Grauert 1979
Grauert 1979
for arbitrary complex spaces. The sheaf !lJ •= .H* j(!)* is called the sheaf of germs of
divisors. In the case of a Riemann surface, every stalk !lJ" is isomorphic to l., and
every non-trivial s e !lJ(U) over an open set U, has discrete support Is I in U. This
is the so-called skyscraper property of !lJ.
For a compact Riemann surface X, the divisor group,
Div X·=~(X1
D= L nxx,
xeX
where nx E l. and n" = 0 for almost all x. Throughout we write Qx(D) instead ofn".
The integer
deg D •= L Qx(D)
xeX
is called the degree of D. The mapping Div X --+ l, which associates deg D to D, is
a group epimorphism.
If Qx(D) ~ 0 for every x eX then the divisor Dis called positive. For divisors
D1o D2 e Div X, we write D 1 ~ D2 when D2 - D1 is positive. The group of divi-
sors is directed with respect to this relation. In other words, given two divisors D 1o
D 2 e Div X, there exists D 3 e Div X so that D 1 ~ D3 and D 2 ~ D 3 •
The set ID I •= {x e X: Qx(D) ::/= 0}, called the support of D, is always finite.
to be the order of s" with respect to F. If g(sx) > 0 (resp. g(sx) < 0) then we call x a
zero (resp. a pole) of s". The situation g(sx) = oo only occurs for the identically
zero germ. It follows that F" = {sx E F':: g(sx) ~ 0} and mxFx = {sx E F'::
g(sx) > 0}.
206 Chapter VII. Compact Riemann Surfaces
is called the divisor (with respect to F) of s. The integer deg(s) is called the degree
(with respect to F) of s.
It follows that (s) is positive if and only if s has no poles, or equivalently, if
s E F(X).
Warning: The order functions o and the divisors (s) of sections s e F 00 (X)
depend heavily on the sheaf with which one starts out. For example, with respect
to(!), the zero divisor is associated to the section 1 e (!) 00 (X). On the other hand, if
(!)is replaced by (!)(D) (see Paragraph 3) for some De Div X then (1) =D. In the
following it will always be completely clear with respect to which sheaf we are
forming divisors.
From the above remarks it follows that every meromorphic function
he (!) 00 (X)* has an associated divisor (h). Such divisors are called principal divi-
sors. The map .H(X)*--+ Div X, which sends h to its divisor (h), is a group homo-
morphism. For all he (!) 00 (X)* and s e F 00 (X)*, it follows easily that
(hs) = (h) + (s ). The image group,
Div X/P(X)
l. The Sheaves F(D). Given a locally free sheaf F and a divisor D, we define an
analytic subsheaf F(D) of F 00 by
xeX
Lemma. Let§",§" 1> and§" 2 be locally free sheaves and D, D~> D 2 divisors on a
compact Riemann surface X. Then
1) Every exact l!/-sequence 0 -+ §" 1 -+ §" -+ §" 2 -+ 0 determines in a natural
way an exact l!/-sequence 0-+ §" 1 (D)-+ §"(D)-+§" 2 (D)-+ 0.
2) If§"= §" 1 + §"2 then §"(D)= §" 1 (D) + §" 2 (D).
3) There is a naturall!/-isomorphism §"(D 1 )(D 2 ) ~ §"(D 1 + D 2 ).
4) If D 1 :::;; D 2 then F(D 1 ) is an analytic subsheaf of F(D 2 ).
The reader can easily carry out the proof. We note here that property 4) will
play an important role in the next section.
3. The Sheaves l!?(D). The above considerations are in particular valid for
§" = l!l. All sheaves l!/(D), De Div X, are locally free of rank 1. Two sheaves, l!I(Dt)
and l!I(D 2 ), are analytically isomorphic if and only if D 1 and D2 are linearly
equivalent.
Proof: The sheaves l!/(Dt) and l!I(D 2) are analytically isomorphic if and only if
l!I(D 1 - D 2) ~ l!l. Let D •= D 1 - D 2. Assuming that D 1 and D 2 are linearly equiv-
alent, D =(h) with he Jt'(X)*. In this case we obtain an l!/-isomorphism
l!i-+ l!?(D) by fx 1--+fx hx. Conversely, given an l!/-isomorphism (!;-+ l!?(D), the image
of 1 e ~(X) in l!I(D)(X) is a meromorphic function he Jt'(X)* with D =(h). 0
By tensoring F with l!/(D), one gets all sheaves §"(D). This is seen from the
natural @-isomorphism F ®~(D)-+ F(D), defined by sx ® hx~--+hxsx.
related to the short exact sequence of sheaves, 0 --+ (9* --+ vlt* --+ f0 --+ 0, is in fact
defined by D ~----+(!)(D). The kernel of~ is just the group P(X) of principal divisors
and therefore one has a natural injection,
1. The Sequence 0--+ ff(D)--+ ff(D')--+ §"--+ 0. Let D, D' be divisors with
D ::::;; D' and let ff be a locally free sheaf of rank r. Then there is a natural exact
sequence
0--+ ff(D)--+ ff(D')--+ §"--+ 0,
where§":= ff(D')/ff(D). Since this sequence plays such an important role in our
considerations, we will now write down its basic properties: From the definitions
it follows that for every x E X
(1)
Consequently the first part of the cohomology sequence associated to (•) is the
exact sequence
Thus
(4) if D::; D' then dime H 1 (X, ff(D)) ~dime H 1 (X, ff(D')).
