Lazarsfeld Notes
Lazarsfeld Notes
T.R. RAMADAS
and one can check that E is indeed coherent, locally free, and of rank r. On
the other hand, given a coherent sheaf E, one can ask if there is a variety E
that “represents its sections” as above. This is certainly true if E is locally
free.
We outline the construction in the case that is most important to us
– when the rank r = 1. In this case we will say that E is “invertible”
and E is a “line bundle”. We will signal our shift in emphasis by using the
notation L and L respectively. (By considering line bundles we avoid having
to deal with matrix-valued functions and having to keep track of order of
multiplication.) First a preliminary
Φ(x, z) = σ0 (x), x ∈ U
Suppose then that L is an invertible sheaf. Let us first deal with the case
when L is globally free. That is, there is an element σX ∈ L(X) such that
the maps
OX (U ) → L(U ), f 7→ f σX |U
LINE BUNDLES 3
are isomorphisms. Consider the trivial line bundle X×C. The corresponding
sheaf of sections is clearly OX , and the sheaf map
OX (U ) → L(U ), f 7→ f σX |U
is an isomorphism. If L is only locally free, there exists an open cover Uα of
X and σα ∈ L(Uα ) such that for each i the map
OUα → L|Uα f 7→ f σα
is an isomorphism of OUα -modules. On the overlaps Uα ∩ Uβ we have
σα = χα,β σβ
where χα,β is a function, regular and nowhere-vanishing on Uα ∩ Uβ . The
collection of functions χα,β satisfy
χα,β χβ,α = 1
χα,β χβ,γ χγ,α = 1
(In other words, w.r.to the cover {Uα } we have a 1-cocycle with values in
the sheaf OX ∗ .)
defines a presheaf. One has to sheafify this to get F1 ⊗OX F2 . This corrects
a statement I made in the lecture.
We can also define the dual Ě of a vector bundle; we have
Ě = HomOX (E, OX )
Under direct sums ranks add, and under tensor products they multiply. In
particular, the tensor product of two line bundles is a line bundle. The
natural map
L ⊗OX Ľ → OX
is an isomorphism, and it is easy to check that the set of isomorphism classes
of line bundles on X is an abelian group. This is why a line bundle is called
an “invertible sheaf”. From now on we will set L−1 = Ľ.
∼ ∼
where the disjoint union tL is over the group of isomorphism classes of line
bundles, the Picard group of X. (The Picard variety, which we shall mostly
avoid, is a subtler object.)
The “traditional” definition of Cartier divisor is essentially as isomor-
phism class of pairs (L, σ) with L a line bundle and σ a meromorphic section,
but reformulated to avoid talking of line bundles, as follows.
Consider the constant sheaf M∗ of nonzero meromorphic functions. This
is the sheaf that associates to any (nonempty) open U the abelian group of
nonzero meromorphic functions on X (and associates to the empty set the
final object in the category of abelian groups, namely the singleton group).
This contains as subsheaf the sheaf OX ∗ of units in the sheaf of algebras O .
X
I claim that there is a bijective homomorphism between global sections of
the quotient sheaf M∗ /OX ∗ and isomorphism class of pairs (L, σ) as above.
Suppose first that a pair (L, σ) is given. Cover X by open sets Uα over
which L is trivialised by meromorphic sections σα . (That is, σα is regular
and non-vanishing on Uα .) Then, for each α, there exists a meromorphic
function φα such that
σ = φα σ α
On overlaps Uα ∩ Uβ , the meromorphic functions φα and φβ differ by a
regular invertible function. In other words, the φα define a global section of
the quotient sheaf M∗ /OX ∗ . Changing the local trivialisations σ does not
α
change this global section, which therefore depends only on (the isomorphism
class of) the pair (L, σ).
Conversely, a global section of M∗ /OX ∗ is, by definition, given by an open
covered (1).
Notation/Definition: If D is a Cartier divisor, we mean by OX (D) the in-
vertible sheaf defined above. Less precisely, we will mean a line bundle L,
together with a meromorphic section σ such that (σ) = D. As noted earlier,
the pair (L, σ) is unique up to isomorphism. Even less precisely, we will
mean just the line bundle L. We will often confound L and L and OX (D).
