then
will be called an extremal weight.
If is a highest weight, then is an extremal weight since + fails to be a weight
for every positive root . We also have (g )V = 0 for every + since V+ = 0. It
follows from this and the commutation relations in g that the g-invariant subspace of V
generated by any non-zero vector v V is contained in the span of v and the spaces V
for < . This implies the following theorem:
169
h, i
h, i
is reflection through the hyperplane determined by . The Weyl group W for g is the
group of transformations of h0 generated by the set of reflections s .
It is clear that the Weyl chambers are in one to one correspondence with the possible
choices of positive root systems. That is, for each Weyl chamber there is exactly one
positive root system for which it is the positive Weyl chamber. It is also easy to prove that
the Weyl group acts transitively on the set of Weyl chambers.
We define a norm in h0 by setting
|||| =
p
h, i.
It turns out, though we shall not prove it here, that each element of the Weyl group arises
from a transformation of h of the form Adg , where g G is in the normalizer of h. In
fact W is isomorphic to the normalizer of h in G mod the centralizer of h in G. It follows
from this and the G invariance of the Killing form, that the Weyl group acts as a group of
isometries relative to the above norm.
We set
1 X
=
2
+
This element plays a special role throughout the theory of semisimple Lie algebras.
17.7 Theorem. Let (, V ) be a finite dimensional irreducible representation of g, h g
a Cartan subalgebra, + a positive root system, h the set of weights for (, V ) and
the highest weight. Then
2h, i
(a)
Z , ;
h, i
(b) is closed under the action of the Weyl group;
170
J. L. TAYLOR
1
(|| + ||2 + || ||2 ) ||||2
2
by part (d) and this implies that |||| < ||||. Thus, only the extremal weights can have
norm equal to ||||. The proof of part (e) will be complete if we can show that the Weyl
group acts transitively on the set of extremal weights, since this will imply that they all
171
have norm equal to that of . Using a Weyl group transformation, we can move any
extremal weight into the closure of the positive chamber and it will still be an extremal
weight. Thus, we will have completed the proof of (e) if we show that is the only extremal
root in the closure of the positive chamber. Thus, suppose that is extremal and lies in
the closure of the positive chamber. Then for each + we have h, i 0 and either
+ or is not a root. However, the sl2 (C) argument of the first paragraph shows
that it must be + that fails to be a root if h, i 0. Hence, is the highest root
and the proof of part(e) is complete.
If is any weight in then
h + , + i = h, i + 2h, i + h, i h, i + 2h, i + h, i
= h + , + i 2h , i < h + , + i
by part (c) and the fact that is a sum of positive roots. This proves part (f).
The elements of h that satisfy the condition in part(a) of the above theorem are called
integral weights.
17.8 Definition. We say that a weight h0 is dominant relative to a system of positive
roots if h, + i > 0 for every positive root . Thus, a weight is dominant if and only
if + belongs to the positive Weyl chamber.
We know that for every finite dimensional irreducible representation there is a unique
highest weight and it is easy to see from part(c) of Theorem 17.7 that highest weights
are dominant. In fact, using the theory of Verma modules one can prove that the finite
dimensional irreducible representations of a complex semisimple Lie algebra are classified
by their highest weights. We wont prove it here, but the theorem that does this is the
following:
17.9 Theorem. Let h g be a Cartan subalgebra and + h a system of positive
roots. Then each dominant integral weight is the highest weight for a unique (up to
isomorphism) finite dimensional irreducible representation of g.
We now introduce the objects of study in the Borel-Weil-Bott Theorem the Gequivariant holomorphic line bundles. Each of these is constructed from an integral weight
in h by an induction process which we shall describe below. In what follows, H will denote
the sheaf of holomorphic functions on either G or X. If it is not clear from the context
which is meant, we shall use the notation G H or X H. We also fix, for the remainder of the
discussion, a Cartan subalgebra h, a system of positive roots + and the Borel subalgebra
b = b spanned by h and the negative root spaces. The corresponding Borel subgroup of
G will be denoted B. We represent the flag manifold X as G/B.
Let (, W ) be a finite dimensional holomorphic representation of B equivalently, a
finite dimensional complex linear representation of b. The space G B W is constructed
from G W by identifying points which lie in the same orbit of the B-action described
by b (g w) gb1 (b)w. Projection on the first factor of G W , followed by the
projection : G G/B, induces a well defined holomorphic projection GB W X. The
inverse image of the typical point under this projection is gBB W , which has a well defined
172
J. L. TAYLOR
b B}
173
174
J. L. TAYLOR
Proof. Since is a highest weight, the weight space V is a complement to the span of the
subspaces (g )W for . Thus, if W is the space of all elements of V which vanish
on this span, then there is a non-zero element w W . Let ( , V ) denote the dual of
the representation (, V ). Clearly, ( , V ) is also a faithful, irreducible representation.