The claim of the lemma will then follow with D' == 0. First we suppose that
D::; D'. The alternating sum of the dimensions of the vector spaces in the exact
cohomology sequence (3) is therefore zero. In other words
If one substitutes r deg(D'- D) for dime ff(X) (see (2) above) then (o) follows
immediately.
If D' is arbitrary, then one chooses D" E Div X with D ::; D" and D' ::; D". Then
it follows from the above that
We have therefore established that every locally free sheaf fF 0 has non- +
identically zero meromorphic sections. This is clear from Theorem 2, since fF(D)(X)
is always contained in fF""(X).
for any p eX, and all n e 7L.. Hence there exists n0 e N so that l!J(np)(X) contains a
non-constant function h for all n > n0 • Since (h) + np ;?: 0, every such function is
non-constant, and holomorphic on X\p, and has a pole of order at most n at p.
This shows the following:
For every p e X there is a non-constant meromorphic function on X which is
holomorphic on X\p. 1 Moreover, X\p is Stein. In particular every compact Riemann
surface can be covered by two Stein domains.
Proof: Let h be as above. Then h: X\p-+ CIs a finite holomorphic map. Thus
X\p is Stein. If p 17 p2 e X are different points, then {X\p 17 X\p2 } is a Stein cover of
X. D
The following is now a consequence of the general theory.
Hq(X, f/') = 0,
For every compact complex space X and every coherent sheaf f/' on X, almost
all of the groups Hq(X, f/') vanish. Thus the Finiteness Theorem allows us to
define the Euler-Poincare Characteristic,
x(f/') = Xo(f/').
1 With a bit more effort one can show at this point that for every p e X there exists n0 e N so that
for every n > n0 there ish e .K{X) which is holomorphic on X\P, and which has a pole of order nat p
(a forerunner of the Weierstrass gap Theorem~
§ 3. The Riemann-Roch Theorem (Preliminary Version) 211
Proof: If D and D' are linearly equivalent then, from Paragraph 1.3, @(D)~
@(D') and consequently Xo(@(D)) = x0 (@(D')). The characteristic formula yields
The reader should note that if X =I= !P' 1, then not every divisor D with deg D = 0
is a principal divisor.
Remark: The degree equation can also be interpreted mapping theoretically, and proved in this way
as well. For this, note that every hE A'(X) defines a branched covering h: X--+ P 1 (the case of h
identically constant is trivial). Certainly, if every point of h- 1 (0) and h- 1 (oo) is counted with its
branching multiplicity, then (h)= h- 1(0)- h- 1 (oo ). For every p E P 1, the sum of the multiplicities of
the points of h- 1 (p) is the sheet number s of the covering h: X --+ P 1. It follows that deg h = s - s = 0.
For linearly equivalent divisors D and D', l(D) = l(D') and i(D) = i(D'). We further
note that the dimension l(D) > 0 if and only if there is a positive divisor D' which is
linearly equivalent to D. In particular l(D) = 0 for every D with deg D < 0.
Proof: The first statement is clear, since the divisors which are linearly equiv-
alent to Dare of the form D +(h) for hE .A*(X). The second statement follows
from the first, since linearly equivalent divisors have the same degree. 0
For the zero divisor @(X)~ C and thus 1(0) = 1. The natural number
is called the genus of X. From this definition it only follows that g is a complex
analytic invariant of X. However in Paragraph 7.1 we will show that g is in fact the
topological genus of X (i.e. H 1 (X, C)~ 1[; 2 9).
For every divisor D it follows that
and in particular
Xo(@) = 1 - g.
212 Chapter VII. Compact Riemann Surfaces
l(D) ~ deg D + 1 - g
This inequality yields the first classical existence theorems. For example, since
l!7(D)(X) contains a non-constant meromorphic function whenever l(D) ~ 2, we
have the following:
For every divisor D with deg D ~ g + 1 there exists a non-constant meromorphic
function h with (h) + D ~ 0.
In particular, given p EX, there always exist non-constant functions which
have poles of order at most g + 1 at p, and which are holomorphic on X\p. One
can state this as a theorem about coverings of 1? 1 : Every compact Riemann surface
X of genus g is realizable as a branched cover with at most g + 1 sheets of the
Riemann sphere 1? 1 . In particular, if g = 0 then X= 1? 1 .
Since l(D) = 0 when deg D < 0, it follows from Theorem 1 that, for every
D E Div X with deg D < 0,
i(D) = g- 1- deg D.
One sees that, with the exception of the case g = 0 and deg D = - 1,
H 1 (X, l!7(D)) ::/= (0) for all divisors of negative degree. Furthermore,
rk ff = rk ff' + rk ff/ff',
is an immediate consequence of the definition. Thus every locally free sheaf !l' of
rank 1 contains only 0 and !l' as locally free subsheaves.
The requirement 1) in the above definition is quite restrictive. For example a
germ tx E ff x always generates a free submodule tx ((} x in ff x• but the quotient
module ff x ftx ((} x is in general not free. For an explicit example, take ff = ((}and tx
a non-unit. On the other hand if tx is a unit then there is no problem: If tx E ff x
and o(tx) = 0 then ff x ftx ((} x is a free ((} x-module.