We end by giving the canonical example of a Weil divisor that is not
Cartier. Consider the cone in C3 :
{(x, y, z)|x2 + y 2 = z 2 )}
This is an affine variety of dimension 2. Let D be the dimension 1 subvariety
defined by the ideal generated by x and z − y. One checks that D is reduced
and irreducible, but D is not defined by the vanishing of a single regular
function.
1http://www.math.umn.edu/ garrett/m/algebra/cech.pdf
LINE BUNDLES 9
P
1.3.
P Line bundles and divisors: some easy remarks. Let D = i ai Di −
a0 D 0 be a Cartier divisor, with a , a0 > 0.
j j j i j
(1) The line bundle O(D) is trivial iff D is a principal divisor.
(2) What are the regular sections of OX (D) on (all of) X? By definition,
these are the meromorphic functions, regular outside ∪i Di and with
poles of order at most ai along Di and vanishing to order at least
order a0j along Dj0 .
(3) If D is an effective
P divisor, that is, a positive linear combination of
divisors (D = i ai Di , with ai > 0), then H 0 (X, O(D)) 6= 0, for
then any constant function is a global section.
(4) On a complete (which in our context means projective) variety, if D
is effective and nonzero, then O(D) is a nontrivial line bundle.
(5) Suppose D is effective as above, and σ ∈ H 0 (X, O(D)) is any nonzero
P
(regular) section, then (σ) = D̃ = ĩ aĩ Dĩ with aĩ > 0 and
X X
aĩ Dĩ − ai Di
ĩ i
1.4. Line bundles on smooth projective curves. Let us see how the
theory plays out on a smooth projective curve X of genus g. Since X
is smooth, Weil divisors are locally principal, and we will drop the prefix
“Cartier”. A divisor is just a formal sum with integer coefficients:
X
D= ax x
x∈X
where the ax are integers and nonzero only for finitely many x. Given a
divisor D, we define its degree by
X
degree D = ax
x∈X
adopted the quotient definition (which requires the above fuss with identify-
ing quotents) for good reason, and (following Lazarsfeld) we will stick to it.
On Grk (V ) the vector spaces Q and S together build up tautological vector
bundles Q of rank k and S of rank (dim V −k) as quotient and (respectively)
sub of the trivial bundle with fibre V . In terms of the corresponding locally
free sheaves, we write:
0 → S → V ⊗C OGrk (V ) → Q → 0
Allowing for abuse of notation, the above sequence could be written:
0 → S → V ⊗C OGrk (V ) → Q → 0
V ⊗C OX −−−−→ QX −−−−→ 0
Set-theoretically the map ψ is described as follows. Let x ∈ X; the surjective
map of bundles V ⊗C OX → QX → 0 yields the surjective map of vector
spaces:
V → (QX )x → 0
which represents the point ψ(x) ∈ Grk (V ). We will prove below that ψ is a
morphism in the case that interests us, that of projective spaces.
ψ is clearly (the open set) Ũ` ⊂ X such that the induced map
` ⊗C OŨ` (,→ V ⊗C OŨ` ) → QX → 0
ψ|Ũ` = φ−1
` ◦ ψ̃`
which shows that ψ|Ũ` is a morphism. Since the sets U` cover P(V ) we are
done.
From now on, we will (unless otherwise flagged) consider projective vari-
eties.
If X is projective and L is a line bundle then H 0 (X, L) is a finite-
dimensional vector space. Any nonzero section defines an effective divisor,
and two sections define the same divisor iff one is a nonzero scalar multiple
of the other. This has the following consequence: given a Cartier divisor D
on a projective variety, the set
{D̃|D̃ is effective and linearly equivalent to D}
can naturally be identified with the projective space Psub (H 0 (X, OX (D))) of
one-dimensional subspaces of H 0 (X, OX (D)). (Of course, it could happen
that OX (D) has no nonzero regular sections, in which case |D| is empty. If
D itself is effective, it will represent a point of the projective space.)
By a linear series, we will mean (interchangeably) a nonzero subspace V ⊂
0
H (X, L) or the set |V | of effective divisors corresponding to such a subspace
V . Note that the set |V | can be identified with Psub (V ) ⊂ Psub (H 0 (X, L)).
If V = H 0 (X, L) we will say that the linear series is complete. If D is a
Cartier divisor, we denote by |D| the complete linear series associated to
OX (D).
