Hilbert Schemes of Dimensional Lie Algebras: 4Smrxwerh-R Rmxi
Hilbert Schemes of Dimensional Lie Algebras: 4Smrxwerh-R Rmxi
Surveys
and
1SRSKVETLW
Volume 228
Hilbert Schemes of
4SMRXWERH-R½RMXI
Dimensional Lie Algebras
Zhenbo Qin
Hilbert Schemes of
Points and Infinite
Dimensional Lie Algebras
Mathematical
Surveys
and
Monographs
Volume 228
Hilbert Schemes of
Points and Infinite
Dimensional Lie Algebras
Zhenbo Qin
EDITORIAL COMMITTEE
Robert Guralnick Benjamin Sudakov
Michael A. Singer, Chair Constantin Teleman
Michael I. Weinstein
The author was supported in part by a Collaboration Grant for Mathematicians (Award
Number: 268702) from the Simons Foundation and by a Research Council Grant (Award
Number: URC-17-084-n) from the University of Missouri.
Copying and reprinting. Individual readers of this publication, and nonprofit libraries acting
for them, are permitted to make fair use of the material, such as to copy select pages for use
in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Permissions to reuse
portions of AMS publication content are handled by Copyright Clearance Center’s RightsLink
service. For more information, please visit: http://www.ams.org/rightslink.
Send requests for translation rights and licensed reprints to [email protected].
Excluded from these provisions is material for which the author holds copyright. In such cases,
requests for permission to reuse or reprint material should be addressed directly to the author(s).
Copyright ownership is indicated on the copyright page, or on the lower right-hand corner of the
first page of each article within proceedings volumes.
2018
c by the American Mathematical Society. All rights reserved.
Printed in the United States of America.
∞ The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at http://www.ams.org/
10 9 8 7 6 5 4 3 2 1 23 22 21 20 19 18
Contents
Preface ix
Since the pioneering work of Grothendieck [Grot], Hilbert schemes, which pa-
rametrize subschemes in algebraic varieties, have been studied extensively as evi-
denced in the survey [Iar2]. Viewed also as the simplest moduli spaces of sheaves,
they are fundamental objects in algebraic geometry, play essential roles in various
important enumerative problems, and provide testing grounds for many interesting
conjectures in algebraic geometry and its interplay with theoretical physics. Apart
from the Hilbert schemes of curves on 3-folds over which the Donaldson-Thomas
theory is defined and investigated [MNOP1, MNOP2], much attention has been
concentrated on Hilbert schemes of points, i.e., those Hilbert schemes parametriz-
ing 0-dimensional closed subschemes. When the dimension of the variety is at least
three, the corresponding Hilbert schemes of points are in general singular. When
the variety is a smooth curve, the corresponding Hilbert schemes of points coincide
with the symmetric products of the curve. When X is a smooth algebraic surface,
the Hilbert scheme X [n] parametrizing length-n 0-dimensional closed subschemes
of X is irreducible and smooth [Bri, Fog1, Iar1]. Moreover, the Hilbert-Chow
morphism ρn from X [n] to the symmetric product X (n) , which sends an element
in X [n] to its support (counted with multiplicities) in X (n) , is a crepant desin-
gularization of X (n) . Many fundamental aspects of X [n] such as its cohomology
groups, Chow groups, motive, cobordism class, and relations with algebraic com-
binatorics, the McKay correspondence and integrable systems have been analyzed
[CoG, dCM, EGL, ES1, Got1, Hai1, Hai2, Mar1].
In the seminal papers [Groj, Nak3] which were motivated by [Nak1, Nak2,
Nak4] regarding the construction of representations of affine Lie algebras on the
homology groups of the moduli spaces of instantons on ALE spaces (equivalently,
on the homology groups of quiver varieties), Grojnowski and Nakajima geomet-
rically constructed Heisenberg algebra actions on the cohomology of the Hilbert
schemes X [n] , where X denotes a smooth algebraic surface. Their geometric con-
structions started a whole new chapter investigating interplays between the Hilbert
schemes of points and infinite dimensional Lie algebras. Subsequently, using the
Hilbert schemes X [n] , Lehn [Leh1] geometrically constructed the Virasoro algebras
and the boundary operator, W.-P. Li, W. Wang and the author [LQW1, LQW4]
constructed the Chern character operators and the W algebras, and Carlsson and
Okounkov [Car1, Car2, CO] constructed the Ext vertex operators. As noted in
[FW, Introduction] and [Mat, Section 5], Lehn’s boundary operator is a version
of the bosonized Calogero-Sutherland operator. The Chern character operators are
vertex operators of higher conformal weights, and the W algebras are higher-spin
generalizations of Lehn’s Virasoro algebras. The Ext vertex operators of Carlsson
and Okounkov are motivated by the study of Nekrasov partition functions (which
ix
x PREFACE
arise from supersymmetric quantum gauge theory) in the setting of Hilbert schemes
of points.
All of these operators and algebras are powerful tools in understanding the
finer geometric properties of the Hilbert schemes X [n] . Indeed, the Heisenberg
algebras serve as a fundamental language in describing the homology and cohomol-
ogy classes of X [n] , which is crucial in studying the Gromov-Witten theory of X [n]
[Cheo2, ELQ, HLQ, LL, LQ1, LQ5, OP3, OP4]. The Virasoro operators and the
Chern character operators determine the cup products and ring structures of the
cohomology of X [n] [CoG, LQW1, LQW2, LQW3, LQW5, LS1, LS2]. The Ext
vertex operators provide a very useful approach in understanding the intersection
theory of X [n] when the tangent bundle of X [n] is involved [Car1, Car2, QY].
Two excellent books on Hilbert schemes of points exist. The first one is
Göttsche’s book [Got2] which includes fundamental facts about X [n] , the Betti
numbers of X [n] when X is a surface, and the computation of the homology and
Chow rings of Hilbert schemes. The second is Nakajima’s book [Nak5] which
contains his Heisenberg algebra constructions, symmetric products (in the Hilbert
scheme X [n] ) of an embedded curve in a surface X, and interactions with singu-
larities, symplectic geometry and the ring of symmetric functions. We also refer to
[EG, Leh2] for surveys on Hilbert schemes of points, the Heisenberg algebras and
the Virasoro algebras, and to [Nak7] for a survey on the equivariant cohomology
of the Hilbert schemes, the Heisenberg algebras, the Virasoro algebras and some
interesting applications to algebraic combinatorics.
The purpose of this book is to present a detailed survey of the developments,
in the context of interactions between Hilbert schemes of points and infinite dimen-
sional Lie algebras, that appeared after the book [Nak5]. This book contains 5
parts consisting of 16 chapters. Part 1 deals with the basics of the Hilbert schemes
of points and some geometry of the Hilbert schemes of points on the projective
plane. Part 2 is devoted to the constructions of various infinite dimensional Lie al-
gebra actions on the cohomology of the Hilbert schemes of points on surfaces, and
to the connections with multiple q-zeta values. It includes Nakajima’s affine Lie
algebra actions on the homology of quiver varieties as a motivation and background
material, the Heisenberg algebras of Grojnowski and Nakajima, the boundary op-
erator and the Virasoro algebras of Lehn, the Ext vertex operators of Carlsson and
Okounkov, and the Chern character operators and the W algebras of W.-P. Li,
W. Wang and the author. Part 3 studies the cohomology ring structure of these
Hilbert schemes, such as the ring generators and the ideals, the approach of Lehn
and Sorger when the canonical divisor of the surface is numerically trivial, the ap-
proach of Costello and Grojnowski in terms of the Calogero-Sutherland operators
and the Dunkl-Cherednik operators, the integral basis for the cohomology group,
and the orbifold cohomology ring of the symmetric product of a surface. Part 4
is about the equivariant cohomology of the Hilbert schemes via Jack polynomials,
the Hilbert/Gromov-Witten correspondence between the equivariant cohomology
of the Hilbert schemes of points on the affine plane and the Gromov-Witten theory
of curves, and applications to the Hurwitz numbers of the projective line. Part 5
includes the cosection localization technique of Kiem and J. Li, the Gromov-Witten
PREFACE xi
theory of the Hilbert schemes of points, the equivariant quantum corrected bound-
ary operator and the quantum differential equations of Okounkov and Pandhari-
pande, the quantum corrected boundary operator of J. Li and W.-P. Li, and the
proof of Ruan’s Cohomological Crepant Resolution Conjecture.
This book is suitable for graduate students and researchers in algebraic ge-
ometry, representation theory, algebraic combinatorics, topology, number theory
and theoretical physics. Furthermore, a semester of an advanced course on Hilbert
schemes of points and infinite dimensional Lie algebras may be organized from se-
lected chapters in this book, e.g., Chapter 1, Chapter 3, Chapter 4, Chapter 7,
Chapter 8, Chapter 11 and Chapter 12.
The author would like to thank all of his collaborators on works related to
the Hilbert schemes of points, without whom this book would be impossible: Dan
Edidin, Jianxun Hu, Wei-Ping Li, Yuping Tu, Weiqiang Wang, Fei Yu and Qi
Zhang. In addition, thanks are due to Lie Fu, Wei-Ping Li, Hiraku Nakajima, Boiar
Qin, Wenzer Qin, Weiqiang Wang and the two anonymous referees for carefully
reading the manuscript and providing useful comments which have greatly improved
the exposition of the book.
The author also thanks Sergei Gelfand, Christine M. Thivierge and the Editorial
Committee of the American Mathematical Society for their editorial guidance.
Zhenbo Qin
Part 1
Hilbert schemes of
points on surfaces
CHAPTER 1
This chapter collects basic definitions and results regarding partitions, the ring
of symmetric functions, symmetric products, Hilbert schemes of points, and inci-
dence Hilbert schemes. Details may be found in Chapter I of [Mac2] and Chapters 1
and 2 of [Got2]. The focuses of this chapter are various symmetric functions and
the Hilbert schemes of points on a smooth projective complex surface. Throughout
the book, all the homology and cohomology are with C-coefficients, unless otherwise
specified.
1.1. Partitions
Let n ≥ 1. A partition λ of n, denoted by λ n, is an infinite sequence
λ = (λ1 ≥ λ2 ≥ λ3 ≥ . . .)
of nonnegative integers λi1 ,...,id−1 indexed by the tuples (i1 , . . . , id−1 ) ∈ (Z≥0 )d−1
such that
λi1 ,...,id−1 = n,
i1 ,...,id−1
One sees immediately that the generating function for P2 (n) is given by
+∞
+∞
1
(1.2) P2 (n)q n =
n=0 n=1
1 − qn
+∞
+∞
1
(1.3) P3 (n)q n = .
n=0 n=1
(1 − q n )n
contains 18 cells. For the cell ♠ ∈ Dλ , we have (♠) = 1, (♠) = 2, a(♠) = 3,
a (♠) = 1, h(♠) = 5, and c(♠) = −1.
that part i ∈ Z has multiplicity mi , mi = 0 for only finitely many i’s, and
such
n = i imi . Define
(λ) = mi ,
i
|λ| = imi ,
i
s(λ) = i2 mi ,
i
!
λ = mi !.
i
1 x2 · · · .
xα = xα1 α2
6 1. BASIC RESULTS ON HILBERT SCHEMES OF POINTS
where α runs over all distinct permutations of the sequence (λ1 , λ2 , . . .).
For each integer r ≥ 1, the r-th power sum is defined to be
+∞
pr = xri .
i=1
where the summation is over certain partitions μ of |λ| + i, which are obtained as
follows:
(i) add i to a part in λ, say λk (possibly 0), and then
(ii) arrange the parts in descending order.
The coefficient aλ,μ is equal to the number of elements in { | μ = λk + i}.
To define the Schur functions, we will work in Z[x1 , . . . , xn ] for the moment.
For a sequence α = (α1 , . . . , αn ) of nonnegative integers, let
α
aα = det(xi j )1≤i,j≤n .
When δ = (n − 1, n − 2, . . . , 1, 0), aδ is the Vandermonde determinant and equal to
i<j (xi − xj ). Note that aα = 0 if two of the integers αi are equal. So up to sign,
we may assume that α1 > . . . > αn ≥ 0. For 1 ≤ i ≤ n, define λi by αi = λi + n − i.
Put λ = (λ1 , . . . , λn ). Then λ is a partition of length at most n, and α = λ + δ.
The determinant aα = aλ+δ is divisible by aδ in the polynomial ring Z[x1 , . . . , xn ],
and
sλ (x1 , . . . , xn ) = aλ+δ /aδ
is a symmetric homogeneous polynomial of degree |λ|. Moreover,
sλ (x1 , . . . , xn , 0) = sλ (x1 , . . . , xn ).
Therefore, there exists a unique element
sλ ∈ Λ
that reduces to sλ (x1 , . . . , xn ) when xn+1 , xn+2 , . . . are all set equal to zero, for
every n ≥ (λ). This symmetric function sλ is the Schur function corresponding
to λ.
1.3. SYMMETRIC PRODUCTS 7
Next, we review the Jack symmetric functions. Fix a positive real number α.
Let Q(α) be the subfield of R generated by α over Q. Let ·, · α be the inner product
on Λ ⊗Z Q(α) defined by declaring
pμ , pν α = α(μ) zμ δμ,ν .
(α)
For a fixed positive real number α, the Jack symmetric functions Pλ form a
basis of Λ ⊗Z Q(α) indexed by partitions λ and are characterized by the following
properties (1.6) and (1.7):
(α)
(α)
(1.6) P λ = mλ + uλ,μ mμ
μ<λ
(α)
where uλ,μ are real numbers and < denotes the dominance ordering, and
(α)
(1.7) Pλ , Pμ(α) α =0
if λ = μ.
It is well-known that each of the sets {eλ }λ , {mλ }λ and {sλ }λ is a Z-linear
basis of Λ. In addition, the set
{ei = m(1i ) }i≥1
generates the integral ring Λ. The set {pλ }λ is a Q-linear basis of Λ ⊗Z Q, but not
(α)
a Z-linear basis of Λ. For a fixed positive real number α, Pλ λ is a linear basis
of Λ ⊗Z Q(α).
Finally, there is an involution ω on Λ satisfying
(1.8) ω(pλ ) = (−1)|λ|−(λ) pλ .
The forgotten symmetric function fλ associated to a partition λ is defined by
(1.9) fλ = ω(mλ ).
The set {fλ }λ is another Z-linear basis of Λ.
by the assumptions on π̃. So g ∗ (α̃(f )) = α̃(f ) for every g ∈ G, and α̃(f ) ∈ A[Y ]G .
Finally, if Y is not affine, then there exists an open cover of Y by G-invariant
affine open subsets Yi , and Y /G is the gluing of the quasi-projective varieties Yi /G.
We refer to [Harr, p. 127] for the proof that Y /G is quasi-projective.
Example 1.9. (i) Note that C[x1 , . . . , xn ]Sn = C[ẽ1 , . . . , ẽn ] where
ẽi = ei (x1 , . . . , xn , 0, . . .)
Example 1.10. The symmetric product (A2 )(2) is not smooth. In fact,
C[u, v, w]
(1.11) (A2 )(2) ∼
= A2 × Spec .
uw − v 2
To prove (1.11), let x1 , y1 and x2 , y2 be the coordinates on the two factors of A2 ×A2 .
Let s = x1 − x2 , t = y1 − y2 , s = x1 + x2 and t = y1 + y2 . Denote the transposition
in S2 by τ . Then, τ (s) = −s, τ (t) = −t, τ (s ) = s , and τ (t ) = t . So
Note that τ (si tj ) = (−1)i+j si tj . Thus, C[s, t]S2 is the subalgebra generated by
s2 , st, t2 . Putting u = s2 , v = st and w = t2 , we see that
C[u, v, w]
C[s, t]S2 ∼
= .
uw − v 2
Hence
C[u, v, w]
(A2 )(2) = Spec C[x1 , y1 , x2 , y2 ]S2 ∼
= A2 × Spec .
uw − v 2
10 1. BASIC RESULTS ON HILBERT SCHEMES OF POINTS
Z = (f × IdX )∗ Z.
Furthermore, if Y is an open subscheme of X, then there is a corresponding
open subscheme HilbP P
Y of HilbX parametrizing closed subschemes in Y . In partic-
P
ular, HilbY is defined for a quasi-projective scheme Y .
If [Z] ∈ HilbP
X is a point corresponding to a closed subscheme Z of X, then the
Zariski tangent space of HilbPX at [Z] is canonically isomorphic to
(1.12) Hom(IZ , OZ ).
Since every homomorphism IZ → OZ of OX -modules factors uniquely through
IZ /IZ2 , the Zariski tangent space of HilbP
X at [Z] is also canonically isomorphic to
ξ form a smooth open (but not necessarily dense) subset of X [n] of dimension
dim(X) · n.
(iii) A length-n 0-dimensional closed subscheme of Ad corresponds to an ideal
I ⊂ C[x1 , . . . , xd ] of colength-n, i.e.,
dim C[x1 , . . . , xd ]/I = n.
Therefore, as sets,
[n]
Ad = I ⊂ C[x1 , . . . , xd ]| I is a colength-n ideal of C[x1 , . . . , xd ] .
Theorem 1.14. Let X be a smooth irreducible variety with dim(X) ≤ 2. Then
X [n] is a smooth irreducible variety of dimension dim(X) · n.
Proof. Since X is connected, X [n] is connected as well. In view of Exam-
ple 1.13 (ii), it remains to prove that for every element ξ ∈ X [n] , the dimension of
the Zariski tangent space Tξ,X [n] of X [n] at ξ is at most dim(X) · n. By (1.12), we
have
Tξ,X [n] = Hom(Iξ , Oξ ).
When dim(X) = 1, we obtain Iξ = OX (−ξ) and Hom(Iξ , Oξ ) = ∼ H 0 (X, Oξ ).
So dim Tξ,X [n] = (ξ) = n and we are done.
Let dim(X) = 2. Applying the functor Hom(·, Oξ ) to the exact sequence
0 → Iξ → OX → Oξ → 0
yields an exact sequence
0 → Hom(Oξ , Oξ ) → Hom(OX , Oξ ) → Hom(Iξ , Oξ ) → Ext1 (Oξ , Oξ ).
Since
dim Hom(Oξ , Oξ ) = (ξ) = n = dim Hom(OX , Oξ ),
we have an injection
0 → Hom(Iξ , Oξ ) → Ext1 (Oξ , Oξ ).
So it remains to prove that
(1.14) dim Ext1 (Oξ , Oξ ) = 2n.
We have
dim Ext2 (Oξ , Oξ ) = dim Hom(Oξ , Oξ ⊗ KX ) = n
by the Serre duality and dim Hom(Oξ , Oξ ) = n. By the Riemann-Roch Theorem,
2
(−1)i · dim Exti (Oξ , Oξ ) = χ(Oξ , Oξ ) = 0.
i=0
is either a set of two distinct points or a point together with a tangent direction at
that point. In fact, the Hilbert scheme X [2] can be described globally as follows.
Let X 2 be the blow-up of X 2 along the diagonal. The action of S2 on X 2 extends
to X 2 . Then X [2] = X2 /S2 .
where
Δi,j = {(x1 , . . . , xn ) ∈ X n |xi = xj }.
Let Δ0 be the open subset of Δ where precisely two of the xi ’s coincide. Let p ∈ Δ0 .
Without loss of generality, we may assume that p ∈ Δ1,2 . By Example 1.10, a
formal neighborhood of the point π(p) in X (n) (i.e., the complete local ring of X (n)
at π(p)) is isomorphic to
C[[u, v, w, s , t , x3 , y3 , . . . , xn , yn ]]
Spec .
uw − v 2
Thus all the points in π(Δ0 ) are singular points of X (n) . So π(Δ0 ) is contained in
the singular locus Sing X (n) . Since Sing X (n) is closed, the closure of π(Δ0 ) is
(n)
contained in Sing X (n) . Since the closure of π(Δ0 ) is X (n) \X0 and
(n)
Sing X (n) ⊂ X (n) \X0 ,
(n)
we conclude that Sing X (n) is equal to X (n) \X0 .
The generating function for the Euler characteristic χ X [n] of the Hilbert
scheme X [n] was calculated in [Che1] (the case d = 1 also follows from Proposi-
tion 1.21 (i) and (1.10) by setting t = −1, and the case d = 2 is due to Göttsche
and follows from Theorem 1.23 below by setting t = −1).
14 1. BASIC RESULTS ON HILBERT SCHEMES OF POINTS
where
+∞
1
M (q) = .
n=1
(1 − q n )n
For the finer invariants of X [n] such as its Betti numbers, we have the following
elegant formula [Got1, GS] when X is a surface.
Theorem 1.23. Let X be a smooth quasi-projective surface. Then,
+∞
+∞ 4 (−1)i bi (X)
[n] n 1
Pt (X )q = .
n=0 n=1 i=0
1 − (−t)2n−2+i q n
It will be clear from Chapter 3 that Theorem 1.23 is important in the interplay
between the Hilbert schemes of points and infinite dimensional Lie algebras.
Let Zn ⊂ X [n] × X be the universal codimension-2 closed subscheme:
(1.24) Zn = (ξ, x) ⊂ X [n] × X | x ∈ Supp (ξ) .
Let p : Zn → X [n] and q : Zn → X be the natural morphisms. For a sheaf F on
X, let
(1.25) F [n] = p∗ q ∗ F.
Since p is a degree-n flat morphism, if F is locally free of rank r, then F [n] is locally
free of rank rn; in this case, F [n] is called a tautological bundle over X [n] . We define
the subset Bn of the Hilbert scheme X [n] by putting
(1.26) Bn = ξ ∈ X [n] | | Supp(ξ)| < n .
and let π2 : X → X be the quotient map. Then, KX
2 [2] ∗
2 = π1 KX 2 + E and
∗ ∗ ∗
KX 2 = π2 KX [2] + E. Thus, π1 KX 2 = π2 KX [2] . Finally, since π2∗ f = β2 , we obtain
Pt (X)
+∞
= · Pt X [n] q n .
1 − t2 q n=0
Formula (1.41) will be important in Chapter 6 which deals with the relations
between the incidence Hilbert schemes of points and infinite dimensional Lie alge-
bras.
We refer to [Can, Che2, Che3] for generalizations from incidence Hilbert
schemes to nested Hilbert schemes of points on surfaces, and to [GSY1, GSY2] for
generalizations to nested Hilbert schemes of points and curves on surfaces which
are related to localized Donaldson-Thomas theory.
CHAPTER 2
19
20 2. THE NEF CONE AND FLIP STRUCTURE OF (P2 )[n]
where the symbol “A1 ∼ A2 ” means that A1 and A2 are homologous. We further
1 , . . . , C
assume that C1 , . . . , Cs , C s are in general position. It follows that
(2.2) βC , D = C i , C j
i Cj
[n]
of OX |Mn (x) over η ∈ Mn (x) ⊂ X [n] is H 0 (X, Oη ) = Oη,x . Since the fiber of
∗
τ Q over η ∈ Mn (O) is Oη,O , we see from the universality of the quotient bundle
2.1. CURVES HOMOLOGOUS TO βn 21
det(Q) = p∗ H yields
[n]
(2.7) γ · Bn = −2γ · c1 (OX )
[n]
= −2γ · c1 (OX |Mn (x) )
= −2γ · c1 (τ ∗ Q)
= −2γ · τ ∗ p∗ c1 (H)
= −2(p ◦ τ )(γ) · c1 (H).
Next, since γ ⊂ Mn (x), we have γ · DC = 0 = βn · DC for every real-surface
C in X. By Lemma 2.1, γ ∼ βn if and only if γ · Bn = −2 = βn · Bn . By (2.7),
γ · Bn = −2 if and only if p ◦ τ (γ) is a line.
Lemma 2.3. Let k = n(n − 1)/2, and L ⊂ G = Grass(R/mn , n) be a curve.
Then, L is mapped to a line by p if and only if L is of the form
(2.8) L = Span(f 1 , . . . , f k−1 , λf k + μf k+1 ) | λ, μ ∈ C, |λ| + |μ| = 0
where f1 , . . . , fk+1 ∈ R are polynomials such that f 1 , . . . , f k+1 are linearly indepen-
dent in the vector space R/mn ∼ = Cn+k .
Proof. From the definition of p, it is clear that if L is of the form (2.8), then
p(L) is a line. Conversely, assume that p(L) is a line. We see from det(Q) = p∗ H
that det(Q|L ) ∼
= OL (1). Restricting (2.5) to L, we obtain an exact sequence:
(2.9) 0 → S|L → (R/mn ) ⊗ OL → Q|L → 0.
So det(S|L ) ∼
= OL (−1). From the injection S|L → (R/mn ) ⊗ OL , we obtain
k
S|L ∼
= OL (ai )
i=1
k
with i=1 ai = −1 and ai ≤ 0. Thus
S|L ∼
⊕(k−1)
= OL ⊕ OL (−1).
⊕(k−1)
Now the injection OL → (R/mn ) ⊗ OL gives rise to linearly independent
elements f 1 , . . . , f k−1 in R/mn . The injection OL (−1) → (R/mn ) ⊗ OL gives rise
to two linearly independent elements f k and f k+1 in R/mn such that
f 1 , . . . , f k−1 , λf k + μf k+1
are linearly independent when |λ| + |μ| = 0. So L is of the form (2.8).
Lemma 2.4. Let x ∈ X and η ∈ Mn+1 (x). Let ξ1 , ξ2 ∈ Mn (x) and ξ1 = ξ2 . If
ξ1 , ξ2 ⊂ η, then Iη,x ⊃ mnx where mx is the maximal ideal of OX,x .
Proof. We have Iη,x ⊂ Iξ1 ,x ∩ Iξ2 ,x ⊂ Iξ2 ,x and Iξ1 ,x ∩ Iξ2 ,x = Iξ2 ,x . Since the
colength of Iη,x is n + 1 and the colength of Iξ2 ,x is n, we obtain Iη,x = Iξ1 ,x ∩ Iξ2 ,x .
Since mnx ⊂ Iξ1 ,x and mnx ⊂ Iξ2 ,x , we get mnx ⊂ Iξ1 ,x ∩ Iξ2 ,x = Iη,x .
Now we are able to characterize all the curves γ in the punctual Hilbert scheme
Mn (x) which is homologous to βn . By Theorem 1.31 (i), the fiber (ψn+1 )−1 (η, x)
over (η, x) ∈ Zn+1 ⊂ X [n+1] × X is isomorphic to P(ωηx ⊗ Cx ).
22 2. THE NEF CONE AND FLIP STRUCTURE OF (P2 )[n]
Since (DC − Bn /2) is very ample, there exists exactly one i such that pi (γ) is a
curve in X [ni ] and pj (γ) is a point in X [nj ] for all j = i. We may assume i = 1.
Let k = n1 and x = x1 . Then γ = γk + ξ where γk ⊂ Mk (x) and ξ ∈ X [n−k] is
fixed with x ∈/ Supp(ξ). It remains to prove γk ∼ βk in X [k] .
Claim. Let x̃ ∈ X be a fixed point, and α ⊂ X [n−1] be a curve such that x̃ ∈
/
Supp(ξ) for every ξ ∈ α. Then, (α + x̃) · Bn = α · Bn−1 .
Proof. Define
= { (ξ, ξ + x̃) | ξ ∈ α} ⊂ X [n−1,n] .
α
Since x̃ ∈/ Supp(α), the restriction gn |α : α → α + x̃ is an isomorphism. So
= α + x̃. Similarly, (fn )∗ α
(gn )∗ α = α. Let
En = {(ξ, η) ∈ X [n−1,n] | Supp(ξ) = Supp(η), ξ ⊂ η}.
Since x̃ ∈ ∩ En = ∅. By (1.40),
/ Supp(α), we obtain α
(gn )∗ Bn − (fn )∗ Bn−1 = 2En .
24 2. THE NEF CONE AND FLIP STRUCTURE OF (P2 )[n]
Conversely, suppose γ = fn+1 (C) where C is a line in (ψn+1 )−1 (η, x) for some
(η, x) ∈ Zn+1 . Assume (ηx ) = k + 1. Let ηk+1 = ηx ∈ X [k+1] . Reversing the
argument in the preceding paragraph and using Lemma 2.6 (ii), we get γ ∼ βn .
(ii) To show the uniqueness of (η, x) and C, let ξ1 , ξ2 ∈ γ with ξ1 = ξ2 . Then,
ξ1 , ξ2 ⊂ η. Since (η) = 1 + (ξ1 ) = 1 + (ξ2 ), we get Iη = Iξ1 ∩ Iξ2 . So η and hence
the point x are uniquely determined by γ. Since
fn+1 |(ψn+1 )−1 (η,x) : (ψn+1 )−1 (η, x) → fn+1 ((ψn+1 )−1 (η, x))
is an isomorphism, C is also uniquely determined by γ.