Proof: Let ffx =((}~and tx = (t 1 , ... , t,), t; E ((}x· Sinceo(tx) = O,some t;,say t 1 ,
is a unit. Let e == t1 1 and define a: ((}~--+ ((}~-l by
Remark: The sheaf !l' = m(D )s is the only locally free subsheaf of ff of rank 1
with s E !l'"' (X). To see this, let !l' be such a sheaf. Then !l' x = (9 x vx for some
Vx E ff X' where Sx = mx Vx and mx E Ax· Let hx := m; 1 . Then Vx = hx Sx and, since
o(vx) ~ 0, hx E m(D)x· Consequently Vx E m(D)xsx or !}X c !l'x. The mx-module
ff xI!l' x contains therefore a submodule which is isomorphic to !l' xI!l' x· Since
ff xI!l' x is free, and since !l'I!l' x is in any case finite, !l' xI!l' x = 0. Hence !l' x = !l' x
for all x and !l' = !l'. D
2. The Existence of Locally Free Subsheaves. The foundation for the study of
the structure of locally free sheaves is the following theorem.
Theorem 3 (Subsheaf theorem). Every locally free sheaf ff =I= 0 contains a locally
free subsheaf which is isomorphic to m(D) for some DE Div X. One can choose
D = (s), where s E ff"'(X)*.
Proof: From Paragraph 2.2 it follows that ff has a non-trivial global merom or-
phic section. Thus the theorem is an immediate consequence of Theorem 2. D
Theorem 4 (Structure theorem for locally free sheaves of rank 1). Every locally
free sheaf !l' of rank 1 is isomorphic to a sheafm(D) with DE Div X. Furthermore
one can choose D = (s) with s E !l'"'(X)*.
This theorem says that in the cohomology sequence which is associated to the
short exact sequence of sheaves 0 --+ (9* --+A* --+ ~ --+ 0,
the homomorphism {> is surjective (see Paragraph 1.3 ). Thus one has a natural
group isomorphism,
Div XIP(X) ~ H 1 (X, (9*),
Theorem 5. There is a unique divisor class on X so that for every divisor K in this
class, Q 1 ~ m(K).
Supplement to Section 4: The Riemann-Roch Theorem for Locally Free Sheaves 215
One calls K a canonical divisor and its class the canonical divisor class on X.
The significance of canonical divisors appears in Section 6.
If X= lfl> 1 then every divisor -2x 0 , x 0 EX, is canonical. This follows from the
fact that, if z is a coordinate on X\x 0 , then dz is a differential form on X which is
holomorphic and nowhere vanishing on X\x 0 and has a pole of order 2 at x 0 . Of
course one must use the fact that on lfl> 1 the degree of a divisor determines its class.
In the case of ellipti<,: curves the zero divisor is canonical.
1. The Chern Function. We denote with LF(X) the set of analytic isomorphism
classes of locally free sheaves over X. A function c: LF(X)--+ 7L is called a Chern
function if
1) ForDE Div X, c(ll?(D)) = deg D.
2) For every exact sequence 0--+ F'--+ F--+ F"--+ 0 oflocally free sheaves, one
has c(F) = c(F') + c(F").
The following is straight forward:
is a Chern function.
where !l' and t'§ have rank 1 and r - 1 respectively. The induction hypothesis and
the additivity imply that y(ff) = c(ff).
2. Properties of the Chern Function. The Chern function behaves nicely with
respect to tensor products:
Proof: Let !l' = l!7(D). Then ff ® !l' = ff(D) and c(!l') = deg D. Thus
When !l' 1 and !l' 2 are locally free sheaves of rank 1, it follows from 4) that
c(!l' 1 ® !l' 2) = c(!l' t} + c(!l' 2 ). Thus the map H 1 (X, l!7*)--+ ?L, defined by
!l' --+ c(!l'), is a group homomorphism from the analytic isomorphism classes of
locally free sheaves of rank 1 to ?L. The reader can check that the map y in the
exact cohomology sequence
is the Chern function, provided H 2 (X, 7L) is identified with 7L in the natural way.
D
We remark without proof that, if ff is locally free of rank r greater than 1,
r
c(ff) = c(det ff), where det ff := 1\ ff is the locally free determinant sheaf of
1
rank 1. Thus, letting F be the vector bundle associated to ff, c(ff) is just the first
Chern class ofF, c 1(F) E H 2 (X, ?L) = 7L.
Proof: On the left hand side we have x(ff(D)) which, by the characteristic
theorem, is the same as r deg D + x(ff). If one writes r · x(l!7) + c(ff) for x(ff),
then, since x(l!7) = 1 - g, the claim follows immediately. D
§ 5. The Equation H 1 (X, .A')= 0 217
has .finite support ((J,~qJ1 )x = 0 for every x ¢ ID I u {p}, wherefx is a unit in l!i x)·
Thus H 1 (X, Jm qJ1 ) = 0 and
every function f =I= 0 in l!i(np)(X) induces a C-vector space epimorphism
is C-linear.