Given a linear series V , its base locus Bs(|V |) is the (Zariski)-closed set
of points where all the nonzero sections in V vanish. If L = O(D) and
V is the corresponding complete linear series, we will also denote this by
Bs(|L|) = Bs(|D|). The notation signifies that the base locus comes with a
natural scheme structure, and one can in fact talk of the base scheme defined
by the sheaf of ideals which is the image of the sheaf map (“evaluate σ ∈ V
at x ∈ X to get σ(x) ∈ Lx , compose with the map Lx ⊗ L−1 x → C”):
V ⊗C L−1 → OX
We will not deal with the base scheme.
14 T.R. RAMADAS
We say that the linear series |V | is (base-point) free if its base locus is
empty. In other words, |V | is free if for every x ∈ X, there is a section
belonging to V which is nonvanishing at x. Given a line bundle L (or a
divisor D), we say that it is free (“generated by global sections”) if the
corresponding complete series is free.
Outside the base locus B of V we have the evaluation morphism e(“evaluate
σ ∈ V at x ∈ X to get σ(x) ∈ Lx ”):
V ⊗C OX → L → 0
e
This yields a morphism φ|V | : X \ B → P(V ) and a commutative diagram:
V ⊗C OX −−−−→ ψ ∗ O(1) −−−−→ 0
=y ∼y
V ⊗C OX −−−−→ L −−−−→ 0
Often we think of φ|V | as a rational map φ|V | : X 99K P(V ).
Let us make this explicit. Let dim V = n. Choose a basis of sections
σ1 , . . . , σn ∈ V , and let σ̌i ∈ V̌ be the dual basis. Let Ui ⊂ X be the open
set
Ui = {x ∈ X|σi (x) 6= 0}
On U1 we have regular functions x1 ≡ 1, x2 , . . . , xn such that
X X
e( ai σi ) = ( ai xi )σ1 (x)
i i
This yields:
H 1 (X, Oan ) ∗
{1} → → H 1 (X, Oan ) → H 2 (X, Z) → {1}
H 1 (X, Z)
∗ ) classifies isomorphism classes of holo-
The cohomology group H 1 (X, Oan
morphic line bundles. The map H 1 (X, Oan ∗ ) → H 2 (X, Z) clearly factors
as:
∗ ∗
H 1 (X, Oan ) → H 1 (X, Ocont ) → H 2 (X, Z)
which shows that topologically trivial line bundles are parametrised by (the
Jacobian)
H 1 (X, Oan )
H 1 (X, Z)
Finally, on an algebraic variety, algebraic line bundles are classified by
H 1 (X, O∗ ) where now O∗ is the Zariski sheaf of invertible regular functions.
It is not entirely obvious how to compare the cohomology groups H 1 (X, O∗ )
and H 1 (X, Oan∗ ), but GAGA assures that on a projective variety analytic and
with ai,ji ∈ Z and the Di,ji reduced, irreducible codimension one subvari-
eties. Then
X
(D1 · · · Dn X) = a1,j1 . . . an,jn #{D1,j1 ∩ · · · ∩ Dn,jn }
j1 ,...,jn
4. Asymptotic Riemann-Roch
We will give a proof of the Theorem, stopping short of pinning down the
dependence on the self-intersection of D. We follow the treatment of O.
Debarre’s “Higher-Dimensional Algebraic Geometry” where a more general
result is proved.
where the limit is taken over open sets containing x, partially ordered by
inclusion. If F is a coherent sheaf on a variety X,
Fx = Γ(U, F) ⊗C[U ] OU,x
for any affine U containing x, where OU,x = OX,x is the local ring of X at
x. The local ring OX,x is itself the stalk of OX .
Let I(x) ⊂ OX denote the ideal sheaf of x, the sheaf of regular functions
vanishing at x. (In the lectures I used the notation mX,x , but this is better
reserved for the maximal ideal in the local ring OX,x .) We have the defining
exact sequence
0 → I(x) → OX → C[x] → 0
where C[x] is the skyscraper sheaf at x:
H 0 (U, C[x] ) = C if x ∈ U and {0} otherwise
and the map OX → C[x] is the evaluation map at z. (Thus C[x] = OX /I(x).)