Our next goal is to give a global description of the moduli space M(βn ) of all
the curves in X [n] homologous to βn . Note that M(βn ) is the union of certain
irreducible components in the Hilbert scheme of curves in X [n] . We will use the
concept of the relative Grassmannian scheme from [Sim].
Theorem 2.8. Let n ≥ 2, X be a simply connected smooth projective surface,
and ωZn+1 be the dualizing sheaf of Zn+1 ⊂ X [n+1] × X. Then, there exists a
bijective morphism from the relative Grassmannian Grass(ωZn+1 , 2) over Zn+1 to
the moduli space M(βn ) of all the curves in X [n] homologous to βn .
= Grass(ωZ , 2) for simplicity. Define
Proof. Put G n+1
Γ = [C], (ξ, η) ∈ G × X [n,n+1] | C is a line in (ψn+1 )−1 (η, x)
and (ξ, η) ∈ C with Supp(Iξ /Iη ) = {x} .
By Theorem 1.31 (i), P(ωZn+1 ) ∼
= X [n,n+1] over Zn+1 . So Γ is the universal sub-
Consider
scheme in G ×Zn+1 P(ωZn+1 ) which is flat over G.
×fn+1
IdG
×Z
α:G × X [n,n+1]
X [n,n+1] ⊂ G −→ × X [n] .
G
n+1
= Grass(ωηx ⊗ Cx , 2) × P(ωηx ⊗ Cx )
= Grass(ωηx ⊗ Cx , 2) × (ψn+1 )−1 (η, x).
Since fn+1 maps (ψn+1 )−1 (η, x) isomorphically to fn+1 ((ψn+1 )−1 (η, x)), we see
that α|Γ : Γ → α(Γ) is bijective. Moreover, assuming that q ∈ G is mapped to
(η, x) via the natural projection G → Zn+1 and that q corresponds to the line
{q} × C = Γ|q ⊂ {q} × P(ωηx ⊗ Cx ) = {q} × (ψn+1 )−1 (η, x)
via the universal subscheme Γ ⊂ G ×Z P(ωZn+1 ), then we see from (2.11) that
n+1
the same point q ∈ G also corresponds to the curve
α(Γ)|q = α(Γ|q ) = α({q}×C) = {q}×fn+1 (C) ⊂ {q} × fn+1 ((ψn+1 )−1 (η, x))
via the subscheme α(Γ) ⊂ G × X [n] . By Theorem 2.7, we conclude that via the
subscheme α(Γ) ⊂ G × X [n] the relative Grassmannian G
which is flat over G,
[n]
parameterizes curves in X homologous to βn . By the universal property of the
Hilbert scheme of curves, we obtain a morphism Λ : G → M(βn ). Moreover, we see
from Theorem 2.7 that the morphism Λ is bijective.
26 2. THE NEF CONE AND FLIP STRUCTURE OF (P2 )[n]
Let θ(x) + ξ ∈ Im(Ψ) where ξ denotes x1 + . . . + xn−2 . Then, (πn+1 )−1 (θ(x) + ξ)
consists of (n − 1) distinct points in Zn+1 :
(θ(x) + ξ, x), (θ(x) + ξ, xi ), 1 ≤ i ≤ n − 2.
Note that
Grass(ωZn+1 , 2)|(θ(x)+ξ,x) = Grass(ωθ(x) ⊗ Cx , 2) ∼
= Grass(C2 , 2) = pt.
In fact, Grass(ωZn+1 , 2)|(θ(x)+ξ,x) is the point in Grass(ωZn+1 , 2) corresponding to
the curve M2 (x) + ξ in X [n] . Also,
Grass(ωZ , 2)|(θ(x)+ξ,x ) ∼
n+1 = Grass(C, 2) = ∅.
i
Nγ⊂X [n] ∼
= Oγ⊕(2n−2) ⊕ Oγ (−2).
Next, we have an exact sequence
0 → Tγ → TX [n] |γ → Nγ⊂X [n] → 0.
2.1. CURVES HOMOLOGOUS TO βn 27
for some λ, μ ∈ C with |λ| + |μ| = 0. So a basis for the subspace Iη,0 /mn ⊂ R/mn
can be chosen as
λu + μv n−1 , u2 , u v, u3 , u2 v, u v 2 , . . . , un−1 , un−2 v, . . . , u v n−2 ,
and the matrix representation of Iη,0 /mn is given by the n2 × n+1
2 -matrix:
⎡ ⎤
0 λ 0 ... 0 0 ... 0 μ
⎢ 0 0 1 ... 0 0 ... 0 0 ⎥
⎢ ⎥
⎢ .. .. .. .. .. .. .. .. ⎥ .
⎣ . . . . . . ... . . ⎦
0 0 0 ... 1 0 ... 0 0
Thus, (p ◦ τ )(η) = [0, . . . , 0, λ, 0, . . . , 0, μ, 0, . . . , 0] where the positions of λ and μ
are independent of η ∈ σn . So the image (p ◦ τ )(σn ) is a line.
Note that the flat limits of the elements (λu + v, v n ), λ ∈ C∗ in Mn (O) as
λ → 0 and λ → ∞ are equal to (v, un ) and (u, v n ) respectively. So in the punctual
Hilbert scheme Mn (O), we have the projective curve:
(2.13) σ̃n = {(λu + v, v n ) | λ ∈ C∗ } ∪ {(v, un ), (u, v n )}.
n
Lemma 2.12. As a curve in X [n] , σ̃n is homologous to 2 βn .
Proof. It suffices to show that σ̃n is homologous to n2 σn in H2 (Mn (O); Z).
By (2.13), if η ∈ σ̃n − {(v, un ), (u, v n )}, then a basis for Iη,0 /mn ⊂ R/mn consists
of
λu + v, λu2 + u v, λu v + v 2 , . . . ,
n−1 n−2
λu +u v, λun−2 v + un−3 v 2 , . . . , λu v n−2 + v n−1 .
As in the proof of Lemma 2.11, we see that the degree of
" #
(p ◦ τ ) σ̃n − {(v, un ), (u, v n )}
β · (nD − Bn /2) = 1
Proof. (i) Note that D is nef and big, but not ample. By Proposition 2.16,
((n − 1)D − Bn /2) is also nef but not ample. By Theorem 1.24 (iii), (−Bn /2) and
D form a basis for A2n−1 (X [n] ) ∼
= H4n−2 (X [n] ; Z). So the nef cone of the Hilbert
scheme X is the cone spanned by D and ((n − 1)D − Bn /2).
[n]
(ii) Since (−Bn /2) and D form a basis for H4n−2 (X [n] ; Z), we see from (2.19)
and the Poincaré duality that {β , βn } is a basis for H2 (X [n] ; Z). By Lemma 2.15,
both βn and (β − (n − 1)βn ) are numerical equivalence classes of effective curves.
By (2.19) again,
βn · D = 0,
(β − (n − 1)βn ) · D = 1,
βn · ((n − 1)D − Bn /2) = 1,
(β − (n − 1)βn ) · ((n − 1)D − Bn /2) = 0.
It follows from (i) that the cone NE(X [n] ) is spanned by βn and (β − (n − 1)βn ).
(iii) Let β = aβn + b(β − (n − 1)βn ). Intersecting β with the nef divisors
D and ((n − 1)D − Bn /2), we see that a and b are nonnegative integers. Since
(nD − Bn /2) is very ample and
Theorem 2.17 can be generalized to the following, and we refer to [QT] for
further details.
t
OX (n − 1) Fi
i=1
t
(2.20) DF1 , ..., DFt , (n − 1) DFi − Bn /2;
i=1
In the following, we will study the normal bundle of a curve γ in X [n] homol-
ogous to β − (n − 1)βn . By Theorem 2.22, there exists a line C in X = P2 such
that γ is a line in C (n) = C [n] ⊂ X [n] . In particular,
Nγ⊂C (n) ∼
= Oγ (1)⊕(n−1) .
So we have the following exact sequence of normal bundles:
(2.22) 0 → Oγ (1)⊕(n−1) → Nγ⊂X [n] → NC (n) ⊂X [n] |γ → 0.
where π̃n∗ : Zn → Pn and q̃n : Pn × P1 → P1 are the projections. It is known that
Zn ⊂ Pn × P1 is defined by the equation
a0 U n + a1 U n−1 V + . . . + an V n = 0
where a0 , a1 , . . . , an and U, V are the homogeneous coordinates on Pn and P1 re-
spectively. So the line bundle OPn ×P1 (Zn ) is of type (1, n) in Pic(Pn × P1 ) ∼
= Z ⊕ Z.
Let p̃n : Pn × P1 → Pn be the projection. Applying p̃n∗ to the exact sequence
0 → q̃n∗ OP1 (1) ⊗ OPn ×P1 (−Zn ) → q̃n∗ OP1 (1) → q̃n∗ OP1 (1)|Zn → 0,
we obtain
⊕2 ∗ ⊕(n−2)
0 → OC n → OC (n) (−1)
(n) → π̃n∗ q̃n OP1 (1)|Z → 0.
Since this exact sequence splits, we conclude the second isomorphism.
(ii) Follows immediately from (2.22) and the isomorphisms in (i).
(iii) By (ii), H 1 (γ, Nγ⊂X [n] ) = 0 for any curve γ homologous to (β −(n−1)βn ).
By deformation theory, M(β − (n − 1)βn ) is unobstructed.
In the next theorem, we further identify M(β − (n − 1)βn ) with a certain
Grassmannian bundle.
Theorem 2.24. Let n ≥ 2, X = P2 , and be a line in X. Then the moduli space
M(β − (n − 1)βn ) of curves in X [n] homologous to (β − (n − 1)βn ) is isomorphic
to the Grassmannian bundle Grass(Symn ((T(P2 )∗ )∗ ), 2) over (P2 )∗ .
Proof. First of all, we give a global description of the union Wn of all the
subsets C [n] ⊂ X [n] with C being lines in X = P2 . Consider
F = {(x, [C]) ∈ P2 × (P2 )∗ | x ∈ C} ⊂ P2 × (P2 )∗ .
We know that F ∼
= P((T(P2 )∗ )∗ ). Then we have
Hilbn (P((T(P2 )∗ )∗ )/(P2 )∗ ) ∼
= Hilbn (F/(P2 )∗ )
⊂ Hilbn (P2 × (P2 )∗ /(P2 )∗ )
∼
= X [n] × (P2 )∗
of relative Hilbert schemes over (P2 )∗ . Let
π : Hilbn (P((T(P2 )∗ )∗ )/(P2 )∗ ) → (P2 )∗
be the projection. Consider the map
α : Hilbn (P((T(P2 )∗ )∗ )/(P2 )∗ ) → X [n] × (P2 )∗ → X [n] .
Note that via the isomorphism
Hilbn (P((T(P2 )∗ )∗ )/(P2 )∗ ) ∼
= Hilbn (F/(P2 )∗ ),
the fiber of the map π over a point [C] ∈ (P2 )∗ is isomorphic to C [n] . So Im(α) =
Wn . Moreover, since n ≥ 2, an element ξ ∈ X [n] is contained in at most one line
C in X = P2 . So if α(ξ ) = α(η ) for ξ , η ∈ Hilbn (P((T(P2 )∗ )∗ )/(P2 )∗ ), then there
exists a unique line C such that α(ξ ) = α(η ) ⊂ C, i.e., ξ and η are contained in
the fiber π −1 ([C]) over the point [C] ∈ (P2 )∗ . Since α maps this fiber isomorphically
to C [n] ⊂ X [n] , we must have ξ = η . Also
∼ P(Symn ((T(P2 )∗ )∗ ))
Hilbn (P((T(P2 )∗ )∗ )/(P2 )∗ ) =
2.4. A FLIP STRUCTURE ON (P2 )[n] WHEN n ≥ 3 35
ϕ1 (X [n] )
satisfying the Flip Conjecture 5-1-10 in [KMM], by proving that (X [n] )+ is smooth,
ϕ+ is a birational morphism, and K(X [n] )+ is ϕ+ -ample. The proof is a little long,
but may be divided into two steps. In step one, we construct two additional bira-
tional morphisms ϕ − and ϕ+ . In step two, we construct the birational morphism
ϕ+ .
Let
:X
ϕ ' [n] → X [n]
−
' '
n ⊂ X
[n] [n]
be the blowing-up of X along the contraction locus Wn of ϕ1 , and W
[n]
be the exceptional divisor. Let N be the normal bundle of Wn in X . Then over
'
W ∗
n = P(N ), we have the tautological surjection
(2.24) − |W
(ϕ n
)∗ (N ∗ ) → OW
n
(1) → 0
where OW
'
(1) is the tautological line bundle over W ∗
n = P(N ). In view of
n
Lemma 2.25 (iii),
N∼ = (ϕ1 |W )∗ M ⊗ OW (−1)
n n
where M is some rank-(n − 2) vector bundle over ϕ1 (Wn ) = ∼ (P2 )∗ , and OW (1)
n
stands for the tautological line bundle over the projective bundle ϕ1 |Wn : Wn →
ϕ1 (Wn ). So from (2.24), we obtain a surjection:
(ϕ1 |Wn ◦ ϕ
− |W
n
)∗ (M ∗ ) → (ϕ
− |W
n
)∗ OWn (−1) ⊗ OW
n
(1) → 0.
(ϕ1 |Wn ◦ ϕ
− |W
): '
W n → ϕ1 (Wn )
n
factors through the natural projection P(M ∗ ) → ϕ1 (Wn ). In fact, the induced
morphism ϕ :W' ∗
n → P(M ) is a P -bundle. So the fibers of
n
(ϕ1 |Wn ◦ ϕ
− |W
): '
W n → ϕ1 (Wn )
n
38 2. THE NEF CONE AND FLIP STRUCTURE OF (P2 )[n]
are naturally Pn × Pn−3 . Moreover, for the projective spaces Pn in the fibers Pn ×
Pn−3 , we have OX '
(W ∼
n )|Pn = OPn (−1), using the facts that OX
'
(W ∼
n )|Pn ×Pn−3 =
[n] [n]
⊗ OPn (1) ∼
⊕(n−2)
OPn = N ∗ |Pn
( )
∼
= − |Pn ×Pn−3 )∗ OX
(ϕ (−W'n )| Pn ×Pn−3
[n]
∼
= OPn
⊕(n−2)
⊗ OPn (−a).
:W
ϕ ' n → P(M ∗
). By the results in [Naka, FN], there exists a morphism
ϕ '
+ : X [n] → (X [n] )
+
for some divisor H+ on (X [n] )+ . Indeed, recall from (2.16) that ((n−1)D −Bn /2) =
(ϕ1 )∗ H for some very ample divisor H on ϕ1 (X [n] ). Consider the divisor
'
Z · [W '
− )∗ Cl(X [n] ) = Cl(X ' + )∗ Cl((X [n] )+ )
n ] ⊕ (ϕ
[n] ) = Z · [W
n ] ⊕ (ϕ
where Cl(·) stands for the divisor class group, we conclude that
(2.26) (ϕ '
− )∗ ((n − 1)D − Bn /2) = λW + )∗ H+
n + (ϕ
for some integer λ and some divisor H+ on (X [n] )+ . Let C ⊂ X ' [n] be a line
'
:W
the projective bundle ϕ ∗
n → P(M ). So λ = 0 and
Since ϕ+ is the blowing-up morphism of the smooth variety (X [n] )+ , we see
∗
+ )∗ |H+ |. So combining with (2.25) and (2.27), we obtain
+ ) H+ | = (ϕ
that |(ϕ
(2.28) − )∗ H| = |(ϕ
|(ϕ1 ◦ ϕ − )∗ ((n − 1)D − Bn /2)|
= |(ϕ+ )∗ H+ |
= (ϕ+ )∗ |H+ |.
Since ϕ1 is the contraction of the extremal ray R = R+ ·(β −(n−1)βn ), it is known
from [KMM, Remark 5-1-1] that ϕ1 (X [n] ) has at most rational singularities. So
− )∗ H| = (ϕ1 ◦ ϕ
|(ϕ1 ◦ ϕ − )∗ |H|.
In view of (2.28),
− )∗ |H| = |(ϕ1 ◦ ϕ
(ϕ1 ◦ ϕ − )∗ H| = (ϕ
+ )∗ |H+ |.
Since H is a very ample divisor on ϕ1 (X [n] ), (ϕ1 ◦ ϕ − )∗ |H| is base-point-free. It
∗ ∗
+ ) |H+ | = (ϕ1 ◦ ϕ
follows from (ϕ − ) |H| that |H+ | is also base-point-free. So |H+ |
induces a morphism
ϕ+ : (X [n] )+ → ϕ+ ((X [n] )+ ).
Moreover, since (ϕ+ )∗ |H+ | = (ϕ1 ◦ ϕ− )∗ |H|, we conclude that ϕ1 ◦ ϕ − = ϕ+ ◦ ϕ +
up to an isomorphism between ϕ1 (X [n] ) and ϕ+ ((X [n] )+ ). So putting ϕ− = ϕ1 , we
obtain a commutative diagram:
'
X [n]
ϕ− ϕ+
[n]
(2.29) X (X [n] )+ .
ϕ−
ϕ+
ϕ1 (X [n] )
Note that ϕ+ : (X [n] )+ − P(M ∗ ) → ϕ1 (X [n] ) − ϕ1 (Wn ) is an isomorphism.
It is standard to check that K(X [n] )+ is ϕ+ -ample. So the commutative triangle
− )−1
+ ◦(ϕ
ϕ
X [n] (X [n] )+
(2.30) ϕ−
ϕ+
ϕ1 (X [n] )
is the flip for ϕ− = ϕ1 .
Remark 2.27. The proof of Theorem 2.26 shows that for n = 3, the morphism
ϕ1 : X [n] → ϕ1 (X [n] )
is the blowing-up along ϕ1 (Wn ) ∼
= (P2 )∗ , and Wn is the exceptional locus.
Part 2
In this chapter, we will survey various infinite dimensional Lie algebra actions
on the cohomology of the Hilbert schemes of points on a smooth projective complex
surface. As a motivation and background material, we will begin with Nakajima’s
affine Lie algebra action on the homology of quiver varieties [Nak1, Nak2, Nak4].
We will review the Heisenberg algebras of Nakajima and Grojnowski [Groj, Nak3]
which provide an important language for describing the homology and cohomology
classes of the Hilbert schemes, determine the geometric interpretations of Heisen-
berg monomial classes, and study the homology classes of curves in the Hilbert
schemes. Then we will present the Virasoro algebras and boundary operator of
Lehn [Leh1] as well as the Ext vertex operators of Carlsson and Okounkov [Car1,
Car2,CO]. As evidenced by [LQW1,LQW2,LQW3,LQW5,LQW7,LS1,LS2],
Lehn’s boundary operator and the Heisenberg algebras of Nakajima and Grojnowski
play essential roles in determining the cohomology ring structures of the Hilbert
schemes. The Ext vertex operators of Carlsson and Okounkov are very powerful
tools in understanding the intersection theory of the Hilbert schemes coupled with
the Chern classes of their tangent bundles, and in investigating the connection be-
tween the multiple q-zeta values and the Hilbert schemes of points [Ok2,QY]. These
operators also have applications in studying the AGT correspondence [AGT, Neg].
In addition, we will prove a structure theorem (Theorem 3.31) for the higher order
derivatives of the Heisenberg operators.
ζ = (ζC , ζR ) ∈ CI ⊕ RI .
43
44 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
Let V, W be I-graded vector spaces such that dim Vi = vi and dim Wi = wi . Put
M(v, w) = Hom(Vout(h) , Vin(h) ) ⊕ Hom(Wi , Vi ) ⊕ Hom(Vi , Wi )
h∈H i∈I
where (h) = 1 and (h) = −1 for h ∈ Ω. Note that CI is canonically identified with
the center of glv . The element ζC = (ζC,i ) ∈ CI is identified with ⊕i∈I ζC,i IdVi ∈ glv .
The quiver variety Mζ (v, w) associated to v, w and ζ is defined to be
μ−1
ss
C (ζC ) //GLv
where the index “ss” means the subset of ζR -semistable objects in μ−1 C (ζC ).
Let A = (ai,j )i,j∈I be the adjacency matrix of the quiver (I, H), that is to say
ai,j is the number of oriented edges in H which are drawn from i to j. Let
C = 2 · Id − A
be the associated Cartan matrix . Then we consider the set of positive roots
R+ = θ ∈ (Z≥0 )I |θ t Cθ ≤ 2 − {0}.
For v ∈ (Z+ )I , define
R+ (v) = θ = (θi ) ∈ R+ |θi ≤ vi ∀i ∈ I .
An element ζ ∈ CI ⊕ RI is called generic with respect to v if for any θ ∈ R+ (v),
ζ ∈ Dθ ⊗ R3 ⊂ RI ⊗ R3 ∼
= CI ⊕ RI
where * +
Dθ = η = (ηi ) ∈ R | I
ηi θi = 0 .
i∈I
Note that
ζ0 = (0, ζR )
is generic if ζR ∈ C0 where
C0 = η = (ηi ) ∈ RI |ηi < 0 ∀i ∈ I .
3.1. AFFINE LIE ALGEBRA ACTION OF NAKAJIMA 45
is connected, Mζ (v, w) and Mζ (v, w) are diffeomorphic for generic ζ and ζ . More-
over, for generic ζ, Mζ (v, w) is a fine moduli space of representations of the quiver
[King, Section 3].
In the rest of this section, let (I, H) be the quiver of type Â−1 with cyclic
orientation. Fix w0 = (1, 0, . . . , 0) ∈ ZI , and ζ0 = (0, ζR ) with ζR ∈ C0 . Let
ei denote the i-th coordinate vector of ZI . For fixed i and v, define the Hecke
correspondence to be the following subvariety of Mζ0 (v, w0 ) × Mζ0 (v + ei , w0 ):
(3.1) Bi (v) = (J1 , J2 )|J1 ∈ Mζ0 (v, w0 ) is
a subrepresentation of J2 ∈ Mζ0 (v + ei , w0 ) .
By [Nak4, Theorem 5.7], Bi (v) is Lagrangian in Mζ0 (v, w0 ) × Mζ0 (v + ei , w0 ).
Next, let sl be the Lie algebra consisting of all complex × -matrices of trace
zero. It has Chevalley generators
ei = Ei,i+1 , fi = Ei+1,i , hi = Ei,i − Ei+1,i+1
with 1 ≤ i ≤ − 1, where Ei,j denotes the matrix which has a 1 on the i-th row
is given by
and j-th column and 0 elsewhere. The affine Kac-Moody algebra sl
= sl [t, t−1 ] ⊕ Cc ⊕ Cd
sl
[Hern, Subsection 4.2]. In [Nak4, Section 9], a geometric representation of sl on
the homology of the quiver varieties Mζ0 (v, w0 ) indexed by v is constructed. The
images of hi and d are certain multiples of the fundamental classes of the diagonals
in the products Mζ0 (v, w0 ) × Mζ0 (v, w0 ). The image of ei is the operator Ei
defined by
Ei (α) = (−1)vi−1 +vi p1∗ p∗2 α · Bi (v)
where pj is the j-th projection on Mζ0 (v, w0 ) × Mζ0 (v + ei , w0 ), while the image
of fi is the operator Fi defined by
Fi (α) = (−1)vi +vi+1 p2∗ p∗1 α · Bi (v) .
The quiver variety Mζ0 (v, w0 ) is related to the Hilbert schemes of points as
follows. Regard the cyclic group Z/ Z as the subgroup of SL2 (C) consisting of the
diagonal matrices diag(i , −i ), 0 ≤ i < where is a primitive -th root of unity.
Then, Z/ Z acts on the affine plane C2 . Let X0 denote the minimal resolution of
C2 /(Z/ Z). Then there exists an element
ζ∞ = (0, ζR ) ∈ CI ⊕ RI
generic with respect to v such that Mζ∞ (v, w0 ) is naturally identified with the
Hilbert scheme (X0 )[n] where n = dim Mζ∞ (v, w0 ) /2. Since both ζ0 and ζ∞ are
generic with respect to v, Mζ∞ (v, w0 ) is diffeomorphic to Mζ0 (v, w0 ). Therefore,
the above geometric action of the affine Lie algebra sl on the homology of the
on the
quiver varieties Mζ0 (v, w0 ) with varying v induces naturally an action of sl
homology of the Hilbert schemes (X0 ) , n ≥ 0.
[n]
The preceding discussion indicates strongly that for an arbitrary complex sur-
face X, there should exist, via certain Hecke correspondence, geometric action of
infinite dimensional Lie algebras on the (co)homology groups of the Hilbert schemes
46 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
X [n] , n ≥ 0 (here we have by-passed the identification of quiver varieties with moduli
spaces of torsion-free sheaves on ALE spaces). Indeed, Nakajima and Grojnowski
[Groj, Nak3] constructed action of Heisenberg algebras on the (co)homology of
these Hilbert schemes. This leads us to the next section.
Remark 3.1. It is unclear whether the Hecke correspondence Bi (v) defined in
(3.1) for the quiver varieties could be transferred directly to the Hilbert schemes
(X0 )[n] . On the other hand, by the theorem in [Nag, Subsection 6.2], the above
induced action of the affine Lie algebra sl on the homology of the Hilbert schemes
(X0 )[n] is related to the action of the Heisenberg algebra (to be constructed in the
next section) via the Frenkel-Kac construction [Kac, Section 5.6].
Let
+∞
HX = HnX = Hn,i
X
n=0 n,i≥0
Definition 3.3. Fix a cohomology class α ∈ H ∗ (X). Let n > 0. The linear
operator a−n (α) ∈ End(HX ) is defined by
" #
a−n (α)(a) = p̃1∗ [Q[m+n,m] ] · ρ̃∗ α · p̃∗2 a
for a ∈ H ∗ (X [m] ), where p̃1 , ρ̃, p̃2 are the projections of X [m+n] × X × X [m] to
X [m+n] , X, X [m] respectively. Define an (α) ∈ End(HX ) to be (−1)n times the
operator obtained from the definition of a−n (α) by switching the roles of p̃1 and
p̃2 . We often refer to a−n (α) (respectively, an (α)) as the creation (respectively,
annihilation) operator. We also set a0 (α) = 0.
Lemma 3.4. Let m ≥ 0 and n > 0. Then, dim Q[m+n,m] = 2m + n + 1.
Proof. Every element (ξ, x, η) ∈ Q[m+n,m] is of the form
η = ηx + η , ξ = ξx + η
where Supp(ηx ) = Supp(ξx ) = {x}, x ∈ Supp(η ) and ηx ⊂ ξx . Fix the length
= (ηx ). If = 0, then these elements (ξ, x, η) ∈ Q[m+n,m] form an open subset
of Q[m+n,m] with dimension equal to
#moduli{η } + #moduli{x} + #moduli{ξx } = 2m + 2 + (n − 1)
= 2m + n + 1
where we have used dim Mn (x) = n − 1 from Theorem 1.18. If > 0, then the
number of moduli of these elements (ξ, x, η) ∈ Q[m+n,m] is at most
#moduli{η } + #moduli{x} + #moduli{ηx } + #moduli{ξx }
≤ 2(m − ) + 2 + ( − 1) + (n + − 1)
= 2m + n.
It follows that the dimension of Q[m+n,m] is equal to 2m + n + 1.
Corollary 3.5. Let n = 0. Then the bi-degree of a−n (α) is (n, 2n − 2 + |α|).
Proof. Follows immediately from Lemma 3.4 and Definition 3.3. Note from
the proof of Lemma 3.4 that Q[m+n,m] has a unique irreducible component of di-
mension 2m+n+1. Although Q[m+n,m] may (or may not) contain other irreducible
components of lower dimension, it does not matter since only the fundamental class
[Q[m+n,m] ] is used in Definition 3.3.
48 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
Remark 3.6. Let n > 0. The geometric meaning of the operator a−n (α)
is roughly the following. Let a ∈ H ∗ X [m] . Represent α and a by the cycles
Γα ⊂ X and Γa ⊂ X [m] respectively such that a generic element x ∈ Γα and a
generic element η ∈ Γa satisfy x ∈ Supp(η). Then a−n (α) is represent by the cycle
Γ ⊂ X [m+n] whose generic elements are of the form η + ξx where η ∈ Γa , x ∈ Γα
but x ∈ Supp(η), and ξx ∈ Mn (x).
