Intersection Theory Notes
Intersection Theory Notes
1. 1/27/20 - Introduction
Announcements. The class will meet on MW in general, but not this Wednesday.
The references are:
• 3264 and all that by Joe Harris and David Eisenbud.
• Intersection Theory by William Fulton.
Next assume that X is a smooth manifold, and that A, B ⊂ X are closed sub-
manifolds. If A and B intersect transversely (which can be arranged by perturbing
them slightly), then
codim(A ∩ B) = codim(A) + codim(B).
So if we restrict our attention to submanifolds of X, then we should be able to
define a graded ring structure, graded by codimension. To accommodate small per-
turbations, we need to consider equivalence classes instead of literal submanifolds.
Exercise. If [ωA ] and [ωB ] are the Poincaré duals of closed oriented submani-
folds A, B ⊂ X meeting transversely, then
ωA ∧ ωB = [ωA∩B ].
The de Rham cohomology ring gives us a model for transverse intersections:
Mn
H ∗ (X, R) = H a (X, R).
a=0
1
2 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
the Poincaré dual of a point to 1. This allows us to count the (oriented) number of
intersection points between A and B when codim(A) + codim(B) = n.
Poincaré duality takes submanifolds to integral cohomology classes, and in fact the
wedge product can be defined directly on the integral cohomology groups, where it
is called the cup product.
This can give funny answers. For example, in a genus 4 surface (with the holes
arranged like a cross), you can draw a pair of loops which meet twice, but their
oriented intersection number is 0. These loops cannot be deformed to disjoint loops.
One way to avoid the orientation issue is to study complex manifolds and their
holomorphic submanifolds. A complex vector space induces a natural orientation
on the underlying real vector space such that any two complementary complex sub-
spaces have positive intersection. Furthermore, the real codimension of a complex
submanifold is always even, so the wedge product is commutative.
If C meets D transversely, then this theorem follows immediately from the dis-
cussion above. If they have non-transverse intersection points, then there is a
multiplicity involved.
We will prove Bézout’s Theorem later on. To finish up, let us work out a sample
application of intersection theory.
Answer. 2! And the answer is the same for lines in R3 (easier to visualize).
To get this answer, we consider the moduli space of all lines in C3 . This space
is non-compact, so to use an intersection product, we compactify it by considering
the space of projective lines in CP3 . The result is a Grassmannian:
Now that we have a nice moduli space, we need to impose the condition of meeting
the fixed lines `1 , `2 , `3 , `4 ⊂ P3 . Let’s define
The count we are after is the size of Σ(`1 ) ∩ Σ(`2 ) ∩ Σ(`3 ) ∩ Σ(`4 ). A dimension
count reveals that each Σ(`i ) is a codimension 1 subvariety1 of G(1, 3). In fact, it is
cut out by the equation `i ∧ α = 0. The count we want is the intersection number
of a quadric hypersurface with 4 hyperplanes in P5 , which is 2 by the multiplicative
property of degree.
To generalize this example, we can study the cohomology ring of the Grassmannian:
Z
Z
H 2∗ (G(1, 3), Z) ' Z2
Z
Z.
The generator of H 2 (G(1, 3), Z) is called σ1 , and it is the class of Σ(`) for any line
` ⊂ P3 . The observations above imply that
Z
(σ1 )4 = 2.
G(1,3)
Question. What are the two generators of H 4 (G(1, 3), Z)? To be continued.
2. Assume that `1 and `2 intersect at p (so they are not in general position).
There are two ways for L to meet both `1 and `2 : either L lies in the plane H
spanned by `1 and `2 , or L contains p. In the first case, L must be the line between
`3 ∩ H and `4 ∩ H. In the second case, projecting from p sends `3 and `4 to a pair
of lines on a screen which meet at a single point q, so L must be pq.
1Σ(`) is a singular subvariety because it is the intersection of G(1, 3) with the hyperplane
tangent to G(1, 3) at [`].
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 5
Theorem. There exists a product structure on A(X) satisfying the condition that
if A, B ⊂ X are generically transverse subvarieties, then
[A] · [B] = [A ∩ B].
Generically here means in A ∩ B. The most intuitive definition of the intersection
product goes through the Moving Lemma.
How do we compute the Chow group of X? If dim(X) = n, then An (X) = Z[X] be-
cause P1 ×X is irreducible. If X is reducible but equidimensional, then An (X) = Zc .
Furthermore A(Xred ) = A(X).
6 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
0 / Z(Y ) / Z(X) / Z(U ) /0
A(Y ) / A(X) / A(U )
0 0 0
the first vertical maps are ∂, defined to be zero on cycles contained in fibers, and
otherwise the difference between the fibers at {0} and {∞}. The theorem follows
from a diagram chase. The Mayer-Vietoris is similar.
Remark. The right exact sequences above can be extended to the left using the
higher Chow groups of Bloch, which are a special case of motivic cohomology.
Remark. In practice, we can avoid appealing to this theorem by checking that the
closed strata in complementary dimension have nonsingular intersection matrix.
Define the pushforward on cycles f∗ : Z(X) → Z(Y ) sending [A] to 0 if dim(A) >
dim(f (A)), and [A] to d[f (A)] otherwise,
d = [R(A) : R(f (A))].
Proposition. The homomorphism f∗ descends to f∗ : A(X) → A(Y ).
Proof. Use the norm map on fields of rational functions, to be introduced later.
Exercise. Find a non-proper map such that f∗ is not well-defined (hint: try
an open immersion).
If X is proper over a point, then point classes are nonzero, which gives the no-
tion of degree for a 0-cycle.
NB: In the case when f is flat, you can just define f ∗ : Z(Y ) → Z(X) and it
descends to Chow; no Moving Lemma is required. Fulton calls this the flat pull-
back, and the more general one the Gysin pullback denoted f ! .
We will describe the Chow group A(G(1, 3)) by giving an affine stratification. Fix
a complete flag p ∈ L ⊂ H ⊂ P3 .
Σ0,0 = G(1, 3)
Σ1,0 = {Λ : Λ ∩ L 6= 0}
Σ2,0 = {Λ : Λ ∩ p 6= 0}
Σ1,1 = {Λ : Λ ⊂ H}
Σ2,1 = {Λ : p ∈ Λ ⊂ H}
Σ2,2 = {Λ : Λ = L}
1. The dense open stratum Σ◦0,0 can be thought of as the locus of V ⊂ C4 com-
plementary to a fixed L ⊂ C4 . Fix one such Ω (as the origin). All others will be
graphs of linear maps from Ω → L, so we get an A4 .
3. The middle two can be thought of as P2 and P2∗ respectively, and you are
removing a P1 = Σ2,1 in both cases. 4. P1 = A1 ∪ {∞}.
We now know the Chow group completely. What is the ring structure? It’s easy
to see that
2 2
σ1,1 = 1, σ2,0 = 1, σ1,1 σ2,0 = 0, σ1 σ2,1 = 1.
Next up, we see that σ1 σ2 = σ2,1 and σ1 σ1,1 = σ2,1 also.
Last time, we saw that A0 (X) = Z[X]. The next case to consider is A1 (X).
Here, rational equivalence of divisors coincides with linear equivalence. By the
correspondence between divisors and line bundles, we have an isomorphism
c1 : Pic(X) → A1 (X).
Remark. If X is singular, then c1 is no longer surjective; in that case A1 (X) '
Cl(X), the Weil divisor class group.
(2) Every component C ⊂ A ∩ B contains points p such that Λ does not meet
Tp X. Indeed, this is a Zariski open condition, and we can choose Λ so that it’s
true for one point on each component. The projection πΛ : X → Pn is nonsingular
at such points p. Hence p ∈/ A ∩ A0 because A
e = A ∪ A0 is singular along A ∩ A0 .
(3) Here we punt a little and just show that A is dimensionally transverse to B ∗ .