If F is a coherent sheaf, tensoring by it yields
I(x) ⊗OX F → F → C[x] ⊗OX F → 0
(In general left-exactness is not preserved.) Denoting by I(x)F the image
of the sheaf morphism I(x) ⊗OX F → F, we get the exact sequence
0 → I(x)F → F → C[x] ⊗OX F → 0
Here C[x] ⊗OX F is a skyscraper sheaf at x with “fibre” the finite-dimensional
vector space H 0 (X, C[x] ⊗OX F). (The term “fibre” is not standard except if
F is locally free in which case this is indeed the fibre at x of the correspond-
ing vector bundle5.) A sequence of sheaves is exact iff the corresponding
sequence of stalks is exact at every point, we have the exact sequence of
modules over the local ring OX,x
0 → (I(x)F)x → Fx → H 0 (C[x] ⊗OX F) → 0
Let mx denote the maximal ideal in OX,x ; then (I(x)F)x = mx Fx . We see
that Nakayama’s Lemma (“if M is a finitely generated module over a local
ring with maximal ideal m, then mM = M iff M = 0”) implies: if F is
coherent, then its fibre at x vanishes iff its stalk FX,x vanishes, that is, iff
F itself restricts to zero on some open neighbourhood of x.
As a consequence:
Lemma 5.2. Let F → G be a morphism of coherent sheaves on a variety
X, and suppose that for some x ∈ X the induced map (of finite-dimensional
vector spaces)
F ⊗OX C[x] → G ⊗OX C[x]
is onto. Then there is an open neighbourhood U of x on which the map of
sheaves is onto.
5If E is a vector bundle, and E the corresponding sheaf, we reserve the right to use
the notation Ex for the fibre and E x for the stalk. In such contexts, we will stick to the
notational distinction between E and E
24 T.R. RAMADAS
Proof. Let K denote the cokernel sheaf, so that we have the exact sequence:
F →G→K→0
Tensoring with C[x] we get the exact equence
F ⊗OX C[x] → G ⊗OX C[x] → K ⊗OX C[x] → 0
which shows that K ⊗OX C[x] = 0. The previous Lemma applies and we see
that K is zero on some neighbourhood of x.
We will need another consequence of Nakayama’s Lemma:
Lemma 5.3. Let f : (A, m) → (A0 , m0 ) be a local homomorphism of Noe-
therian local rings, with A0 finitely generated as an A-module and suppose
that f induces an isomorphism of residue fields A/m = A0 /m0 ≡ k. For
f to be surjective, it is (necessary and) sufficient that the induced map of
“Zariski cotangent spaces”
m/m2 → m0 /m02
is surjective.
Proof. We prove sufficiency. (My earlier proof had an error, as pointed out
by Akashdeep; this one is adapted from Joe Harris: Algebraic Geometry:
A First Course.) We will use the follwing corollary of Nakayama’s Lemma
repeatedly: if A is a Noetherian local ring with maximal ideal m, M a finite
A-module, and N a submodule such that N + mM = M , then N = M .
By replacing A by its image in A0 , we can assume that the f is an injection.
By assumption, the inclusion m → m0 induces a surjection m/m2 → m0 /m02 ,
so
mA0 + m02 = m0
Applying the above Corollary to the inclusion of A0 -modules mA0 ⊂ m0
yields mA0 = m. Consider now the inclusion of A-modules A ⊂ A0 . Since
mA0 + A = m0 + A = A0 , we see that A = A0 .
We can now turn to the proof of the Theorem.
Proof. (1) implies (2): Assume first that L is very ample. Then
X ,→ P(V ),
ι
such that ι∗ O(1)= L. (We will need to talk about the fibre of L, so for a
while we will keep the distinction between L and L.) Then H i (X, F ⊗OX
Lm ) = H i (P(V ), ι∗ (F ⊗OX Lm )) = H i (P(V ), ι∗ F ⊗OP (V ) O(m)). We have
used the extension by zero operation ι∗ ; this takes a coherent sheaf on a
closed subvariety to a coherent sheaf on the ambient variety. This is a
special case of a direct image by a morphism. We have also used a special
case of the projection formula: given a morphism π : X → Y ,
π∗ (F ⊗OX π ∗ V) = π∗ F ⊗OY V
LINE BUNDLES 25
provided V is locally free6. Now, given any coherent sheaf G on P(V ), there
exists a m(G) such that for i > 0,
H i (P(V ), G ⊗OX O(m)) = 0 provided m ≥ m(G)
Applying this to ι∗ F gives the desired result, with m1 (F) = m(ι∗ F). If L
is only ample, with LM very ample, let ι be the corresponding projective
embedding. We have, for 0 ≤ l ≤ M − 1 and i > 0,
H i (X, F ⊗OX Ll ⊗OX LmM ) = 0 provided m ≥ m(ι∗ (F ⊗OX Ll )),
With a little book-keeping, this gives the desired result.