Lemma 3.7. a−n (α)† = (−1)n · an (α).
Proof. The argument is completely formal. We may assume that n > 0. Let
a ∈ H ∗ X [m] and b ∈ H ∗ X [m+n] . We must prove that
(3.7) a−n (α)(a), b = (−1)(2n−2+|α|) |a| · a, (−1)n · an (α)(b)
noting that the bi-degree of a−n (α) is (n, 2n − 2 + |α|).
Let notations be the same as in Definition 3.3. By Definition 3.3,
,
a−n (α)(a), b = a−n (α)(a) · b
[m+n]
,X " #
= p̃1∗ [Q[m+n,m] ] · ρ̃∗ α · p̃∗2 a · b
[m+n]
,X " #
(3.8) = p̃1∗ [Q[m+n,m] ] · ρ̃∗ α · p̃∗2 a · p̃∗1 b
X [m+n]
where we have used the projection formula in the last step. Similarly, for the
right-hand-side of (3.7), we see from the definition of an (α) that
(−1)(2n−2+|α|) |a| · a, (−1)n · an (α)(b)
, " #
|α| |a|
= (−1) · p̃2∗ p̃∗2 a · [Q[m+n,m] ] · ρ̃∗ α · p̃∗1 b
X [m]
, " #
= p̃2∗ [Q[m+n,m] ] · ρ̃∗ α · p̃∗2 a · p̃∗1 b .
X [m]
Combining this with (3.8), we conclude that (3.7) holds.
The following theorem is proved by Nakajima [Nak3, Section 3] and Gro-
jnowski [Groj, Section 3].
Theorem 3.8. The operators an (α) ∈ End(HX ) with α ∈ H ∗ (X) and n ∈ Z
satisfy the following Heisenberg algebra commutation relation:
(3.9) [am (α), an (β)] = −m · δm,−n · (α, β) · IdHX
where we have used δm,−n to denote 1 if m = −n and 0 otherwise. Moreover,
the space HX is an irreducible module over the Heisenberg algebra generated by the
operators an (α) with a highest weight vector |0 = 1 ∈ H 0 (X [0] ).
It follows that HX is linearly spanned by all the Heisenberg monomial classes:
(3.10) a−n1 (α1 ) · · · a−nk (αk )|0
where k ≥ 0, n1 , . . . , nk > 0, and the cohomology classes α1 , . . . , αk run over a fixed
linear basis of H ∗ (X). In view of Corollary 3.5, we have
(3.11) a−n1 (α1 ) · · · a−nk (αk )|0 ∈ H m (X [n] )
where n = n1 + . . . + nk and m = ki=1 (2ni − 2 + |αi |).
3.2. HEISENBERG ALGEBRAS OF NAKAJIMA AND GROJNOWSKI 49
Next, we will introduce the notion am1 · · · amk (τk∗ α) and study its properties.
This notion will be useful in writing down the Chern character operators and the
higher derivatives of the Heisenberg operators.
Definition 3.9. For k ≥ 1, define τk∗ : H ∗ (X) → H ∗ (X k ) to be the map
induced by the diagonal embedding τk : X → X k , and am1 · · · amk (τk∗ α) to be
am1 (αj,1 ) · · · amk (αj,k )
j
when
(3.12) τk∗ α = αj,1 ⊗ · · · ⊗ αj,k
j
Therefore,
cj,1,1 · · · cj,k,k = ds,1,1 · · · ds,k,k
j s
for all 1, . . . , k. It follows that
am1 (αj,1 ) · · · amk (αj,k )
j
⎛ ⎞
= ⎝ cj,1,1 · · · cj,k,k ⎠ am1 (γ1 ) · · · amk (γk )
1 ,...,k j
= ds,1,1 · · · ds,k,k am1 (γ1 ) · · · amk (γk )
1 ,...,k s
= am1 (βs,1 ) · · · amk (βs,k ).
s
i
k
,
τ(k−1)∗ (αβ) = (−1)|β| =j+1 |αi, |
αi,j β · ⊗1≤s≤k,s=j αi,s
i X
" j−1 #
τ(k+u−1)∗ (α) = ⊗s=1 αi,s ⊗ (τu∗ αi,j ) ⊗ ⊗kt=j+1 αi,t .
i
Proof. The basic idea is to use the projection formula. Let pj be the projec-
tion of X k to the j-th factor. Then, we have
k " #
(−1)|β|· =j+1 |αi, |
· ⊗j−1
s=1 αi,s ⊗ (αi,j β) ⊗ ⊗t=j+1 αi,t
k
i
= αi,1 ⊗ . . . ⊗ αi,k · p∗j (β)
i
= τk∗ (α) · p∗j (β)
= τk∗ (α · (pj ◦ τk )∗ (β))
= τk∗ (αβ).
k
,
(−1)|β| =j+1 |αi, |
αi,j β · ⊗1≤s≤k,s=j αi,s
i X
" #
|β|· k=j+1 |αi, |
= p̃j∗ (−1) · ⊗j−1
s=1 αi,s ⊗ (αi,j β) ⊗ ⊗kt=j+1 αi,t
i
= p̃j∗ τk∗ (αβ)
= τ(k−1)∗ (αβ)
where we have used the first formula in the second step. This proves the second
formula.
For the third formula, let
Then, we obtain
" j−1 #
⊗s=1 αi,s ⊗ (τu∗ αi,j ) ⊗ ⊗kt=j+1 αi,t
i
= p̂j∗ αi,1 ⊗ . . . ⊗ αi,k
i
= p̂j∗ (τk∗ (α))
= τ(k+u−1)∗ (α).
In the lemma below and throughout the book, the products of Heisenberg
operators are understood to be in the increasing order of the parametrizing indices
from the left to the right, e.g.,
ans = an1 an2 . . . anj−1
1≤s<j
and
anu = an1 . . . ant−1 ant+1 . . . ank .
1≤u≤k,u=t
Proof. To simplify the signs, we will assume that all the cohomology classes
have even degrees.
(i) Let
τk∗ α = αa,1 ⊗ · · · ⊗ αa,k ,
a
τs∗ β = βb,1 ⊗ · · · ⊗ βb,s .
b
52 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
By (3.5), we have
s j−1
s
= am (βb, ) · [an1 · · · ank (τk∗ α), amj (βb,j )] · am (βb, ).
b j=1 =1 =j+1
k ,
= − nt δnt ,−mj · αa,t βb,j · anu (αa,u ).
t=1 a X 1≤u≤k,u=t
Combining with (3.14), we see that [an1 · · · ank (τk∗ α), am1 · · · ams (τs∗ β)] is equal to
k
s j−1
− nt δnt ,−mj · am (βb, )
t=1 j=1 b =1
⎛ ⎞
s
·⎝ anu ⎠ τ(k−1)∗ (αβb,j ) · am (βb, ).
1≤u≤k,u=t =j+1
(ii) Let
τ(k−1)∗ α = αa,1 ⊗ · · · ⊗ αa,k−1 ,
a
τ2∗ αa,j = βa,j,b,1 ⊗ βa,j,b,2
b
3.2. HEISENBERG ALGEBRAS OF NAKAJIMA AND GROJNOWSKI 53
a,b
Note that
βa,j,b,1 βa,j,b,2 = eX αa,j .
b
Therefore,
an · · · ank (τk∗ α)
⎛1 ⎞
= ⎝ ans · anj+1 anj · ans ⎠ (τk∗ α)
1≤s<j j+1<s≤k
,
−nj δnj ,−nj+1 eX αa,j · ans (αa,s ) · ans+1 (αa,s )
a X 1≤s<j j<s≤k−1
⎛ ⎞
= ⎝ ans · anj+1 anj · ans ⎠ (τk∗ α)
1≤s<j j+1<s≤k
−nj δnj ,−nj+1 ans (τ(k−2)∗ (eX α))
1≤s≤k
s=j,j+1
where at the last step, we have used the second formula in Lemma 3.11.
In the following, we will define some convenient operation regarding Heisenberg
monomial classes. It enables us to decompose cohomology classes. In Section 15.3,
we will see that this operation plays an interesting role in describing the structure
of the 3-point genus-0 extremal Gromov-Witten invariants of the Hilbert scheme
X [n] .
Definition 3.13. Let A = a−n1 (α1 ) · · · a−n (α )|0 where n1 , . . . , n > 0.
(i) If B = a−m1 (β1 ) · · · a−ms (βs )|0 with m1 , . . . , ms > 0, then we define
(3.15) A ◦ B = a−n1 (α1 ) · · · a−n (α )a−m1 (β1 ) · · · a−ms (βs )|0 .
54 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
(ii) We use the symbol B ⊂ A if B = a−ni1 (αi1 ) · · · a−nis (αis )|0 with 1 ≤
A
i1 < . . . < is ≤ . In this case, we use A/B or AB −1 or to denote the
B
cohomology class obtained from A by deleting the factors
a−ni1 (αi1 ), . . . , a−nis (αis ).
Note that when B ⊂ A, then we have the identity (A/B) ◦ B = A.
Lemma 3.14. Assume that the odd Betti numbers of X are equal to zero. Let
A = a−n1 (α1 ) · · · a−n (α )|0 where n1 , . . . , n > 0. Let p be a monomial of Heisen-
berg creation operators. Then,
A
(3.16) p† A = p|0 , B · .
B
B⊂A
Proof. Let
p = a−p1 (β1,1 ) · · · a−p1 (β1,k1 ) · · · a−ps (βs,1 ) · · · a−ps (βs,ks )
where 0 < p1 < . . . < ps . We use induction on s. Let s = 1. If p1 = ni for every
i = 1, . . . , , then both sides of (3.16) are equal to 0. Next, let p1 = ni for some
i ∈ {1, . . . , }. We may assume that i = 1 with
p1 = n1 = . . . = nt < nt+1 ≤ nt+2 . . . ≤ n .
By Lemma 3.7 and Theorem 3.8,
p† A
= (−1)k1 p1 ap1 (β1,1 ) · · · ap1 (β1,k1 )a−n1 (α1 ) · · · a−n (α )|0
(−n1 )k1 · β1,1 , ασ(j1 ) · · · β1,k1 , ασ(jk1 ) · A
= (−1)k1 p1
1≤j <...,<j ≤
a−nj1 (αj1 ) · · · a−njk (αjk1 )|0
1 k1 1
σ∈Per{j1 ,...,jk }
1
A
= p|0 , B ·
B
B⊂A
where Per{j1 , . . . , jk1 } denotes the permutation group of {j1 , . . . , jk1 }. Therefore,
(3.16) is true when s = 1. Next, let s > 1 and assume that (3.16) holds for
p̃ = a−p1 (β1,1 ) · · · a−p1 (β1,k1 ) · · · a−ps−1 (βs−1,1 ) · · · a−ps−1 (βs−1,ks−1 ).
†
Note from (3.2) that p† = a−ps (βs,1 ) · · · a−ps (βs,ks ) p̃† . By induction,
p† A
†
= a−ps (βs,1 ) · · · a−ps (βs,ks ) p̃† A
† A
= a−ps (βs,1 ) · · · a−ps (βs,ks ) p̃|0 , B1 ·
B1
B1 ⊂A
A/B1
= a−ps (βs,1 ) · · · a−ps (βs,ks )|0 , B2 · p̃|0 , B1 ·
B2
B1 ⊂A B2 ⊂A/B1
If a−ps (βs,1 ) · · · a−ps (βs,ks )|0 , B2 · p̃|0 , B1 = 0, then B1 and B2 do not share
any common Heisenberg factors, and
a−ps (βs,1 ) · · · a−ps (βs,ks )|0 , B2 · p̃|0 , B1 = p|0 , B1 ◦ B2 .
3.3. GEOMETRIC INTERPRETATIONS OF HEISENBERG MONOMIAL CLASSES 55
k
dimR (W c ) < dimR (W 0 ) = 2ni − 2 + dimR (Xi ) .
i=1
It follows that
k
dimR (W c ) < 2ni − 2 + dimR (Xi ) = dimR (W 0 ).
i=1
the conclusion is trivially true. In the following, assuming that the conclusion is
true whenever i=1 isi + ki=1 ni < n, we will prove that it still holds when
k
isi + ni = n.
i=1 i=1
There are two cases: si > 0 for some i, or ni > 0 for some i. Since the proof of
these two cases are similar, we only prove the first case.
So let si > 0 for some i. Without loss of generality, we may assume that
i = , i.e., s > 0. Denote (3.18) and (3.19) by As and Ws0 respectively. By the
induction hypothesis, As −1 is represented by the closure Ws0 −1 . By Definition 3.3,
a− (1X )(As −1 ) is represented by
" #
] ∩ p∗2 Ws0 −1
[n,n−]
p1∗ [QX
W (1, X; . . . ; 1, X ; . . . ; , X; . . . ; , X ; n1 , X1 ; . . . ; nk , Xk ; , X)
/ 01 2 / 01 2
s1 times s −1 times
si
k
ξ= ξi,j + ξi
i=1 j=1 i=1
where (ξi,j ) = i, (ξi ) = ni , Supp(ξi,j ) = xi,j , Supp(ξi ) = xi , all the points xi,j , xi
are distinct, xi ∈ Xk if and only if i = k, and xi,j ∈ Xk for all i, j, k. Now,
" #
(p1 )−1 (ξ) ∩ QX ∩ p−1
[n,n−]
2 Ws0 −1
Let 1X [n] be the fundamental class of X [n] . Recall the notations Bn and DC
from (1.26) and (1.27) respectively.
58 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
Let π̃1 , π̃2 be the restrictions to Zn of the two natural projections on X [n] × X.
The following lemma and its proof are from [Obe, Lemma 1] (similar statement
also appeared in [HLQ, Lemma 5.1]).
Lemma 3.19. Let n ≥ 2. Let Γ be an irreducible curve in the Hilbert scheme
X [n] . Then,
(3.27) [Γ] = βπ̃2∗ [ZΓ ] + dβn 2 (X [n] ))
(mod H
for some integer d.
Proof. Since π̃1 is flat, π̃2∗ [ZΓ ] = π̃2∗ π̃1∗ [Γ]. By (3.25), it remains to prove
π̃2∗ π̃1∗ βn = 0,
π̃2∗ π̃1∗ βγ = γ,
π̃2∗ π̃1∗ PD a−1 (αj )a−1 (αk )a−1 (x)n−2 |0 = 0
for γ ∈ H2 (X) and |αj | = |αk | = 3. Representing βn by (1.34), we see that
π̃2∗ π̃1∗ βn = 0.
Next, to prove π̃2∗ π̃1∗ βγ = γ, represent the homology class γ by a real smooth
surface S in X. Let y1 , . . . , yn−1 ∈ X be distinct with {y1 , . . . , yn−1 } ∩ S = ∅. Let
U = {(x1 , . . . , xn ) ∈ X n |x1 , . . . xn are mutually distinct},
and let σ : U → X [n] be the morphism sending (x1 , . . . , xn ) to x1 + . . . + xn . Define
= U ×X [n] Zn . Let π̃1 : U
U → U be the natural map. Then we have a diagram of
morphisms:
U → Z
π̃
→2 X
⏐ ⏐n
⏐ ⏐
%π̃1 %π̃1
σ
U → X [n] .
→ Zn →2 X be the composition. Then, βγ = σ∗ [y1 × · · · × yn−1 × S]. So
Let f : U
π̃
Next, we study the homology classes of curves in C (n) ⊂ X [n] where C denotes
a smooth curve in X. Let gC be the genus of C. Assume that gC ≥ 1. We recall
60 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
some standard facts about C (n) from [ACGH, BT]. For a fixed point p ∈ C, let Ξ
denote the divisor p + C (n−1) ⊂ C (n) . Let
AJ : C (n) → Jacn (C)
be the Abel-Jacobi map sending an element ξ ∈ C (n) to the corresponding degree-n
divisor class in Jacn (C). For an element δ ∈ Jacn (C), the fiber AJ−1 (δ) is the
complete linear system |δ|. Let
Zn (C) ⊂ C (n) × C
be the universal divisor, and let π̃1 , π̃2 be the two projections on C (n) × C. By the
Lemma 2.5 on p.340 of [ACGH] and the Proposition 2.1 (iv) of [BT], we have
(3.28) c1 π̃1∗ OZn (C) = (1 − gC − n)Ξ + Θ
where Θ is the pull-back via AJ of a Theta divisor on Jacn (C).
Lemma 3.20. Let n ≥ 2 and X be a smooth projective surface. Let C ⊂ X be
a smooth curve with genus gC ≥ 1, and Γ ⊂ C (n) be a curve. Then, [Γ] is equal to
(3.29) (Ξ · Γ)βC + − (n + gC − 1)(Ξ · Γ) + (Θ · Γ) βn 2 (X [n] )).
(mod H
In addition, for every line Γ0 in a positive-dimensional fiber AJ−1 (δ), we have
(3.30) [Γ0 ] = βC − (n + gC − 1)βn 2 (X [n] )).
(mod H
2 (X [n] ) from (3.24). By Lemma 3.19,
Proof. (i) Recall the subspace H
(3.31) [Γ] = βγ + dβn 2 (X [n] ))
(mod H
for some effective cycle γ and some integer d.
To determine d and γ, let Zn be the universal codimension-2 subscheme of
X [n] × X, and let π1 , π2 be the two projections on X [n] × X. Then,
π1∗ OZn |C (n) = π̃1∗ OZn (C) .
By Theorem 1.24 (i), c1 π1∗ OZn = −Bn /2. Combining with (3.31) and (3.28), we
obtain
(3.32) d = (−Bn /2) · Γ
= c1 π1∗ OZn |C (n) · Γ
= c1 π̃1∗ OZn (C) · Γ
= (1 − gC − n)(Ξ · Γ) + (Θ · Γ).
Next, let α ∈ H (X). Then, Dα |C (n) = (C · α)Ξ. Thus, we get
2
Dα · Γ = (C · α)(Ξ · Γ).
By (3.31),
α · γ = Dα · Γ = (C · α)(Ξ · Γ).
It follows that γ = (Ξ · Γ) C.
(ii) Note that Θ · Γ0 = 0. Also, it is known that Ξ|AJ−1 (δ) = OAJ−1 (δ) (1). So
Ξ · Γ0 = 1. Now (3.30) follows from (3.29).
Remark 3.21. In view of the proof of Lemma 2.19 (i), we conclude that when
the genus gC of the smooth curve C is equal to 0, Lemma 3.20 still holds if we
delete the term (Θ · Γ) in (3.29).
3.5. VIRASORO ALGEBRAS OF LEHN 61
where the class (−Bn /2) acts on HnX = H ∗ (X [n] ) by the cup product. For a linear
operator f ∈ End(HX ), define its derivative f by
f = ad(f) = [d, f].
The higher order derivative f(k) of f is defined inductively by
f(k) = [d, f(k−1) ].
The operators Ln (α) and d and the following theorem are due to Lehn [Leh1,
Theorems 3.3 and 3.10].
Theorem 3.24. Let X be a smooth projective complex surface with the canon-
ical class KX and the Euler class eX . Let n, m ∈ Z and α, β ∈ H ∗ (X). Then,
(i) [Lm (α), an (β)] = −n · am+n (αβ).
n(|n| − 1)
(ii) an (α) = n · Ln (α) − · an (KX α).
2
(iii) The operators Ln (α) satisfy the Virasoro algebra commutation relation:
,
m3 − m
[Lm (α), Ln (β)] = (m − n) · Lm+n (αβ) + δm,−n (eX αβ) · IdHX .
12 X
62 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
Proof. We refer to [Leh1] for the proofs of (ii) and (iii) since the argu-
ments are very technical. To prove (i), assume first that m = 0. By (3.33) and
Lemma 3.12 (i),
1
[Lm (α), an (β)] = − · [ai am−i (τ2∗ α), an (β)]
2
i∈Z
1
= · iδi,−n am−i (αβ) + (m − i)δm−i,−n ai (αβ)
2
i∈Z
1
= · − nam+n (αβ) − nam+n (αβ)
2
= −n · am+n (αβ).
Next, we address the issue of how to identify two linear operators in End(HX ).
We begin with a lemma.
[f, a−1 (1X )] = [f , a−1 (1X )] + [f, a−1 (1X ) ]
= [f, a−1 (1X ) ]
= −[a−1 (1X ) , f]
[a−1 (1X ), [f, a−n (α)]] = [[a−1 (1X ), f], a−n (α)] + [f, [a−1 (1X ), a−n (α)]]
= −[[f, a−1 (1X )] , a−n (α)] + [f, [−L−1 (1X ), a−n (α)]]
= −[[f, a−1 (1X )] , a−n (α)] − [f, na−(n+1) (α)]
where Theorem 3.24 (ii) and (i) have been used. So the lemma holds.
Proof. Recall that the space HX is linearly spanned by the Heisenberg mono-
mial classes (3.10). Therefore, it suffices to show that
We end this section with a lemma which will play an important role in the proof
of Lemma 3.29 below. It enables us to use induction on the order of derivatives.
3.6. HIGHER ORDER DERIVATIVES OF HEISENBERG OPERATORS 63
Lemma 3.27. Let k ≥ 1 and α ∈ H ∗ (X). Then (an1 · · · ank (τk∗ α)) is equal to
k
ni
− · an1 · · · ani−1 : am1 am2 : ani+1 · · · ank (τ(k+1)∗ α)
i=1
2 m1 +m2 =ni
k
ni (|ni | − 1)
− · an1 · · · ank (τk∗ (KX α)).
i=1
2
Proof. Let
τk∗ α = αj,1 ⊗ · · · ⊗ αj,k
j
k i−1
k
= ans (αj,s ) · ani (αj,i ) · ans (αj,s ).
i=1 j s=1 s=i+1
k
ni (|ni | − 1)
i−1 k
− · ans (αj,s ) · ani (KX αj,i ) · ans (αj,s )
i=1
2 j s=1 s=i+1
k
ni
= − · an1 · · · ani−1 : am1 am2 : ani+1 · · · ank (τ(k+1)∗ α)
i=1
2 m1 +m2 =ni
k
ni (|ni | − 1)
− · an1 · · · ank (τk∗ (KX α))
i=1
2
where we have used the third and first formulas in Lemma 3.11 at the last step.
(ii) Define
AX = α ∈ H ∗ (X)|KX α = 0
which is an ideal in the cohomology ring H ∗ (X).
(k)
The following lemma determines an (α) when α ∈ AX . The main idea in its
proof is to use induction on k, Lemma 3.27, and Lemma 3.12 (ii).
(k)
Lemma 3.29. Let k ≥ 0, n ∈ Z, and α ∈ AX . Then, an (α) is equal to
⎛ ⎞
1 s(λ) − 1
(−n)k k! ⎝ aλ (τ∗ α) − aλ (τ∗ (eX α))⎠ .
λ! 24λ!
(λ)=k+1,|λ|=n (λ)=k−1,|λ|=n
(k)
for some constant dμ . By induction, the coefficient of aλi,j (τ∗ α) in an (α) is
(k+1)
Combining all of these, we see that the coefficient of aλ (τ∗ α) in an (α) is
(−n)k k! mi (mj − δi,j )
i+j
· − (mi+j + 1)(2 − δi,j )
λ! mi+j + 1 2
i≤j
⎛ ⎞
(−n)k k! ⎝
= − mi mj (i + j) − mi (mi − 1)i⎠
λ! i<j i
⎛ ⎞
(−n)k k! ⎝ 1
= − mi mj (i + j) + mi i⎠
λ! 2 i,j i
(−n)k+1 (k + 1)!
=
λ!
where we have used i mi = (λ) = k + 2 and i imi = |λ| = n in the last step.
(ii) Before going into the proof, let us use one example to illustrate the idea.
Take k = 2, n = −3, and λ = ((−5)2). To figure out the coefficient of aλ (τ∗ (eX α))
(3) (2)
in a−3 (α), we apply the induction hypothesis to a−3 (α) and take the derivative
one more time. In view of Lemma 3.12 (ii) and Lemma 3.27, there are exactly two
ways to get aλ (τ∗ (eX α)). The first one is similar to (i) above, and comes from the
term {aλ−5,2 (τ∗ (eX α))} where λ−5,2 = ((−3)), i.e., the generalized partition λ−5,2
is obtained from λ by subtracting 1 from the multiplicities of (−5) and 2 and by
adding 1 to the multiplicity of (−3) = (−5) + 2. The second source comes from
three terms {aλ̃ (τ∗ α)} where λ̃ = ((−5)12 ), ((−5 + j)(−j)2) with j = 1, 2. For
instance, when λ̃ = ((−5 + j)(−j)2) with j = 1 or 2, {aλ̃ (τ∗ (α))} contains
a−5 aj a−j a2 (τ4∗ α) = a−5 a−j aj a2 (τ4∗ α) + (−j)a−5 a2 (τ2∗ (eX α))
= a−5 a−j aj a2 (τ4∗ α) + (−j)aλ (τ∗ (eX α)).
Keeping track of the coefficients in these derivatives
{aλ−5,2 (τ∗ (eX α))} , {aλ̃ (τ∗ α)} ,
(3)
we obtain the coefficient of aλ (τ∗ (eX α)) in a−3 (α).
Now we go back to the proof and compute f˜eX (λ). In this case, (λ) = k. We
(k+1)
will prove that the coefficient of aλ (τ∗ (eX α)) in an (α) is equal to
(−n)k+1 (k + 1)!
·c
λ!
where
1 − s(λ) 1 − i i2 mi
c= = .
24 24
In view of Lemma 3.27 and Lemma 3.12 (ii), there are exactly two sources con-
(k+1)
tributing to the term aλ (τ∗ (eX α)) in an (α) from the derivatives of the terms in
(k)
an (α). In the following, we handle these two sources separately.
The first is similar to (i) above, and comes from the derivatives {aλi,j (τ∗ (eX α))}
where i ≤ j, mi ≥ 1, mj ≥ 1. For fixed i and j, {aλi,j (τ∗ (eX α))} is equal to
i+j
− (mi+j + 1)(2 − δi,j )aλ (τ∗ (eX α)) + dμ aμ (τ∗ (eX α))
2 μ=λ
(μ)=k,|μ|=n
66 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
(k)
for some dμ . By induction, the coefficient of aλi,j (τ∗ (eX α)) in an (α) is
(−n)k k! 1 − s(λi,j ) (−n)k k! mi (mj − δi,j ) ij
· = · · c − .
(λi,j )! 24 λ! mi+j + 1 12
Combining these and noting i mi = k and i imi = n, we see that the total
(k+1)
contribution of the first source to the coefficient of aλ (τ∗ (eX α)) in an (α) is
The second source comes from the derivatives {aλ̃i,j (τ∗ α)} where i = 0, 0 <
2j ≤ |i|, mi ≥ 1, and λ̃i,j is defined as follows. There are two cases, depending on
whether i > 0 or i < 0. For i > 0, define λ̃i,j to be the generalized partition obtained
from λ by subtracting 1 from the multiplicity of i and by adding 1 to the multiplic-
ities of j and (i − j). Then, {aλ̃i,j (τ∗ α)} contains a term (. . . aj . . . ai−j . . .) which
contains a term (. . . aj . . . a−j ai . . .) after expanding the derivative ai−j . When we
(k+1)
switch aj with a−j , we get the term aλ (τ∗ (eX α)) in an (α). For i < 0, define λ̃i,j
to be the generalized partition obtained from λ by subtracting 1 from the multiplic-
ity of i and by adding 1 to the multiplicities of −j and (i + j). Then {aλ̃i,j (τ∗ α)}
contains a term (. . . ai+j . . . a−j . . .) which contains a term (. . . ai aj . . . a−j . . .) after
expanding ai+j . Switching aj with a−j , we get aλ (τ∗ (eX α)) in an
(k+1)
(α).
More precisely, for fixed i > 0 and j, the derivative {aλ̃i,j (τ∗ α)} is
(k)
for some constant dμ . By induction, the coefficient of aλ̃i,j (τ∗ (eα)) in an (α) is
(−n)k k! (−n)k k! mi
= !
· .
(λ̃i,j ) ! λ (m i−j + 1)(m j + 1 + δi,2j )
3.6. HIGHER ORDER DERIVATIVES OF HEISENBERG OPERATORS 67
It follows that the total contribution of the second source with i > 0 to the coeffi-
(k+1)
cient of the term aλ (τ∗ (eX α)) in an (α) is equal to
(−n)k k! mi
(3.38) !