Proof: The claim becomes clear when one looks at how qJ1 is defined via the
Cech complex. Let U = {Uj}, i E /,be a cover of X with cochain groups
and
Associated to every l!i-homomorphism qJ1 ,fe l!i(np), are the following homomor-
phisms at the section level:
218 Chapter VII. Compact Riemann Surfaces
which induces the homomorphism 1/11 : H 1 (X, l!J(D))-+ H 1 (X, l!J(D + np)). From
the definition of tp J{i 0 , i 1 ) in (*) it is clear that
3. The Equation H 1 (X, .,H)= 0. We can now prove the following fundamental
theorem:
is likewise a representative of~. Since Vj ~ U 1for all i, each function ~ 1011 has finitely
many zeros and poles on Vj011 • Thus one can find a divisor De Div X so that
'' e Z 1 (m, l!I(D)). Let p eX\ ID I· Then, since &(D) c &(D + np) for all n ~ 0, it
follows that ~' c Z 1 (m, l!I(D + np)) for all such n. Since &(D + np) c vH and
H 1 (X, &(D + np)) = 0 for n ~ n0 , it follows that~, is cohomologous to zero as an
element of Z 1 (m, vH). Thus ~ = 0, and consequently H 1 (X, vH) = 0. D
§ 6. The Duality Theorem of Serre 219
The most important applications of the equation H 1 (X, .H)= 0 are found in the next section. In
closing this section we note in passing a rather simple consequence. Let 0"' denote the sheaf of germs
of merom orphic 1-forms. The differential d: M -+ 0"', which is defined in local coordinates by
dhx•=dhx/dz dz, is C-linear. The C-sheaf d.H c 0"' (it is not an (I)-sheaf!) consists of all residue free
germs of meromorphic 1-forms (i.e. germs of abelian differentials of the second kind; for the idea of a
residue see Paragraph 6.5). Since H 1 (X, .H)= 0, the short exact C-sequence 0-+ C-+ .H-+ d.H-+ 0
induces the following exact cohomology sequence:
In words,
the cohomology group H 1 (X, C) is isomorphic to the quotient space of global abelian differentials of the
second kind modulo differentials of global meromorphic functions.
It follows immediately that Jt(D), like ::0, is soft. In fact, every section over an
open set U has discrete support in U. Since every stalk Jt(D)x is the quotient
module A x/@(D)x, this implies that the C-vector space Jt(D)(X) of global distribu-
tions of principal parts is canonically isomorphic to the direct sum ffi Ax j(!)(D )x:
xeX
2. The Equation H 1 (X, (I)(D)) = J(D). Let R be the set of all maps F = (fx)
which assign to every x E X a germfx E .Ax so that almost allfx are holomorphic.
Clearly R is a C-vector space. Further, given D E Div X,
is a subspace of R. Every element of the direct sum Et> .Ax (resp. Et> (I)(D)x) is a
xeX xeX
family Ux} x EX withfx E Ax (resp.fx E (I)(D)x), where almost allfx vanish. Thus
we have the natural C-linear injection Et> .Ax-+ R which maps Et> (I)(D)x into
xeX xeX
R(D), and which induces a C-isomorphism
Proof: Associated to the short exact sequence (1) we have the exact cohomo-
logy sequence
where Im e ~ .A(X)/(I)(D)(X). The claim now follows from (3) and (4) above.
0
The reader should note that the spaces R, R(D) and .A(X) have very large
infinite dimensions, but that the finiteness theorem implies that J(D) is finite
dimensional.
§ 6. The Duality Theorem of Serre 221
3. Linear Forms. For maps F = Ux) and G = (g,J E R, we define the product
FG == Ux gx). Equipped with this product, R is of course a ~::-algebra, but it is more
importantly also an algebra over the field vH(X) cR. Let tX: R-+ C be a C-linear
form and h E vH(X). Then one defines the C-linear form htX: R-+ C by htX(F) =
tX(hF). Thus Homc(R, C) becomes an vH(X)-vector space. We explicitly state two
simple, but important, properties:
a) If vH(X) c ker tX then vH(X) c ker htX.
b) If R(D) c ker tX then R(D +(h)) c ker htX.
Proof: The statement a) is trivial. IfF= Ux) E R(D + (h)) thenfx E (I)(D + (h ))x
and therefore u(hxfx) ~ -ux(D). Hence hF E R(D). This proves b). 0
We denote by J(D) the dual space of J(D). It follows from a) and b) that every
meromorphic function hE vH(X)* determines a natural C-linear mapping,
J(D)-+J(D +h),
Proof: Every A. E J(D) is a C-linear form A.: R/(R(D) + vH(X))-+ C and is there-
fore liftable to a C-linear form tX: R-+ C such that tX vanishes on R(D) + vH(X).
From a) and b) it follows that htX: R-+ C vanishes on R(D +(h))+ vH(X). Clearly
htX induces a C-linear form hA. E J(D + (h)), which is uniquely determined by A. and
h. It is obvious that the map A.-+ hA. is C-linear. 0
J==U J(D),
D
the union of all J(D)'s for DE Div X. For two divisors D 1 ::; D 2 , we have
R(D 1 )::; R(D 2 ). Thus J(D 1 ) ~ J(D 2 ) when D 1 ::; D2 . This immediately implies
that every finite subset of J is contained in some J(D).
The set J is thus a C-vector space which is filtered by the subspaces J(D),
DE Div X. The above defined map, J(D)-+ J(D +(h)), gives us a mapping
vH(X) x J-+ J. Since Homc(R, C) is an vH(X)-vector space, the following remark
is obvious.
Theorem 2. The set J is, with respect to the operation vH(X) x J-+ J, a vector
space over vH(X).
4. The Inequality DimAt(Xl J ::; 1. The critical point of the proof of the duality
theorem is the following surprising dimension estimate. It is obtained by taking a
limit of the preliminary form of the Riemann-Roch formula.
Proof: Let A., Jl e J. We choose De Div X with A., Jl e J(D). Let p eX be fixed.
For every f e lP(np)(X) it follows that, since D- np ~ D +(!),fl. e J(D + (!)) c
J(D- np). Similarly gp. e J(D- np) for all g e lP(np)(X). The C-linear map
Jt(X) $ Jt(X)--+ J, (J, g)~-+fl.+ gp.,
for all n e 7L. If A. and J1. were linearly independent then both maps (o) and(~) would
be injective. This would mean that
for all n e 7L. From the Riemann inequality (Paragraph 3.2) it follows that
dime J(D- np) =dime I(D- np) = i(D- np) = g- 1 - deg(D- np).