(2) implies (3): For x ∈ X, let m1 (x) be such that for i > 0,
H i (X, I(x)F ⊗OX Lm ) = 0 provided m ≥ m1 (x)
As a consequence, for m ≥ m1 (x), the evaluation map
H 0 (F ⊗OX Lm ) ⊗C OX → F ⊗OX Lm
is onto at x and therefore in an open neighbourhood Ux . By (quasi-)compactness
we can cover X by finitely many of these open sets. Taking m1 to be the
supremum of the corresponding m1 (x)’s we are done.
(3) implies (4): For x ∈ X, consider the sheaf I(x)L; by (3) there exists
m(x) such that I(x)L ⊗OX Lm is globally generated for m ≥ m(x). By
quasi-compactness we can m̃3 such that for m ≥ m̃3 + 1, the evaluation map
(5) H 0 (I(x)Lm ) ⊗C OX → I(x)Lm
is onto for every x ∈ X. (Since L is invertible, I(x)L ⊗OX Ll = I(x)Ll+1 .)
Now I(x)Lm is the sheaf of sections of Lm vanishing at x. So, given distinct
x, x0 ∈ X, there is a section of Lm that is nonvanishing at x0 and vanishing
at x. In other words, the linear system of Lm is base-point free and the
corresponding map φ|Lm | : X → P(H 0 (X, Lm )) is injective. Since X is
compact, the image is closed.
All that is left is to make sure that every (local) regular function on X is
the pull back of a (local) regular function. Explicitly, let f ∈ OX,x ; this is
a regular function on U ∩ X, with U open in P(V ). We need to exhibit a f˜
regular on U , possibly at the cost of shrinking U . In other words, we have to
ensure that for every x ∈ X, the map of local rings OP(V ),x → OX,x is onto.
In turn, it suffices to show (Lemma 5.3) that mP(V ),x /m2P(V ),x → mX,x /m2X,x
is onto. Consider the map
H 0 (X, Lm ) ⊗OX L−m → O .
This restricts to
H 0 (X, I(x)Lm ) ⊗OX L−m → I(x) .
6As Pramath pointed out, if π is an affine morphism (as ι is) this holds even if V is
not locally free. Proof: given a ring homomorphism A → B, an A-module MA and a
B-module MB , we have an isomorphism of A-modules MB ⊗B B ⊗A MA → MB ⊗A MA .
26 T.R. RAMADAS
which is also onto, although we seem not to need this (see below). But we
have a commutative diagram:
H 0 (P(V ), IP(V ) (x)O(−1)) ⊗C O(−1)x −−−−→ mP(V ),x /m2P(V ),y
=y
y
onto
H 0 (X, I(x)Lm ) ⊗C L−m
x −−−−→ mX,x /m2X,x
which shows that the map mP(V ),x /m2P(V ),x → mX,x /m2X,x is onto, provided
we justify the equality H 0 (P(V ), IP(V ) (x)O(1)) = H 0 (X, I(x)Lm ). To see
this, consider the commutative diagram
0 −−−−→ H 0 (P(V ), IP(V ) (x)O(1)) −−−−→ H 0 (P(V ), O(1))
=
y y
0 −−−−→ H 0 (X, I(x)Lm ) −−−−→ H 0 (X, Lm )
Finally, (4) implies (1) by definition.
We will join Lazarsfeld’s losing battle and refer to the property of ample-
ness as “amplitude”.