· ·
i>0
λ (mi−j + 1)(mj + 1 + δi,2j )
0<2j≤i
(mi−j + 1)(mj + 1 + δi,2j )
·j(i − j) ·
1 + δi,2j
(−n)k k! j(i − j)
= mi
λ! i>0
1 + δi,2j
0<2j≤i
⎛ ⎞
(−n)k k! 1
= mi ⎝ j(i − j)⎠
λ! i>0
2 0<j<i
(−n)k k! i3 − i
= m i ·
λ! i>0
12
(−n)k k! 1 3
= · (i mi − imi ).
λ! 12 i>0
Similarly, we see that the total contribution of the second source with i < 0 to the
(k+1)
coefficient of aλ (τ∗ (eX α)) in an (α) is equal to
(−n)k k! 1 3
(3.39) · (i mi − imi ).
λ! 12 i<0
Next, we see from (3.37), (3.38) and (3.39) that the coefficient of the term
(k+1)
aλ (τ∗ (eX α)) in the derivative an (α) is equal to
(−n)k k! n 2 1 3
nc − knc + i mi − i mi
λ! 12 i 12 i
(−n)k k! 1 3
+ · (i mi − imi ).
λ! 12 i
(k+1)
Since n = i imi , the coefficient of the term aλ (τ∗ (eX α)) in an (α) is equal to
(−n)k k! n 2 n
nc − knc + i mi −
λ! 12 i 12
(−n)k k!
= (nc − knc − 2nc)
λ!
1 − s(λ)
= (−n)k+1 (k + 1)! .
24λ!
(k+1)
Combining (i) and (ii), we have proved the lemma for an (α).
68 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
Remark 3.30. Let α ∈ AX . From the proof of Lemma 3.29, we can read that
⎛ ⎞
1 s(λ) − 1
⎝ aλ (τ∗ α) − aλ (τ∗ (eX α))⎠
λ! 24λ!
(λ)=k+1,|λ|=n (λ)=k−1,|λ|=n
⎛ ⎞
1 s(λ) − 1
= −n(k + 1) ⎝ aλ (τ∗ α) − aλ (τ∗ (eX α))⎠ .
λ! 24λ!
(λ)=k+2,|λ|=n (λ)=k,|λ|=n
Note that when n = 0, this formula is not covered by Lemma 3.29. Applying this
formula twice in the second equality below, we obtain that for a fixed constant d,
⎛ ⎞
1 s(λ) + d
⎝ aλ (τ∗ α) − aλ (τ∗ (eX α))⎠
λ! 24λ!
(λ)=k+1,|λ|=n (λ)=k−1,|λ|=n
⎛ ⎞
1 s(λ) − 1
= ⎝ aλ (τ∗ α) − aλ (τ∗ (eX α))⎠
λ! 24λ!
(λ)=k+1,|λ|=n (λ)=k−1,|λ|=n
⎛ ⎞
d+1⎝ 1
− aλ (τ∗ (eX α))⎠
24 λ!
(λ)=k−1,|λ|=n
⎛ ⎞
1 s(λ) − 1
= −n(k + 1) ⎝ aλ (τ∗ α) − aλ (τ∗ (eX α))⎠
λ! 24λ!
(λ)=k+2,|λ|=n (λ)=k,|λ|=n
n(k − 1)(d + 1) 1
+ · aλ (τ∗ (eX α)).
24 λ!
(λ)=k,|λ|=n
Therefore, we have
⎛ ⎞
1 s(λ) + d
⎝ aλ (τ∗ α) − aλ (τ∗ (eX α))⎠
λ! 24λ!
(λ)=k+1,|λ|=n (λ)=k−1,|λ|=n
⎛ ⎞
1 s(λ) + d
= −n(k + 1) ⎝ aλ (τ∗ α) − aλ (τ∗ (eX α))⎠
λ! 24λ!
(λ)=k+2,|λ|=n (λ)=k,|λ|=n
n(d + 1) 1
− · aλ (τ∗ (eX α)).
12 λ!
(λ)=k,|λ|=n
Proof. We see from Lemma 3.27 and Lemma 3.12 (ii) that
f (λ)
(3.40) a(k)
n (α) = aλ (τ∗ (α))
λ!
∈{1X ,eX ,KX ,KX } (λ)=k+1−| |/2,|λ|=n
2
Comparing with (3.40) and noting that the set of all the Heisenberg monomials
aλ (τ∗ α), aλ̃ (τ∗ (eX α)) with (λ) = k + 1, |λ| = n, (λ̃) = k − 1, |λ̃| = n is linearly
independent, we see that f1X (λ) = (−n)k k! and
(−n)k k!(s(λ) − 1)
feX (λ) = − .
24
This completes the proof of the theorem.
Proof. We will show that the coefficients of q n on both sides of (3.44) are
equal. Let {ej }j be a linear basis of H ∗ (X [n] ). Then the fundamental class of the
diagonal Δn in X [n] × X [n] is given by
(3.45) [Δn ] = (−1)|ej | ej ⊗ e∗j
j
where {e∗j }j is the linear basis of H ∗ (X [n] ) dual to {ej }j in the sense that ej , e∗j =
δj,j . By the definition of W(Lm , z),
Tr q n W(Lm , z) = q n (−1)|ej | W(Lm , z)ej , e∗j
j
,
|ej |
= q n
(−1) (ej ⊗ e∗j ) c2n (ELm )
j X [n] ×X [n]
,
= qn [Δn ] c2n (ELm )
X [n] ×X [n]
,
= qn c2n (ELm |Δn ).
Δn
The following two results are proved in [Car1,CO], but we follow the notations
in [Car1, Theorem 1] adjusted in view of Remark 3.34.
Theorem 3.36. Let L be a line bundle over the smooth surface X. Then,
(3.49) W(L, z) = Γ− (L − KX , z) Γ+ (−L, z).
We outline the proof of Theorem 3.36, and refer to [Car1,Car2,CO] for further
details. The first step is to use certain universality results and reduce to the special
case of the equivariant cohomology of a smooth toric surface X. Then for a smooth
toric surface X, choose a torus action on X such that the fixed points are isolated.
Some decomposition argument with respect to the isolated fixed points further
reduces the proof to the cases of X = C2 and P1 × P1 . Finally, the cases of X = C2
and P1 × P1 can be verified directly by using equivariant localization.
72 3. HILBERT SCHEMES AND INFINITE DIMENSIONAL LIE ALGEBRAS
Corollary 3.37. Let TX [n] ,m and Lm be the tangent bundle TX [n] and the
trivial line bundle on X respectively with a scaling action of C∗ of character z m .
Let q be a real number with |q| < 1. Then,
+∞
Lm ,KX −Lm −χ(X)
(3.50) q n χ TX [n] ,m = (q; q)∞
n=0
where (a; q)n = (1 − a)(1 − aq) · · · (1 − aq n ).
Proof. By Lemma 3.33, it suffices to prove that
Lm ,KX −Lm −χ(X)
(3.51) Tr q n W(Lm , z) = (q; q)∞ .
By Theorem 3.36, we obtain
Tr q n W(Lm , z) = Tr q n Γ− (Lm − KX , z) Γ+ (−Lm , z)
= Tr Γ− (Lm − KX , zq) q n Γ+ (−Lm , z)
= Tr q n Γ+ (−Lm , z) Γ− (Lm − KX , zq).
By (3.48) and since |q| < 1, we get
Tr q n W(Lm , z)
= (1 − q)Lm ,KX −Lm Tr q n Γ− (Lm − KX , zq) Γ+ (−Lm , z)
Lm ,KX −Lm
= (q; q)∞ Tr q n Γ− (Lm − KX , 0) Γ+ (−Lm , z)
Lm ,KX −Lm
= (q; q)∞ Tr q n
Lm ,KX −Lm −χ(X)
= (q; q)∞
where we have used (1.22) in the last step.
CHAPTER 4
When γ = 1X , we have
G(1X , n) = p1∗ (ch(OZn ) · p∗2 td(X)).
So we see from the Grothendieck-Riemann-Roch Theorem [Hart, Theorem 5.3 in
Appendix A] that
(4.3) G(1X , n) = ch(p1∗ OZn ) = ch (OX )[n] .
Recall that (OX )[n] is a locally free sheaf of rank-n on X [n] . By Theorem 1.24 (i),
we have c1 (OX )[n] = −Bn /2. It follows that
(4.4) G0 (1X ) = G0 (1X , n) ∪ = (n · IdX [n] ) = L0 (1X ),
n n
(4.5) G1 (1X ) = G1 (1X , n) ∪ = c1 (p1∗ OZn ) ∪ = d.
n n
Proof. For i ≥ 1, let pi,1 and pi,2 be the projections of X [i] × X to X [i] and
X respectively. From the definition of G(γ, n), we see that
gn∗ G(γ, n) = gn∗ (pn,1 )∗ (ch(OZn ) · (pn,2 )∗ td(X) · (pn,2 )∗ γ)
∗
= p1∗ gX (ch(OZn ) · (pn,2 )∗ td(X) · (pn,2 )∗ γ)
∗
= p1∗ (gX ch(OZn ) · p∗2 td(X) · p∗2 γ)
where p2 is the projection of X [n,n−1] × X to X. Similarly,
fn∗ G(γ, n − 1) = p1∗ (fX
∗
ch(OZn−1 ) · p∗2 td(X) · p∗2 γ).
So combining these with (4.8) above, we conclude that
gn∗ G(γ, n) − fn∗ G(γ, n − 1)
∗ ∗
= p1∗ ((gX ch(OZn ) − fX ch(OZn−1 )) · p∗2 td(X) · p∗2 γ)
= p1∗ (ρ∗X ch(OΔX ) · p∗1 exp(−[En ]) · p∗2 td(X) · p∗2 γ)
= p1∗ (ρ∗X ch(OΔX ) · p∗2 td(X) · p∗2 γ) · exp(−[En ])
= p1∗ ρ∗X (ch(OΔX ) · (p1,2 )∗ td(X) · (p1,2 )∗ γ) · exp(−[En ])
= ρ∗ (p1,1 )∗ (ch(OΔX ) · (p1,2 )∗ td(X) · (p1,2 )∗ γ) · exp(−[En ]).
Applying the Grothendieck-Riemann-Roch Theorem to the diagonal embedding
τ2 : X → X × X, we get
ch(OΔX ) · (p1,2 )∗ td(X) = τ2∗ [X].
So
(p1,1 )∗ (ch(OΔX ) · (p1,2 )∗ td(X) · (p1,2 )∗ γ) = (p1,1 )∗ (τ2∗ [X] · (p1,2 )∗ γ) = γ.
It follows that
gn∗ G(γ, n) − fn∗ G(γ, n − 1) = ρ∗ γ · exp(−[En ]).
This completes the proof of the claim.
We continue the proof of the theorem. By Definition 3.3, for a ∈ H ∗ (X [n−1] ),
a−1 (α)(a) = p̃1∗ ([Q[n,n−1] ] · ρ̃∗ α · p̃∗2 a).
Define ι : X [n,n−1] → X [n] × X × X [n−1] by putting ι(ξ, η) = (ξ, ρ(ξ, η), η). Then,
ι is an embedding. Moreover, ι∗ [X [n,n−1] ] = [Q[n,n−1] ]. Thus, we have
a−1 (α)(a) = p̃1∗ (ι∗ [X [n,n−1] ] · ρ̃∗ α · p̃∗2 a)
= p̃1∗ ι∗ ([X [n,n−1] ] · (ι∗ ◦ ρ̃∗ )α · (ι∗ ◦ p̃∗2 )a)
= gn∗ ([X [n,n−1] ] · ρ∗ α · fn∗ a).
Combining this with the above Claim, we conclude that
G(γ)a−1 (α)(a) = G(γ, n) · gn∗ ([X [n,n−1] ] · ρ∗ α · fn∗ a)
= gn∗ (gn∗ G(γ, n) · [X [n,n−1] ] · ρ∗ α · fn∗ a)
= gn∗ (fn∗ G(γ, n − 1) · [X [n,n−1] ] · ρ∗ α · fn∗ a)
+gn∗ (ρ∗ γ · exp(−[En ]) · [X [n,n−1] ] · ρ∗ α · fn∗ a)
= gn∗ ([X [n,n−1] ] · fn∗ G(γ, n − 1)ρ∗ α · fn∗ a)
+gn∗ ((exp(−[En ]) · [X [n,n−1] ]) · ρ∗ (γα) · fn∗ a).
76 4. CHERN CHARACTER OPERATORS
Thus,
[G(γ), a−1 (α)](a) = gn∗ ((exp(−[En ]) · [X [n,n−1] ]) · ρ∗ (γα) · fn∗ a)
which is equal to exp(ad(d))(a−1 (γα))(a) by the Lemma 3.9 in [Leh1].
Recall from Definition 3.28 (ii) that
AX = α ∈ H ∗ (X)|KX α = 0 .
The next lemma determines Gk (α) when α ∈ AX .
Lemma 4.3. Let k ≥ 0, and α ∈ AX . Then, Gk (α) is equal to
1 s(λ) − 2
(4.9) − !
aλ (τ∗ α) + aλ (τ∗ (eX α)).
λ 24λ!
(λ)=k+2,|λ|=0 (λ)=k,|λ|=0
So our theorem is true for k = |α| = 0. In the following, assume k + |α| > 0.
we write Gk (α) ∈ End(HX ) as a linear combination
Next,
of Heisenberg mono-
mials (ai (αi,1 ) · · · ai (αi,mi )). For degree reasons, i imi = 0 and
i
(4.10) −2 mi + |αi,j | = 2k + |α|.
i i,j
where for each fixed generalized partition λ = (· · · (−2)m−2 (−1)m−1 1m1 2m2 · · · )
with |λ| = 0, the operator
gλ stands for the component in Gk (α) containing all the
expressions of the form (ai (αi,1 ) · · · ai (αi,mi )).
i
Fix n ∈ Z and β ∈ H ∗ (X). By Lemma 4.5 and Lemma 3.12 (ii),
(4.12) [Gk (α), an (β)] = d(, μ)aμ (τ∗ (αβ))
∈{1X ,KX ,K 2 ,eX }
X
(μ)=k+1−||/2,|μ|=n
where all the coefficients d(, μ) ∈ Q are independent of the surface X and the
cohomology classes α, β ∈ H ∗ (X). It follows that the generalized partitions λ in
(4.11) must satisfy k ≤ (λ) ≤ (k + 2). Moreover, for a fixed generalized partition
λ with |λ| = 0 and k ≤ (λ) ≤ (k + 2), and for every i with m−i > 0, we have
⎧
⎨ d(1X , λ−i )aλ−i (τ∗ (αβ)), if (λ) = k + 2
(4.13) [gλ , ai (β)] = d(KX , λ−i )aλ−i (τ∗ (KX αβ)), if (λ) = k + 1
⎩
aλ−i (τ∗ ω), if (λ) = k
where λ−i is the generalized partition obtained from λ by subtracting 1 from
2 2
the multiplicity of (−i), and ω = d(eX , λ−i )eX + d(KX , λ−i )KX αβ. Note that
d(, λ−i ) is independent of X, α and β.
By choosing special surfaces X and α, β ∈ H ∗ (X), we now claim that
2
d(1X , λ−i ) d(KX , λ−i ) d(eX , λ−i ) d(KX , λ−i ) 2
(4.14) , , eX + KX
im−i im−i im−i im−i
are independent of i as long as m−i > 0. Indeed, to prove that the 3rd item in (4.14)
is independent of i as long as m−i > 0, we may assume that gλ = 0 by (4.13). Note
that |λ| = 0 and (λ) = k. By (4.10), i,j |αi,j | = 4k + |α| for every Heisenberg
monomial (ai (αi,1 ) · · · ai (αi,mi )) contained in gλ . On the other hand, |αi,j | ≤ 4
i
and thus i,j |αi,j | ≤ 4 (λ) = 4k. It follows that |α| = 0 and |αi,j | = 4 for every
i and j. So gλ = b · aλ (τ∗ [x])where b is a nonzero number, and [x] ∈ H 4 (X) is
4.1. CHERN CHARACTER OPERATORS 79
We claim that Gk (α) = f. Indeed, for every n ∈ Z − {0} and β ∈ H ∗ (X), we have
[f, an (β)]
d(, λ−i )
= nm−n · aλ−n (τ∗ (αβ))
2 ,e } (λ)=k+2−| |/2,|λ|=0
im−i
∈{1X ,KX ,KX X
= d(, λ−n ) · aλ−n (τ∗ (αβ))
2 ,e } (λ)=k+2−| |/2,|λ|=0
∈{1X ,KX ,KX X
Finally, since the numbers d(, λ−i )/(im−i ) are independent of X and α ∈
H ∗ (X), by using Lemma 4.3 and an argument similar to that in the proof of
Theorem 3.31, we conclude that d(1X , λ−i )/(im−i ) = −1/λ! and
d(eX , λ−i ) s(λ) − 2
= .
im−i 24λ!
In particular, d(eX , λ−i )/(im−i ) is independent of i whenever m−i > 0. So by
(4.14), d(K 2 , λ−i )/(im−i ) is also independent of i whenever m−i > 0. Now denoting
d(, λ−i )/(im−i ) by g (λ)/λ! for = KX and KX 2
, we have proved the theorem.
It turns out that the operator G1 (α) can be determined completely. To state
the result, we need the notation : a3 :0 (τ3∗ α). So fix α ∈ H ∗ (X). Define a vertex
operator (i.e. a field) a(α)(z) by putting
(4.15) a(α)(z) = an (α)z −n−1 .
n∈Z
80 4. CHERN CHARACTER OPERATORS
If
τp∗ (α) = αi,1 ⊗ αi,1 ⊗ . . . ⊗ αi,p ∈ H ∗ (X)⊗p ,
i
then we define
: a(z)p : (τ∗ α) = : a(αi,1 )(z)a(αi,2 )(z) · · · a(αi,p )(z) : .
i
the m-th Fourier component of the field : a(z)p : (τ∗ α)), and maps H ∗ (X [n] ) to
H ∗ (X [n+m] ). We have
(4.17) : a3 :0 (τ3∗ α) = a−k a− ak+ + a−k− ak a (τ3∗ α).
k,>0
Proof. Observe that both sides of (4.18) annihilate the vacuum vector |0 . To
prove (4.18), it suffices to show that the commutators of both sides of (4.18) with the
operators an (β), n ∈ Z, β ∈ H ∗ (X), coincide. By Lemma 4.5 and Theorem 3.24 (ii),
[G1 (α), an (β)] = [G1 (1X ), an (αβ)]
= an (αβ)
n(|n| − 1)
= n · Ln (αβ) − · an (KX αβ).
2
Using (4.17) and Lemma 3.12 (i), we can easily compute the commutator of the
right hand side of (4.18) with an (β). These two commutators indeed coincide.
In particular, setting α = 1X , we obtain from (4.18) that
1 n−1
(4.19) d = G1 (1X ) = − : a3 :0 (τ3∗ 1X ) − : an a−n : (τ2∗ KX ).
6 n>0
2
This is equivalent to Theorem 3.24 (ii) of Lehn.
By the definition of Chern character operators (Definition 4.1 (iii)), we have
Gk (α, n) = Gk (α)1X [n] where 1X [n] denotes the fundamental class of the Hilbert
scheme X [n] . It follows that the cohomology class Gk (α, n) can be understood via
Theorem 4.7. This leads us to the next section.
where for each fixed i with 0 ≤ i ≤ a, σi runs over all the maps
{ 1, . . . , i } → { 1, . . . , b }
satisfying σi (1) < · · · < σi (i). Moreover,
σi0 = { | 1 ≤ ≤ b, = σi (1), . . . , σi (i)},
σa1 = { | 1 ≤ < σa (a), = σa (1), . . . , σa (a)},
σa2 = { | σa (a) < ≤ b}.
Proof. Note that for all i with 0 ≤ i < a and for all the above σi , we move
[· · · [g, a−nσi (1) (βσi (1) )], · · · ], a−nσi (i) (βσi (i) )]
all the way to the right. This produces (4.20). In doing so, we obtain (4.21) by
repeatedly applying the fact that
g1 g2 = [g1 , g2 ] + (−1)s1 s2 g2 g1
for two operators g1 , g2 ∈ End(HX ) of bi-degrees ( 1 , s1 ), ( 2 , s2 ) respectively.
Theorem 4.10. Let n ≥ 1, k ≥ 0, and α ∈ H ∗ (X). Then, Gk (α, n) is equal to
(−1)|λ|−1
· 1−(n−j−1) a−λ (τ∗ α)|0
λ! · |λ|!
0≤j≤k
λ(j+1)
(λ)=k−j+1
(−1)|λ| g (λ + (1j+1 ))
+ · 1−(n−j−1) a−λ (τ∗ (α))|0
λ! · |λ|!
∈{KX ,K 2 } λ(j+1)
X (λ)=k−j+1−||/2
0≤j≤k
where the universal function g is from Theorem 4.7, and λ + (1j+1 ) is the partition
obtained from λ by adding (j + 1) to the multiplicity of 1.
Proof. By (3.21),
1
· a−1 (1X )n |0 .
1X [n] =
n!
From the definition of Chern character operators, we see that Gk (α)|0 = 0 and
Gk (α, n) = Gk (α)1X [n] . Also, by Theorem 4.7 and Lemma 3.12 (i), we have
[. . . [Gk (α), an1 (α1 )], . . .], ank+2 (αk+2 )] = 0
82 4. CHERN CHARACTER OPERATORS
j
s
απi = sign(α, π) · αi .
i=1 i=1
m
i −ri
0 < ni,1 ≤ . . . ≤ ni,mi −ri , ni,j ≤ (kj + 1) for each i, and
j=1 j∈πi
⎛ ⎞
(π) i −ri
m
s
(4.25) ⎝mi − 2 + ni,j ⎠ = ki .
i=1 j=1 i=1
⎛ ⎛ ⎞ ⎞
(σ) i −ri
m
sign(α, σ) · 1−(n−ñ) ⎝ ⎝ a−ni,j ⎠ (τ(mi −ri )∗ (i ασi ))⎠ |0
i=1 j=1
⎛ ⎛ ⎞ ⎞
m i −ri(σ)
sign(α, σ)
= · a−1 (1X )n−ñ ⎝ ⎝ a−ni,j ⎠ (τ(mi −ri )∗ (i ασi ))⎠ |0
(n − ñ)! i=1 j=1
ri = |i |/2 ≤ mi ≤ 2 + kj ,
j∈σi
m
i −ri m
(σ) i −ri
0 < ni,1 ≤ . . . ≤ ni,mi −ri , ni,j ≤ (kj + 1), and ñ = ni,j . More-
j=1 j∈σi i=1 j=1
over, the coefficients in the linear combination are independent of X, α2 , . . . , αs and
s
n. Now apply Gk1 (α1 ) to Gki (αi , n), and move Gk1 (α1 ) to the right by using
i=2
Lemma 4.9. Note that |τ(mi −ri )∗ (i ασi )| ≡ |ασi | (mod 2). By Theorem 4.7,
s
s
Gki (αi , n) = Gk1 (α1 ) Gki (αi , n) ,
i=1 i=2
84 4. CHERN CHARACTER OPERATORS
s
we see from Lemma 4.9 that Gki (αi , n) is a universal linear combination of
i=1
expressions:
|α1 |
|ασi |+
u
|ασi ||ασw |
sign(α, σ) n − ñ v
(4.26) · (−1) 1≤i≤(σ),i∈U v=1 w>iv ,w∈U
·
(n − ñ)! t
⎛ ⎛ ⎞ ⎞
i −ri
m
⎜ ⎝ ⎟
·a−1 (1X )n−ñ−t ⎝ a−ni,j ⎠ (τ(mi −ri )∗ (i ασi ))⎠ ·
1≤i≤(σ) j=1
i∈U
⎡
mi1 −ri1
⎣
· · · · [Gk1 (α1 ), a−1 (1X )], · · · ], a−1 (1X )], a−ni1 ,j (τ(mi1 −ri1 )∗ (i1 ασi1 ))],
/ 01 2
j=1
t times
⎛ ⎞ ⎤
miu −riu
· · · ], ⎝ a−niu ,j ⎠ (τ(miu −riu )∗ (iu ασiu ))⎦ |0
j=1
s
In view of (4.26) and (4.27), Gki (αi , n) is a linear combination of
i=1
⎛ ⎛ ⎞ ⎞
i −ri
m
(4.28) sign(α, π) · 1−(n−ñ−t) ⎝ ⎝ a−ni,j ⎠ (τ(mi −ri )∗ (i απi ))⎠ ·
1≤i<(π) j=1
⎡
mi1 −ri1
⎣
· · · · [Gk1 (α1 ), a−1 (1X )], · · · ], a−1 (1X )], a−ni1 ,j (τ(mi1 −ri1 )∗ (i1 ασi1 ))],
/ 01 2
j=1
t times
⎛ ⎞ ⎤
miu −riu
· · · ], ⎝ a−niu ,j ⎠ (τ(miu −riu )∗ (iu ασiu ))⎦ |0 .
j=1
with all the coefficients being independent of X, α1 , . . . , αs and n. Also notice that
in the expression (4.28), the only factor depending on n is 1−(n−ñ−t) .
4.2. CHERN CHARACTERS 85
∈ {1X , KX , KX
2
, eX }, r = ||/2 ≤ m, 0 < n1 ≤ . . . ≤ nm−r , and
i −ri
m
n1 + . . . + nm−r = t + ni,j ≤ (kj + 1)
i∈U j=1 j∈π(π)
The following lemma determines the leading term in the cup product
s
Gki (αi , n).
i=1
s
Lemma 4.14. Let notations be the same as in Lemma 4.13, and n0 = (ki +1).
i=1
m
(π) i −ri
(i) If an expression from ( 4.24) satisfies ni,j = n0 , then it is equal
i=1 j=1
to
s
1−(n−n0 ) a−(ki +1) (αi ) · |0 .
i=1
s
s
(ii) The coefficient of 1−(n−n0 ) a−(ki +1) (αi ) · |0 in Gki (αi , n) is
i=1 i=1
s
(−1)ki
.
i=1
(ki + 1)!
where |πi | stands for the number of elements in the subset πi . Note that for every
i with 1 ≤ i ≤ (π), we have (mi − ri ) ≥ 1 since
i −ri
m
ni,j = (kj + 1) ≥ 1.
j=1 j∈πi
(ii) The idea is to use induction on s and track the proof of Lemma 4.13 more
carefully. When s = 1, the statement is true by Theorem 4.10. Next, let s≥ 2
s s
and ñ0 = (kj + 1). Assume that the coefficient of 1−(n−ñ0 ) a−ni (αi ) |0
j=2 i=2
s s
(−1)ki
in the cup product Gki (αi , n) is equal to . Tracking the proof
i=2 i=2
(ki + 1)!
of Lemma 4.13 and applying Theorem 4.7, we conclude that the coefficient of
s s s
(−1)ki
1−(n−n0 ) a−(ki +1) (αi ) |0 in Gki (αi , n) is equal to .
i=1 i=1 i=1
(ki + 1)!
We now apply Lemma 4.13 to study certain intersection numbers in the Hilbert
scheme X [n] . For this purpose, we establish the notation
,
(4.30) ω1 , . . . , ωk = ω1 · · · ωk .
Y
(π)
is a finite linear combination of sign(α, π) · i απi where π runs over all par-
i=1
titions of {1, . . . , s}, i ∈ {1X , KX , KX
2
, eX }. Moreover, all the coefficients in this
linear combination are independent of X, α1 , . . . , αs and n.
Proof. Note that a−1 (x)n |0 is a generator of H 4n (X [n] ) ∼
< C.
= So in view =
s
of Lemma 4.13, an expression (4.24) nontrivially contributing to Gki (αi , n)
i=1
must satisfy:
(i) ni,j = 1 for all 1 ≤ i ≤ (π) and 1 ≤ j ≤ mi − ri ,
(ii) i απi = i απi · x for all 1 ≤ i ≤ (π), and
(π)
(iii) n − (mi − ri ) = 0.
i=1
Since
/ ⊗ ·01
τk∗ (x) = x · · ⊗ x2
k times
for all k ≥ 0, our conclusion follows immediately from (4.24).
s
Remark 4.16. Assume that (ki + 2) = 2n. By Theorem 4.2,
i=1
Reading the top Chern classes cn L[n] from (4.34), we recover the following formula
of Nakajima [Nak5, Formula (9.16)]:
⎛ ⎞
+∞
[n] (−1)r−1
(4.35) C = exp ⎝ a−r ([C])⎠ · |0 .
n=0
r
r≥1
Proof. We modify the proof of (4.34) in [Leh1] by inserting the formal vari-
able in suitable positions. First of all, for a class u ∈ K(X), put
c (u) = ci (u)i .
i
In particular, [a−1 (1X ), R (w)] commutes with R (w). It follows that
[a−1 (1X ), R (w)m ] = mR (w)m−1 [a−1 (1X ), R (w)].