With another page of work, we could show that they are generically transverse, but
with the theory of intersection multiplicities the weaker statement suffices. Consider
Ψ = {(Λ, p, q) ∈ G(N − n − 1, N ) × A × B ∗ : Λ ∩ pq 6= ∅}
The fiber over a point (a, b) ∈ A × B ∗ is the set Λ meeting a fixed line, which is
codimension n inside G. Thus Ψ is irreducible, and
dim Ψ = dim A + dim B + dim G − n.
e ∩ B ∗ = A0 ∩ B ∗ . That fiber has
Now, a general fiber of Ψ → G surjects onto A
dimension dim A + dim B − n, so
e ∩ B ∗ ) ≤ dim A + dim B − n.
dim(A
On the other hand, dim(A e ∩ B ∗ ) ≥ dim A + dim B − n because codimension is
subadditive, so we are done.
section of X (a smooth conic C), which meets B at a single point. On the other
hand, A0 already meets B in a single point. We see that A · B = 0 using
A ∼ C − A0 .
3.3. Applications. Now we return to the Grassmannian G(1, 3) to see what the
intersection product can do for us.
Recall that Σ2 is the locus of lines through p. The linear projection πp : P3 99K P2
sends C to a degree d plane curve with b nodes. A smooth plane curve always
has genus d−1
2 by the Riemann-Hurwitz formula. The number of nodes is the
difference between the arithmetic and geometric genera of πp (C) ⊂ P2 :
d−1
b= − g(C).
2
The genus g(C) must be part of the problem; it can take any value from 0 to the
Castelnuovo bound, which is quadratic in d.
12 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
4.1. Grassmannians. Today we will discuss the Chow ring of a general G(k, n).
As in the case G(2, 4) from last week, G(k, n) admits an affine stratification where
the strata are indexed by sequences a of integers of length k such that:
n − k ≥ a1 ≥ a2 ≥ · · · ≥ ak ≥ 0.
This can be viewed as a Young tableau contained in a box of height n−k and width
k, with the bars decreasing in height from left to right. We will suppress trailing
zeroes from the notation.
Remark. There are nk such tableaux, by counting possibilities for the jagged
With this in hand, we can encode the full ring structure in terms of structure
constants:
c
γa,b = σa · σb · σc∗ .
c
Remark. The γab are always positive because they are actually the same as
Littlewood-Richardson coefficients. Irreducible representations of GLk are classified
by partitions of length k (Schur functors applied to the standard representation).
If you tensor together two irreps, you can decompose the result into irreps. This
gives a ring called the representation ring, Rep(GLk ) which can also be expressed
in terms of characters as polynomial functions in the entries of a matrix, invariant
under conjugation. We have a sequence of isomorphisms
Rep(GLk ) → Sym∗ (gl∨
k)
GLk
→ H 2∗ (BGLk )
(closure followed by Chern-Weil). The latter has a natural map to → H 2∗ (G(k, n))
via the tautological bundle on G(k, n). A particular Schur irrep maps to the corre-
sponding Schubert class via the composition.
Proof. Similar to our proof of the fact that σ12 = σ1,1 + σ2 . Move the Pn−2
defining Σ1 so that it intersects P2−a in codimension 1, and P3−b in codimension
2. There are two ways for Λ to lie in Σ1 ∩ Σa,b : either it meets Pn−2 ∩ P2−a , or it
lies inside the span of P2−a and Pn−2 ∩ P3−b , which is a P3−b−1 .
Corollary. In G(1, n), the top intersection σ12n−2 has degree Catalan(n+1).
14 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
Proof. Stacking unit boxes inside the (n + 1) × 2 box such that each step is a
Young tableau is the same as placing n+1 pairs of parentheses.
Theorem. (Pieri) In G(k, n), σc · σa is equal to the sum over all tableaux ob-
tained by adding c blocks to a, at most one block in each row.
These σc are called special Schubert classes. Actually, any class can be expressed
as a polynomial in these special classes:
Proof. Use the Pieri formula with induction by expanding along the last column.
For example:
σ σa1 +1
σa1 ,a2 = a1 = σa1 σa2 − σa1 +1 σa2 −1 .
σa2 −1 σa2
Pieri and Giambelli together give an algorithm to compute any intersection prod-
uct. Vakil gave a more efficient algorithm for computing LR coefficients in terms
of checkers puzzles. Positivity is clear from Vakil’s approach.
Remark. The term “Fano variety” has two meanings. In this context it means
the variety of lines contained in X. More broadly, a Fano variety is a variety
whose anticanonical line bundle is ample. The meanings do not match; F (X) is
often not a Fano variety in the second sense. To add to the confusion, the hyper-
surfaces appearing in the Debarre-de Jong Conjecture are Fano in the second sense.
We can compute the class of F (X) ∈ Ad+1 (G(1, n)) as the top Chern class of
a vector bundle E on G(1, n) = G(2, n + 1). There is a tautological sub-bundle
S ⊂ O⊕n+1 , and we set E = Symd (S ∨ ). The fiber of E at a line L ' P1 ⊂ Pn is
canonically identified with H 0 (L, O(d)). The equation defining X ⊂ Pn is a degree
d polynomial, an element of H 0 (Pn , O(d)). For any line L, we can restrict the
polynomial to H 0 (L, O(d)), so there is a section σX ∈ H 0 (E). The zero locus of
σX consists of lines contained in X, which is precisely the locus F (X).
To be precise, dualizing and taking the symmetric power of the tautological in-
clusion S ,→ O⊕n+1 , we have
∨⊕n+1
OG(2,n+1) ⊗ H 0 (Pn , O(d)) ' Symd (OG(2,n+1) ) → Symd (S ∨ ) → 0
Taking global sections, we have a map H 0 (Pn , O(d)) → H 0 (E) sending the equation
defining X to σX .
To answer this, we need a general version of the adjunction formula. Recall that if
Y is a smooth projective variety, and D ⊂ Y is a divisor, then
KD = (KY + D)|D .
This can be proved by taking c1 of the terms in the normal bundle short exact
sequence:
0 → TD → TY |D → ND/Y → 0,
together with the observation that ND/Y ' O(D)|D . The latter is true because
∨ 2
ND/Y ' ID /ID ' ID ⊗ OY /ID ' OY (−D) ⊗ OD = O(−D)|D .
If instead D ⊂ Y is subvariety of codimension r cut out by a section of a rank
r vector bundle E, then we have ND/Y ' E|D . To prove this, take the Koszul
complex for OY → OD , and restrict it to D. Putting all this together, we deduce
the following formula for the genus g of F (X) in the case where 2n − d − 3 = 1:
2g − 2 = deg KF (Y ) = deg KG(1,n) + c1 (Symd (S ∨ ))
F (X)
d ∨ d ∨
= deg KG(1,n) + c1 (Sym (S )) · ctop (Sym (S )).
The splitting principle can be used to prove several identities, which we leave as
exercises. For vector bundles E and F or ranks e and f , respectively:
ck (E ∨ ) = (−1)k ck (E);
c1 (E ⊗ F ) = f c1 (E) + ec1 (F );
k
X r−k+i
ck (E ⊗ L) = c1 (L)i ck−i (E);
i=0
i
cef (E ⊗ F ) = Resultant(ct (E), ct (F )).
Given any sequence of symmetric analytic functions
f1 (x1 ), f2 (x1 , x2 ), f3 (x1 , x2 , x3 ), . . .
one can define a characteristic class of algebraic vector bundles. If E is a bundle
of rank r, rewrite fr as a power series in the elementary symmetric polynomials
pk , and then replace each pk with ck . For particular choices of fr the resulting
characteristic class may have nice geometric properties. For example:
r
Y
fr = (1 + xi ) (Total Chern class)
i=1
r
X
fr = exi (Chern character)
i=1
r
Y xi
fr = (Todd class).
i=1
1 − e−xi
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 19
ch(E ⊗ F ) = ch(E)ch(F ).
ch : K0 (X) → A∗ (X) ⊗ Q.
f∗
K0 (Y ) / K0 (X)
ch ch
∗
A∗ (Y )Q
f
/ A∗ (X)Q .
For covariant (pushfoward) functoriality, we must define a candidate for the proper
pushforward of a vector bundle E. The pushforward of the locally sheaf E is not
necessary locally free. For example, the pushforward OZ via a closed immersion
Z ,→ X will be a torsion sheaf. The correct setting to define the proper pushfor-
ward is on the (larger) abelian category of coherent sheaves on X.