5.2. Projective schemes. For the most part we want to stick to varieties,
but we will have to deal with more general schemes occasionally. Instead of
biting the bullet and defining amplitude on schemes, we will manage with
the following
Lemma 5.4. Let Y ⊂ P(V ) be a subscheme, F a coherent sheaf on Y ,
and L an invertible sheaf on Y such that L|Y1 is ample for every reduced,
irreducible subscheme Y1 ⊂ Y . Then for large enough m we have
H i (Y, F ⊗OX Lm ) = 0, i > 0 .
Proof. We use the fact that F has a filtration
F = FN ⊃ FN −1 ⊃ · · · ⊃ F0 = {0}
7Given finite-dimensional vector spaces V, W , the bilinear map (v̌, w̌) 7→ v̌ ⊗ w̌ induces
an embedding Psub (V̌ ) × Psub (W̌ ) → Psub (V̌ ⊗ W̌ ), and dually P(V ) × P(W ) → P(V ⊗ W ).
28 T.R. RAMADAS
We claim that
(1) ∅ =
6 Amp(X) ⊂ N ef (X)
(2) Amp(X) is open, and
(3) N ef (X) is closed.
In fact, we will see later that the nef cone is the closure of the ample cone
and the ample cone is the interior of the nef cone.
Let us pause to justify the above claims. Since X is projective, the pull-
back of the hyperplane bundle by any projective embedding is very ample,
so ample classes exist. An ample class restricted to any curve C is ample
and hence has positive degree. Thus
∅=
6 Amp(X)Z ⊂ N ef (X)Z
and (1) follows. That N ef (X) is closed is clear because it is the intersection
of closed “half-spaces”. It remains to show:
Proposition 7.2. Amp(X) is open.
Proof. Before going further, we fix a basis of generators {[Li ]|i = 1, . . . , ρ(X)}
for the abelian group N 1 (X). This defines an L∞ -norm on the vector space
N 1 (X)R which we will use to topologise it from now on.
We prove first: let [L] be an integral ample class. Then there exists (L) >
0 such that every rational class [L0 ] with ||[L0 ] − [L]|| < (L) is ample. In
fact, there exists m0 > 0 such that Lm Li is ample for all i and m ≥ m0 . Let
[L0 ] = [L] + i ai [Li ], with ai rational numbers. Then
P
1 X
[L0 ] = {[L] + ai ρ(X)[Li ]}
ρ(X)
i
is ample provided |ai ρ(X)| ≤ m10 , i.e., ||[L0 ] − [L]|| < (L) ≡ ρ(X)m
1
0
.
We prove next: let [L], [M ] be integral classes, with L ample. Suppose give
δ > 0 such that [L] + d[M ] is ample for every rational d such that |d| < δ.
Then [L] + dM ] is ample for every real d such that |d| < δ. Proof: Pick
rational numbers d1 , d2 such that −δ < d2 < d < d1 < δ. Then
d − d2 d1 − d
[L] + t[M ] = ([L] + d1 [M ]) + ([L] + d2 [M ])
d1 − d2 d1 − d2
which expresses [L] + d[M ] as a convex (real) linear combination of ample
rational classes.
8Note that {P b [L ]|b ≥ 0, P b > 0, L nef } ⊂ N ef (X), but apparently this can
j j j j j j j
be a strict inclusion.
LINE BUNDLES 33
P
Finally suppose [L] = j tj [Hj ] with tj real and positive and [Hj ] integral
and ample. Let 0 < t < t1 be rational. Then
X X X
[L] + ai [Li ] = t[H1 ] + ai [Li ] + (t1 − t)[H1 ] + tj [Hj ]
i i j>1
P
It is enough to show that t[H1 ]+ i ai [Li ] is ample for |ai | small enough. By
the preceding paragraphs, there exists 1 > 0 such that for each i, [H1 ]+bi [Li ]
for bi real and |bi | < 1 . Now note that
X t X ρ(X)
t[H1 ] + ai [Li ] = {[H1 ] + ai [Li ]}
ρ(X) t
i i
We will prove below that Amp(X) is the interior of N ef (X) and N ef (X)
the closure of Amp(X). This will be one of the corollaries of
Theorem 7.3. (Kleiman) Let L ne a nef line bundle on a projective variety
X. Then
c1 (L)dim V [V ] ≥ 0
for any reduced, irreducible subvariety V ⊂ X.
Proof. The proof will be by induction on n = dim X. The claim is true for
n = 1. By induction, it suffices to prove the above inequality for V = X.