Therefore, we have
[C (L), S (w)] = [a−1 (1X ), exp (R (w))]
= exp (R (w)) · [a−1 (1X ), R (w)]
= S (w) · (−)r a−r−1 (c (L))wr
r≥1
where we have used (4.42) in the last step. Combining with a−1 (1X )|0 = 0, we
obtain
C (L)S (w)|0
= S (w)C (L)|0 + S (w) · (−)r a−r−1 (c (L))wr · |0
r≥1
= S (w)a−1 (c (L))|0 + S (w) · (−)r a−r−1 (c (L))wr · |0
r≥1
= S (w) · (−) a−r−1 (c (L))wr · |0
r
r≥0
d
= S (w)|0
dw
where we have used (4.41) in the last step. Hence, S (w)|0 is also a solution to
(4.40), and is equal to (4.39).
Corollary 4.20. Let L be a line bundle on the surface X, and let L[n]∗ be the
dual of L[n] . Then,
⎛ ⎞
+∞ 1
(4.43) c(L[n]∗ )wn = exp ⎝ a−r (c(L∗ ))wr ⎠ · |0 .
n=0
r
r≥1
and similar generating series for the tangent bundles TX [n] of the Hilbert schemes
X [n] have been further investigated in [Boi1, BN, QY].
(here Z+ stands for the set of all nonnegative integers). We assume that A affords an
algebra structure which is super commutative and compatible with the Z2 -grading.
We further assume that there exists a linear operator Tr : A → C which is zero on
the graded subspace An unless n is the top degree. This defines a supercommutative
bilinear form ·, · : A × A → C by
α, β = Tr(αβ).
d
Let t be an indeterminate and let ∂t = . Let Was be the associative algebra
dt
of regular differential operators on the circle S . Denote by W(A)as the associative
1
where D = t∂t . Note that f (D)t = tf (D + 1) for every f (w) ∈ C[w]. Hence,
Jkp (α) = −tk D(D − 1) · · · (D − p + 1) ⊗ α.
Let W(A) denote the Lie superalgebra obtained from the associative superal-
gebra W(A)as by taking the usual super bracket of operators:
[X ⊗ α, Y ⊗ β] = (XY − Y X) ⊗ (αβ).
Assuming that α and β are of homogeneous degree, we see that
[X ⊗ α, Y ⊗ β] = (−1)1+|α|·|β| [Y ⊗ β, X ⊗ α].
The commutation relation in W(A) is given by
[tr f (D) ⊗ α, ts g(D) ⊗ β]
(4.44) = tr+s (f (D + s)g(D) − f (D)g(D + r)) ⊗ (αβ).
When A = C (hence d = 0), we will simply write the Lie superalgebra W(A)
as W, which is the usual Lie algebra of differential operator on the circle [FKRW,
Kac]. The algebra W affords a universal central extension which is usually referred
to as the W1+∞ algebra. In W1+∞ , the operators L0k generate a Heisenberg algebra,
while the operators L1k generate a Virasoro algebra.
The algebra W(A) has a natural weight filtration
W(A)0 ⊂ W(A)1 ⊂ W(A)2 ⊂ . . . ⊂ W(A)
by letting the weight of Lpk (α) be p. We have
>
The W (super)algebra W(A) is a central extension of the Lie superalgebra
GW(A) by a one-dimensional center with a specified generator C:
(4.45) >
0 −→ CC −→ W(A) −→ GW(A) −→ 0,
>
such that the commutators in W(A) = GW(A) + CC are given by:
3
mδm,−n Tr(αβ) · C, if p = q = 0,
(4.46) [Lpm (α), Lqn (β)] =
(qm − pn) · Lp+q−1
m+n (αβ), otherwise.
>
In this section, we will be mainly interested in the W (super)algebra W(A)
∗
when A is (a subring of) the cohomology ring H (X) of a projective surface X,
with the trace defined by
,
Tr(α) = − α
X
for α ∈ H ∗ (X). First of all, we introduce the following definitions.
4.4. W ALGEBRAS AND HILBERT SCHEMES 93
(ii) We define WX to be the linear span of the identity operator IdHX and the
operators Jpn (α) in End(HX ), where p ≥ 0, n ∈ Z and α ∈ H ∗ (X).
Some of the operators Jpn (α) can be identified with the familiar ones, by using
the above definition, Theorem 4.7 and Theorem 3.31. For example, for α ∈ H ∗ (X),
we see that
J0n (α) = −an (α),
J1n (α) = Ln (α).
For α ∈ AX , we obtain
Jp0 (α) = p! · Gp−1 (α),
(p)
Jp−1 (α) = −a−1 (α).
In general, we have the following.
Lemma 4.22. For p ≥ 0, α ∈ AX and β ∈ H ∗ (X), the following identity holds:
n p
[Gp (α), an (β)] = · J (αβ).
p! n
Proof. Follows from Definition 4.21 (i), Theorem 4.7, and Lemma 3.12 (i).
By Lemma 4.22, we may regard the linear operators Jpn (α) ∈ End(HX ) as
being geometric since both the Chern character operators Gp (α) and the Heisenberg
operators an (β) are geometric by their constructions in Section 4.1 and Section 3.2
respectively.
We are interested in the commutation relation among the operators Jpm (α). To
this end, it is useful to adopt the vertex algebra language [Bor, FB, Kac] for our
present setup. Our convention for fields is to write them in a form
φ(z) = φn z −n−Δ
n
If ψ(z) is another vertex operator, we define a new vertex operator, which is called
the normally ordered product of φ(z) and ψ(z), to be:
: φ(z)ψ(z) : = φ+ (z)ψ(z) + (−1)φψ ψ(z)φ− (z)
94 4. CHERN CHARACTER OPERATORS
where (−1)φψ is −1 if both φ(z) and ψ(z) are odd (i.e. fermionic) fields and 1
otherwise. Inductively we can define the normally ordered product of k vertex
operators φ1 (z), φ2 (z), . . . , φk (z) from right to left by
For α ∈ H ∗ (X), recall the field a(α)(z) from (4.15), and the notations : a(z)p :
(τ∗ α) and : ap :m (τ∗ α) from (4.16). For r ≥ 1, we can similarly define the field
: (∂ r a(z))a(z)p−1 : (τ∗ α), and define the operator : (∂ r a)ap−1 :m (τ∗ α) as the
coefficient of z −m−r−p in : (∂ r a(z))a(z)p−1 : (τ∗ α).
We remark that when the variable z or the cohomology class α is clear or
irrelevant in the context, we will drop z or α from the notations of fields. For
instance, we will sometimes use : ap : (τ∗ α) or : ap :, : (∂ r a)ap : (τ∗ α) or : (∂ r a)ap :,
etc., to stand for : a(z)p : (τ∗ α), : (∂ r a(z))a(z)p : (τ∗ α), etc., respectively.
Lemma 4.23. In terms of fields, the operator Jpm (α) can be rewritten as:
1
(4.47) − : ap+1 :m (τ∗ α)
(p + 1)
1
+ p(m2 − 3m − 2p) : ap−1 :m (τ∗ (eX α))
24
1
+ p(p − 1) : (∂ 2 a) ap−2 :m (τ∗ (eX α)).
24
Proof. First of all, by the definition of : ap+1 :m (τ∗ α), we have
1
p! · aλ (τ∗ α)
λ!
(λ)=p+1,|λ|=m
1
= : ai1 · · · aip+1 : (τ∗ α)
(p + 1) i
1 +...+ip+1 =m
1
= : ap+1 :m (τ∗ α).
(p + 1)
s(λ) + m2 − 2
p! · aλ (τ∗ (eX α))
λ!
(λ)=p−1,|λ|=m
p−1
= p· i2b + m2 − 2 : ai1 · · · aip−1 : (τ∗ (eX α))
i1 +...+ip−1 =m b=1
p−1
= p· (ib + 1)(ib + 2) + m − 3m − 2p ·
2
i1 +...+ip−1 =m b=1
Now the lemma follows from the definition of the operator Jpm (α).
4.4. W ALGEBRAS AND HILBERT SCHEMES 95
Remark 4.24. More explicitly, we can rewrite the operator Jpm (α) as the m-th
Fourier component of a vertex operator as follows:
Jpm (α)
1 p
= − : ap+1 :m (τ∗ α) + (∂ 2 : ap−1 :)m (τ∗ (eX α))
(p + 1) 24
(p + 1)p p(p2 − p − 2) p−1
+ (∂ : ap−1 :)m (τ∗ (eX α)) + :a :m (τ∗ (eX α))
12 24
1
+ p(p − 1) : (∂ 2 a)ap−2 :m (τ∗ (eX α)).
24
In practice, (4.47) suffices for the purpose of computations below. Also, since
Jp+1
0 (α) = (p + 1)! · Gp (α)
for α ∈ AX , we conclude that the Chern character operator Gp (α) with α ∈ AX is
the zero-mode of a vertex operator.
Note that our algebra WX contains as subalgebras the Heisenberg algebra (Sec-
tion 3.2) of Nakajima and Grojnowski generated by the operators
J0m (α) = −am (α),
and the Virasoro algebra (Section 3.5) of Lehn generated by the operators
J1m (α) = Lm (α).
We denote
BX = {α ∈ H ∗ (X)|eX α = KX α = 0}
which is an ideal in the ring H ∗ (X). Obviously, BX ⊂ AX . Denote by
WX
B
96 4. CHERN CHARACTER OPERATORS
the linear span of Jpn (α), where p ≥ 0, n ∈ Z and α ∈ BX . From the commutation
relation in Theorem 4.25 we observe that WX B
is a Lie subalgebra of the Lie (su-
>
per)algebra WX . Recall the (super)algebra W(A) introduced in (4.45) for a general
ring A. The next theorem follows from comparing the commutator (4.46) and the
ones given in Theorem 4.25.
Theorem 4.26. Let X be a smooth projective complex surface. Then, the
> X ) is isomorphic to the Lie (super)algebra W B by sending
Lie (super)algebra W(B X
C → IdHX and Ln (α) → Jpn (α), where p ≥ 0, n ∈ Z, and α ∈ BX .
p
Our goal is to compute the commutator [Jpm (α), Jqn (β)] by using the OPE tech-
nique. To this end, we need to compute the following OPEs
(4.52) (: a(z)p+1 : (τ∗ α))(: a(w)q+1 : (τ∗ β)),
(4.53) (: a(z)p+1 : (τ∗ α))(: a(w)q−1 : (τ∗ (eX β))),
(4.54) (: a(z)p+1 : (τ∗ α))(: ∂ 2 a(w)a(w)q−2 : (τ∗ (eX β))),
which correspond to the first three commutators on the right-hand-side of (4.49)
respectively. Then we will be able to rewrite the terms appearing in these OPEs as
combinations of (the derivative fields of) the four fields
: a(z)p+q : (τ∗ (αβ)),
: a(z)p+q−2 : (τ∗ (eX αβ)),
: ∂ 2 (a(z))a(z)p+q−3 : (τ∗ (eX αβ)),
: ∂ 3 (a(z))a(z)p+q−3 : (τ∗ (eX αβ)),
by using the following simple lemma.
Lemma 4.28. As in the previous section, write the fields ar (z), ∂ s a(z), etc.,
simply as ar , ∂ s a, etc., respectively. Then for N ≥ 0, we have the following:
(i) ∂ 2 : aN := N : (∂ 2 a)aN −1 : +N (N − 1) : (∂a)2 aN −2 :
(ii) ∂ 3 : aN : is equal to
N N
N : (∂ 3 a)aN −1 : +6 : (∂a)(∂ 2 a)aN −2 : +6 : (∂a)3 aN −3 :
2 3
(iii) ∂ : aN ∂ 2 a :=: (∂ 3 a)aN : +N : (∂a)(∂ 2 a)aN −1 :
(iv) ∂ : (∂a)2 aN := 2 : (∂a)(∂ 2 a)aN : +N : (∂a)3 aN −1 :.
Proof. Using the Leibnitz rule, we obtain
∂ 2 : aN : = ∂(N : (∂a)aN −1 :)
= N : (∂ 2 a)aN −1 : +N (N − 1) : (∂a)2 aN −2 : .
This proves part (i). The proofs of the remaining formulas are similar.
Another ingredient in calculating various OPEs is the Wick Theorem [Kac,
Theorem 3.3]. The Wick Theorem in the vertex algebra literature is a very powerful
tool for calculating the OPE of two normally ordered products of free fields.
Theorem 4.29. (Wick Theorem) Let a1 (z), . . . , aM (z) and b1 (z), . . . , bN (z) be
two collections of free fields such that the following properties hold:
(i) [ai (z)− , bj (w)], ck (z) = 0 for all i, j, k, and c = a or b,
(ii) [ai (z)± , bj (w)± ] = 0 for all i and j.
Let [ai bj ] = [ai (z)− , bj (w)] be the contraction of ai (z) and bj (w). The one has the
following OPE in the domain |z| > |w|:
: a1 (z) · · · aM (z) : : b1 (w) · · · bN (w) :
min(M,N )
"
= ± [ai1 bj1 ] · · · [ais bjs ]
s=0 i1 <···<is
j1 =···=js
#
: a1 (z) · · · aM (z)b1 (w) · · · bN (w) :(i1 ,...,is ;j1 ,...,js )
98 4. CHERN CHARACTER OPERATORS
,
L1 n [n] [n]
(5.1) chk1 · · · chL
kN
N
= q chk1 (L1 ) · · · chkN (LN ) · c(TX [n] )
n≥0 X [n]
where 0 < q < 1, and c TX [n] is the total Chern class of the tangent bundle TX [n] .
By (1.22),
1 = (q; q)−χ(X)
∞
where
n
(a; q)n = (1 − aq i ).
i=0
The reduced series chL
k1 · · · chL
kN is defined by
L1 L1
(5.2) chk1 · · · chLN
kN = chk1 · · · chL
kN / 1
N
= (q; q)χ(X)
∞ · chLk1 · · · chkN .
1 LN
where the classes Gki (αi , n) are defined in Definition 4.1 (iii). Note that
,
F (q) = 1 = qn c TX [n] = χ(X [n] )q n = (q; q)−χ(X)
∞
n X [n] n
by (1.22). Recall the line bundle Lm and Ext operator W(L, z) from Section 3.7.
The following lemma is similar to Lemma 3.33.
Lemma 5.2. Let n be the number-of-points operator, i.e., n|H ∗ (X [n] ) = n Id.
Then,
N
(5.4) Fkα11,...,k
,...,αN
N
(q) = Tr q n W(L1 , z) Gki (αi ).
i=1
Proof. We use the same arguments as in the proof of Lemma 3.33. We will
show that the coefficients of q n on both sides of (5.4) are equal. Let {ej }j be a
linear basis of H ∗ (X [n] ). By (3.45),
[Δn ] = (−1)|ej | ej ⊗ e∗j
j
where {e∗j }j is the linear basis of H ∗ (X [n] ) dual to {ej }j in the sense that ej , e∗j =
δj,j . By the definitions of W(L, z) and Gk (α), we have
N
Tr q n W(L1 , z) Gki (αi )
i=1
< N =
= qn (−1)|ej | W(L1 , z) Gki (αi ) ej , e∗j
j i=1
, N
|ej |
= q n
(−1) Gki (αi ) ej ⊗ e∗j c2n (EL1 )
j X [n] ×X [n] i=1
, N
|ej |
= q n
(−1) Gki (αi , n) ej ⊗ e∗j c2n (EL1 )
j X [n] ×X [n] i=1
, N
|ej |
= q n
(−1) (ej ⊗ e∗j ) p∗1 Gki (αi , n) c2n (EL1 )
j X [n] ×X [n] i=1
, N
= q n
[Δn ] p∗1 Gki (αi , n) c2n (EL1 )
X [n] ×X [n] i=1
where p1 : X [n] × X [n] → X [n] denotes the first projection. By (3.42), we have
c2n EL1 |Δn = c TX [n] . Therefore,
, N
N
n n
Tr q W(L1 , z) Gki (αi ) = q Gki (αi , n) c TX [n] .
i=1 X [n] i=1
where λ(1) , . . . , λ(N ) are generalized partitions, and the notation aλ(i) (τ∗ αi ) is de-
fined in (3.36). Then, the structure of the generating series Fkα11,...,k
,...,αN
N
(q) will follow
from Lemma 5.2, Theorem 4.7 and the structure of the trace (5.5).
We begin with four technical lemmas. To explain the ideas behind these lem-
mas, note from (3.49) that the trace (5.5) is equal to
N
a λ(i) (τ∗ αi )
(5.6) Tr q n Γ− (L1 − KX , z) Γ+ (−L1 , z) .
i=1
λ(i)!
n
aλ (τ∗ α) z
Lemma 5.3 deals with the commutator between and exp a−n (γ) .
λ! n
It enables us to eliminate Γ− (L1 − KX , z) from (5.6) in Lemma 5.4, and allows
us to eliminate Γ+ (−L1 , z) from (5.6) in Lemma 5.5. Lemma 5.7 determines the
structure of the trace
N
aλ(i) (τ∗ αi )
Tr q n .
i=1
λ(i)!
The proofs of these lemmas are standard but lengthy.
Recall from Definition 1.6 that P denotes the set of generalized partitions. If
λ = · · · (−2)s−2 (−1)s−1 1s1 2s2 · · · and μ = · · · (−2)t−2 (−1)t−1 1t1 2t2 · · · , let
Proof. Note that the adjoint of am (β) is equal to (−1)m a−m (β). So (5.8)
follows from (5.7) by taking the adjoint on both sides of (5.7) and by making
suitable adjustments. To prove (5.7), put
n
aλ (τ∗ α) z
A= exp a−n (γ) .
λ! n
α ,...,αN
5.2. THE SERIES Fk 1,...,k (q) 103
1 N
Then,
t
aλ (τ∗ α) 1 zn
A = a−n (γ)
λ! t! n
t≥0
t t
1 1 zn t t−i
= a−n (γ) · · · [aλ (τ∗ α), a−n (γ)], . . . , a−n (γ) .
λ! t! n i / 01 2
t≥0 i=0
i times
Let λ = · · · (−2)
(−1) 1 2 · · · . We conclude
s−2 s−1 s1 s2
from Lemma 3.12 (i) that
the commutator · · · [aλ (τ∗ α), a−n (γ)], . . . , a−n (γ) is equal to
/ 01 2
i times
(i,s)
N a (i)
λ − s≥1 μ(i,s) τ∗ ((1X − KX )
mn
s,n≥1 αi )
·Tr q n Γ+ (−L1 , z) !
i=1 λ(i) − s≥1 μ(i,s)
where μ(i,s) = 1m1
(i,s) (i,s)
· · · n mn + for 1 ≤ i ≤ N and s ≥ 1.
··· ∈ P
Proof. For simplicity, denote the trace (5.9) by Q1 . By (3.49),
N
a λ(i) (τ∗ αi )
Q1 = Tr q Γ− (L1 − KX , z) Γ+ (−L1 , z)
n
i=1
λ(i)!
N
a λ(i) (τ∗ αi )
(5.11) = Tr Γ− (L1 − KX , zq) q n Γ+ (−L1 , z)
i=1
λ(i)!
N
aλ(i) (τ∗ αi )
= Tr q n Γ+ (−L1 , z) · Γ− (L1 − KX , zq).
i=1
λ(i)!
104 5. MULTIPLE q-ZETA VALUES AND HILBERT SCHEMES
N
aλ(i) −μ(i,1) τ∗ ((1X − KX ) n≥1 m(i,1)
n αi )
· !
.
i=1 λ(i) − μ(i,1)
Since L1 , L1 − KX = 0, we see from (3.48) that Q1 is equal to
(−(zq)n )mn(i,1)
Tr q Γ− (L1 − KX , zq)Γ+ (−L1 , z) ·
n
(i,1)
·
μ(i,1) ∈P +
1≤i≤N mn !
n≥1
1≤i≤N
N
aλ(i) −μ(i,1) τ∗ ((1X − KX ) n≥1 m(i,1)
n αi )
· !
.
i=1 λ(i) − μ(i,1)
Repeat the above process beginning at line (5.11) s times. Then, Q1 is equals to
(−(zq r )n )mn
(i,r)
s
m(i,r)
N
aλ(i) −sr=1 μ(i,r) τ∗ ((1X − KX ) r=1 n≥1 n αi )
· !
i=1 λ(i) − sr=1 μ(i,r)
(i,r) (i,r)
where μ(i,r) = 1m1 · · · n mn · · · . Letting s → +∞ proves our lemma.
Lemma 5.5. Let λ̃(1) , . . . , λ̃(N ) ∈ P be generalized partitions, and α̃1 , . . . , α̃N ∈
∗
H (X). Then, the trace
N
a λ̃(i) (τ∗ α̃i )
(5.12) Tr q n Γ+ (−L1 , z)
i=1
λ̃(i)!
α ,...,αN
5.2. THE SERIES Fk 1,...,k (q) 105
1 N
is equal to
(z −1 q t−1 )nm̃n(i,t)
(i,t)
N (i) |−
m̃n !
i=1 (|λ̃ t≥1 |μ̃(i,t) |)=0 1≤i≤N
t,n≥1
N a (i)
λ̃ − t≥1 μ̃(i,t) (τ∗ α̃i )
·Tr q n
(i) − (i,t) !
i=1 λ̃ t≥1 μ̃
μ̃(i,1) ∈P −
1≤i≤N m̃n !
n≥1
1≤i≤N
N
aλ̃(i) −μ̃(i,1) (τ∗ α̃i )
·Tr !
· q n Γ+ (−L1 , z)
i=1 λ̃(i) − μ̃(i,1)
z −nm̃(i,1)
n N (i,1)
|−|λ̃(i) |)
= (i,1)
·q i=1 (|μ̃
μ̃(i,1) ∈P −
1≤i≤N m̃n !
n≥1
1≤i≤N
N
aλ̃(i) −μ̃(i,1) (τ∗ α̃i )
·Tr q n Γ+ (−L1 , z) !
.
i=1 λ̃(i) − μ̃(i,1)
By degree reasons, Tr q n Γ+ (−L1 , z)aμ (β) = 0 if |μ| > 0. If |μ| = 0, then we have
Tr q n Γ+ (−L1 , z)aμ (β) = Tr q n aμ (β).
So Q2 is equal to
z −nm̃n(i,1) N (i,1)
|−|λ̃(i) |)
(i,1)
·q i=1 (|μ̃
N
aλ̃(i) −μ̃(i,1) (τ∗ α̃i )
·Tr q n Γ+ (−L1 , z) !
i=1 λ̃(i) − μ̃(i,1)
z −nm̃(i,1)
n
N
aλ̃(i) −μ̃(i,1) (τ∗ α̃i )
+ (i,1)
· Tr q n !
.
N (i) |−|μ̃(i,1) |)=0 1≤i≤N m̃n ! i=1 λ̃(i) − μ̃(i,1)
i=1 (|λ̃ n≥1
106 5. MULTIPLE q-ZETA VALUES AND HILBERT SCHEMES
N a
λ̃(i) − tr=1 μ̃(i,r) (τ∗ α̃i )
(5.14) ·Tr q Γ+ (−L1 , z)
n
t
(i) − (i,r) !
i=1 λ̃ r=1 μ̃
N a
λ̃(i) − tr=1 μ̃(i,r) (τ∗ α̃i )
·Tr q n t .
(i) − (i,r) !
i=1 λ̃ r=1 μ̃
(t). Since
Denote line (5.13) by U
N
t
(|λ̃(i) | − |μ̃(i,r) |) < 0
i=1 r=1
and |μ̃(i,r) | < 0, U (t) is a polynomial in q with coefficients being bounded in terms of
N (t). Line (5.14) is bounded in terms of the generalized
− i=1 |λ̃ |. Moreover, q t |U
(i)
that Q2 equals
⎛ ⎞
⎜ z −nm̃n(i,t) N t (i) ⎟
· q i=1 ( =1 |μ̃ |−|λ̃ |) ⎠ ·
(i,)
⎝ (i,t)
N m̃n !
(|λ̃(i) |−
i=1 |μ̃(i,t) |)=0 t≥1
t≥1
1≤i≤N
n≥1
N a (i)
λ̃ − t≥1 μ̃(i,t) (τ∗ α̃i )
·Tr q n .
(i) − (i,t) !
i=1 λ̃ t≥1 μ̃
N t N
Replacing q i=1 ( =1 |μ̃(i,) |−|λ̃(i) |)
by q − i=1 ≥t+1 |μ̃(i,) |
proves our lemma.
We remark that the assumption 0 < q < 1 was used in the above proof.
However, for most parts of this chapter (e.g. in (5.4)), q can also be viewed as a
formal variable.
Let
N
a λ(i) (τ∗ αi )
(5.15) AN = Tr q n
i=1
λ(i)!
α ,...,αN
5.2. THE SERIES Fk 1,...,k (q) 107
1 N
the partition {π1 , . . . , πu } of {1, . . . , N } depend only on λ(1) , . . . , λ(N ) , and Sign(π)
depends on the parity of the degrees of α1 , . . . , αN and is the sign defined by
u
(5.17) αj = Sign(π) · α1 · · · αN .
i=1 j∈πi
Case 1: |λ(i) | = 0 for every 1 ≤ i ≤ N . Then, (λ(i) ) ≥ 2 for every i. Since the
bi-degree of aλ(i) (τ∗ αi ) is given by (5.18), AN = 0 unless (λ(i) ) = 2 and |αi | = 0
for all 1 ≤ i ≤ N . Assume that (λ(i) ) = 2 and |αi | = 0 for all 1 ≤ i ≤ N . Then
108 5. MULTIPLE q-ZETA VALUES AND HILBERT SCHEMES
for every 1 ≤ i ≤ N , we have λ(i) = ((−ni )ni ) for some ni > 0. We further assume
that n1 = . . . = nr for some 1 ≤ r ≤ N and ni = n1 if r < i ≤ N . Let α1 = a1X
and
τ2∗ 1X = (−1)|βj | βj ⊗ γj
j
N
= aq n1 (−1)|βj | Tr a−n1 (βj ) q n an1 (γj ) aλ(i) (τ∗ αi )
j i=2
N
= aq n1 Tr q n an1 (γj ) aλ(i) (τ∗ αi ) · a−n1 (βj ).
j i=2
Moving the operator a−n1 (βj ) to the middle, we see that AN is equal to
N
aq n1 Tr q n an1 (γj )a−n1 (βj ) aλ(i) (τ∗ αi )
j i=2
r
i−1
+ aq n1 Tr q n an1 (γj ) aλ(k) (τ∗ αk )
j i=2 k=2
N
·[aλ(i) (αi ), a−n1 (βj )] · aλ(k) (τ∗ αk ).
k=i+1
where 0 ≤ k1 ≤ r −1, {j1 , . . . , jk1 } ⊂ {2, . . . , r}, every factor in (−n1 )k1 comes from
a commutator of type [an1 (·), a−n1 (·)], and the coefficients of this linear combination
depend only on λ(1) , . . . , λ(N ) . In particular, we have
(−n1 )q n1
(5.20) A1 = Tr q n aλ(1) (α1 ) = (q; q)−χ(X)
∞ · eX , α1 · .
1 − q n1
Combining with (5.19), we see that our lemma holds in this case.
N
i=1 |λ | = 0 but |λ | = 0 for some i0 . Then N ≥ 2, and we may
(i) (i0 )
Case 2:
assume that |λ | < 0. To simplify the expressions, we further assume that every
(i0 )
0 −1
i
r−1 8 9
aλ(i) (τ∗ αi ) aλ(r) (τ∗ αr ) aλ(i0 ) (τ∗ αi0 ) aλ(i) (τ∗ αi )
+ Tr q n
· , · .
r=1 i=1
λ(i)! λ(r)! λ(i0 )! r+1≤i≤N
λ(i)!
i=i0
α ,...,αN
5.2. THE SERIES Fk 1,...,k (q) 109
1 N
Since
q n aλ(i0 ) (τ∗ αi0 ) = q −|λ |
(i0 )
aλ(i0 ) (τ∗ αi0 )q n ,
we see that AN is equal to
aλ(i0 ) (τ∗ αi0 ) n aλ(i) (τ∗ αi )
q −|λ |
(i0 )
Tr q
λ(i0 )! 1≤i≤N
λ(i)!
i=i0
0 −1
i
r−1 8 9
aλ(i) (τ∗ αi ) aλ(r) (τ∗ αr ) aλ(i0 ) (τ∗ αi0 ) aλ(i) (τ∗ αi )
+ Tr q n · , · .
r=1 i=1
λ(i)! λ(r)! λ(i0 )! r+1≤i≤N
λ(i)!
i=i0
Since
aλ(i0 ) (τ∗ αi0 ) n aλ(i) (τ∗ αi ) aλ(i) (τ∗ αi ) a (i0 ) (τ∗ αi )
Tr q = Tr q n
· λ (i )! 0 ,
λ(i0 )! 1≤i≤N
λ(i)! 1≤i≤N
λ(i)! λ 0
i=i0 i=i0
AN is equal to
a (i) (τ∗ αi ) a (i0 ) (τ∗ αi )
q −|λ |
(i0 )
Tr q n λ
(i)!