Proof. It suffices to show that every coherent sheaf admits a resolution by lo-
cally free sheaves. This follows from the Hilbert Syzygy Theorem.
The higher pushforwards Ri f∗ (E) are explicitly given by sheafifying the presheaf
(U ⊂ Y ) 7→ H i (f −1 (U ), E|f −1 (U ) ).
The fact that f∗ is well-defined on K0 follows from the long exact sequence. The fact
that it is a covariant functor follows from the spectral sequence for compositions.
20 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
K0 (X)
f∗
/ K0 (Y )
ch ch
A∗ (X)Q
f∗
/ A∗ (Y )Q .
But alas, this is still not so. Consider the case when X is a curve and Y is a point.
h0 (L) − h1 (L) 6= deg(L).
The Riemann-Roch Theorem (for curves) says that
h0 (L) − h1 (L) = deg(L) + 1 − g = deg(L).
The Riemann-Roch Theorem (for surfaces) says that
c1 (L)2 + c1 (L)c1 (TX ) c1 (TX )2 + c2 (TX )
h0 (L) − h1 (L) + h2 (L) = + .
2 12
Notice that this is the codimension 2 part of
c1 (L)2 c1 (TS ) c1 (TX )2 + c2 (TX )
1 + c1 (L) + 1+ +
2 2 12
The right hand factor is the 2-truncated part of the Todd class. To see a bit more:
c1 c2 + c2 c1 c2 −c41 + 4c21 c2 + c1 c3 + 3c22 − c4
1+ + 1 + + + ...
2 12 24 720
Lemma. If E has rank r, then
Xr
(−1)i ch(∧i E ∨ ) = cr (E) · Td(E)−1 .
i=0
Proof. This is a straightforward application of the splitting principle:
r r r
X
i i ∨
Y
−αi
Y 1 − e−αi
(−1) ch(∧ E ) = (1 − e ) = (α1 α2 . . . αr ) .
i=0 i=1 i=1
αi
Theorem. (Grothendieck) Let f : X → Y be a morphism of smooth projective
varieties and E a coherent sheaf on X. Then in A∗ (Y ) we have the equality:
ch(f∗ E) · Td(Y ) = f∗ (ch(E) · Td(X)) .
K0 (X)
f∗
/ K0 (Y )
ch(−)·Td(X) ch(−)·Td(Y )
A∗ (X)Q
f∗
/ A∗ (Y )Q .
Step 2: Check that the identity holds for X = Pm × Y → Y . This can be reduced
to checking it for Y = pt: the box product K0 (Pm ) ⊗ K0 (Y ) → K0 (Pm × Y ) is
surjective, by induction on m with the localization sequence for K0 . If the outer
and upper squares commute, then lower square commutes:
K0 (Pm ) ⊗ K0 (Y ) / K0 (pt) ⊗ K0 (Y )
∼
K0 (Pm × Y )
f∗
/ K0 (Y )
The proof for ch(f∗ E) is similar; the Koszul complex above tensored with p∗ E gives
a resolution of f∗ E and the rest follows.
Putting everything together, we use the fact that any morphism f : X → Y can be
factored X → Γf ⊂ Pm × Y as a closed immersion followed by a simple projection
with fiber Pm . The deformation to the normal cone trick (next lecture) allows us
to reduce the statement for closed embeddings to the setting of Step 3.
22 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
Y
f
/ PE
π p
}
X
and sub-line bundles L ⊂ π ∗ E over Y . To go from a commutative triangle to a line
bundle, simply pull back the tautological line bundle S via f :
f ∗ S ⊂ f ∗ p∗ E = π ∗ E
To go the other direction, cover X by open sets trivializing both E and L, and
define f |U using the universal property for projective space.
In what follows, ζ = c1 (OPE (1)) is the relative hyperplane class, which depends
on the choice of E, not just on the projective bundle PE. We will assume that E
has rank r + 1 for now so that PE has fiber Pr .
24 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
Proof. We start with the case of projective space Pr (so X = pt) and dim(Z) = n.
It’s possible to find coordinates xi on Pr such that (x0 = x1 = · · · = xn = 0) is
disjoint from Z. Define
I 0
gt = n+1 ∈ P GL(r + 1).
0 tIr−n
The flat limit of gt (Z) as t = 0 will be the nonreduced cycle deg(Z) · Pn , so
Z ∼ ζ r−n deg(Z).
The case of PE is similar. If L is ample on X then for N 0, E ∨ ⊗ L⊗N is globally
generated. Replacing E with E 0 = E ⊗ L−⊗N does not affect the projectivization.
Fix a point x ∈ W ⊂ X. For a general choice of global sections τ0 , τ1 , . . . , τr ,
(1) (τ0 )x , (τ1 )x , . . . , (τn )x forms a basis for the fiber Ex∨ .
(2) (τ0 = τ1 = · · · = τk−l = 0) is disjoint from Zx .
In fact, both conditions are Zariski open in X, so they are true on some U ⊂ X.
Using condition (1),
PE|U = U × Pr ,
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 25
Theorem. As a ring,
A∗ (PE) ' A∗ (X)[ζ]/(ζ r+1 + c1 (E)ζ r + · · · + cr+1 (E))
Proof. It suffices to prove that the monic polynomial relation above is satisfied.
There can be no other relations by our description of the group structure. From
0 → S → p∗ E → Q → 0,
we have c(S)c(Q) = c(p∗ E), so c(Q) = c(p∗ E)(1 + ζ + ζ 2 + . . . ). Since rk(Q) = r,
0 = cr+1 (Q) = ζ r+1 + c1 (E)ζ r + · · · + cr+1 (E).
Remark. Fulton uses this result as the definition of (algebraic) Chern classes.
⊕r+1
Remark. Applying the theorem to the trivial bundle OX , we obtain
A∗ (X × Pr ) ' A∗ (X) ⊗ A∗ (Pr ).
This is called a Chow-Künneth formula. No such formula holds for general products.
For the rest of the lecture, we will assume that E has rank r because the main
player will be P(E ⊕ OX ), which has fiber Pr .
Proof. To compute, [POX ] · [POX ] we write one of them in terms of ζ and leave
the other alone. Since zero and ∞ are disjoint, [POX ] · ζ = 0.
[POX ]2 = [POX ] · (ζ r + c1 (E)ζ r−1 + · · · + cr (E)
= [POX ] · p∗ cr (E)
= j∗ (cr (E)).
Proposition. (?) More generally, for any cycle α ∈ A∗ (X), j ∗ j∗ (α) = α · cr (E).
Proof. For any β ∈ A∗ (P(E ⊕ OX )), by the Moving Lemma j ∗ β = p∗ (β · [POX ]).
j ∗ j∗ (α) = p∗ (j∗ (α) · [POX ])
= p∗ (p∗ α · [POX ] · [POX ])
= p∗ (p∗ α · j∗ cr (E))
= α · p∗ j∗ cr (E)
= α · cr (E).
Definition. Let Z ⊂ X a closed subscheme defined by an ideal sheaf I. The
blow-up of X along Z is a new scheme given by Proj of the Rees algebra:
BlZ X := Proj OX ⊕ I ⊕ I 2 ⊕ . . . → X.
If we restrict to Z ⊂ X, the fibered product is obtained by tensoring the Rees
algebra with OX /I:
EZ X := Proj OX /I ⊕ I/I 2 ⊕ I 2 /I 3 ⊕ . . .
If Z is smooth (or more generally l.c.i.), then I k /I k+1 ' Symk NZ/X
∨
. In that case,
EZ X ' Proj Sym∗ NZ/X ∨
= PNZ/X .
By a dimension count, we find that EZ X ⊂ BlZ X is a divisor. In fact, the blow up
is the minimal way to replace Z with a Cartier divisor.
Fulton and MacPherson had the idea to specialize cycles in X to cycles in NZ/X in
M ◦ := M r BlZ X → P1 .