Remark: There are divisors D such that H 1 (X, lV(D)) =I= (0). In other words,
J(D)' =1= (0). Thus J is a 1-dimensional Jt(X)-vector space.
<w, F) •= L Resx(fxwx) E C
XEX
and
n 1 1 1
L Resxwx= L ~ J w= -~ J w= - - J dw=O,
21ti X\uH,
XEX v=l 1tl oH, 1tl o(X\uH,)
defined by F~-+(ro, F). The C-vector spaces 0 00 (X) and Homc(R, C) are ..R(X)-
vector spaces as well. In fact, the mapping 0 00 (X)-+ Homc(R, C), defined by
ro 1-+ ro* is ..R (X)-linear.
From the residue theorem we see that ..R(X) c ker ro*. Moreover, if
roe Q(D)(X) then R( -D) c ker ro* (Theorem 4, 1)). Thus, if roe Q(D)(X~ ro*
induces a C-linear form,
The reader should note that, if roe Q(D) n Q(D')(X), the elements E>D(ro) and
E)D.(ro) agree in J.
The following lemma is the preparatory step in the duality theorem:
Theorem 7. The maps 9: Q 00 (X)-+ J and 9D: Q(D)(X)-+ J( -D) are bijective.
In this form the duality theorem is a classical theorem in the subject of alge-
braic curves. The more general form (for complex manifolds of higher dimension
which are not necessarily compact) was first formulated and proved by J.-P. Serre
in 1954 (Un theoreme de dualite, Comm. Math. Helv. 29, 9-26 (1955)).
1. The Equation i(D) = l(K- D). Since every finite dimensional vector space
has the same dimension as its dual space, the duality theorem implies the
following:
For every positive D, H 0 (X, Q( -D)) c H 0 (X, n). Thus using Theorem 3.1,
we can estimate i(D) and l(D) from above: If the divisor class of D contains a
positive divisor then i(D) ~ g and l(D) ~ deg D + 1.
226 Chapter VII. Compact Riemann Surfaces
Since Q ~ ll'!(K) (see Theorem 4.5), it follows that 0( -D)~ ll'!(K- D) for all
De Div X. Thus the above dimension equation can be written in the form
2. The Formula of Riemann-Roch. It follows from Theorem 3.1 that, for all
D e Div X, l(D)- i(D) = deg D + 1 -g. If one writes l(K - D) instead of i(D)
then one obtains the formula of Riemann-Roch.
degK=degw=2g-2.
§ 7. The Riemann-Roch Theorem (Final Version) 227
This equation contains for example the fact that H 1 (1P' 1 , C)= (0). One sees this by
noting that the differential dz, where z is an inhomogeneous coordinate, has degree
- 2. Thus, since 2g - 2 = - 2, IP' 1 has genus 0.
From the above it follows that every non-trivial w e O"'(X) has degree - x(X), where x(X) is the
topological Euler-Poincare characteristic of X. Consequently, if IX: X-+ X' is an s-sheeted ramified
covering map between compact Riemann surfaces X and X', and We Div X is the ramification divisor
of IX, then
x(X) + deg w = s · x(X')
Proof: Let w' e O"'(X'), and define w = IX*(w'). A direct calculation shows that deg(w) =
s · deg(w') + deg W. Since deg(w) = -x(X) and deg(w') = -x(X), the claim follows immediately.
0
In particular this shows that deg W is always even, and, when X' = IP' 1 with x(X') = 2, it follows
that
deg W = 2(s + g - 1)
4. Theorem A for Sheaves lD(D ). Let 2 be a locally free sheaf of rank 1 over X,
and let x eX. Then the following are equivalent:
i) The module of sections 2(X) generates the stalk 2" as an (!)"-module.
ii) There is a section s e 2(X) with o(sx) = 0.
iii) There is a sections e 9'(X) with 9'" = lD"s".
iv) dime H 1 (X, 9'( -x)) ~dime H 1 (X, 9').
Proof: i)::;. ii): Lett" e 9'" be such that 9'" = (!)" · t". By i) there exist sections
m
s,. e 9'(X) and germs f"" e (!)"' 1 ~ Jl ~ m, so that t" = "f.f,.xs,.". Thus some
1
s,."(x) e C must differ from 0.
Hence for that section lD(s"") = lD.
ii)::;. iii): Every germ t" e 9'" with o(tx) = 0 generates 9'" as an (!)"-module.
228 Chapter VII. Compact Riemann Surfaces
iii)=>i): Trivial.
ii)~iv): The sheaf ff which is defined by 0-+ !l'( -x)-+ !l'-+ ff-+ 0 has sup-
port only at x where its stalk is C. Thus we have the following
cohomology sequence:
Theorem 4 (Theorem A). Let D E Div X with deg D ;e:: 2g. Then for every x < X
there exists a sections E &(D)(X) with &(D)x = (Dxsx.
Proof: Let !l' :=&(D). If H 1(X, !l'( -x)) = 0, then the above equivalences guar-
antee the existence of such a section. Now !l'(- x) = &(D - x ). But by assump-
tion, deg(D- x) ;e:: 2g- 1. Hence Theorem B implies that H 1(X, &(D- x)) = 0.