We will suppose that the equality fails. i.e., that
c1 (L)n [X] < 0
and derive a contradiction.
Let M be a very ample line bundle, and consider, for t ∈ R,
P (t) = (c1 (L) + tc1 (M ))n [X]
Expanding the expression on the right we get
P (t) = c1 (L)n [X] + tc1 (L)n−1 c1 (M )[X] + · · · + tn c1 (M )n [X]
The coefficient of the tn−k term is c1 (L)k c1 (M )n−k [X]; provided k < n, we
can (by Bertini) compute this by taking n − k generic hyperplane sections
H1 , . . . , Hn−k such that their intersection is a reduced, irreducible variety of
dimension k and evaluating c1 (L)k on this intersection:
c1 (L)k c1 (M )n−k [X] = c1 (L)k [H1 ∩ · · · ∩ Hn−k ]
By our inductive hypothesis these coefficients are all non-negative, and the
coefficient of tn (when k = 0) is positive (being the degree of X w.r.to the
embedding given by M ). If the constant term is negative, the polynomial
P will have precisely one real root t0 > 0.
Now write P (t) = Q(t) + R(t) where
Q(t) = c1 (L)(c1 (L) + tc1 (M ))n−1 [X]
34 T.R. RAMADAS
and
R(t) = tc1 (M )(c1 (L) + tc1 (M ))n−1 [X]
Now R(t) is a polynomial with no constant term, and non-negative coeffi-
cients, and with the coefficient of the top degree term strictly positive. So
R(t0 ) > 0.
We will now show next that Q(t0 ) ≥ 0, which yields the desired contra-
diction since Q(t0 ) + R(t0 ) = P (t0 ) = 0.
To prove that Q(t0 ) ≥ 0 it suffices to show that Q(t) ≥ 0 for any rational
number ab ≡ t > t0 (where a, b are coprime positive integers, of course).
Note that
1 1
c1 (L) + tc1 (M ) = {bc1 (L) + ac1 (M )} = c1 (Lb M a )
b b
so that Now given any k-dimensional subvariety V ⊂ X, with k < n,
c1 (Lb M a )k [V ] = (bc1 (L)+ac1 (M ))k [V ] = ak c1 (M )k [V ]+non − negativeterms > 0
since M is ample and therefore the first term is positive. On the other hand,
a
c1 (Lb M a )n [X] = (bc1 (L) + ac1 (M ))n [X] = bk P ( ) > 0
b
b a
Hence, by Nakai’s criterion, L M is ample. Then
1
Q(t) = ( )n−1 c1 (L)c1 (Lb M a )n−1 [X] ≥ 0
b
because L is nef.
This is generated by mx /m2x and therefore defines a cone in the Zariski tan-
gent space Tx X ≡ HomC (mx /m2 ). This affine scheme (“tangent cone”) need
not be reduced or irreducible (even if V itself is), but its (Krull) dimension
is equal to the dimension of X at x. In the projective space P(mx /m2x ) this
defines a sub-scheme (“projectivised tangent cone”) of dimension dim X −1.
The multiplicity multx (X) of X at x is defined to be its degree:
multx (X) = c1 (O(1))dimx X−1
[projectivised tangent cone]
Even if X itself is reduced and irreducible, the tangent cone is likely to be
neither, so our definition of intersection numbers has to be suitably extended.
LINE BUNDLES 35
where the I(C, D, x) are intersection multiplicities which we will not define.
The key fact which we will use is:
I(C, D, x) ≥ multx (C)multx (D)
8.3. Seshadri’s Criterion.
Theorem 8.1. Let X be a projective variety, and L a line bundle on X.
Then L is ample iff there exists (L) > 0 such that for every (reduced,
irreducible) curve C and every point x ∈ C, we have
(6) deg L|C ≥ (L) multx C
(In other words, the degree of L restricted to any curve C should be bounded
below uniformly in terms of the “maximum singularity” of C.)