At the level of sets, a subvariety B ⊂ X specializes to the normal cone
CB∩Z/B ⊂ NZ/X ,
and this descends to a map on Chow groups:
A∗+1 (NZ/X )
i∗
/ A∗+1 (M ◦ ) / A∗+1 (X × A1 ) /0
O
i∗ ∼
w
A∗ (NZ/X ) A∗ (X).
The dotted arrow exists because for i : NZ/X ,→ M ◦ the inclusion, i∗ i∗ = 0.
This allows us to give a definition for Gysin pullback without appealing to the
Moving Lemma. If j : Z → X is the inclusion, then
j ∗ := (π ∗ )−1 ◦ σ.
Remark. Fulton uses this construction to define the intersection product.
With a few formal observations, we can get a lot of mileage out of this. Recall
28 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
that the zero section PO ⊂ P(N ⊕ O) is cut out by a section of the tautological
quotient bundle Q. If β ∈ Ak (N ), then we have
(π ∗ )−1 β = p∗ β · [PO]
= p∗ β · cr (Q)
= p∗ β · c(Q) k−r
= p∗ β · c(p∗ (N ⊕ O)) · (1 + ζ + ζ 2 + . . . ) k−r
= c(N ) · p∗ β · (1 + ζ + ζ 2 + . . . ) k−r
The factor p∗ β · (1 + ζ + ζ 2 + . . . ) is an example of a total Segre class.
Proof. Apply p∗ to the identity c(Q) = c(p∗ E)(1 + ζ + ζ 2 + . . . ) and use the
push-pull formula.
The total Segre class contains the same information as the total Chern class, but
it has the advantage of being definable using only Chow groups. For thisL reason,
the definition can be generalized to cones. Any graded OX -algebra S = i≥0 S i
such that OX S 0 and S generated by S 1 gives rise to a cone scheme C =
Spec (S) → X. The Segre classes of C are defined in the analogous way, using
p : P(C ⊕ OX ) → X and ζ = c1 (OP(C⊕O) (1)).
s(C) = p∗ 1 + ζ + ζ 2 + . . . ∈ A∗ (X).
a normal bundle (in our case Z) which is equivalent to it being regularly embedded,
that is cut out by a regular sequence of length r. In that case
Symk (IZ /IZ2 ) ' IZk /IZk+1 .
If Z, B, and B ∩ Z are all smooth, then using the normal bundle sequence
0 → NB∩Z/B → NB∩Z/X → NB/X |B∩Z → 0
we can rewrite the excess intersection formula more symmetrically:
c(NZ/X |B∩Z ) · c(NB/X |B∩Z )
[B ∩ Z] = .
c(NB∩Z/X ) k−r
The most general form of the excess intersection formula is stated for Z → X
a regular embedding and B → X an arbitrary morphism. The fibered product
W = B ×X Z fits into a Cartesian square
W
i /B
g f
Z
j
/ X.
Theorem. For any class α ∈ Ak (Z), we have
f ∗ j∗ (α) = i∗ g ∗ (α · c(NZ/X )) · s(CW/B ) k+dim(B)−dim(X) .
Lastly, let us return to the reduction step in the proof of the Grothendieck-Riemann-
Roch theorem. To relate the case of an arbitrary closed embedding to the zero
section of the completed normal bundle, we cited the deformation to the normal
cone trick. The key fact is that M → P1 is flat, by the following “Miracle Flatness:”
where Γ is the finite set that we want to count. The correct answer is 65 minus the
excess contribution supported on S ' P2 . We will use h = c1 (OP2 (1)).
5
Y
excess = deg c(NZi /P5 |S ) · s(CT /P5 ).
i=1
A second approach to the five conics problem involves blowing up the Veronese
surface S to tease apart the scheme-theoretic behavior there. Luckily for us,
M ' BlS P5
has a nice moduli theoretic description. It happens to be the space of complete
conics. This is the closure in P5 × P5∗ of the locus U consisting of (C, C ∗ ) where C
is a smooth conic and C ∗ its dual. It also happens to be isomorphic to the coarse
moduli scheme of the Kontsevich stack of stable maps:
M0 (P2 , 2h) = {f : C → P2 | C is conn. nodal genus 0, f∗ [C] = 2h, |Aut(f )| < ∞}.
The fibers of the exceptional E → S are P2 ' Sym2 P1 , encoding the branch
points of the double cover C → P1 . From these models, it is clear that the proper
transforms Zei intersect only in a finite set away from the exceptional divisor:
5
\
Z
ei = Γ.
i=1
The class of Zei in the blow up is 6H − 2E = 2(3H − E), because Zi vanishes with
multiplicity 2 along S. To prove this, note that a general pencil of conics containing
a double line meets Zi at only 4 other points, by the Riemann-Hurwitz argument.
(3H − E)5 = 243H 5 − 5 · 81H 4 E + 10 · 27H 3 E 2 − 10 · 9H 2 E 3 + 5 · 3HE 4 − E 5
= 243[pt] − 90 · j∗ (2h)2 ζ 2 + 15 · j∗ (2h)(−ζ 3 ) − j∗ (ζ 4 )
T
such that F∗ [Ct ] = dh. This pre-stack turns out to be a smooth algebraic stack. To
make it proper, we allow the curves Ct to be nodal, and require that any contracted
component have ≥ 3 nodes. This makes the isotropy groups finite. A coarse space
for a stack X is a morphism m : X → X to an algebraic space such that
m k :X k →X k
is a bijection, and any morphism to an algebraic space factors through m.
Theorem. (Keel-Mori) Any stack with finite inertia has a coarse moduli space.
The map m looks like [Tx X/G] → Tx X/G formally locally around a smooth point
x ∈ X with isotropy group G. By a theorem of Chevalley-Shephard-Todd, if G is
generated by pseudo-reflections, i.e. elements fixing a hyperplane, then the quotient
is smooth (and conversely). This is true when d = 2, but fails when d ≥ 3.
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 33
for all i. This follows from the Hodge Index Theorem, which says that the inter-
section matrix of the Ei is negative definite.
For such D, we can define D · Z for any cycle [Z] ∈ Ak (X) by pushing forward
1
D|Z := c1 (O(mD)|Z ) .
m
This recovers Mumford’s construction, since a Cartier divisor D satisfies
∗ D · E = 0.
More generally, Chern classes of vector bundles can be defined and pulled back via
morphisms of schemes. The correct setting for a ring structure on the full Chow
group is that of smooth algebraic stacks.
34 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
Axioms (1) and (2) assert that X is a 2-sheaf on the site Sch/k; isomorphisms
glue, and objects 2-glue. Axiom (3) implies that any morphism from a scheme to X
is representable, via the Magic Square (below), which is Cartesian in any category
with fibered products. This allows us to describe maps from schemes to X.
X ×Z Y / X ×Y
Z
∆ / Z × Z.
Representable morphisms between stacks can be modified by any adjective that
modifies morphisms of schemes, is local on both the domain and target, and is
preserved under base change (e.g. flat, smooth, finite, étale). Hence, Axiom (4)
only makes sense given Axiom (3).
NB: The hypothesis on ∆ for a DM stack is slightly weaker than the hypothesis of
the Keel-Mori theorem (existence of coarse spaces) which was finiteness of inertia
stack IX → X, defined as IX = X × X, but this distinction rarely matters.
X×X
Objects of IX are pairs of isomorphisms from x → y. This category is equivalent
to one whose objects are pairs (x, φ) where φ is an automorphism of x.
Absolute properties of X like being reduced or finite type are defined from the
atlas scheme Y . Smoothness and properness can also be defined via lifting criteria.
Definition. A stack X is proper if for every discrete valuation ring R with fraction
field K, any morphism Spec K → X extends to Spec R → X.
Theorem. (Kresch) Artin stacks have Chow groups. If they are smooth and
(1) Deligne-Mumford, or (2) stratified by global quotients stacks, then there is an
intersection product defined over Z.
where B̂k (Y) is a union over all stacks T with pairs of projective morphisms p1 , p2 :
T → Y such that f ◦ p1 ' f ◦ p2 of the following abelian group:
{p1∗ β1 − p2∗ β2 | (β1 , β2 ) ∈ Âpk1 (T) ⊕ Âpk2 (T) satisfies ιp1 (β1 ) = ιp2 (β2 )}.