0
Theorem 4 is the optimal form a Theorem A in the sense that for deg D < 2g
the sheaf &(D) may not be generated by its global sections. For example, let
D = K + {p). Then &(D)= !l(p). A section in &(D)(X) is just a differentialform w
which is holomorphic on X\P, and, if it would generate &(D)P, would have a pole
of order 1 at p. This would contradict the residue theorem.
There are divisors with deg D < 2g, and for which Theorem A, however, is
valid, namely K: If g =I= 0 then Theorem A holds for the sheafQ ~ &(K). In other
words, given x E X there exists a holomorphic differentia/form w E !l(X) which does
not vanish at x.
Proof: Since !l(D) ~ &(K +D) and deg(K +D) ;e:: 2g, the result follows from
Theorem 4. 0
In particular if D := mp, m ;e:: 2, then we have the following:
Let p E X and m ;e:: 2 be given. Then there exists a meromorphic differential form
§ 7. The Riemann-Roch Theorem (Final Version) 229
Proof: Define D := p 1 + p2 • Then there exists w E Q(D )(X) which generates the
stalk Q(D)p 1 • This form must be holomorphic on X\{Pt. p 2 }, have a pole of order 1
at p 1 and have a pole of at most order 1 at p2 • But the residue theorem requires
that it has a pole of exactly order 1 at p 2 . D
the space ff(X) = §"P = @((v + l)p)p/@(vp)P is !-dimensional. This implies that
1. :-::;; 1.+ 1 :-::;; 1. + 1. By definition w is a gap value if and only if every hE @(wp)(X)
is already in @((w- l)p)(X). That is, @(wp)(X) = @((w- l)p)(X) or equivalently
lw=lw-1• D
If v ~ 2g - 1 it follows from Theorem 3 that 1. = v + 1 - g. Thus for v ~ 2g,
1. = 1. _ 1 + 1 and there are no gap values greater than 2g - 1. One can improve
this remark:
7. Theorems A and B for Locally Free Sheaves. We call a locally free sheaf
"Stein" if Theorems A and B are valid:
A) H 0 (X, ff) generates every stalk ff x as an (!) x-module.
B) H 1 (X, ff) = (0).
230 Chapter VII. Compact Riemann Surfaces
Proof: Since H 1 (X, !l') = H 1 (X, <'§) = (0), it follows immediately from the
exact cohomology sequence that H 1 (X, !F)= (0). Further we consider the com-
mutative diagram
where the rows are exact and the maps in the columns associate to a section its
germ at x. By assumption, the images of Ax and 'l'x generate (as lllx-modules) !l'x
and <§x respectively. It follows therefore that the image of H 0 (X, !F) under cpx
generates !F X as an llJ X-module. D
Theorem 7. Given a locally free sheaf!F over a Riemann surface X, there exists a
natural number n+ so that,for every DE Div X with deg D ~ n+, the sheaf!F(D) is
Stein.
Proof: (by induction on the rank r of !F). Since every locally free sheaf of
rank 1 is isomorphic to a sheaf of the type lP(D), the case of r = 1 handled by
Theorem 3 and 4. Now we assume that r > 1 and that the statement holds for
sheaves of rank at most r - 1. By the "subsheaf theorem" (Theorem 4.3 ~ there
exist locally free sheaves !l' and <§ of rank 1 and r - 1 so that the following
sequence is exact for every D E Div X:
By the induction assumption and Lemma 6 above, there exists n+ E 7L. so that,
for deg D ~ n+, !F(D) is Stein. D
The following is analogous to the fact that if deg D < 0, then
H0 (X, C9(D)) = (0).
Proof: (by induction on the rank r of !F). If the rank of !F is 1 then !F = lli(D')
for some D' E Div X and n- •= - deg D' has the desired property. If rank !F > 1
then, as in the proof of Theorem 7, we choose locally free sheaves !l' and <§ of rank
§ 7. The Riemann-Roch Theorem (Final Version) 231
1 and r- 1 respectively so that, for every De Div X, we have the exact sequence
The claim follows immediately from the induction hypothesis and the exact
cohomology sequence. 0
Since H 0 (X, C)= H 0 (X, l!J) = C and H 2 (X, l!J) = (0), the associated cohomology
sequence yields
But H 2 (X, C)= H 1 (X, 0) =C. Hence p is bijective and we therefore have the
exact sequence
The map oc can be explicitely described in the following way: Let ro e O(X) and
U = {U;} a covering of X by contractable neighborhoods. Then there exists
J; e l!J(U;) so that dJ; = ro luc
The function ocii == jj- J; is therefore constant on Uii (we take these intersec-
tions to be connected). The family {oc;i(w)} forms a 1-cocycle in Z 1 (U, C). This
cocycle represents the cohomology class oc(ro) e H 1 (X, C).
Since l!;i c C, H 1 (X, i~;i) is an l!;i-vector subspace of H 1 (X, C). We now show
Proof: Suppose that oc(ro) e H 1 (X, i~;i). From the above description of oc it fol-
lows that there is a covering {U;} of X and functionsJ; e l!J(U;) such that dJ; = ro lui
and every function jj - J; is constant and real on Uii· Let g; == exp(2n.J=Tt;).
Then lg;j 2 = l9il 2 on U;i and therefore {lg;j 2} determines a real valued contin-
uous function, g, on X. By the maximum principle g is identically constant and
thus ro = 0. 0
There are of course many C-vector spaces Vin H 1 (X, C) so that V ffi Im oc =
H 1 (X, C). However there is one particular V that is quite natural. For this we
consider the "conjugate" resolution of C,
- d=B -
o - c ---+l!J- n-o.