For the corresponding representation of G, we have
( (g)w)(v) = w((g 1 )v) w V
It follows that W is the lowest weight space of ( , V ) and has weight . Also
(g )W = 0 if . On the other hand, (g )W 6= 0 for each + . For
if g and g both annihilate W , then it follows from the commutation relations among
the root spaces and from the fact that W generates V , that g and g both annihilate all of V . This P
is impossible since is a faithful representation. Thus, the Borel
subalgebra b = h +
g is the stabilizer of W in g. That is,
b = { g : ()W W }.
It follows from this that the corresponding Borel subgroup B is the stabilizer of the one
dimensional subspace W under the action of G on V . Thus, under the action of G on
the complex projective space P (V ) of V induced by , the point p0 P (V ) determined
by W has B as stabilizer. Hence, the orbit of p0 under this action of G is compact, hence
closed, and is a copy of X = G/B. Since the action of G on P (V ) is holomorphic, it
follows from the complex implicit function theorem that the orbit is a compact, complex
submanifold of P (V ). Thus, we have a closed holomorphic embedding : X P (V ).
Now a typical section of the sheaf H(1) on an open subset of P (V ) is determined by
a homogeneous function of degree one on the corresponding open subset of V that is,
by a linear functional on V , i. e. a vector v V . If U is an open subset of X, then the
pullback of this section to U may be identified with the function f on 1 (U ) G defined
by g (g)w(v), where w is a non-zero element of W . This function satifies
f (gb1 ) = (gb1 )w(v) = e (b) (g)w(v) = e (b)f (g)
and, hence, determines a section of H . It follows that 1 H(1) = H .
The above result shows why we choose to depart from the standard convention in the
theory of group representations which assigns to a Borel b the system of positive roots for
which b = b+ . Instead, we assign to b the system of positive roots for which b = b . With
this convention, the of the above theorem is a highest weight rather than a lowest weight.
This is a better convention because then highest weights induce positive line bundles rather
than negative line bundles. This is the convention we will use throughout this section.
As an immediate corollary of the above theorem, we have:
17.14 Corollary. The flag manifold X is a non-singular projective variety.
The Casimir operator is a particular element of the center Z(g) of the enveloping algebra
U (g). It may be described as
X
X
X
X
(2 + [ , ])
( + ) =
i2 +
=
i2 +
+
175
where the i form a self dual basis for h relative to the Killing form, so that hi , j i = ij .
The elements g are chosen so that h , i = 1. For each the element =
[ , ] belongs to h and has the property that ( ) = h, i for each h , where h, i
also stands for the form on h which is dual to the Killing form. Thus, we have
X
X
X
=
i2 +
+ 2
.
+
P
From the definition of the dual form on h , it follows that
(i )2 = h, i. We also have
( ) = h, i. Thus,
X
()v = h, i +
h, i = h, i + h, 2i = h, + 2i
+
d
d
`( )f (exp(t )b)|t=0 = e (b1 exp(t ))|t=0 `( )f (e) = 0
dt
dt
and
`()f (b) =
d
d
f (exp(t)b)|t=0 = e (b1 exp(t))|t=0 f (e)
dt
dt
1
= ()e (b )f (e) = ()f (b)
for h. That `()f (b) = h, + 2if (b) now follows as in the previous theorem from
the identity
X
X
X
.
+ 2
=
i2 +
+
176
J. L. TAYLOR
This says that the sections `()f and h, +2if of the line bundle corresponding to H()
agree at the point x. However, since x was arbitrary, this completes the proof.
Note that
h, + 2i = h + , + i h, i = || + ||2 ||||2
177
The elements p are necessarily weights which occur in the representation V and, hence,
are weights dominated by . When we tensor by the one dimensional representation of
B with character e , we still have the same filtration but each weight p is changed to
p = p + .
It follows from Theorem 17.11 that the sheaf I( ) is filtered by a sequence of subsheaves
{I( p )}, where p is restricted to V p and the subquotient I( p )/I( p+1 ) is isomorphic
p
p
p
to Hp . By Theorem
Q 17.16, pthe pCasimir acts on H as the scalar h , + 2i. From
this it follows that p (h , +2i) acts as zero on I( ). This and Problem 17.3 then
imply that I( ) decomposes as a direct sum of subsheaves which are the - eigenspaces
for the eigenvalues hp , p + 2i).