To make sense of this, we need to define
Âpk (T) := lim Ak+rk(E) (p∗ E); ιp : Âpk (T) → Âk (T).
→ BY
(a0 , . . . , an ). The approximations to Chow are given by (Pn ×(Cl −0))/C× → Pl−1 ,
which is isomorphic to:
P(O(a0 ) ⊕ · · · ⊕ O(an )) → Pl−1
The Chow ring of this variety is given by:
Z[h, ζ]/(hl , ζ n+1 + σ1 ζ n + · · · + σn+1 ),
where σi is the ith symmetric polynomial in the variables a0 h, a1 h, . . . , an h. As we
send l to infinity, we are left with
Z[h, ζ]/(ζ n+1 + σ1 (a)hζ n + · · · + hn+1 σn+1 (a)).
In particular, if C× acts on P1 with weights ±1, we get Z[h, ζ]/(ζ 2 − h2 ).
Theorem. (Weyl) Let G be a semisimple Lie group, g its Lie algebra, and h ⊂ g
the Cartan subalgebra. The irreducible representations Π of G are classified by
their highest weight, λ. For such an irrep, the character of eH ∈ T is given by
(w) ew(λ+ρ)(H)
P
chΠ (e ) = Q w∈W α(H)/2
H .
α∈∆+ e − e−α(H)/2
The character formula is enough to determine chΠ on all of G because any semi-
simple element can be conjugated into T , and semisimple elements are Zariski dense.
Setting H = 0, we recover the dimension formula for irreps.
Proof. In the case of SL2 , the torus is C× . The highest weight is a nonneg-
ative integer n, and the representation can be realized as H 0 (P1 , O(n)). Weyl’s
formula in this case reads
ei(n+1)θ − e−i(n+1)θ
iθ
e 0 sin((n + 1)θ)
tr πn −iθ = =
0 e eiθ − e−iθ sin(θ)
This will be a consequence of EGRR for the action of C× on P1 with weights ±1.
×
χC
K0 [P1 /C× ] / K0 [pt/C× ]
τ τ
Â∗Q [P1 /C× ]
π∗
/ Â∗ [pt/C× ]
Q
40 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
The equivariant Euler characteristic of OP1 (n) is an element of K0 (BC× ) = Z[t, t−1 ].
Since OP1 (n) has no higher cohomology, we get the representation Πn = H 0 (P1 , O(n)).
1 nζ 2ζ
τ (Πn ) = π∗ (ch(OP (n)) · Td(P )) = π∗ e
1
1 − e−2ζ
Lemma. For p(ζ) ∈ A∗Q [P1 /C× ] = Q[h, ζ]/(ζ 2 − h2 ),
p(h) − p(−h)
π∗ (p(ζ)) =
2h
and the same is true after completion.
Proof. From the P1 -bundle model, any polynomial in h only gets killed by π∗ .
Since ζ 2 = h2 , all even degree terms in p(ζ) go to 0, whereas an odd degree term
a2k+1 ζ 2k+1 = a2k+1 ζh2k goes to a2k+1 h2k .
The morphism [X/G] → pt is not representable, so GRR does not apply directly.
However, there is a generalization of Hirzebruch-Riemann-Roch in the case when
[X/G] is proper and Deligne-Mumford. We would like a formula for the Euler
characteristic of a bundle V on [X/G] in terms of intersection numbers (degrees of
0-cycles) coming from Chern characters and Todd classes. A 0-dimensional substack
of [X/G] is a gerbe for a finite group Gx , and we set:
deg(BGx ) = 1/|Gx |.
Occasionally we write deg as an integral over [X/G], in analogy with cohomology.
Recall that K0 [X/G] is a ring, but it also a module over K0 [pt/G] via the pull-
back through [X/G] → [pt/G]. Geometrically, K0 [X/G]C is a coherent sheaf over
Spec K0 [pt/G]C . If G is a diagonalizable group (subgroup of a torus), then
Spec K0 [pt/G]C = Spec Rep(G)C ' G.
For a more general linear algebraic group,
Spec K0 [pt/G]C = Spec Rep(G)C ' G ad G,
whose points are conjugacy classes of semisimple elements of G. Both of these
isomorphisms send a representation to its character. The augmentation ideal in
Rep(G)C corresponds to the identity element 1 ∈ G, since the trace 1 recovers the
dimension of a representation.
Lemma. (Thomason) If h ∈ G does not fix any points in X, then the localization
(K0 [X/G]Q )h = 0.
If [X/G] is Deligne-Mumford, then only finite order elements h ∈ H can have fixed
points in X. This implies that K0 [X/G]C is supported over finitely many points hi :
k
Y
K0 [X/G]Q ' (K0 [X/G]Q )hi .
i=1
Proof. EGRR says that τ is an isomorphism after taking the completion K̂0
at 1, but τ it is only covariant for representable morphisms. There exists a smooth
G-variety X 0 with a finite G-equivariant surjection f : X 0 → X such that [X 0 /G] is
represented by a smooth variety. Now use the covariance of τ and the surjectivity
of f∗ on Chow groups, which implies surjectivity of f∗ on K̂0 :
K0 (X 0 /G)Q
f∗
/ K̂0 [X/G]Q / K0 (pt)Q = Q
∼ τ ∼ ∼
A∗ (X 0 /G)Q
f∗
/ Â∗ [X/G]Q / A∗ (pt)Q = Q.
The left square commutes by EGRR, and large square commutes by HRR.
To compute the full χ([X/G], α), we must add up contributions from the remaining
sectors, closed points in the support of K0 [X/G]Q over Rep(G)Q . Each contribution
will be the degree of a Hirzebruch-Riemann-Roch type expression. Our main tools
will be localization in K-theory, followed by twist operators.
42 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
as modules over K0 [pt/G] ' Rep(G). This is referred to as the localization theorem.
Now, the localization theorem follows from geometry! If the weights are distinct
and h is general, then there are n + 1 fixed points, so
n
Y Yn
K0 [(Pn )h /C× ] = K0 (BC× ) = Z[t, t−1 ]
i=0 i=0
m−1
If m > 1 weights coincide, then there is a P in the fixed locus which contributes
×
K0 (BC ) ⊗ Z[x]/(x − 1) m
= Z[t, t−1 ][x]/(x − 1)m .
When you localize at h, the two Artinian rings match geometrically (graph x = t−a ),
and this correspondence can be refined for h a root of unity.
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 43
Using the inverse formula, along with functoriality of [X h /G] → [X/G] → pt,
we find that for αh ∈ (K0 [X/G])h ,
i∗ αh
h
χ([X/G], αh ) = χ [X /G], .
λ(N ∨ )
44 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
There is a slicker way to write this formula as a single integral over the inertia stack
I[X/G] if you care to unwind the statement. Let
I(X, G) = {(x, h) : hx = x} ⊂ X × G,
with the G-action given by g · (x, h) = (gx, ghg −1 ). The inertia stack is a global
quotient in this case: I[X/G] = [I(X, G)/G]. Note that I(X, G) admits a finite G-
equivariant decomposition into connected components indexed by h ∈ supp K0 [X/G],
so one can define a global twist operator tw.
Example. Consider the stack P(4, 6) = (C2 r {0})/C× , acting with weights (4, 6).
To compute K0 , we use
K0 [{0}/C× ] → K0 [C2 /C× ] → K0 (P(4, 6)) → 0.
The first two terms are both isomorphic to Z[t, t−1 ], since all vector bundles on C2
are trivial. The pushforward is multiplication by
λ(T0∨ ) = (1 − t−4 )(1 − t−6 ).
Hence, K0 (P(4, 6)) ' Z[t, t−1 ]/(t4 − 1)(t6 − 1). As a sheaf over Spec Z[t, t−1 ] it is
supported at ±1, ±i, η = e±πi/3 , ω = e±2πi/3 . Using the formula above,
1 (−1)k ik + (−i)k ω k + ω −k + η k + η −k
χ(P(4, 6), tk ) = (k + 5) + (k + 5) + + .