· λ (i )! 0
1≤i≤N
λ λ 0
i=i0
0 −1
i
r−1 8 9
aλ(i) (τ∗ αi ) aλ(r) (τ∗ αr ) aλ(i0 ) (τ∗ αi0 ) aλ(i) (τ∗ αi )
+ Tr q n
· , · .
r=1 i=1
λ(i)! λ(r)! λ(i0 )! r+1≤i≤N
λ(i)!
i=i0
Note that
a (i) (τ∗ αi ) a (i0 ) (τ∗ αi )
Tr q n λ
(i)!
· λ (i )! 0
1≤i≤N
λ λ 0
i=i0
is equal to
AN
N 8 9 N
aλ(i) (τ∗ αi ) aλ(r) (τ∗ αr ) aλ(i0 ) (τ∗ αi0 ) aλ(i) (τ∗ αi )
+ Tr q n
· , · .
r=i0 +1 1≤i≤r−1
λ(i)! λ(r)! λ(i0 )! i=r+1
λ(i)!
i=i0
N aλ(i) (τ∗ αi )
q −|λ |
(i0 )
Tr q n
r=i0 +1 1≤i≤r−1,i=i0
λ(i)!
8 9 N
aλ(r) (τ∗ αr ) aλ(i0 ) (τ∗ αi0 ) aλ(i) (τ∗ αi )
· , ·
λ(r)! λ(i0 )! i=r+1
λ(i)!
0 −1
i a (i) (τ∗ αi ) 8 a (r) (τ∗ αr ) a (i0 ) (τ∗ αi ) 9
r−1 aλ(i) (τ∗ αi )
+ Tr q n λ
(i)!
· λ (r)! , λ (i )! 0 · .
r=1 i=1
λ λ λ 0 r+1≤i≤N
λ(i)!
i=i0
110 5. MULTIPLE q-ZETA VALUES AND HILBERT SCHEMES
q n0
N aλ(i) (τ∗ αi )
Tr q n
1 − q n0 r=i0 +1
λ(i)!
1≤i≤r−1,i=i0
8 9 N
aλ(r) (τ∗ αr ) aλ(i0 ) (τ∗ αi0 ) aλ(i) (τ∗ αi )
· , ·
λ(r)! λ(i0 )! i=r+1
λ(i)!
1 0 −1
i aλ(i) (τ∗ αi ) 8 aλ(r) (τ∗ αr ) a (i0 ) (τ∗ αi ) 9
r−1
(5.21) + Tr q n
· , λ (i )! 0
1 − q n0 r=1 i=1
λ (i)! λ (r)! λ 0
aλ(i) (τ∗ αi )
· .
r+1≤i≤N,i=i
λ(i)!
0
By Lemma 3.12 (i) and (ii), our lemma holds in this case as well.
is equal to
N
N
(i)
|λ(i) |
z i=1 · (q; q)−χ(X)
∞ · (1X − KX ) n≥1 mn , αi
i=1
(−1)mn
(i) (i)
q nmn 1 1
· ,
+W
(i) (i) (i)
1≤i≤N,n≥1 mn ! (1 − q n )mn m̃(i)
n ! (1 − q ) n
n m̃
and the lower weight term W is a linear combination of expressions of the form:
< =
N
u
z i=1 |λ | · (q; q)−χ(X)
(i) r
(5.23) ∞ · Sign(π) · ri i
KX eX , αj
i=1 j∈πi
v
q ni wi pi
·
i=1
(1 − q ni )wi
v N
i=1 (λ ), the integers u, v, ri , ri ≥ 0, ni > 0, wi > 0, pi ∈
(i)
where i=1 wi <
{0, 1} and the partition π = {π1 , . . . , πu } of {1, . . . , N } depend only on the general-
ized partitions λ(1) , . . . , λ(N ) , and Sign(π) is from (5.17). Moreover, the coefficients
of this linear combination are independent of q, α1 , . . . , αN and X.
α ,...,αN
5.2. THE SERIES Fk 1,...,k (q) 111
1 N
Proof. For simplicity, denote the trace (5.22) by Trλ . Combining Lemma 5.4
and Lemma 5.5, we conclude that Trλ is equal to
(−(zq s )n )mn(i,s) (z −1 q t−1 )nm̃n(i,t)
(i,s)
· (i,t)
N
(|λ(i) |− s≥1 |μ(i,s) |− t≥1 |μ̃(i,t) |)=0 1≤i≤N mn ! 1≤i≤N m̃n !
i=1 s,n≥1 t,n≥1
μ (i,s)
∈P+ , μ̃ (i,t)
∈P −
(i,s)
N a (i)
λ − s≥1 μ(i,s) − t≥1 μ̃(i,t) τ∗ ((1X − KX )
s,n≥1 mn αi )
·Tr q n !
i=1 λ(i) − s≥1 μ(i,s) − t≥1 μ̃(i,t)
(i,s) (i,s) (i,t) (i,t)
where μ(i,s) = 1m1 · · · nmn · · · and μ̃(i,t) = · · · (−n)m̃n · · · (−1)m̃1 . The
N
sum of all the exponents of z is i=1 |λ(i) |. So Trλ is equal to
N
z i=1 |λ | ·
(i)
N
(|λ(i) |− s≥1 |μ(i,s) |− t≥1 |μ̃(i,t) |)=0
i=1
, μ̃(i,t) ∈P
μ(i,s) ∈P
+ −
(i,s)
N a (i)
λ − s≥1 μ(i,s) − t≥1 μ̃(i,t) τ∗ ((1X − KX )
s,n≥1 mn αi )
(5.24) ·Tr q n
!
.
i=1 λ(i) − s≥1 μ(i,s) − t≥1 μ̃(i,t)
By our convention,
μ(i,s) + μ̃(i,t) ≤ λ(i)
s≥1 t≥1
for every 1 ≤ i ≤ N . In the rest of the proof, we group the possible terms in the
above summand into two cases: Case A and Case B.
Case A: s≥1 μ
(i,s)
+ t≥1 μ̃(i,t) = λ(i) for every 1 ≤ i ≤ N . Then line (5.24) is
equal to
N
(i,s)
Tr q n · (1X − KX ) s,n≥1 mn
, αi
i=1
N
(i,s)
= (q; q)−χ(X)
∞ (1X − KX ) s,n≥1 mn , αi .
i=1
q (i,s)
(s−1)nmn q (t−1)nm̃n(i,t)
· (i,s)
· (i,t)
.
(i,s) (i) 1≤i≤N mn ! (i,t) (i) 1≤i≤N m̃n !
s≥1 mn =mn
s,n≥1 t≥1 m̃n =m̃n
t,n≥1
1≤i≤N, n≥1 1≤i≤N, n≥1
Since
(q (s−1)n )is,n 1 1
= ,
is,n ! in ! (1 − q n )in
s≥1 is,n =in , n≥1
s,n≥1 n≥1
C1 is equal to
N
N
|λ(i) | m(i)
(5.25) z i=1 · (q; q)−χ(X)
∞ (1X − KX ) n≥1 n , αi ·
i=1
(−1)mn
(i)
q nmn
(i)
1 1
· (i) (i)
· (i) (i)
.
1≤i≤N,n≥1 mn ! (1 − q n )mn 1≤i≤N,n≥1 m̃n ! (1 − q n )m̃n
Case B: s≥1 μ(i,s) + t≥1 μ̃(i,t) < λ(i) for some 1 ≤ i ≤ N . Without loss of
generality, we may assume that
μ(i,s) + μ̃(i,t) = λ(i)
s≥1 t≥1
An argument similar to that in the previous paragraph shows that for the fixed
generalized partitions λ̃(i) with N1 + 1 ≤ i ≤ N , the contribution C2 of this case to
Trλ is equal to
N
N1
(i)
|λ(i) |
(5.26) z i=1 · (1X − KX ) n≥1 mn , αi
i=1
(−1) m(i)
n q nmn
(i)
1 1
· (i) (i) (i) (i)
1≤i≤N1 mn ! (1 − q n )mn m̃n ! (1 − q n )m̃n
n≥1
(i)
(−1)pn q npn
(i)
1 1
· (i) (i) (i) (i)
N1 +1≤i≤N pn ! (1 − q n )pn p̃n ! (1 − q n )p̃n
n≥1
N
aλ(i) −λ̃(i) τ∗ ((1X − KX ) n≥1 p(i)
n αi )
·Tr q n
!
.
i=N1 +1 λ(i) − λ̃(i)
α ,...,αN
5.2. THE SERIES Fk 1,...,k (q) 113
1 N
(−1)pn
(i) (i)
q npn 1 1
· (i) (i) (i)
·
N1 +1≤i≤N pn ! (1 − q n )pn p̃(i)
n ! (1 − q ) n
n p̃
n≥1
< =
u (j)
v
q ni
(q; q)−χ(X)
∞ · Sign(π) · em
X ,
i
(1X − KX ) n≥1 pn αj ·
i=1 j∈πi i=1
1 − q ni
N
where v < i=N1 +1 (λ
(i)
− λ̃(i) ), ni > 0, mi ≥ 0, {π1 , . . . , πu } is a parti-
tion of {N1 + 1, . . . , N }, and Sign(π) compensates the formal difference between
u
i=1 j∈πi αj and αN1 +1 · · · αN . The coefficients of this linear combination are
independent of q, α1 , . . . , αN and X, and depend only on the partitions λ(i) − λ̃(i) .
Note that for nonnegative integers a and b, the pairing
eaX , (1X − KX )b β = eaX (1X − KX )b , β
is a linear combination of eaX KX
c
, β , 0 ≤ c ≤ b. In addition, we have
N
m(i) (i)
n + m̃n + p(i) (i)
n + p̃n +v < (λ(i) )
1≤i≤N1 ,n≥1 N1 +1≤i≤N,n≥1 i=1
Remark 5.9. When N = 1, we can work out the lower weight term W in
Theorem 5.8 by examining its proof more carefully and by using (5.20). To state
For n1 ≥ 1 with
the result, let λ = (· · · (−n)m̃n · · · (−1)m̃1 1m1 · · · nmn · · · ) ∈ P.
mn1 · m̃n1 ≥ 1, define mn1 (n1 ) = mn1 − 1, m̃n1 (n1 ) = m̃n1 − 1, and mn (n1 ) = mn
and m̃n (n1 ) = m̃n if n = n1 . Then,
aλ (τ∗ α)
Tr q n W(L1 , z)
λ!
is equal to the sum
(5.27) z |λ| · (q; q)−χ(X)
∞ · (1X − KX ) n≥1 mn , α ·
(−1) n m
q nmn 1 1
·
mn ! (1 − q n )mn m̃n ! (1 − q n )m̃n
n≥1
n1 q n1
+ z |λ| · (q; q)−χ(X)
∞ · eX , α · ·
1 − q n1
n1 ≥1 with mn1 ·m̃n1 ≥1
(−1)mn (n1 ) q nmn (n1 )
1 1
·
mn (n1 )! (1 − q n )mn (n1 ) m̃n (n1 )! (1 − q n )m̃n (n1 )
n≥1
114 5. MULTIPLE q-ZETA VALUES AND HILBERT SCHEMES
where the first two lines come from Case A in the proof of Theorem 5.8, and the
last two lines come from Case B in the proof of Theorem 5.8.
Example 5.10. Let λ = ((−5)(−4)5). Then, we see from (5.27) with n1 = 5
that Tr q n W(L1 , z) a((−5)(−4)5) (τ3∗ α) is equal to
q4 −q 5 q5
z −4 · (q; q)−χ(X)
∞ · 1X − KX , α · ·
1 − q 1 − q 1 − q5
4 5
5q 5 1
+ z −4 · (q; q)−χ(X)
∞ · eX , α · ·
1 − q 1 − q4
5
where the first line is the leading term and the second line is the lower weight term.
The next lemma is used to organize the leading term in Theorem 5.8.
Lemma 5.11. For α ∈ H ∗ (X) and k ≥ 0, define Θαk (q) to be
(5.28) − (1X − KX ) n≥1 in , α
(λ)=k+2,|λ|=0
(−1)in q nin 1 1
·
in ! (1 − q n )in ĩn ! (1 − q n )ĩn
n≥1
a a
(−1)si 1
b
(5.29) − (1X − KX ) i=1 si , α ·
a,s1 ,...,sa ,b,t1 ,...,tb ≥1 i=1
si ! t !
j=1 j
a
s + b t =k+2
i=1 i j=1 j
a
(qz)ni si
b
z −mj tj
· · .
n1 >···>na >0 i=1
(1 − q ni )si m (1 − q mj )tj
1 >···>mb >0 j=1
Proof. Put
A = (1X − KX ) n≥1 in , α
which implicitly depends on n≥1 in . Rewrite |λ| and (λ) in terms of the integers
in and ĩn . Then, Θαk (q) is equal to
(−1)in q nin 1 1
(5.30) − A .
in ! (1 − q n )in ĩn ! (1 − q n )ĩn
in +
n≥1 ĩn =k+2
n≥1 n≥1
n≥1 nin = n≥1 nĩn >0
and the lower weight term W1 is an infinite linear combination of the expressions:
< = v
u
q ni wi pi
−χ(X) ri ri
(5.32) (q; q)∞ · Sign(π) · KX eX , αj ·
i=1 j∈π i=1
(1 − q ni )wi
i
v N
where i=1 wi < i=1 (ki + 2), and the integers u, v, ri , ri ≥ 0, ni > 0, wi > 0, pi ∈
{0, 1} and the partition π = {π1 , . . . , πu } of {1, . . . , N } depend only on the integers
ki . Moreover, the coefficients of this linear combination are independent of q, αi , X.
Proof. By Lemma 5.2,
N
Fkα11,...,k
,...,αN
N
(q) = Tr q n W(L1 , z) Gki (αi ).
i=1
N
a λ(i) (τ∗ αi )
(5.34) (−1)N · Tr q n W(L1 , z) .
i=1
λ(i)!
(λ(i) )=ki +2,|λ(i) |=0
1≤i≤N
i=1
N
= (q; q)−χ(X)
∞ · Coeff z10 ···zN
0 Θα
ki (q, zi )
i
+ W1,2
i=1
In order to relate the lower weight term W1,2 in (5.35) and (5.36) to multiple q-
zeta values (with additional variables z1 , . . . , zN inserted), we will assume eX αi = 0
for all 1 ≤ i ≤ N . The following lemma strengthens Lemma 5.7, and follows from
the same arguments.
and α1 , . . . , αN ∈ H ∗ (X) be homoge-
Lemma 5.13. Let λ(1) , . . . , λ(N ) ∈ P,
neous. Assume that eX αi = 0 for every 1 ≤ i ≤ N , and N i=1 |λ | = 0. Put
(i)
N
aλ(i) (τ∗ αi )
AN = Tr q n .
i=1
λ(i)!
where
N
˜= (λ(i) )/2 = wi ,
i=1 i=1
p̃i ∈ {0, 1}, 0 ≤ pi ≤ wi , the partition π = {π1 , . . . , πu } of {1, . . . , N } de-
pend only on λ(1) , . . . , λ(N ) , the integers ñ1 , . . . , ñ˜ are the positive parts
(repeated with multiplicities) in λ(1) , . . . , λ(N ) , the integers n1 , . . . , n de-
note the different integers in ñ1 , . . . , ñ˜, and each ni appears wi times in
ñ1 , . . . , ñ˜.
Theorem 5.14. For 1 ≤ i ≤ N , let ki ≥ 0 and αi ∈ H ∗ (X) be homogeneous.
Assume that eX αi = 0 for every 1 ≤ i ≤ N . Then,
N
(5.37) F 1
α ,...,αN
(q) = (q; q)−χ(X)
k1 ,...,kN ∞ · Coeff z0 ···z0
1
α
Θ i (q, zi ) + W1,2 ,
N ki
i=1
χ(X)
and (q; q)∞ · W1,2 is a linear combination of the coefficients of z10 · · · zN 0
in some
N
multiple q-zeta values (with variables z1 , . . . , zN inserted) of weights < i=1 (ki +2).
Moreover, the coefficients in this linear combination are independent of q.
We now sketch the proof of Theorem 5.14, and refer to the proof of [QY,
Theorem 4.10] for further details. Recall that Fkα11,...,k
,...,αN
N
(q) is defined in (5.34),
and that (5.37) is just (5.36). From the proofs of (5.36) and Theorem 5.8, we
see that the lower weight term W1,2 in (5.37) is the contribution of Case B in the
proof of Theorem 5.8 to the right-hand-side of (5.34). A careful analysis of these
contributions together with Lemma 5.13 completes the proof.
We will end this section with three propositions about Fkα (q), which provide
some insight into the lower weight term W1 in Theorem 5.12. Proposition 5.15
deals with F0α (q) for an arbitrary α ∈ H ∗ (X). Proposition 5.17 calculates F1α (q)
by assuming eX α = 0. Proposition 5.18 computes Fkα (q), k ≥ 2 by assuming
eX α = KX α = 0.
α ,...,αN
5.2. THE SERIES Fk 1,...,k (q) 117
1 N
Proof. We have
F1α (q) = Tr q n W(L1 , z) G1 (α).
By (4.18),
aλ (τ3∗ α) n − 1
G1 (α) = − − (a−n an )(τ2∗ (KX α)).
λ! n>0
2
(λ)=3,|λ|=0
Expanding Coeff z0 Θα
1 (q, z) yields the weight-3 terms in our proposition.
118 5. MULTIPLE q-ZETA VALUES AND HILBERT SCHEMES
a
(qz)ni si
b
z −mj tj
(5.39) · · .
n1 >···>na >0 i=1
(1 − q ni )si m (1 − q mj )tj
1 >···>mb >0 j=1
As in the proof of Proposition 5.17, Remark 5.9 and the proof of Theorem 5.12
yield
Fkα (q) = (q; q)−χ(X)
∞ · Coeff z0 Θα
k (q, z).
By the definition of Θα k (q, z) in (5.29), we see that (i) holds and that our formula
for Fkα (q) with |α| = 4 and k ≥ 0 holds. Note that line (5.39) can be rewritten as
a
(qz 2 )ni si /2 b
(qz −2 )mj tj /2
· .
n1 >···>na >0 i=1
(1 − q ni )si m (1 − q mj )tj
1 >···>mb >0 j=1
In this section, using the calculations from the previous section, we will prove
Conjecture 5.1 modulo the lower weight term. Moreover, for abelian surfaces, we
will verify Conjecture 5.1. The main idea is to express the Chern character of the
tautological bundle L[n] in terms of the classes Gi (α, n).
Let L be a line bundle on the smooth projective surface X. We see from (4.1)
and the Grothendieck-Riemann-Roch Theorem that
(5.40) ch(L[n] ) = p1∗ (ch(OZn ) · p∗2 ch(L) · p∗2 td(X))
= p1∗ (ch(OZn ) · p∗2 (1X + L + L2 /2) · p∗2 td(X))
= G(1X , n) + G(L, n) + G(L2 /2, n).
Since the cohomology degree of Gi (α, n) is 2i + |α|, we have
(5.41) chk (L[n] ) = Gk (1X , n) + Gk−1 (L, n) + Gk−2 (L2 /2, n).
Next, we recall the generating series chL k1 · · · chkN
LN
1
and its reduced version
L1
chk1 · · · chL
kN
N
from (5.1) and (5.2) respectively.
L1 L
5.3. THE REDUCED SERIES chk · · · chk N 119
1 N
and the lower weight term W is an infinite linear combination of the expressions:
u @ A v
ri ri q ni wi pi
KX eX , L1i,1 · · · LNi,N ·
i=1 i=1
(1 − q ni )wi
v N
where i=1 wi < i=1 (ki + 2), and the integers u, v, ri , ri , i,j ≥ 0, ni > 0, wi >
0, pi ∈ {0, 1} depend only on k1 , . . . , kN . Furthermore, all the coefficients of this
linear combination are independent of q, L1 , . . . , LN and X.
Proof. We conclude from (5.1), (5.2), (5.41) and (5.3) that
L
chk1 · · · chL
kN = (q; q)χ(X)
∞ · Fk11X,...,k
,...,1X
N
(q) + (q; q)χ(X)
∞ ·A
where A is the sum of the series Fkα1,...,k
,...,αN
such that for every 1 ≤ i ≤ N ,
1 N
Theorem 5.19 proves Conjecture 5.1, modulo the lower weight term W . Note
that the leading term
N
1X
Coeff z10 ···zN
0 Θki (q, zi )
i=1
in chL1
k1 · · · chLN
kN has weight
N
(ki + 2),
i=1
by Lemma 5.2, Theorem 4.7 and (5.34). By Theorem 5.14 and the proof of Theo-
rem 5.19, our theorem follows.
120 5. MULTIPLE q-ZETA VALUES AND HILBERT SCHEMES
We remark that some of the multiple q-zeta values mentioned in Theorem 5.20
are in the generalized sense, i.e., in the following form:
(−ni )wi q ni pi fi (z1 , . . . , zN )ni
n1 >···>n >0 i=1
(1 − q ni )wi
Proof. Our formula follows from (5.43), (5.38) and Proposition 5.17.
Proposition 5.22. Let L be a line bundle over an abelian surface X. If 2 k,
then chL
k = 0. If 2|k, the generating series chL
k is the coefficient of z 0 in
L, L a
(−1)si 1
b
− · ·
2 ≥1 i=1
a,s1 ,...,sa ,b,t1 ,...,tb
si ! t !
j=1 j
a
s + b t =k
i=1 i j=1 j
a
(qz)ni si
b
z −mj tj
· · .
n1 >···>na >0 i=1
(1 − q ni )si m (1 − q mj )tj
1 >···>mb >0 j=1
121
122 6. LIE ALGEBRAS AND INCIDENCE HILBERT SCHEMES
[m+n,m] = 2m + n + 3.
Combining (6.2), (6.3) and (6.4), we get dim Q
X
This allows us to define the adjoint f̃† ∈ End H X . As in
for f̃ ∈ End H
Lemma 3.7, we have
(6.6) ãn (α) = (−1)n · ã−n (α)† .
Let n = 0. By (6.5) and Lemma 6.1, ã−n (α) has bi-degree (n, 2n − 2 + |α|), i.e.,
ã−n (α) : H r X [m,m+1] → H r+2n−2+|α| X [m+n,m+n+1] .
X is an H ∗ (X)-module. The module structure is induced
Next, notice that H
by the natural morphisms ρ̃n : X [n,n+1] → X.
X → H
Lemma 6.3. The map ã−n (α) : H X is an H ∗ (X)-module homomor-
phism.
Proof. We will only prove the lemma for n > 0 since the proof for n < 0 is
∈ H ∗ X [m,m+1] . We need to show that
similar. Let n > 0, β ∈ H ∗ (X), and A
By definition,
ã−n (α) ρ̃∗m β · A [m+n,m] ] · ρ̃∗ α · p̃∗2 (ρ̃∗m β · A)
= p̃1∗ [Q
= (−1)|α||β| p̃1∗ [Q
[m+n,m] ] · p̃∗2 ρ̃∗m β · ρ̃∗ α · p̃∗2 A
= (−1)|α||β| p̃1∗ τ∗ ((ρ̃m ◦ p̃2 ◦ τ )∗ β) · ρ̃∗ α · p̃∗2 A
[m+n,m] → X [m+n,m+n+1] × X × X [m,m+1] is the inclusion map.
where τ : Q
[m+n,m] ,
From the proof of Lemma 6.1, we see that for (ξ, ξ ), x, (η, η ) ∈ Q
Supp(Iη /Iξ ) = Supp(Iη /Iξ ) = {x}.
124 6. LIE ALGEBRAS AND INCIDENCE HILBERT SCHEMES
The map πi1 ,...,ik in (6.18) denotes the projection of this ambient space to the
product of its i1 -th, . . ., ik -th factors.
In the sequel, the ambient spaces in similar situations will not be explicitly
presented since they can be written out easily from the context.
From Lemma 6.1, we know that the expected dimension of the intersection W
should be 2m + n + k + 4. The subset W may have many irreducible components.
Those that are mapped by π1245 to subsets of dimension less than 2m + n + k + 4
will not contribute to the cycle w. The aim of the computations below is to pick
out those components with dimension no less than the expected dimension.
Any element ((ξ, ξ ), x, (η, η ), t, (ζ, ζ )) in W must satisfy the conditions:
ζ ⊂ ζ , η ⊂ η , ξ ⊂ ξ , ζ ⊂ η ⊂ ξ, ζ ⊂ η ⊂ ξ ,
Supp(Iη /Iξ ) = {x}, Supp(Iζ /Iη ) = {t},
Supp(Iζ /Iζ ) = Supp(Iη /Iη ) = Supp(Iξ /Iξ ) = {p}
for some p ∈ X. In the following, we consider four different cases separately. We
use Wi to denote the subset of W consisting of all the points satisfying Case i, and
put
wi = dim π1245 (Wi )
(similar notations such as Ui , ui , Vi , vi will be used throughout the rest of this
chapter).
Case 1: x, t, p are distinct. We have the following decompositions:
ζ = ζ0 + ζx + ζt + ζp , ζ = ζ0 + ζx + ζt + ζp ,
η = ζ0 + ζx + ηt + ζp , η = ζ0 + ζx + ηt + ζp ,
ξ = ζ0 + ξx + ηt + ζp , ξ = ζ0 + ξx + ηt + ζp ,
where Supp(ζ0 ) ∩ {x, t, p} = ∅, (ζx ) = i, (ζt ) = j, (ζp ) = , and ζx , ζt , . . . are
supported at {x}, {t}, . . . respectively. Then,
w1 = #(moduli of ζ0 ) + #(moduli of ζx ) + #(moduli of ζt ) +
+ #(moduli of ηt ) + #(moduli of ξx ) + #(moduli of ζp ⊂ ζp )
= 2(m − i − j − ) + max(i − 1, 0) + 2 + max(j − 1, 0) + 2 +
+ (j + k − 1) + (i + n − 1) + + 2
= (2m + k + n + 4) − i + max(i − 1, 0) − j + max(j − 1, 0) − .
Case 1.1: If i = j = = 0, then we have w1 = (2m + k + n + 4).
Case 1.2: If one of the integers i, j, is positive, then w1 < (2m + k + n + 4).
The three remaining cases are listed by:
Case 2: x = t = p;
Case 3: x = p = t (by symmetry, this also covers the case t = p = x);
Case 4: x = p = t.
For these three cases, we skip the arguments which are in the same style as in
Case 1. They all have dimension less than the expected dimension (2m + n + k +
4). Therefore, only Case 1.1 has contribution to the cohomological operation. In
this subcase, it is not difficult to show that the intersection (6.19) along W1 is
transversal. Moreover, π1245 (W1 ) consists of all the points of the form:
(ζ0 + ξx + ηt , ζ0 + ξx + ηt + p), x, t, (ζ0 , ζ0 + p)
6.1. HEISENBERG ALGEBRA ACTIONS FOR INCIDENCE HILBERT SCHEMES 127
In these three cases, all the dimensions are smaller than the expected dimension
(2m + n + k + 4). So only Case 1.1 contributes to the class u in (6.20).
Next we consider the operator ãk (β)ã−n (α). This is the only case where there
are two components for V below with the expected dimension. One of the com-
ponents will cancel out with the one from ã−n (α)ãk (β), and the other is a non-
transversal intersection which carries a multiplicity . Using Lemma 6.4 which
compares ãn (α) with the Heisenberg operators an (α) on HX , we determine the
multiplicity.