232 Chapter VII. Compact Riemann Surfaces
Here fi •= 0 1 (see 11.2.3, in particular the diagram). For reasons similar to those
above, the associated cohomology sequence is
Proof: Since H 1 (X, C)~ C 29 and Im ex~ Im &~ C9 , it is enough to show that
lm ex n Im &= (0). The conjugation C -+ C, defined by c1--+ c, determines an ~
linear involution a: H 1 (X, C)-+ H 1 (X, C) which has H 1 (X, ~) as a fixed space.
Obviously a leaves every element of Im ex n lm & fixed. Thus Im ex n Im &c:
Im ex n H 1 (X, ~). Hence, by Lemma 9,
Im ex n Im &= (0). 0
1. The Number JL(§"). Let !l' be a locally free sheaf of rank 1 over a compact
Riemann surface X. Then, for every s E !l''"'(X)*, deg s only depends on !l' (see
Paragraph 4.2 ). For locally free sheaves of rank r > 1 this is not in general true.
For example, consider§"= (!')(n 1 p) E9 · · · E9 (!')(n,p) for some p EX. Then§" has
sections of degree n~> n2, ... , n,.
In order to define an integer which can be used in place of the "degree", we
make the following observations:
Every homomorphism n: §"-+ f§ between locally free sheaves induces a homomor-
phism n: §"'"'(X)-+ ~'"'(X). If
§ 8. The Splitting of Locally Free Sheaves 233
is an exact sequence of non-zero locally free sheaves, then, for every section
s E .?" 00 (X)*, either
a) n(s) = 0, in which cases= i(s') with s' E .?' (X)* and deg s' = deg s,
00
or
b) n(s) =I= 0, in which case deg(s):::;; deg(n(s)).
Proof: If .?~ is complementary to i(.?~) in .?"" then for every germ tx E .?':
there exist uniquely determined germs t~ E .?~00 , t~ E .?~ 00 so that
If n(s) = 0, then s E i(.?'(X)) (i.e. s = i(s') with s' E .?' 00 (X)*). But o(sx) = o(s~) for
all x EX. Hence deg s = deg s'.
Now suppose that n(s) =I= 0, and let sx = i(t~) + t~. Since nx maps.?~ isomor-
phically onto '§x, it follows that o(n(s)x) = o(t~). Hence o(sx):::;; o(n(s)x) for all
x EX, and consequently deg(s):::;; deg(n(s)). D
For every locally free sheaf.? over X, we define J.l(.?) as follows:
Proof: (by induction on the rank r of.?). For r = 1 the statement is trivially
true. There exists an exact sequence of locally free sheaves 0-+ 2 -+ .? -+ '§ -+ 0,
where 2 and'§ have rank 1 and r - 1 respectively. The proof follows immediately
by the induction hypothesis and the above estimates. D
Lemma. Let ff be locally free of rank 2 and .!l' a maximal subsheaf which is
isomorphic to@. Then, defining~:= ff/.!1', c(~) ~ 2g.
Proof: Since H 1 (X, r2) ~ U, the cohomology sequence associated to 0 .__. (2 .__.
0 starts out like
ff .__. ~ .__.
Now the sheaf~ is locally free of rank 1. Thus if c(~) > 2g, then, by the Riemann
inequality,
We now show
Thus fF' is a locally free sheaf of rank 2. Clearly!£ is a maximal subsheaf of fF',
and therefore
Consequently
4. The Splitting Criterion. One says that the short exact (!)-sequence,
i
0 ---+ f/ 1 -----+ f/ -----+ f/ 2 -----+ 0
e:
splits if there exists an (!)-homomorphism f/ 2 - f/ with 1t e
= id. Thus, in this
0
case, f/ = i(f/ t) ffi e(f/ 2);;;;;; f/ 1 ffi f/ 2· There is a simple formal splitting criterion:
An exact (!)-sequence of locally free sheaves 0 --. f/ 1 --. f/ --. t'§ --. 0 splits when
Proof: Since t'§ is locally free, the above short exact sequence induces the exact
(!)-sequence of sheaves,
If 1t* is surjective, then there exists a global section e E Hom(~, 9') with 1t*(e) =
e
1t o = id. o
Remark: The splitting criterion is valid for any complex space. In the case of Stein spaces the
relevant cohomology group is always zero. Thus we have the following observation.
Every locally free subsheqf of a locally free sheaf over a Stein space is a direct summand.
For !-dimensional Stein manifolds (i.e. non-compact Riemann surfaces~ it follows that every
locally free sheaf of rank r is isomorphic to the sheaf (!J•. (Recall that H 1 (X, (!J) = H 2 (X, Z) = 0. Thus
H 1 (X, tr*) = 0, and thus every locally free sheaf of rank 1 is free.)
Thus
r
H 1 (X, Jf Mn(~, 2)) ~ E9 H 1 (X, Jf om(2i, 2)) = (0)
i=2
@(n) == @(np)
for every n E 7!... Then n is the Chern number of @(n) and @(n) ;;; @(m) if and only if
n = m. Furthermore, if ff' is a locally free sheaf of rank 1 over 1P 1 then ff' ;;; @(n ),
where n = c(ff').
Proof: Certainly ff' ;;; @(D) for some D E Div 1P 1 . On IP 1, however, two divisors
are linearly equivalent if and only if they have the same degree. Hence @(D) is
isomorphic to @(deg D) and, since deg D = c(ff'), we have the desired result.