24 24 8 12
This recovers the dimension formula for spaces of classical modular forms because
P(4, 6) is isomorphic to the moduli space of elliptic curves, M 1,1 . In this case,
χ(tk ) = h0 (tk ) because the higher cohomologies vanish.
0
k odd
dim Mk (SL2 (Z)) = bk/12c + 1 k 6≡ 2(12)
bk/12c k ≡ 2(12).
46 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
Our first application of principal parts will be counting singular elements in linear
series. A pencil of plane curves is given by a line P1 ' hF, Gi ⊂ PH 0 (OP2 (d)) ' PN .
A general pencil will contain curves with at worst ordinary double points, so we
consider P 1 (OP2 (d)), a bundle of rank 3. The global sections F, G of OP2 (d) induce
sections τF , τG of P 1 (OP1 (d)), and we want to count the points where the latter
two sections become linearly dependent. This is given by c2 :
deg c2 P 1 (OP2 (d) = 3(d − 1)2 .
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 47
Proposition. The total Chern class for the bundle P k (OPn (d)) is given by
n+k
c P k (OPn (d)) = (1 − (d − k)h)( n )
Proof. The Euler sequence for Pn tensored with OPn (d) reads:
0 → ΩPn (d) → OPn (d − 1)⊕n+1 → OPn (d) → 0.
The middle term is not isomorphic to P 1 (OPn (d)), but they have the same Chern
classes! This handles the case k = 1. We leave the general case as an exercise.
Proof. Using the Whitney sum formula, c(P 1 (L)) = c(L)c(L ⊗ ΩX ). The degree
n part of the product becomes
cn P 1 (L) = cn (L ⊗ ΩX ) + c1 (L)cn−1 (L ⊗ ΩX )
n
X n−1
X
= c1 (L)i cn−i (ΩX ) + c1 (L) (i + 1)c1 (L)i cn−1−i (ΩX )
i=0 i=0
n
X
= (i + 1)c1 (L)i cn−i (ΩX ).
i=0
Next, we can ask for the degree of the locus of plane curves with triple points,
which has codimension 4 in PN . Consider P 2 (OP2 (d)), a bundle of rank 6. Take a
general hF0 , F1 , F2 , F3 , F4 i ' P4 ⊂ PN , and consider the associated sections τFi of
P 2 (OP2 (d)). The locus where they become linearly dependent has size:
deg c2 P 2 (OP2 (d)) = 15(d2 − 4d + 4).
For example, the locus of asterisks in the space of plane cubics (P9 ) has degree 15.
Taking this argument to the extreme, the locus of cones inside the space PN of
degree d hypersurfaces in Pn has degree:
n+d−1
d−1
n
deg cn P (OP (d)) =
n .
n
PS of the tautological sub-bundle on G(2, 4), and let L be the pull-back of OP3 (d)
via U → P3 . We form the bundle of relative principal parts using ∆U ⊂ U ×G U :
k
(L) = pr2∗ pr1∗ (L) ⊗ OU ×G U /I∆
k+1
PU/G .
Its fiber at a point (M, p) ∈ U is given by
k
PU/G (L)(M,p) = {germs of sections of L|M at p}/{those vanishing to order k + 1}.
These bundles fit into the same short exact sequences as before, only with the
cotangent sheaf ΩU replaced by its relative version, ΩU/G . The same relative con-
struction works for any smooth proper morphism X → Y .
Question. What is the maximum number M (d) of lines that can appear on a
smooth surface Sd ⊂ P3 with d ≥ 4?
For reasons of Kodaira dimension, the number of lines must be finite, and indeed
for general Sd there are no lines at all. However, the Fermat surface
V (wd + xd + y d + z d = 0) ⊂ P3
has 3d2 lines on it (exercise). This gives a uniform lower bound on M (d), although
it can be improved for certain d. To produce an upper bound, look at the curve Γ
in U of lines meeting Sd at a point with multiplicity ≥ 4. It has class
Y 3
3
c4 PU/G (L) = ((d − 2i)ζ + iσ1 ) ∈ A4 (PS).
i=0
Except for M0 and M1 , all of these moduli spaces are Deligne-Mumford stacks.
Recall that the coarse space morphism induces an isomorphism on Chow groups
A∗ (M)Q ' A∗ (M )Q . The integral Chow groups of a moduli stack are typically
hard to compute; here is a sample of what is known.
Theorem. (Larson)
A∗ (M2 ) ' Z[λ1 , λ2 , δ1 ]/(24λ21 − 48λ2 , 20λ1 λ2 − 4δ1 λ2 , δ13 + δ12 λ1 , 2δ12 + 2δ1 λ1 ).
The first result that we will prove is Mumford’s relation in A1 (Mg )Q :
The LHS is given by (g − 1)[B] + λ. The inside of the expression on the RHS has
codimension 2 part given by:
c1 (ω)2
(1) + c1 (ω) · Td1 (S/B) + Td2 (S/B)
2
50 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
Let us describe A1Q = PicQ for some moduli spaces. First, I claim that the Picard
group is discrete in each of the standard examples. By the localization sequence
for Chow groups, we have
A1 (∆) → A1 (M) → A1 (M) → 0,
so it suffices to show that Pic(M) ' H 1 (M, O× ) is discrete. By the exponential
long exact sequence
c
· · · → H 1 (M, O) → H 1 (M, O× ) −→
1
H 2 (M, Z) → . . . ,
it suffices to prove that the irregularity H 1 (M, O) = 0. By Hodge theory, it suffices
to show that H 1 (M, C) = 0. To see this, note that over C each of the examples
above is the quotient of a contractible space by a discrete group.
• M1,1 (C) ' BG where G is SL2 (Z).
• Mg (C) ' BG where G is the mapping class group MCGg .
• Ag (C) ' BG where G is the arithmetic group Sp2g (Z).
• Kg (C) ' BG where G is the arithmetic group O0 (Λ2g−2 ).
Each of these groups G has finite abelianization (MCGg is perfect for g ≥ 3, and
MCGab 2 = Z/10). Passing from M to M only makes π1 smaller, so H1 (M, C) = 0.
To build a moduli space, we need to embed all the varieties X into the same
ambient space P , consider the appropriate Hilbert scheme, and then take the quo-
tient by the automorphism group of P . If every aut of X is induced by an aut of
the ambient space, then we have successfully constructed a moduli space. Canon-
ical embeddings are nice because any automorphism of X preserves the canonical
bundle. Since not every canonical map is an embedding, pluricanonical maps are
used to prove existence, but the canonical map is easier to study (see below).
The hyperelliptic, theta-null, and trigonal loci can be used to produce an affine
stratication of Mg for 2 ≤ g ≤ 5. This gives a Q-basis for all the Chow groups.
To complete the description of Pic(Ag )Q for g ≤ 3, note that A1 ' M1,1 , A2 ' M2ct
and A3 ' M3ct , so they have Picard ranks 0, 1, and 2, respectively.
Looijenga proved (1995) that for all Mg , a polynomial in the kappa classes κi
of total degree g − 1 vanishes in A∗ (Mg ). This is consistent with S. Diaz’s theorem
(1984) that any complete subvariety of Mg has dimension ≤ g − 2.
Is there a closed manifold with R∗ (Mg ) as its even cohomology? Wide open.
isomorphic to Σ glued to a rational nodal tail, a curve whose moduli point lies in
∆0 ∩ ∆1 . To construct the stable family explicitly, consider
S = Bl(q,q) (Σ × Σ) → Σ,
with the proper transforms of ∆Σ and Σ × {q} identified, to produce a family of
nodal curves over Σ ' C0 . The normal bundles of the two section curves have
degrees 2 − 2(g − 1) − 1 and −1, respectively, which gives deg C0 · ∆0 = 2 − 2g.
Theorem. (O’Grady) For any N , there exists g such that dim Pic(Kg )Q > N .