More precisely, the operator ãk (β)ã−n (α) is induced by the class:
(6.22) " ∗ [m+k+n,m+k] #
∗
v = π1245∗ π123 τm+n,m+k+n (Q [m+k+n,m+n] ) · π345
Q
in A2m+k+n+4 (V ), where V = π1245 (V ) and V is given by
(6.23) −1
V = π123 τm+n,m+k+n (Q [m+k+n,m+k] .
[m+k+n,m+n] ) ∩ π −1 Q
345
Any element ((ξ, ξ ), x, (η, η ), t, (ζ, ζ )) in (6.23) must satisfy the conditions:
ζ ⊂ ζ , η ⊂ η , ξ ⊂ ξ , ξ ⊂ η ⊃ ζ, ξ ⊂ η ⊃ ζ ,
Supp(Iξ /Iη ) = {x}, Supp(Iζ /Iη ) = {t},
Supp(Iζ /Iζ ) = Supp(Iη /Iη ) = Supp(Iξ /Iξ ) = {p}
for some p ∈ X. In the following, we consider four different cases separately.
Case 1 : x, t, p are distinct. We have the following decompositions:
ζ = ζ0 + ζx + ζt + ζp , ζ = ζ0 + ζx + ζt + ζp ,
η = ζ0 + ζx + ηt + ζp , η = ζ0 + ζx + ηt + ζp ,
ξ = ζ0 + ξx + ηt + ζp , ξ = ζ0 + ξx + ηt + ζp ,
where Supp(ζ0 ) ∩ {x, t, p} = ∅, (ξx ) = i, (ζt ) = j, and (ζp ) = . Then,
v1 = #(moduli of ζ0 ) + #(moduli of ζx ) + #(moduli of ζt ) +
+ #(moduli of ηt ) + #(moduli of ξx ) + #(moduli of ζp ⊂ ζp )
= 2(m − i − j − ) + (i + k − 1) + max(j − 1, 0) +
+ (j + n − 1) + max(i − 1, 0) + + 6
= (2m + k + n + 4) − i + max(i − 1, 0) − j + max(j − 1, 0) − .
Case 1 .1: If i = j = = 0, then we have v1 = (2m + k + n + 4).
Case 1 .2: If one of the integers i, j and is positive, then v1 < (2m + k + n + 4).
Case 2 : x = t = p. We have the following decompositions:
ζ = ζ0 + ζx + ζp , ζ = ζ0 + ζx + ζp ,
η = ζ0 + ηx + ζp , η = ζ0 + ηx + ζp ,
ξ = ζ0 + ξx + ζp , ξ = ζ0 + ξx + ζp ,
where Supp(ζ0 ) ∩ {x, p} = ∅, (ξx ) = i, (ζx ) = j, and (ζp ) = . Note that
(6.24) k + i = (ηx ) = n + j.
Case 2 .1: i = j = = 0. By (6.24), we have k = n. Thus,
v2 = 2(m + k) + 4 = 2m + k + n + 4.
6.2. A TRANSLATION OPERATOR FOR INCIDENCE HILBERT SCHEMES 129
Case 2 .2: If one of the integers i, j and is positive, then v2 < (2m + k + n + 4).
The two remaining cases are listed by:
Case 3 : x = p = t (by symmetry, this also covers the case t = p = x);
Case 4 : x = p = t.
In these cases, the dimensions are smaller than (2m + k + n + 4). Therefore these
two cases have no contributions to the class v in (6.22).
Finally, note that the contribution of Case 1.1 to the class u in (6.20) and the
contribution of Case 1 .1 to the class v in (6.22) cancel out in the commutation:
(6.25) [ã−n (α), ãk (β)].
So [ã−n (α), ãk (β)] = 0 when n = k. This proves (6.17) when n = k.
When n = k, Case 2 .1 also contributes to the operator ãk (β)ã−n (α), and hence
to (6.25). Note from Case 2 .1 that π1245 (V2 ) consists of all the points of the form:
(ζ0 , ζ0 + p), x, x, (ζ0 , ζ0 + p)
in X [m+n,m+n+1] × X × X × X [m+n,m+n+1] , where x = p and p ∈ Supp(ζ0 ). Thus,
,
[ã−n (α), ãn (β)]|H ∗ (X [m+n,m+n+1] ) = c · (αβ) · IdH ∗ (X [m+n,m+n+1] )
X
for some constant c. Now we conclude from Lemma 6.4 and (3.9) that c = n. This
completes the proof of the commutation relation (6.17) when n = k.
Proof. We skip the proof of (ii) since it is similar to the proof of Lemma 6.3.
In the following, we prove (i). Note that the operator t̃† t̃ is induced by the class:
" ∗ #
(6.27) w = π13∗ π12∗ m ) · π23
τ (Q m ∈ A2m+2 (W )
Q
It follows that w = c[Δ] for some constant c, where Δ denotes the diagonal in
X [m,m+1] × X [m,m+1] . Hence, we get
t̃† t̃ = c IdH ∗ (X [m,m+1] ) .
To determine c, note that we can split off the factor ξ0 from our consideration,
i.e., we can simply consider the case m = 0. Then we have the morphism:
π13 : X × X [1,2] × X → X × X.
Fix a point p ∈ X. Since [Δ] · [{p} × X] = 1, we see from (6.27) that
c = w · [{p} × X]
" ∗ #
= π13∗ π12 ∗ 0 ) · π23
τ (Q 0 · [{p} × X]
Q
∗
= π12 0 ) · π23
τ (Q ∗
Q0 · [{p} × X [1,2] × X]
= [{p} × Up × X] · [{p} × Q 0 ]
where Up = (p, ξp )| ξp ∈ M2 (p) . It follows that
(6.29) 0 ] = [Up ] · φ1∗ [Q
c = [Up × X] · [Q 0 ] = [Up ] · [φ1 (Q
0 )]
where φ1 : X [1,2] × X → X [1,2] is the projection. We have
0 ) = {(x, ξx )| x ∈ X and ξx ∈ M2 (x) .
φ1 ( Q
Recall the natural morphism g2 : X [1,2] → X [2] . By (1.40),
0 )] = 1 g2∗ [M2 (X)]
[φ1 (Q
2
where M2 (X) = ∪x∈X M2 (x). Hence, we see from (6.29) that
1
c = [Up ] · g2∗ [M2 (X)]
2
1
= g2∗ [Up ] · [M2 (X)]
2
1
= [g2 (Up )] · [M2 (X)]
2
1
= [M2 (p)] · [M2 (X)]
2
= −1
where we have used the fact that [M2 (p)] · [M2 (X)] = −2 from (1.35).
The translation operator t̃ may look similar to the creation operators at the
first glimpse. However, we now show that it differs from the creation operators in
the essential way in that it commutes with all the annihilation operators.
Proposition 6.8. The translation operator t̃ and its adjoint t̃† commute with
the Heisenberg operators ã−n (α) for all n and α, i.e.,
[t̃, ã−n (α)] = [t̃† , ã−n (α)] = 0.
Proof. Let n > 0. Then Proposition 6.8 is decomposed into two parts:
(6.30) [t̃, ã−n (α)] = 0,
(6.31) [t̃, ãn (α)] = 0.
We prove them separately, and will compare the proof with that of Proposition 6.5.
132 6. LIE ALGEBRAS AND INCIDENCE HILBERT SCHEMES
Proof of ( 6.30):
Let n > 0. This part is similar to the proof of (6.16).
The operator ã−n (α)t̃ is induced by the class:
" [m+n+1,m+1] ∗ #
(6.32) w = π124∗ π123∗
Q m ∈ A2m+n+4 (W )
· π34 Q
Next, we compare the Heisenberg operators ãn (α) with the pull-back of the
Heisenberg operators an (α) via the morphism
gm+1 : X [m,m+1] → X [m+1] .
The comparison of annihilation operators is similar to Lemma 6.4. The proof of
the following lemma is omitted since it is similar to the proof of Lemma 6.4.
Lemma 6.9. Let n > 0 and α ∈ H ∗ (X). Then, we have a commutative diagram:
an (α)
H ∗ X [m+1] ←−−−− H ∗ X [m+n+1]
⏐ ⏐
⏐g ∗ ⏐g∗
% m+1 % m+n+1
ãn (α)
H ∗ X [m,m+1] ←−−−− H ∗ X [m+n,m+n+1] .
134 6. LIE ALGEBRAS AND INCIDENCE HILBERT SCHEMES
Lemma 6.9 will not hold for the creation operators. However, Proposition 6.11
below provides an explicit formula relating the creation operators. In order to prove
Proposition 6.11, we begin with a technical lemma.
Let m ≥ 0. Define Q m,0 to be the diagonal of X [m,m+1] × X [m,m+1] . For n ≥ 1,
define Qm,n to be the closed subset of X [m+n,m+n+1] × X [m,m+1] :
Qm,n = (ξ, ξ ), (η, η ) | ξ ⊃ ξ ⊃ η ⊃ η,
Supp(Iξ /Iξ ) = Supp(Iη /Iξ ) = Supp(Iη /Iη ) .
m,n has dimension (2m + n + 2), and contains exactly one irreducible
Note that Q
component of dimension (2m + n + 2) whose generic elements are of the form:
(6.36) (η + ξx , η + ξx ), (η, η + x) , x ∈ Supp(η).
Lemma 6.10. The restriction of t̃n to H ∗ (X [m,m+1] ) is given by the cycle
m,n ].
[Q
Cohomology rings of
Hilbert schemes of points
CHAPTER 7
t
(7.1) Gmj (βj , n)
j=1
t
satisfies (mj + 1) ≤ n0 . In fact, we see from the proof of Lemma 7.1 that an
j=1
t
expression (7.1) satisfies the upper bound (mj + 1) = n0 if and only if it is equal
j=1
s
s
to Gni −1 (αi , n), whose coefficient is ((−1)ni −1 ni !) by Lemma 4.14 (ii).
i=1 i=1
7.1. TWO SETS OF RING GENERATORS FOR THE COHOMOLOGY 143
t
whose coefficients are independent of X, α and n. Here mj ≤ k, and β1 , . . . , βt
j=1
t
depend only on eX , KX and α. In addition, mj = k if and only if the product
j=1
t
Bmj (βj , n) is equal to Bk (α, n), whose coefficient is equal to (−1)k /(k + 1)!.
j=1
u
(nj + 1) < (k + 1),
j=1
The following corollary asserts that the cohomology ring of the Hilbert scheme
X [n] is uniquely determined by the cohomology ring of the surface X and its canon-
ical class KX .
Corollary 7.6. Let X and Y be smooth projective complex surfaces. Assume
that there exists a ring isomorphism Φ : H ∗ (X) → H ∗ (Y ) with Φ(KX ) = KY . Then
for every n ≥ 1, the cohomology rings H ∗ (X [n] ) and H ∗ (Y [n] ) are isomorphic.
144 7. THE COHOMOLOGY RINGS OF HILBERT SCHEMES OF POINTS
whose coefficient is 1.
7.1. TWO SETS OF RING GENERATORS FOR THE COHOMOLOGY 145
ki
Proof. Put Ni = j=1 ni,j for 1 ≤ i ≤ s. For each i, we see from Lemma 7.1
that ⎛ ⎞
ki
1−(n−ki ⎝ a−ni,j (αi,j )⎠ |0
j=1 ni,j )
j=1
ti
ti
is a finite linear combination of products Gmi,j (βi,j , n) where (mi,j + 1) ≤
j=1 j=1
Ni , and βi,1 , . . . , βi,ti depend only on eX , KX , αi,1 , . . . , αi,ki and τj∗ with 1 ≤
j ≤ Ni . Moreover, the coefficients in the linear combinations are independent of
X, αi,1 , . . . , αi,ki and n. By Remark 7.2, the product satisfies the upper bound
ti ki
(mi,j + 1) = Ni if and only if it is equal to Gni,j −1 (αi,j , n). Furthermore,
j=1 j=1
ki
the coefficient of Gni,j −1 (αi,j , n) in
j=1
⎛ ⎞
ki
1−(n−ki ⎝ a−ni,j (αi,j )⎠ |0
j=1 ni,j )
j=1
ki
is equal to ((−1)ni,j −1 ni,j !).
j=1
So (7.3) is a universal finite linear combination of products
s
ti
Gmi,j (βi,j , n)
i=1 j=1
where
s
ti
s
s
ki
(mi,j + 1) ≤ Ni = ni,j .
i=1 j=1 i=1 i=1 j=1
Also,
s
ti
s
ki
(mi,j + 1) = ni,j
i=1 j=1 i=1 j=1
is equal to
s
ki
Gni,j −1 (αi,j , n)
i=1 j=1
N
s
ki
mp = ni,j
p=1 i=1 j=1
Theorem 7.7 indicates that the cup product on the Hilbert scheme X [n] is
independent of n in an appropriate sense. Furthermore, it provides an explicit
form of the leading term in the cup product. This result enables us to construct
the Hilbert ring in the next section.
S = S0 ∪ S1
then we put
(ρ) = (ρ(c)) = mr (c),
c∈S c∈S,r≥1
ρ = |ρ(c)| = r · mr (c),
c∈S c∈S,r≥1
Pn (S) = {ρ ∈ P(S) | ρ = n}.
Given
ρ = (ρ(c))c∈S = (r mr (c) )c∈S,r≥1 ∈ P(S)
7.2. THE HILBERT RING 147
and n ≥ 0, we define
a−ρ(c) (c) = a−r (c)mr (c) = a−1 (c)m1 (c) a−2 (c)m2 (c) · · ·
r≥1
aρ (n) = 1−(n−ρ) a−ρ(c) (c) · |0 ∈ H ∗ (X [n] )
c∈S
where we fix the order of the elements c ∈ S1 appearing in once and for all.
c∈S
Note from Definition 3.18 that aρ (n) = 0 for 0 ≤ n < ρ .
As ρ runs over all partition-valued functions on S with ρ ≤ n, the correspond-
ing aρ (n) linearly span H ∗ (X [n] ) as a corollary to Theorem 3.8. By Theorem 7.7
(for s = 2), we have the cup product
(7.5) aρ (n) · aσ (n) = aνρσ aν (n)
ν
∗
in H (X ), where ν ≤ ρ + σ and the structure coefficients aνρσ are indepen-
[n]
Note that the Hilbert ring does not depend on the choice of a linear basis S of
H ∗ (X) containing 1X since the operator an (α) depends on the cohomology class
α ∈ H ∗ (X) linearly. It follows from the super-commutativity and associativity
of the cohomology ring H ∗ (X [n] ) that the Hilbert ring HX itself is also super-
commutative and associative.
Proof. (i) Follows from Definition 7.9 and the observations in the preceding
paragraph.
(ii) Recall that as ρ runs over all partition-valued functions on S with ρ ≤ n,
the corresponding classes aρ (n) linearly span H ∗ (X [n] ). Given the Hilbert ring HX ,
the cup products (7.5) of these cohomology classes aρ (n) ∈ H ∗ (X [n] ) can be read
from the multiplications (7.7).
We also have the following result on the structure of the Hilbert ring HX . For
convenience, in the case when (ρ) = 1, that is, when the partition ρ(c) is a one-part
partition (r) for some element c ∈ S and is empty for all the other elements in S,
we will simply write aρ = ar,c and aρ (n) = ar,c (n).
Theorem 7.11. The Hilbert ring HX is isomorphic to the tensor product P ⊗E,
where P is the polynomial algebra generated by
ar,c , c ∈ S0 , r ≥ 1,
ar,c , c ∈ S1 , r ≥ 1.
By Lemma 7.1 and Lemma 7.4, the ring HX is generated by the elements ar,c ,
where c ∈ S = S0 ∪ S1 and r ≥ 1. By the super-commutativity of HX , we have
a2r,c = 0 for c ∈ S1 and r ≥ 1. It remains to show that as ρ = (r mr (c) )s∈S,r≥1 runs
over P(S), the monomials
am r (c)
r,c
c∈S,r≥1
i
where ai ∈ C and ρi = (r mr (c) )c∈S,r≥1 runs over a finite set I of distinct elements
in P(S). Since 1−(n−r) a−r (c)|0 = ar,c (n), we obtain
mir (c)
(7.8) ai · 1−(n−r) a−r (c)|0
i∈I c∈S,r≥1
i
= ai · ar,c (n)mr (c)
i∈I c∈S,r≥1
= 0
by the definition of the structure constants in HX .
Take an integer n large enough such that
def
n ≥ ni = rmir (c)
r,c
for all i ∈ I. From Theorem 7.7, we see that (7.8) can be rewritten as
⎛ ⎞
ai ⎝1−(n−ni ) (a−r (c))mr (c) · |0 + wi ⎠ = 0
i
(7.9)
i∈I c∈S,r≥1
where i satisfies ni = max{nj |j ∈ I}. Since all the integers ni in (7.10) are equal,
the Heisenberg monomials in (7.10) are linearly independent as a corollary to The-
orem 3.8. Thus all the coefficients ai in (7.10) are zero. By repeating the above
argument, we conclude that ai = 0 for all i ∈ I.
In Section 7.1 and the present section, the cohomology ring structure of the
Hilbert scheme X [n] has been studied and described. However, these descriptions
are very implicit in terms of the multiplication operators or explicit generators
with implicitly given relations. In [LS2, CoG], two new approaches to study the
cohomology ring of the Hilbert scheme X [n] appeared. In the next two sections, we
will survey these two papers respectively.
To generalize this to get a graded ring homomorphism ϕ∗ : A⊗I → A⊗J for any
surjective map ϕ : I → J of finite sets of cardinality n = |I| and k = |J|, choose a
bijection g : [k] → J and let ni = |ϕ−1 (g(i))|. Then there is a bijection f : [n] → I
such that for each i ∈ [k], we have
ϕ−1 (g(i)) = f (n1 + · · · + ni−1 + 1), . . . , f (n1 + · · · + ni ) .
The composition
f −1 ϕn g
ϕ∗ : A⊗I −→ A⊗n −→
•
A⊗k −→ A⊗J
,k
by
(7.15) mπ,ρ (a ⊗ b) = fπ,ρ,πρ f π,π,ρ (a) · f ρ,π,ρ (b) · eg(π,ρ) .
where g(π, ρ) is the graph defect defined in (7.13). The next result is [LS2, Propo-
sitions 2.13 and 2.14], and follows from (7.15) and repeated applications of (7.11).
Proposition 7.13. (i) The product A{Sn } × A{Sn } → A{Sn } defined
by
aπ · bρ = mπ,ρ (a ⊗ b)πρ
is associative, Sn -equivariant, and homogeneous of degree nd.
(ii) For any two homogeneous elements aπ, bρ ∈ A{Sn }, we have the following
(non)commutativity relation:
aπ · bρ = (−1)|a|·|b| π ∗ (b)πρπ −1 · aπ = (−1)|a|·|b| π̃ ∗ (bρ) · aπ.
By Proposition 7.13, A[n] is a subring of the center of A{Sn }. Let
T : A{Sn } → C
be defined by *
T (a), if π = 1,
T(aπ) =
0, else.
Then T induces a non-degenerate and Sn -invariant bilinear pairing on A{Sn }.
Therefore, the restriction of T to A[n] defines the structure of a graded Frobenius
algebra of degree nd on A[n] .
Let
V(A) = Sym(A ⊗ t−1 C[t−1 ])
be the bosonic Fock space modeled on the graded vector space A. Then V(A) is
bi-graded by degree and weight, where an element a ⊗ t−m ∈ A ⊗ t−m is given
degree |a| and weight m. The component of V(A) of weight n is the graded vector
space C
V(A)n ∼= Symαi A
αn i
where α = (1α1 2α2 · · · ) runs through all partitions of n. Let
f : {1, . . . , N } → π \[n]
be an enumeration of the orbits of π ∈ Sn , and let i = |f (i)|. Define Φ : A⊗N →
V(A) by
1
a1 ⊗ · · · ⊗ aN → (a1 ⊗ t−1 ) · · · (aN ⊗ t−N ),
n!
and let
+∞
Φ: A{Sn } → V(A)
n=0
⊗π\[n]
be given on the summand A · π by the composition
f −1 Φ
A⊗π\[n] −→ A⊗N −→ V(A).
Note that Φ is surjective and invariant under the Sn -action, so that its restriction
to A[n] is also surjective. Moreover, the vector spaces A[n] and V(A)n have the
same dimension. These observations lead to the following [LS2, Proposition 2.11].
154 7. THE COHOMOLOGY RINGS OF HILBERT SCHEMES OF POINTS
∂
∂=x .
∂x
Via (7.18), F n is linearly identified with H ∗ (X [n] ). Cup product with an element
+∞
a∈F = Fn
n=0
to be the centralizer in Diff H,K F of LK and the energy operator ∂. Since IMK
commutes with ∂, it projects to give a subalgebra of End(F n ) for each n ≥ 1.
An explicit construction of IMK as a polynomial algebra is given. The algebra
IMKX is obtained by degenerating
K = λ + KX
to KX as λ → 0. A description of IMKX in terms of degenerating families of
Dunkl-Cherednik operators is then obtained.
The following is the main result [CoG, Theorem 1.0.1].
Theorem 7.16. IMKX is a commutative algebra, and for every n the image
of IMKX in End(F n ) is precisely the algebra of left multiplication operators (i.e.,
the image of the natural map H ∗ (X [n] ) → End(F n ), a → a∪).
Theorem 7.16 together with the explicit structure of IMKX provides a new
description of the cohomology ring of X [n] . We refer to [CoG] for further details.
CHAPTER 8
4
I= H (X),
=1
the quotient ring H ∗ (X [n] )/I [n] is isomorphic to the cohomology ring H ∗ ((C2 )[n] ) of
the Hilbert schemes of points on the affine plane C2 . For this purpose, we will review
the cohomology ring H ∗ ((C2 )[n] ), following [ES2,Leh1,LS1,Vas]. Moreover, when
I = H 4 (X), we will show that modulo the ideal I [n] , the structure constants in
certain cup products of the Hilbert schemes X [n] are independent of n. The main
results and their proofs in this chapter are from [LQW5].
which is defined by
H∗BM (X) = H∗ (X̂, ∞)
where X̂ = X ∪ {∞} is the one-point compactification of X, and H∗ (X̂, ∞) is the
ordinary relative homology of the pair (X̂, ∞). Then the creation operators are
modeled on the Borel-Moore homology H∗BM (X), while the annihilation operators
are modeled on the ordinary homology H∗ (X). The Fock space of the Heisenberg
algebra is taken to be the direct sum
+∞
H∗BM (X [n] )
n=0
of the Borel-Moore homology groups H∗BM (X [n] ). Let Hc∗ (X) be the cohomology
with compact support. Using the the Poincaré dualities
PD : H 4−i (X) → HiBM (X)
PD : Hc4−i (X) → Hi (X),
we can regard the creation operators a−n (α) with n > 0 as being modeled on H ∗ (X)
(i.e., α ∈ H ∗ (X)), while we can regard the annihilation operators an (β) with n > 0
as being modeled on Hc∗ (X) (i.e., β ∈ Hc∗ (X)). Accordingly, with the help of the
Poincaré duality
PD : H 4n−i (X [n] ) → HiBM (X [n] ),
from now on we can take the Fock space to be the direct sum
+∞
(8.1) HX = H ∗ (X [n] )
n=0
In the rest of this section, we consider the affine plane X = C2 . Since H ∗ (C2 ) =
C · 1C2 , we see that the Fock space
+∞
HC2 = H ∗ (C2 )[n]
n=0
(ii) For each n, the cohomology ring H ∗ ((C2 )[n] ) is generated over C by the
Chern characters chk (OC2 )[n] .
Theorem 8.2 (i) is from [Leh1, Theorem 4.10], while Theorem 8.2 (ii) is a
classical result that follows from [ES2, Theorem 1.1]. Intuitively, Theorem 8.2 (i)
also follows from Theorem 4.7. Indeed, since KC2 = eC2 = 0, Theorem 4.7 says
that Gk (1C2 ) is equal to
1
(8.7) − aλ (τ∗ 1C2 ).
λ!
(λ)=k+2,|λ|=0
Note that aλ (τ∗ 1C2 ) is not zero only when the generalized partition λ has exactly
one negative part. Moreover, when λ has exactly one negative part, via the linear
isomorphism Ψ1 , the operator aλ (τ∗ 1C2 ) is of the form
qn0 +...+nk (−∂n0 ) · · · (−∂nk )
where n0 , . . . , nk > 0 (see (8.6)). Recalling the definition of λ! from Definition 1.6
and keeping track of the coefficient of qn0 +...+nk (−∂n0 ) · · · (−∂nk ), we conclude that
Theorem 8.2 (i) is a consequence of (8.7).
160 8. IDEALS OF THE COHOMOLOGY RINGS OF HILBERT SCHEMES
where π runs through all permutations with cycle type λ, i.e., those having a disjoint
cycle decomposition with cycle lengths λ1 , . . . , λ .
The class algebra C(Sn ) is the center of C[Sn ], and has a linear basis
{Cλ |λ n}.
The product on C(Sn ) is the cup product ∪ defined by
*
σ π, if d(σ) + d(π) = d(σπ),
σ∪π =
0, otherwise
where d(π) is defined in (7.12). Put
+∞
(8.10) C= C(Sn ).
n=0
for λ = (1 m1 m2
2 · · · ). The following is from [LS1, Theorem 1.1].
Theorem 8.3. The composite isomorphism of Fock spaces
Ψ Ψ
C −→
2
C[q1 , q2 , . . .] −→
1
HC2
induces an isomorphism of C-algebras
C(Sn ) −→ H ∗ (C2 )[n]
for each n ≥ 0.
The main idea in the proof of Theorem 8.3 is to use the alternating character
εn ∈ C(Sn ) defined by
εn = (−1)d(π) π = sgn(π) π.
π∈Sn π∈Sn
8.2. IDEALS IN H ∗ (X [n] ) FOR A PROJECTIVE SURFACE X 161
It is known that
m
n m−1 z
(8.13) Ψ2 (εn )z = exp (−1) qm .
m>0
m
n≥0
where γn denotes the total Chern class of the tautological bundle (OC2 )[n] . There-
fore, Ψ1 Ψ2 maps the alternating character εn ∈ C(Sn ) to the total Chern class of
(OC2 )[n] in H ∗ (C2 )[n] . In addition, via some technical computation, one proves
that
Ψ1 Ψ2 (εn ∪ w) = γn ∪ Ψ1 Ψ2 (w)
for every w ∈ C(Sn ). Now Theorem 8.3 follows since by Theorem 8.2 (ii), the
cohomology ring H ∗ (C2 )[n] is generated by the Chern classes of (OC2 )[n] .
such that for each fixed j, there exists some with αj, ∈ I.
Proof. First of all, note from the third formula in (3.11) that if
τk∗ α = αj,1 ⊗ αj,2 ⊗ · · · ⊗ αj,k ,
j
then
τ(k+1)∗ α = (τ2∗ αj,1 ) ⊗ αj,2 ⊗ · · · ⊗ αj,k .
j
Therefore, by induction, it suffices to prove the lemma for k = 2. Now the case
k = 2 follows from the first formula in (3.11) that if we write
τ2∗ (1X ) = αi ⊗ βi ,
i
then
τ2∗ (α) = (ααi ) ⊗ βi
i
with (ααi ) ∈ I.
In view of the preceding lemma, we introduce the following important definition.
162 8. IDEALS OF THE COHOMOLOGY RINGS OF HILBERT SCHEMES
So 1 = σi (1) since σi (1) < · · · < σi (i). By Theorem 4.7 and Lemma 3.12 (i), (8.15)
is a linear combination of expressions of the form
⎛ ⎞
(8.16) ⎝ a−n (α )⎠ a−λ (τ∗ (αασi (1) · · · ασi (i) ))|0
∈σi0
where ∈ {1X , eX , KX , KX2
}, λ ij=1 nσi (j) , and (λ) = k + 2 − ||/2 − i. By
Lemma 8.4, the expression (8.16) is contained in I [n] since ασi (1) = α1 ∈ I. It
follows that (8.15) is contained in I [n] . This proves (8.14).
(ii) This follows from Theorem 4.10 and Lemma 8.4.
The following technical definition is similar to Definition 4.4.
Definition 8.7. Let X be a smooth projective complex surface, s ≥ 1, and
t1 , . . . , ts ≥ 1. Fix mi,j ≥ 0 and βi,j ∈ H ∗ (X) for 1 ≤ i ≤ s and 1 ≤ j ≤ ti . Then, a
ti
universal linear combination of Gmi,j (βi,j , n), 1 ≤ i ≤ s is a linear combination
j=1
of the form
s
ti
fi Gmi,j (βi,j , n)
i=1 j=1
where the coefficients fi are independent of X and n.