0
Every sheaf @(nt) EB @(n 2 ) EB · · · EB @(n,), nt> ... , n, E 7!.., is locally free of rank r.
The splitting theorem of Grothendieck says that one obtains all locally free
sheaves over IP 1 in this way:
6. Existence of the Splitting. Using the following lemma, both the existence and
uniqueness are proved by successively "splitting off" maximal subsheaves.
Splitting Lemma 6. Let !F be a non-zero locally free sheafover 1P t> and let ff' be a
locally free subsheaf of rank 1 with ~.-\ff');;::: Jl(ff/ff')- 1. Then ff' is a direct sum-
mand of !F.
In particular every maximal subsheaf of !F is a direct summand.
238 Chapter VII. Compact Riemann Surfaces
In particular, c(!l')- c(!l'1) ~ -1, 2 ;:5; i ;:5; r. By Theorem 4 it follows that !l' is a
direct sum of F.
If !l' is a maximal subsheaf ofF, then from Theorem 3 it follows that c(!l') =
p,(F) ~ p,(F/!l'). (The genus is zero!). D
7. Uniqueness of the Splitting. For every locally free sheaf F on lfJ 1 we let
m •= p,(F). Then
F( -m)(lfJt) = {s e F 00 (1F» 1 )Ideg(s) = m} u {0} +0,
so
Uniqueness Lemma 7. Let F = !l' 1 ffi · · · ffi !l' r = !l'J. E9 · · · E9 !l'~ be two split-
tings ofF with !l'1 ~ l!l(n1), !l'J. ~ l!l(n1), 1 ;:5; i ;:5; r. Assume that n 1 ~ n2 ~ • • • ~ nr,
and n't ~ n2 ~ ··· ~ n~. Then, with d•=dimc F(-m)(lfJ 1 ) ~ 1,
1) !l' 1 $ · · · ffi !l'd = !l'J. $ · · · ffi !l'4 = .#"; !l'; ~ l!l(m) ~ !l'i, for i = 1, ... , d,
and
2) n1 = ··· = nd = n1 = ··· = n:, = m; n1 = nifor i = d + 1, ... , r.
Proof: Since F=!l' 1 $···E9!l'r, it follows that F(-m)(IP 1 )=
r
EB !l'1( - m)(IP 1 ~ where !l'1( - m)~ l!l(l1) with 11 •= n1 - m ;:5; 0 (by the definition
i=1
ofm). Now
l!l(l)(!Pt) = 0 for l < 0, and l!l(IP 1 ) =C.
Thus there must be d equations n1 = m. Since the n;'s are monotonically decreas-
ing, n 1 = · · · = nd = m > nd+ 1 and !l' 1( - m) ~ 0, 1 ;:5; i ;:5; d.
For every i ;:5; d, !l'1( -m)(IP 1 ) generates the sheaf !l'1( -m). Thus
§ 8. The Splitting of Locally Free Sheaves 239
d d
E9
i= 1
..2"1( -m)(IPt} generates the sheaf E9 ..2"1( -m). Since §"( -m)(IPt) =
i= 1
d d d
E9 ..2"1( -m)(IPt), it follows that :T = E9 ..2"1( -m). Hence # = :T(m) = E9 ..2"1•
i=1 i=1 i=1
The above can obviously be repeated for the splitting §" = ..2'~ EB · · · EB ..2"~.
Thus 1) has been proved as well as n1 = · · · = n;, = m. It now follows that
Since §"j# has rank r- d < r, and since n1, n; are monotonically decreasing, it
follows by induction that n; = ni for i = d + 1, ... , r. D
Corollary. Let§"= C§ EB Yl', where C§, Yl' are non-zero locally free sheaves over
IP 1 . Then
Proof: Since§"~ @(nt) EB · · · EB @(n,), iffollows that J.L(§") = max{n 1, ... , n,}.
D
We see now that every locally free subsheaf ..2' of rank 1 in §"with c(..2') ~
J.L(§"/..2') is necessarily maximal, because the Splitting Lemma implies that
§" = C§ EB 9', and the above corollary shows that J.L(§") = max(J.L(C§),
J.L(..2')) = c(..2').
The following remark is also quite easy to see.
Let § = C§ EB Yl', where C§, Yl' are non-zero locally free sheaves with J.L(§") >
J.L(C§). Let ..2' be a locally free subsheafof rank 1 in§ so that c(9') > J.L(C§). Then ..2' is
a locally free subsheaf of Yl'.
Proof: Let ..2' = (!)s with s E §" 00 (1Pd*, and deg(s) = c(..2'). Let n: ff--+ C§
denote the natural sheaf projection. Then by the rematlk in Paragraph 1, either
n(s) = 0 or deg(s)::;; deg(n(s)). The latter is not possible, because deg(n(s))::;; J.L(C§)
and c(..2') > J.L(C§). Thus s E (~ e-t n)(IP t) = Yl'(IP1 ). Hence ..2' = (!)s c Yl'. D
In particular we have the following:
A sheaf§";;:; m(nt) EB · · · EB m(n,) with n 1 > max{n 2 , ••. , n,} has only one locally
free subsheaf ..2' which is isomorphic to m(n 1 ), and there are no such with ..2' ~ m(n),
n 1 > n > max{n 2 , ... , n,}.
However, for every n::;; max{n 2 , •.. , n,}, there are in the above setting locally
free subsheaves ..2' ;;:; @(n) in§". For this, see Prop. 2.4 in "On holomorphic fields
of complex line elements with isolated singularities", Ann. Inst. Fourier 14,99-130
(1964), by A van de Van.