Proof. Recall that points of Kg are pairs (S, L) where S is a K3 surface, meaning
that H 1 (OS ) = 0 and KS ' OS , and L ∈ Pic(S) is primitive and ample, with
c1 (L)2 = 2g − 2. To construct examples of polarized K3 surfaces, take the double
cover of a surface T branched along a smooth curve B ∈ | − 2KT |. If T = P1 × P1 ,
then ϕ : S → T is branched along curve of bidegree (4, 4). For any integers a, b > 0,
L = ϕ∗ (ah1 + bh2 )
is ample with c1 (L)2 = 4ab, and it is primitive as long as (a, b) = 1. Hence, for
each pair a, b of coprime positive integer we have a divisor Da,b ⊂ K1+2ab . Indeed,
counting dimensions, h0 (T, 4h1 + 4h2 ) = 25 so we have dim P24 − dim Aut(T ) = 18.
For each Da,b , we will construct a test curve Ca,b ⊂ Da,b such that:
• deg Ca,b · Da,b < 0,
• deg Ca,b · Da,b = 0 if min{c, d} > 2 min{a, b} or min{c, d} < 12 min{a, b}.
The second property comes from the fact that Ca,b and Dc,d are disjoint under the
given inequalities. For g = 1 + 2p1 p2 p3 . . . p2N (with pi distinct primes), we can
produce a diagonal intersection matrix (of size N ) with nonzero determinant.
By Noether-Lefschetz theory (next time), the normal bundle NDa,b /Kg is isomor-
phic to the restriction of the dual Hodge line bundle to Da,b . The Hodge bundle
is ample. For disjointness, assume WLOG that a = min{a, b}, and suppose that
we have found an automorphism f of a K3 surface S such that f ∗ ϕ∗ (ch1 + dh2 ) '
ϕ∗ (ah1 + bh2 ). Replacing ϕ∗ with ∼ for readability, we have
f∗ (h̃2 ) · (ch̃1 + dh̃2 ) = h̃2 (ah̃1 + bh̃2 ) = 2a
On the other hand, using the fact that h2 is nef, f∗ (h̃2 )·(ch̃1 +dh̃2 ) ≥ min{c, d} > 2a,
contradiction. Swapping the roles of {a, b} with {c, d} gives a contradiction for the
second inequality too.
54 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
There is a natural Hodge line bundle λ on D, whose fiber is H 2,0 , and it descends
to the quotient D/Γ. Baily-Borel proved that some power of λ has enough sections
to embed D/Γ into projective space (λ is ample). The closure of D/Γ in projective
space is a singular projective variety, often denoted (D/Γ)BB .
classes in A1 (K
e g ) as coefficients:
X X
f (q) = −λ + [x⊥ ]q n .
n∈Q>0 x∈Λ∨
g /Γ
(x,x)=−2n
Although there are infinitely many divisors [x⊥ ], their span is finite-dimensional
in A1 (K e g )∨ , we have
e g )Q . Indeed, for each element of µ ∈ A1 (K
Q
Remark. The growth of dim Pic(Mg )Q came from the boundary divisors, which
were related to lower genus curves. The growth of dim Pic(Kg )Q comes from the
NL-divisors, which are themselves locally symmetric spaces for the group O(2, 18).
Corollary. (BKPS) Any complete family with constant Picard number is isotrivial.
Proof. The crucial fact is ampleness of the Hodge line bundle λ (Baily-Borel).
It suffices to express aλ as a combination of the NL-divisors x⊥ i . Let θ(q) be the
Siegel theta function of weight 1/2, and let E10 (q) be the Eisenstein series of weight
10. If f (q) ∈ Mod(21/2, 2g − 2) with constant term c0 , then
f (q) − c0 θ(q)E10 (q) ∈ Mod(21/2, 2g − 2)
is a cusp form. The coefficients of a cusp form grow like O(nwt/2+ ) = O(n11/4+ ).
On the other hand, the coefficients of θ(q)E10 (q) are bounded from below by O(n9 ).
These estimates imply that the coefficients of f (q) are eventually nonzero. Next
consider the composition:
α
M
Qei → Mod(21/2, 2g − 2)∨ → Pic(Kg )Q ,
i
2The Borcherds result is slightly more refined; you get a vector-valued modular form for the
Weil representation on C[Λ∨
g /Λg ]. This is why we use the trace formula to compute the dimension.
56 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
To set this up, let G be a semi-simple R-group, and let K be a maximal compact
subgroup such that D = G/K has a complex structure. The standard examples
are SL(2, R)/SO(2), more generally Sp(2n, R)/U (n), O(2, n)/SO(2) × SO(n), and
U (1, n)/U (1) × U (n). The compact dual of D is a flag variety Y = GC /P , where
P = KC ·P+ is a parabolic subgroup with unipotent radical P+ such that P ∩G = K.
Our main application of the principle will be to Ag , the moduli space of ppav’s,
and its toroidal compactifications. Here D = Hg is the Siegel upper-half space, and
the compact dual is Yg = Sp(2g, C)/P is the Lagrangian Grassmannian of (C2g , ω).
To compute the cohomology of Yg , fix a flag of isotropic subspaces:
0 ⊂ Z1 ⊂ Z2 ⊂ · · · ⊂ Zg ⊂ C2g ,
⊥
and extend it by taking Zg+i = Zg−i . Given a partition µ satisfying µi − 1 ≤
µi+1 ≤ µi fitting inside the g × g box (there are 2g such partitions), we define
Σµ (Z• ) = {L ∈ Yg : dim(L ∩ Zg+i−µi ) ≥ i} ⊂ Yg .
These are the only Schubert cycles in G(g, 2g) which lie in Yg because of the fact
that dim(L ∩ Zi⊥ ) = dim(L ∩ Zi ) + g − i.
Remark. A more concise way of writing the relation is ch2k (S) = 0 for k ≥ 1.
The restriction of the bundle S from Yg to the open subset Hg coincides with the
Hodge bundle E = π∗ (Ω1Xg /Ag ) pulled back from Ag . It is the automorphic bundle
associated to the standard representation of U (n). We define the tautological ring
R∗ (Ag ) ⊂ A∗ (Ag )Q to be generated by the Chern classes of the Hodge bundle.
Σ
Since E extends over any toroidal compactification Ag , we define the tautological
ring of smooth compactifications in the same way. The Proportionality Principle
implies (sending ui 7→ λi ) that
Σ
R∗ Ag ' A∗ (Yg ),
but the degrees of 0-cycles differ by the constant multiple:
g
1 Y
c(Γ) = (−1)g(g+1)/2 g ζ(1 − 2k).
2
k=1
This is the reciprocal of a large integer (1/24 for g = 1 is familiar).
The proof uses a technique called theta lifting, which is a way of passing between
automorphic forms for symplectic and orthogonal groups. The key fact is that if
W is a symplectic vector space and V is an orthogonal vector space, then W ⊗ V is
naturally symplectic, and we have Sp(W ) × O(V ) ⊂ Sp(W ⊗ V ). The Siegel theta
function on Sp(W ⊗ V ) restricts to a function Θ, which serves as a correspondence
between automorphic forms for the two groups:
Sp(W ) ← Sp(W ) × O(V ) → O(V ).
A second interaction between Kg and Ag stems from the exceptional isomorphism
SO(2, 3)+ ' Sp(4, R).
5We have only given the statement for K with coefficients in Chow groups, but Kudla-Millson
g
prove a statement for all O(p, q) symmetric spaces with coefficients in cohomology.
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 59
This implies that higher NL-loci of dimension 3 are Shimura varieties for Sp(4, R).
Using the structure theorem for R∗ (A2 ) from last time, we have
λ21 = 2λ2 = 0 ∈ R2 (A2 ).
Theorem. (v.d.Geer-Katsura) If Γ\D is a Shimura variety for O(2, n), then
λn−1 = 0 ∈ A∗ (Γ\D)Q
In particular, any complete subvariety of Kg has dimension ≤ 17.
MO 0 /N
O
M 0 ×N N 0 / N0 / L,
where the vertical arrows are monos and the horizontal arrows are epis.
Note: 0 is the final object of C, but not of QC because you lose uniqueness.
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 61
Proof. By the long exact sequence of homotopy groups, it suffices to show that
the homotopy fiber of BQB → BQC is contractible. Recall that the homotopy
fiber of a continuous map of pointed spaces f : B → C is the space of pairs
(b ∈ B, γ : f (b) → 0). To avoid alternating arrow orientations, Quillen proves that
a certain homotopy fiber category is contractible for any choice of basepoint.