Note that
p + = p + + = p + w = w(w1 p + )
so that each p + is of the form w( +), where is a weight for the representation (, V ).
It follows from Theorem 17.7(f) that either ||p + || < || + || or p + = w( + ).
In the latter case, p + = w + w + = w + + , so that p = w + .
Since h, + 2i = || + ||2 ||||2 for any , this implies that the only p for which
hp , p + 2i = h, + 2i is p = w + . From this it follows that the summand of I( )
on which h, + 2i vanishes must be a copy of Hw+ since this sheaf appears with
multiplicity one as a subquotient of I( ). Then
H p (X, I( )) = H p (X, Hw+ ) H p (X, J )
Where J is a summand of I( ) on which h, + 2i is non-vanishing. However,
we proved above that the Casimir acts as the scalar h, + 2i on the left side of this
equality. It acts as this same scalar on the first term on the right side but it acts on the
second term as an operator with eigenvalues all distinct from h, + 2i. It follows that
H p (X, J ) = 0 for all p and
H p (X, H ) V = H p (X, I( )) = H p (X, Hw+ )
for all p. This completes the proof.
The above is a very strong result. In particular, it implies the following:
17.18 Corollary. The set of integers p such that H p (X, H ) is non-vanishing is constant
as + varies over the integral weights in a given Weyl chamber.
Proof. If + belongs to the Weyl chamber which is the image of the positive chamber
under w W , then there is a dominant root such that + = w( + ). Thus,
= w + where = w . It then follows from the previous lemma that H p (X, H )
is non-vanishing if and only if H p (X, H ) is non-vanishing. Since, depends only on the
chamber determined by w, the proof is complete.
The Borel-Weil Theorem follows easily from the Lemma 17.17 and the above corollary
applied to the case where w is the identity, so that w + = .
178
J. L. TAYLOR
179
Now for the proof of the Theorem. If + belongs to a wall, then its Weyl group orbit
consists entirely of elements which belong to walls. If H p (X, H ) 6= 0 for some p, then
it will be a g-module with infinitesimal character by Theorem 17.16. However, there
is no finite dimensional representation with infinitesimal character since every finite
dimensional irreducible representation has infinitesimal character for dominant and
the Weyl group orbit of a dominant weight does not meet any walls.
We need one more lemma before proving the Borel-Weil-Bott Theorem:
17.21 Lemma. Let (, V ) be a finite dimensional irreducible representation of g and let
be its set of weights. Let + be a system of positive roots and let be an element
of + . If h0 satisfies h, i = 0 and h, i > 0 for + distinct from , then the
maximal value of || + || for is achieved at exactly two points, = and = s (),
where is the highest weight in .
Proof. Since ||||2 = h, i is a convex function of in h0 , the maximum clearly can occur
only at weights + for which is an extremal weight in . Given two extremal weights
and , we have |||| = |||| by Theorem 17.7 and, hence,
|| + ||2 || + ||2 = 2h, i
If is the highest weight in then is a sum of positive roots and so || + ||2 || +
||2 > 0 except in the case where involves only the root , i. e. has the form n for
some n. The only extremal weight of this form is s ().
Two Weyl chambers are adjacent if their closures have a wall in common. The reflection
through that wall will then interchange the two chambers. If is a positive root defining
the wall (that is, the wall is { h0 : h, i = 0}), then h, i and hs (), i will have
opposite signs, while h, i and hs (), i will have the same sign for other positive roots
. Thus, for in one of the two chambers there will be exactly one more negative number
in the set {h, i : + }. The distance from a Weyl chamber (or one of its elements)
to the positive chamber is the minimal number of such wall crossings needed to pass from
the positive chamber to the given chamber. Thus, it is the number of negative numbers in
the set {h, i : + } for in the chamber. The length of a Weyl group element w is
the distance from w to the positive chamber.
17.22 Borel-Weil-Bott Theorem. Let be an integral weight. Then
(a) if + = w(+) for a dominant weight and w W of length d, then H d (X, H )
is isomorphic to the irreducible g module of highest weight ;
(b) H p (X, H ) vanishes in all other cases.
Proof. We already know from Theorem 17.20 that H p (X, H ) = 0 if + lies in a wall.
Thus, we may assume that + lies in a Weyl chamber i. e. that = w + for a
dominant integral weight , where = w . By Theorem 17.17 we have
H p (X, H ) = H p (X, H ) V
where V is the irreducible finite dimensional representation of highest weight . Thus, the
theorem is true for a given with + in the chamber determined by w if and only if it
180
J. L. TAYLOR
is true for the weight = w that is, if and only if H p (X, H ) = 0 unless p is the
length of w, in which case H p (X, H ) = C. Of course, this means that the theorem is true
for all weights with + in a given chamber if and only if it is true for one such weight.