36 T.R. RAMADAS
Proof. Suppose first that L is ample. Then some power Lm has a section
such that E ≡ (σ) such that σ|C 6= 0 but σ(x) = o. Thus E is effective,
passes through x, and “meets C properly” (i.e., C * E.) Then
X
m deg L|M = i(E, C, y) ≥ i(E, C, x) ≥ multx C
y∈C∩E
Clearly,
N ef (X) = {z ∈ N 1 (X)R |z(C) ≥ 0, ∀ C ⊂ X}
curve
1
= {z ∈ N (X)R |z(α) ≥ 0, ∀ α ∈ N E(X)}
= {z ∈ N 1 (X)R |z(α) ≥ 0, ∀ α ∈ N E(X)
The theory of dual cones yields
N E(X) = {α ∈ N1 (X)R |z(α) ≥ 0 ∀ nef z}
We finish with Kleiman’s criterion for amplitude via cones.
Proposition 8.5. A class w ∈ N 1 (X)R is ample iff w(α) > 0 for all
nonzero classes in N E(X). That is, iff w 6= 0 and N E(X) \ 0 is contained
in the open half-space defined by w:
N E(X) \ 0 ⊂ {α|w(α) > 0} ⊂ N1 (X)R
Proof. If w is ample, clearly w(α) ≥ 0 for α ∈ N E(X). Suppose w(α) = 0
for some nonzero α ∈ N E(X). Let z ∈ N 1 (X)R such that z(α) < 0.
Then z + tw is ample for large t, so z(α) = {z + tw}(α) ≥ 0, yielding a
contradiction.
Conversely, suppose w is a class such that
N E(X) \ 0 ⊂ {α|w(α) > 0}
Consider the function α 7→ w(α) on the closed compact set:
N E(X) ∩ {α|||α|| = 1}
Let be the minimum value. By assumption > 0, which yields, for any
curve C
w(C) > ||[C]||
where [C] is the class of C in N1 (X)R . Suppose that the norm in question
is X
||α|| = |wi (α)|
i
⊂
Amp(X) N ef (X)
⊂
(That Amp(X) ⊂ N E(X) is clear, and the other inclusions follow by taking
closures.)
(1) When is N ef (X) ⊂aP N E(X) not an equality?
P Clearly iff there is an
real effective divisor i ai Ci such that i ai Ci .C < 0 for some curve
C. Since C.C 0 ≥ 0 for distinct irreducible curves, this can happen
iff C = Ci for some i and C 2 < 0.
(2) Suppose there is no curve C with negative self-intersection. Then
N ef (X) = N E(X) ⊂ C. Since the interior of N ef (X) consists of
ample classes, any curve with C with zero self-intersection yields
a nonzero class [C] ∈ ∂N E(X) ∩ {z ∈ C|z 2 = 0}. Conversely, if
∂N E(X) ∩ {z ∈ C|z 2 = 0} is not empty, there exists such a curve.
(3) Suppose there does exist an irreducible curve C with negative self-
intersection. Let C≥0 ⊂ N ( X) denote the closed half-space of classes
whose intersection with C is non-negative. Claim: N E(X) is the
cone spanned by N E(X) ∩ C≥0 and C, and the half-line spanned by
40 T.R. RAMADAS
8.9. More on the cone of curves. Can it happen that N E(X) is not
closed? As we saw above, the closure N E(X) contains the ξ-axis, but we
will show that there is no curve C such that [C] lies on this axis, provided
E is suitably chosen. If this were the case, we would have
O(C) = O(m) ⊗ π ∗ A
for some m > 0 and a line bundle of degree zero on C. As we argued above
this implies that the bundle S m E contains the degree zero line sub-bundle
(Ǎ). If we can find E such that S m E is stable for every m > 0 (and such
bundles exist for genus C > 1), this cannot happen. The existence of such
an E follows from the Narasimhan-Seshadri Theorem.
Here is a proof supplied by Narasimhan. It is a fact that there exist two
elements A1 , B1 in SU (2) that generate a dense subgroup. (In fact this
is true for SU (N ), N ≥ 2.) Let A2 , B2 , . . . , Ag , Bg be elements of SU (2)
defined by A2 = B1 , B2 = A1 and Ai = Bi = I, i > 2. Then
Y
Ai Bi (Ai )−1 (Bi )−1 = I
i
so we have a representation of the fundamental group of C in SU (2). Since
the symmetric powers of the standard two-dimensional representation (on
C2 ) of SU (2) are irreducible representations of SU 2), the induced represen-
tations of the fundamental group in SU (m+1) = SU (S m C2 ) are irreducible.