For a scheme X, let Mr (X) ⊂ Coh(X) be the Serre subcategory of coherent sheaves
whose support has codimension ≥ r. Then the localization Mr (X)/Mr+1 (X) has
a dévissage over all points x ∈ X of codimension r:
M M
··· → Kn+1 (F (x)) → Kn (Mr+1 (X)) → Kn (Mr (X)) → Kn (F (x)) → . . .
x∈X (r) x∈X (r)
Here we have fixed the codimension r. If instead we use the full filtration M• (X)
of Coh(X), the long exact sequences above combine to produce an exact couple
whose spectral sequence6 relates the K-theory of functions fields with the K-theory
of Coh(X), filtered by K-theory of Mr (X).
M
E1pq = K−p−q (F (x)) ⇒ K−q (X).
x∈X (p)
Proof. The vanishing of the differential implies that the E2 = E∞ in the spectral
sequence, and furthermore the coniveau filtration on Kn (Coh(X)) is the stupid one.
This means that the only nonvanishing terms in E2 are E20q ' Kq (X). This is the
0-th cohomology of the E1 -page complex which is otherwise exact.
Theorem. (Bloch’s Formula) Given a smooth variety X, consider the sheaf Kn (X)
of abelian groups associated to the presheaf U 7→ Kn (O(U )). The classical Chow
groups are computed by sheaf cohomology:
An (X, 0) = H n (X, Kn (X)).
6The construction of the spectral sequence is very analogous to the Serre spectral sequence.
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 63
The algebraic K-theory of fields is far simpler than the general case: K0 (F ) ' Z
and K1 (F ) ' F × . So the map in question goes from
M
F (x)× → Z n (X),
x∈X (n−1)
and is actually the map div from Lecture 3. Quillen reduces to the case of a DVR,
where K1 (F ) → K0 (R/m) is the ord function.
The Steinberg group is the universal central extension of E(R) ⊂ GL(R), and its
center is K2 (R) = H2 (E(R), Z): 0 → K2 (R) → St(R) → E(R) → 1.
a
More explicitly, St(R) is the free group on Xij , where i, j ∈ N and a ∈ R, modulo
the relations satisfied by elementary matrices:
a b a+b
Xij Xij = Xij ;
a b ab
[Xij , Xjk ] = Xik , i 6= k;
a b
[Xij , Xkl ] = 1, i 6= l, j 6= k.
The kernel of St(R) → E(R) is a quotient of R∗ ⊗ R∗ , by the Steinberg relations:
{a, b} = {b, a}−1 ; {a, b}{a0 , b} = {aa0 , b}; {a, 1 − a} = 1.
For X/k a regular scheme with function field F , there is a well-defined pairing
d log ∧ d log : F × ⊗ F × → Ω2F/k .
It satisfies the Steinberg relations, so it descends to K2 (F ). When restricted to
K2 (X) via the Gersten resolution, it lifts to a map K2 (X) → Ω2X/k . Bloch proves
that the induced map
H 2 (X, K2 (X)) → H 2 (X, Ω2X/k )
is the cycle class map via his isomorphism H 2 (X, K2 (X)) ' A2 (X).
If ρ is injective, ρ̃ is called a face map (facet if m = n − 1). The higher cycle groups
Z(X, n) ⊂ Z(X × ∆n ) are the subgroups generated by cycles which intersect each
face X ×∆m dimensionally transversely. If ∂i denotes the pullback to the i-th facet,
we can define an algebraic boundary map
Xn
∂= (−1)i ∂i : Z(X, n) → Z(X, n − 1).
i=0
Definition. The higher Chow groups A(X, n) are defined by taking the homology:
Z(X, n) = Hn (Z(X, ∗))
Bloch verifies the functoriality properties (proper pushforward, flat pullback) and
the property that A∗ (X, n) ' A∗ (X × A1 , n) directly from the definitions.
Corollary. For any vector bundle E → X, the flat pullback A∗ (X, n) → A∗ (E, n)
is an isomorphism.
There is a higher Chern character map ch(n) : Kn (X) → A∗ (X, n) due to Gillet.
It is completely formal, and uses the Quillen +-construction of algebraic K-theory,
adapted to the world of simplicial schemes. Multiplying by the Todd class of X,
τ (n) : Kn (X) → A∗ (X, n) satisfies a higher Grothendieck-Riemann-Roch for proper
morphisms X → Y , and gives an isomorphism over Q.
66 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
Remark. It is important to use cycles over Q in the definition for the Chow
groups; over C they are often infinite rank.
AJ R(2q,q) =
0 / Jac(X) / H 2q (X, Z(q)) / Hdg(q) / 0.
D
INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 67
The Bloch-Kato conjecture gives the leading coefficient of LX (s) at the special
points. It is too complicated to state here, but in the case where X = OF for a
number field F it recovers the class number formula:
2r1 · (2π)r2 · RF · hF
lim (s − 1)ζF (s) = √ ,
s→1 ωF DF
as well as the Dirichlet unit theorem:
rk OF× = r1 + r2 − 1.
68 INTERSECTION THEORY IN ALGEBRAIC GEOMETRY
References
[1] S. Bloch. K2 and algebraic cycles. Ann. Math., 99:349–379, 1974.
[2] S. Bloch. Algebraic cycles and higher K-theory. Adv. Math., 61:267–304, 1986.
[3] D. Edidin. Riemann-Roch for Deligne-Mumford stacks. In B. Hassett, J. McKernan, J. Starr,
and R. Vakil, editors, A Celebration of Algebraic Geometry, volume 18, pages 241–266. Clay
Mathematics Proceedings, 2013.
[4] D. Edidin and W. Graham. Equivariant intersection theory. Invent. Math., 131:595–634, 1998.
[5] D. Edidin and W. Graham. Riemann-Roch for equivariant Chow groups. Duke Math. J.,
102(3):567–594, 2000.
[6] D. Eisenbud and J. Harris. 3264 and All That: A Second Course in Algebraic Geometry.
Cambridge University Press, 2016.
[7] C. Fontanari and E. Looijenga. A perfect stratification of Mg for g ≤ 5. Geom. Dedicata,
136:133–143, 2008.
[8] W. Fulton. Intersection theory. Springer-Verlag, 1984.
[9] H. Gillet. Riemann-Roch theorems for higher K-theory. Adv. Math., 40:203–289, 1981.
[10] J. Harris and I. Morrison. Moduli of curves. Springer-Verlag, 1998.
[11] A. Kresch. Cycle groups for Artin stacks. Invent. Math., 138:495–536, 1999.
[12] S. Kudla and J. Millson. Intersection numbers of cycles on locally symmetric spaces and
Fourier coefficients of holomorphic modular forms in several complex variables. Publ. Math.
IHÉS, 71:121–172, 1990.
[13] D. Maulik. Supersingular K3 surfaces for large primes. Duke Math. J., 163(13):2357–2425,
2014.
[14] D. Mumford. Hirzebruch’s proportionality theorem in the non-compact case. Invent. Math.,
42:239–272, 1977.
[15] H. Nielsen. Diagonalizably linearized coherent sheaves. Bull. Soc. Math. Fr., 102:85–97, 1974.
[16] K. O’Grady. On the Picard group of the moduli space for K3 surfaces. Duke Math. J.,
53(1):117–124, 1986.
[17] D. Quillen. Higher algebraic K-theory I. In H. Bass, editor, Higher K-theories, pages 85–147.
Springer, 1972.
[18] R.W. Thomason. Une formula de Lefschetz en K-théorie equivariante algébrique. Duke Math.
J., 68:447–462, 1992.
[19] G. van der Geer. The cohomology of the moduli space of abelian varieties. In The Handbook
of Moduli, Advanced Lectures in Mathematics, pages 415–458. Int. Press, 2011.
[20] G. van der Geer and T. Katsura. Note on tautological classes of moduli of K3 surfaces. Mosc.
Math. J., 5(4):775–779, 2005.
[21] A. Vistoli. Intersection theory on algebraic stacks and on their moduli spaces. Invent. Math.,
97:613–670, 1989.