We prove the theorem by induction on the distance d from our chamber to the positive
chamber. The case d = 0 is the Borel-Weil theorem. Thus, we suppose the theorem is true
for Weyl chambers at distance d 1 from the positive chamber and consider a chamber C
at a distance d, obtained by applying w of length d to the positive chamber. We choose
an integral weight which is in the positive chamber and is the highest weight of an
irreducible finite dimensional representation (, W ). Then w is in the our chamber C.
Now C is adjacent to a chamber C 0 at distance d 1 from the positive chamber. Let
be the positive root defining the wall separating the two chambers. Then w0 = S w is
the Weyl group element of length d 1 which maps the positive chamber to C 0 . We set
= w0 () C 0 .
As in the proof of Lemma 17.17, let be the representation restricted to the Borel
B and let be the tensor product of with the one dimensional representation of B
determined by an integral weight . Here we choose to be a weight that satisfies
h + , i = 0,
h + , i > 0 for
+ , 6=
For example, = w + w0 has this property. We then consider the induced bundle
I( ) and its sheaf of sections I( ). As in the proof of Lemma 17.17, I( ) = I(e ) I()
as G-equivariant bundles, I( ) = I(e ) W as sheaves of g-modules, and for each p
H p (X, I( )) = H p (X, H ) W
as g-modules. Since, + is in a wall, it follows from Theorem 17.20 that
H p (X, H ) = 0 and, hence,
H p (X, I( )) = 0
for all p.
Also, as in the proof of Lemma 17.17, the sheaf H decomposes into eigenspaces for the
action of the Casimir and the possible eigenvalues are of the form h + , + + 2i =
|| + + ||2 ||||2 where is a weight of the representation (, W ). Now we apply
Lemma 17.21 in the case where the positive root system is the one for which + is in
the positive chamber and is + . It implies that the maximal value for the expression
|| + + ||2 ||||2 , as ranges over the weights of , is achieved only for = and
= s ().
We have
h, i
s () = 2
h, i
and, since h, i > 0, this implies that s () < in our ordering of integral weights. We
may define a B-submodule W 0 of W to be the span of all weight spaces for weights < .
If W 00 is the quotient W/W 0 , then we have a short exact sequence of B-modules
0 W 0 W W 00 0
181
with W 0 containing the weight space for weight s () and W 00 containing the weight
space for weight . If we tensor this sequence with the one dimensional representation
with character e and then induce, we are led to a corresponding short exact sequence of
sheaves of g-modules:
0 A I( ) B 0
If we then apply the projection onto the -eigenspace for eigenvalue t = || + + ||2
||||2 = ||s () + + ||2 ||||2 , we obtain a short exact sequence of sheaves:
0 At I( )t Bt 0
From the construction, it is clear that Bt = H( + ) and At = H(s () + ). Since I( )
has vanishing cohomology in all degrees, the same thing is true of its direct summand
I( )t . From the long exact sequence of cohomology, we conclude that
H p+1 (X, s () + ) ' H p (X, + )
for all p. Since + is in the wall determined by , it is fixed by s and so, since is in
the chamber C 0 , it follows that + + is also in C 0 . Then
s () + + = s ( + + )
is the corresponding element of the chamber C. Since we have assumed the theorem true
for all + in C 0 and, hence, for + = + + , the above identity for cohomology
shows that the theorem is also true when + is the element s ( + + ) of the chamber
C. As noted above, the theorem is true for all + in a chamber if it is true for one. This
completes the proof.
17. Problems
1. Let (, W ) be a finite dimensional holomorphic representation of a Borel subgroup B.
Prove that the induced bundle I() on X is a trivial G-equivariant vector bundle if and
only if is the restriction to B of a holomorphic representation of G.
2. Prove that if (, W ) is a finite dimensional holomorphic representation of B then there
is a holomorphic action of G on H p (X, I()) for each p.
3. Suppose that A is a linear operator on a vector space V (not necessarily
finite dimenQ
sional) and suppose that 1 , . . . , n are distinct scalars such that i (A i ) = 0. Then
prove that V decomposes as a direct sum of eigenspaces of A with eigenvalues 1 , . . . , n .
2h, i
2h, i
4. Let and be roots. Use the fact that
and
are integers (Theorem
h, i
h, i
2h, i
17.7) to prove that if |||| |||| then
can only have the values 0, 1, or 1.
h, i
5. Use the result of Problem 4 to prove that if and are non-proportional roots and
h, i > 0 then is a root.
182
J. L. TAYLOR
183
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.