8.2. IDEALS IN H ∗ (X [n] ) FOR A PROJECTIVE SURFACE X 163
The next result deals with the generators of the ideal I [n] .
Proof. Note that every Heisenberg monomial class in I [n] can be written as
s
A = 1−(n−n0 ) a−ni (αi ) |0
i=1
s
where s ≥ 1, n1 , . . . , ns ≥ 1, n0 = ni , and α is contained in I and homogeneous
i=1
for some . By Lemma 8.6 (ii), it suffices to show that A ∈ I [n] is a universal finite
linear combination of expressions of the form
t
(8.17) Gmj (βj , n)
j=1
t
where (mj + 1) ≤ n0 , and β ∈ I for some .
j=1
Use induction on n0 . When n0 = 1, s = n1 = 1. So
(defined to be the leading term) plus a universal finite linear combination of expres-
sions
s̃
1−(n−ñ0 ) a−ñi (α̃i ) |0
i=1
s̃
s
ñi = ñ0 < (ki + 1) = n0 .
i=1 i=1
Remark 8.9. The assumption in Theorem 8.8 that the ideal I ⊂ H ∗ (X) is
homogeneous can be dropped when the surface X is simply connected.
164 8. IDEALS OF THE COHOMOLOGY RINGS OF HILBERT SCHEMES
Remark 8.10.
(i) Let X be a smooth projective complex surface, and I be a homogeneous
ideal in the cohomology ring H ∗ (X). Assume further that for every k ≥ 1
and α ∈ I, the pushforward τk∗ α can be written as
αj,1 ⊗ · · · ⊗ αj,k ∈ H ∗ (X k )
j
s
where s ≥ 1, n1 , . . . , ns ≥ 1, n0 = ni , and α1 , . . . , αs ∈ I. Then the
i=1
same proof of Theorem 8.8 shows that A can be written as a polynomial
of the classes Gk (α, n), k ≥ 0 and α ∈ I. Moreover, the coefficients in the
polynomial are independent of n.
(ii) Under certain conditions, the conclusion in (i) can be sharpened. For
instance, let λ n0 , α ∈ H ∗ (X) with |α| = 2, and m ≥ 1.
(ii-a) The class 1−(n−n0 ) a−λ (x)|0 can be written as a polynomial of the
classes Gk (x, n), k ≥ 0. Moreover, the coefficients are independent of
n and X.
(ii-b) If the odd Betti numbers of the surface X are equal to zero, then
1−(n−n0 −m) a−λ (x)a−m (α)|0
= α, KX · F1 (n) + Gki (α, n) · F2,i (n)
i
where F1 (n) and F2,i (n) are polynomials of the classes Gk (x, n), k ≥
0. Moreover, the coefficients of F1 (n) and F2,i (n) are independent of
n, α and X.
The statements (ii-a) and (ii-b) follow from the same proof of Theorem 8.8
by setting I = C · x ⊂ H ∗ (X) and I = C · x + C · α ⊂ H ∗ (X) respectively.
A natural problem to consider is to interpret the quotient ring
H ∗ (X [n] )/I [n]
for an ideal I of the cohomology ring H ∗ (X). In the next two sections, we will do
this for I = ⊕4=1 H (X) and I = H 4 (X) respectively.
8.3. Relation with the cohomology ring of the Hilbert scheme (C2 )[n]
Let X denote a smooth projective complex surface. In this section, we will
study the quotient ring H ∗ (X [n] )/I [n] when
4
I= H (X),
=1
and prove that it is isomorphic to the cohomology ring H ∗ ((C2 )[n] ) which has been
dealt with in Section 8.1.
8.3. RELATION WITH THE COHOMOLOGY RING OF (C2 )[n] 165
First of all, note that H ∗ (X [n] )/I [n] has a linear basis consisting of Heisenberg
monomials of the form
a−n1 (1X )r1 · · · a−nk (1X )rk |0
k
where r1 , . . . , rk ≥ 1, and 0 < n1 < . . . < nk with =1 r n = n. Therefore, we
have an isomorphism of vector spaces
(8.18) Φ: H ∗ (X [n] )/I [n] → C[q1 , q2 , . . .]
n≥0
defined by
Φ a−n1 (1X )r1 · · · a−nk (1X )rk |0 = qnr11 · · · qnrkk .
Setting the degree of the variable qi to be i, we see that Φ maps H ∗ (X [n] )/I [n] to
the homogeneous component of C[q1 , q2 , . . .] of degree n.
Lemma 8.11. Let
4
I= H (X).
=1
∗ [n] [n]
Then, the quotient ring H (X )/I is generated by the Chern character classes
Gk (1X , n), k = 0, 1, . . . , n − 1. Moreover,
(−1)k
(8.19) Gk (1X , n) ≡ · 1−(n−k−1) a−(k+1) (1X )|0 (mod I [n] ).
(k + 1)!
Proof. Since the ring H ∗ (X [n] ) is generated by the classes Gk (α, n) with
0 ≤ k < n and α ∈ H ∗ (X), the first statement follows from Lemma 8.6 (ii). To
prove (8.19), we note from Theorem 4.10 that the leading term in Gk (1X , n) is
(−1)k
· 1−(n−k−1) a−(k+1) (1X )|0
(k + 1)!
corresponding to j = k, λ (j + 1) = k + 1, and (λ) = k − j + 1 = 1. The other
terms in Gk (1X , n) contain τi∗ () with either i ≥ 2 or = eX , KX , KX
2
∈ I, and
hence are contained in I . This proves (8.19).
[n]
Theorem 8.12. Let X be a smooth projective complex surface, and
4
I= H (X).
=1
∗
Then, the quotient H (X [n]
)/I [n]
is isomorphic to the cohomology ring H ∗ ((C2 )[n] ).
Proof. By Lemma 8.11, the quotient ring H ∗ (X [n] )/I [n] is generated by the
classes Gk (1X , n), k = 0, 1, . . . , n − 1. So by Theorem 8.2, it suffices to show that
via the isomorphism Φ in (8.18), the linear operator gk on C[q1 , q2 , . . .] induced by
the operator Gk (1X ) on H ∗ (X [n] ) is given by
n≥0
(−1)k
(8.20) gk = qn1 +...+nk+1 ∂n1 · · · ∂nk+1 .
(k + 1)! n
1 ,...,nk+1 >0
Indeed, let
A = a−n1 (1X ) · · · a−nb (1X )|0 ∈ H ∗ (X [n] )
166 8. IDEALS OF THE COHOMOLOGY RINGS OF HILBERT SCHEMES
where n1 , . . . , nb > 0 with n = n. By Lemma 4.9, Gk (1X )(A) is equal to
(8.21)
k+1
a−n (1X ) · [[· · · [Gk (1X ), a−nσi (1) (1X )], · · · ], a−nσi (i) (1X )]|0
i=0 σi ∈σi0
where for each fixed i, σi runs over all the maps { 1, . . . , i } → { 1, . . . , b } satisfying
σi (1) < · · · < σi (i), and
σi0 = { | 1 ≤ ≤ b, = σi (1), . . . , σi (i)}.
Note from Lemma 3.12 that
[[· · · [at1 · · · atr (τr∗ α), a−nσi (1) (1X )], · · · ], a−nσi (i) (1X )]|0 ∈ I [n]
if i ≤ r − 2 or α ∈ I. Hence by (8.21) and Theorem 4.7, Gk (1X )(A) equals
⎛ ⎞
⎝ a−n (1X )⎠ ·
σk+1 0
∈σk+1
⎡⎡ ⎡ ⎤ ⎤ ⎤
1
⎣⎣· · · ⎣− aλ (τ∗ 1X ), a−nσk+1 (1) (1X )⎦ , · · · ⎦ , a−nσk+1 (k+1) (1X )⎦ |0
λ!
(λ)=k+2,|λ|=0
k+1 ⎛ ⎞
− (−nσk+1 () ) ⎝ a−n (1X )⎠ a−nσk+1 (1) −...−nσk+1 (k+1) (1X )|0 .
σk+1 =1 0
∈σk+1
1−(n−ρ−(ρ(1X )))
(8.23) = mr (ρ̃(1X )) m (ρ̃(1 ))!)
a−r (c)mr (ρ̃(c)) · |0 .
r≥2 (r r X c∈S,r≥1
c=1X or r>1
where we fix the order of the elements c ∈ S1 appearing in the product once
c∈S
and for all. For 0 ≤ n < ρ + (ρ(1X )), we set bρ (n) = 0. This is consistent with
(8.23) and Definition 3.18. We remark that the only part in bρ (n) involving n is
the factor 1−(n−ρ−(ρ(1X ))) in (8.23) when n ≥ ρ + (ρ(1X )).
Remark 8.13. The Heisenberg monomial class bρ (n) is different from the
Heisenberg monomial class aρ (n) defined in Section 7.2.
By Theorem 3.8, H ∗ (X [n] ) has a linear basis consisting of the classes:
(8.24) bρ (n), ρ ∈ P(S) and ρ + (ρ(1X )) ≤ n.
Fix a positive integer n and choose ρ, σ ∈ P(S) satisfying ρ + (ρ(1X )) ≤ n and
σ + (σ(1X )) ≤ n. Then we can write the cup product bρ (n) · bσ (n) as
(8.25) bρ (n) · bσ (n) = bνρσ (n) bν (n)
ν∈P(S)
In the following, we will study various cup products in the cohomology of X [n]
modulo the ideal I [n] with I = H 4 (X). We begin with an elementary topological
lemma.
Lemma 8.15. If k ≥ 2 and α is homogeneous, then
τk∗ α = αj,1 ⊗ · · · ⊗ αj,k
j
where for each j, either |αj, | = 4 for some , or 0 < |αj, | < 4 for every .
Proof. Assume |αj, | < 4 for every . If |αj, | = 0 for some , then
k
4(k − 1) + |α| = |τk∗ α| = |αj,i | ≤ 3(k − 1).
i=1
So (k − 1) + |α| ≤ 0, contradicting k ≥ 2.
s
Lemma 8.16. Let I = H 4 (X), s ≥ 0, n1 , . . . , ns > 0, ñ = =1 n , and
n ≥ ñ. Let α, α1 , . . . , αs ∈ H ∗ (X) be homogeneous. Assume that k + |α| ≥ 1 and
n + |α | ≥ 2 for every . Then modulo I [n] , the cup product
Gk (α, n) · 1−(n−ñ) a−n1 (α1 ) · · · a−ns (αs )|0
is a universal linear combination of the basis ( 8.24).
Proof. By Lemma 8.6 (i) and (ii), the statement is trivial if one of the classes
α, α1 , . . . , αs ∈ H ∗ (X) is contained in I. So in the rest of the proof, we assume
that none of the classes α, α1 , . . . , αs is contained in I.
Our argument is similar to the proof of Lemma 8.6 (i). Put g = Gk (α) and
B = Gk (α, n) · 1−(n−ñ) a−n1 (α1 ) · · · a−ns (αs )|0 .
Then, B is equal to
1
ga−1 (1X )n−ñ a−n1 (α1 ) · · · a−ns (αs )|0 .
(n − ñ)!
8.4. PARTIAL n-INDEPENDENCE OF STRUCTURE CONSTANTS 169
·[[· · · [[[[g, a−1 (1X )], . . .], a−1 (1X )], a−nσi (1) (ασi (1) )], · · · ], a−nσi (i) (ασi (i) )]|0
/ 01 2
j times
in (8.28) does not contain a−1 (1X ). Since n +|α | ≥ 2 for every , this is equivalent
to show that modulo I [(λ)] , the part a−λ (τ∗ (αασi (1) · · · ασi (i) ))|0 in (8.28) does
not contain a−1 (1X ). By Lemma 8.15, this is true if (λ) ≥ 2. So let (λ) = 1.
Then, we have
a−λ (τ∗ (αασi (1) · · · ασi (i) ))|0 = a−t (αασi (1) · · · ασi (i) )|0
where t = |λ| = j + nσi (1) + · · · + nσi (i) and k + 2 − ||/2 − (j + i) = 1.
If a−λ (τ∗ (αασi (1) · · · ασi (i) ))|0 contains a−1 (1X ), then we must have t = 1 and
|αασi (1) · · · ασi (i) | = 0. So j + nσi (1) + · · · + nσi (i) = 1, = 1X , and
|α| = |ασi (1) | = . . . = |ασi (i) | = 0.
Thus, either j = 0, i = 1, nσ1 (1) = 1 and |ασ1 (1) | = 0, or j = 1 and i = 0. The first
case contradicts nσ1 (1) + |ασ1 (1) | ≥ 2. In the second case, we see from k + 2 − ||/2 −
(j + i) = 1 that k = 0, contradicting k + |α| ≥ 1. So a−λ (τ∗ (αασi (1) · · · ασi (i) ))|0
can not contain a−1 (1X ).
170 8. IDEALS OF THE COHOMOLOGY RINGS OF HILBERT SCHEMES
The next lemma studies the basis element bρ (n) modulo I [n] when I = H 4 (X).
Lemma 8.18. Let I = H 4 (X). Then modulo I [n] , the basis element bρ (n) is
t
a universal finite linear combination of products of the form Gmj (βj , n) where
j=1
mj + |βj | ≥ 1 for each j, and
t
(mj + 1) ≤ ρ + (ρ(1X )).
j=1
where m1,j + |β1,j | ≥ 1 for every j. Therefore, bρ (n) · bσ (n) is a universal finite
linear combination of expressions of the form
t
Gmj (βj , n)
j=1
where mj + |βj | ≥ 1 for every j. Now our result follows from Lemma 8.17 (i).
statement is true. Let X = P(L1 ⊕ L2 ) where L1 and L2 are two invertible sheaves
over C. Let σ (respectively, σ ) be the section of X → C corresponding to the
natural surjection L1 ⊕ L2 → L1 → 0 (respectively, L1 ⊕ L2 → L2 → 0). Put
X = X − σ. Then X satisfies the S-property. To see this, let X = X − σ , and
notice that X and X are affine bundles over C. Hence X is homotopic to C, and
H i (X) ∼= H i (C) for every i. Therefore, to verify the S-property of X, it remains
to verify the surjectivities of the induced homomorphisms ri : H i (X) → H i (X) for
i = 1, 2. Consider the relative cohomology group H 2 (X, X). We have
H 2 (X, X) ∼= H 2 (X − σ , X − σ ) = H 2 (X , X − σ)
by the excision theorem. By the Thom isomorphism, since X is an affine bundle
over C with σ being the zero section, we have
H 2 (X , X − σ) ∼
= H 0 (C) ∼
= C.
Hence H 2 (X, X) ∼
= C. Now consider the exact sequence
r δ r
H 1 (X) −→ 1
H 1 (X) −→ H 2 (X, X) −→ H 2 (X) −→2
H 2 (X).
Since H 2 (X) = ∼ H 2 (C) =
∼ C and H 2 (X) ∼
= C ⊕ C, we conclude that the map δ must
be zero and the map r2 must be surjective. Therefore, r1 is also surjective.
In the following, let X be a smooth quasi-projective surface satisfying the S-
property. So X can be embedded in a smooth projective surface X such that the
induced ring homomorphism ι∗ : H ∗ (X) → H ∗ (X) is surjective where ι : X → X is
the inclusion map. Let I = ker(ι∗ ). Fix a linear basis S of H ∗ (X) as in Section 8.4
such that S contains a linear basis SI of I and ι∗ (S −SI ) is a linear basis of H ∗ (X).
Put
SX = ι∗ (S − SI )
By Lemma 8.1, ker(ι∗n ) = I [n] which is defined in Definition 8.5. Also, a linear basis
of H ∗ (X [n] ) is given by
(8.30) bρX (n), ρX ∈ P(SX ) and ρX + (ρX (1X )) ≤ n
where bρX (n) is defined in a similar way as in (8.22) and (8.23). So for ρX , σX ∈
P(SX ), we can write the cup product bρX (n) · bσX (n) as
(8.31) bρX (n) · bσX (n) = bνρX
X σX
(n) bνX (n)
νX ∈P(SX )
where bνρX
X σX
(n) ∈ C stands for the structure constants.
Theorem 8.24. Let X be a smooth quasi-projective surface satisfying the S-
property. Then the structure constants bνρX
X σX
(n) in ( 8.31) are independent of n.
Proof. Let notations be as above. Note that ι∗ (S − SI ) = SX . Define ρ ∈
P(S) by *
mr (ρX (ι∗ c)), if c ∈ (S − SI ),
mr (ρ(c)) =
0, if c ∈ SI .
Similarly, define σ ∈ P(S) by
*
mr (σX (ι∗ c)), if c ∈ (S − SI ),
mr (σ(c)) =
0, if c ∈ SI .
8.5. APPLICATIONS TO QUASI-PROJECTIVE SURFACES 173
By Theorem 8.19,
[n]
bρ (n) · bσ (n) ≡ bνρσ (n) bν (n) (mod I )
ν∈P(S−{[x]})
where I = H 4 (X) and all the bνρσ (n) are independent of n. Since I ⊂ I,
(8.32) bρ (n) · bσ (n) ≡ bνρσ (n) bν (n) (mod I [n] )
ν∈P(S−SI )
where all the structure constants bνρσ (n) are independent of n. By Lemma 8.1,
ι∗n (bρ (n)) = bρX (n) and ι∗n (bσ (n)) = bσX (n). Therefore, applying ι∗n to (8.32), we
see that all the structure constants bνρX X σX
(n) in (8.31) are independent of n.
Definition 8.25. Let X be a smooth quasi-projective surface satisfying the
S-property. We define the FH ring GX associated to X to be the ring with a linear
basis given by the symbols bρX , ρX ∈ P(SX ), with the multiplication given by
bρX · bσX = bνρX
X σX
bνX
νX ∈P(SX )
The cohomology H ∗ (X [n] ) of the Hilbert scheme X [n] with complex coefficients
has been investigated in previous chapters. By Theorem 3.8, it is linearly spanned
over C by the Heisenberg monomial classes
a−n1 (α1 ) · · · a−nk (αk )|0
where n1 , . . . , nk > 0, n1 + . . . + nk = n, and the cohomology classes α1 , . . . , αk run
over a fixed linear basis of H ∗ (X). Its ring structure was discussed in Chapters 7
and 8. A natural problem is to study the integral cohomology of the Hilbert scheme
X [n] , its integral linear basis, and its integral generators. This chapter is devoted to
these issues. We will define integral operators, show that some familiar operators are
integral, and construct some integral operators by using integral classes in H 2 (X).
Moreover, we will prove that the intersection matrix of certain integral classes in
the middle cohomology H 2n (X [n] ) has determinant ±1. Then we will write down
a linear basis of the integral cohomology of X [n] , modulo the torsion part. As an
application, we will compare two integral bases of the integral cohomology of the
Hilbert scheme (P2 )[n] of points on the complex projective plane P2 , and prove that
when the basis elements are arranged properly, the transition matrix is a triangular
matrix. Further applications can also be found in [Kap, KaM]. Theorem 9.12,
Theorem 9.22 and their proofs are from [QW2], while Theorem 9.14, Theorem 9.23
and their proofs are from [LQ3]. The proof of Theorem 9.35 follows the presentation
of [Mur], while we refer to [Mur, Subsection 3.3] for the proof of Proposition 9.32.
A linear basis of H ∗ (X [n] ) is integral if and only if its members are integral
classes and the matrix formed by the pairings of its members is unimodular. In the
next two lemmas, we prove that the Heisenberg operators an (α) and the operators
1−n are integral.
Lemma 9.2. (i) If f ∈ End(HX ) is integral, then so is its adjoint f† ;
(ii) The Heisenberg operators an (α), n ∈ Z are integral if α ∈ H ∗ (X) is integral.
Proof. (i) Note that a class A ∈ HX is integral if and only if A, B is an
integer whenever B ∈ HX is an integral class. It follows that f ∈ End(HX ) is
integral if and only if its adjoint operator f† ∈ End(HX ) is integral.
(ii) Recall that a0 (α) = 0. Next, fix n > 0. By Definition 3.3, a−n (α) is
integral. By Lemma 3.7 and (i), an (α) is integral as well.
Lemma 9.3. For n ≥ 0, the operator 1−n is integral.
Proof. Recall from Definition 3.18 that 1−n = 1/n! · a−1 (1X )n . Fix any
integer m ≥ 0 and an integral class A ∈ H j (X [m] ). For i = m, . . . , m + n − 1,
recall the subscheme Q[i+1,i] of X [i+1] × X × X [i] defined by (3.6), and let Qi be
the image of Q[i+1,i] under the natural projection
X [i+1] × X × X [i] → X [i+1] × X [i] .
Set-theoretically,
Qi = {(ξi+1 , ξi ) ∈ X [i+1] × X [i] | ξi+1 ⊃ ξi }.
Let φi,1 and φi,2 be the two projections of X [i+1] × X [i] . Let
Y = X [m+n] × · · · × X [m+1] × X [m] .
By Definition 3.3,
a−1 (1X )(A) = (φm,1 )∗ ([Qm ] · φ∗m,2 A).
Repeating this process and using the projection formula, we conclude that
a−1 (1X )n (A) = (φ̃m+n )∗ ([Q] · φ̃∗m A)
where φ̃i denotes the projection of Y to X [i] and
Q = {(ξm+n , . . . , ξm+1 , ξm ) ∈ Y | ξm+n ⊃ · · · ⊃ ξm+1 ⊃ ξm }
(the scheme structure on Q can be described similarly as that on Qi ). To show that
1−n (A) ∈ H j (X [m+n] ) is integral, it suffices to prove that the intersection number
(φ̃m+n )∗ [Q] · φ̃∗m A · B
is divisible by n! for any integral class B ∈ H 4(m+n)−j (X [m+n] ).
Represent the integral class A by a piecewise smooth cycle WA ⊂ X [m] . Let
QA = {(ξm+n , . . . , ξm+1 , ξm ) ∈ Y | ξm+n ⊃ · · · ⊃ ξm+1 ⊃ ξm and ξm ∈ WA }.
Then, the integral class [Q] · φ̃∗m A is represented by QA . Note that
φ̃m+n |QA : QA → φ̃m+n (QA )
is generically finite, and a generic element in φ̃m+n (QA ) is of the form ξm + x1 +
. . . + xn where ξm ∈ WA is generic, the points x1 , . . . , xn are distinct, and
Supp(ξm ) ∩ {x1 , . . . , xn } = ∅.
9.1. INTEGRAL OPERATORS 177
k
k
aA (B) = aA 1−mi (Bi ) = 1−mi aA (Bi ) .
i=1 i=1
Proof. Let
n
c (L[n] ) = ci (L[n] )i
i=0
be the generating function for the total Chern classes of L[n] , where is a formal
variable. By Proposition 4.19,
⎛ ⎞
+∞ (−)r−1
(9.4) c (L[n] )wn = exp ⎝ a−r (c (L))wr ⎠ · |0
n=0
r
r≥1
9.1. INTEGRAL OPERATORS 179
Proof. (ii) follows from (i) and Proposition 9.5. To prove (i), we let n = |λ|
and use induction on n. Our result is trivially true when n = 0, 1. In the following,
assume that 1/zμ · a−μ (1X )|0 is integral whenever
|μ| < n.
Let λ = (1m1 2m2 3m3 · · · ) so that n = r rmr . First of all, assume mr > 0
for at least two different r’s. Then, rmr < n for every r. Putting μr = (r mr )
for r ≥ 1 and applying induction to the partitions μr , we obtain integral classes
1/zμr · a−μr (1X )|0 . Put
1
Ar = · a−μr (1X )|0 .
zμ r
By Proposition 9.5, the operators aAr = 1/zμr · a−μr (1X ) are integral. Thus,
1/zλ · a−λ (1X )|0 = 1/zμr · a−μr (1X ) · |0
r≥1
is integral.
We are left with the case when mr > 0 for a unique r. In this case, |λ| = rmr =
n and (λ) = mr . Applying (9.3) to L = OX and i = n − mr , we have
[n]
a−ν (1X )|0
(9.6) ci (OX ) = (−1)i ·
zν
|ν|=rmr , (ν)=mr
Note that for any partition ν = (1t1 2t2 · · · ) satisfying ν = λ, |ν| = rmr and (ν) =
mr , there are at least two i’s with ti > 0. By the previous paragraph, 1/zν ·
a−ν (1X )|0 is integral. Hence, 1/zλ · a−λ (1X )|0 is integral by (9.6).
180 9. INTEGRAL COHOMOLOGY OF HILBERT SCHEMES
Next, we recall some notations from Section 1.2. Let Λ be the ring of symmetric
functions in infinitely many variables, and
ΛC = Λ ⊗Z C.
Let Λn and ΛnC be the degree-n parts in Λ and ΛC respectively. For a partition λ,
let pλ , mλ and sλ be the power-sum symmetric function, the monomial symmetric
function and the Schur function respectively.
In [Nak5, (9.13)], Nakajima defined a linear isomorphism
(9.11) ΦC : ΛC = ΛnC → HC
n≥0
Our goal in this section is to prove that mλ,α ∈ End(HX ) is an integral operator
if α is a divisor. In view of (9.13) and Proposition 9.5, it suffices to show that [Lλ α]
is an integral class. In the next three lemmas, we will study the properties of the
class [Lλ α].
Lemma 9.9. Let α ∈ H 2 (X) and i > 0. Then, we have
(9.17) a−i (α)[Lλ α] = aλ,μ [Lμ α]
μ
Next, recall from (1.8) the involution ω on Λ and from (1.9) the forgotten sym-
metric function fλ = ω(mλ ) associated to a partition λ. The forgotten symmetric
function fλ is an integral linear combination of the monomial symmetric functions
mμ , μ |λ|. Thus ΨC,α ◦ ΦC (fλ ) is an integral linear combination of the classes
ΨC,α ◦ ΦC (mλ ) = [Lμ α], μ |λ|, and hence is an integral class. By (9.19),
Combining (9.20) and (9.21), we conclude that [Lλ (−α)] is integral as well.
Proof. We start with some notations. For the partitions μ obtained from λ
as in (9.17), we will denote μ = λ ↑i . If we specify further that such a μ is obtained
from adding i to a part of λ equal to j (here j is allowed to be 0), then we denote
μ = λ ↑ij . Given a partition λ, we denote by mk (λ) the multiplicity of the parts of
λ equal to k. In these notations, the coefficient aλ,μ in Lemma 9.9 for μ = λ ↑ij is
simply equal to mi+j (μ). Denote the right-hand-side of (9.22) by Rλ (α).
Claim. a−i (α)Rλ (α) = aλ,μ Rμ (α).
μ=λ↑i
Note that the partitions ρ1 ∪λ2 and λ1 ∪ρ2 associated to λ1 , λ2 , ρ1 and ρ2 appearing
above are of the form λ ↑i . Thus, the same types of terms appear on both sides of
the Claim. It remains to identify the coefficients of a given term.
Fix μ = λ ↑ij for some part j of λ, and fix μ1 , μ2 such that μ1 ∪μ2 = μ. From the
above computation, the contributions to the term mμ1 ,α1 mμ2 ,α2 |0 in the left-hand-
side of the Claim come from two places: the term in (9.24) for ρ1 = μ1 , λ2 = μ2
whose coefficient is mi+j (μ1 ), and the term in (9.25) for λ1 = μ1 , ρ2 = μ2 whose
coefficient is mi+j (μ2 ). Therefore in view of (9.23), the coefficients of the term
mμ1 ,α1 mμ2 ,α2 |0 in both sides of the Claim coincide thanks to
(9.26) mi+j (μ1 ) + mi+j (μ2 ) = mi+j (μ1 ∪ μ2 ) = mi+j (μ) = aλ,μ .
This completes the proof of the above Claim.
Next, we continue the proof of (9.22) by using induction on n and the reverse
dominance ordering of partitions λ of n (see Definition 1.2 for the dominance or-
dering). For n = 1, formula (9.22) is clear. Assume that formula (9.22) holds for
all partitions of size less than n.
For λ = (n), (9.22) holds since mλ,α = a−n (α) and [Lλ α] = a−n (α)|0 .
the
For a general partition λ of n with a part equal to, say, i, we denote by λ
partition obtained from λ with a part equal to i removed. Now replacing λ in (9.17)
and the above Claim by λ respectively, we obtain
(9.27) a−i (α)[Lλ α] = aλ,μ μ
[L α],
i
μ=λ↑
(9.28) a−i (α)Rλ (α) = aλ,μ μ
R (α).
i
μ=λ↑
Note that λ appears among the above μ’s as the maximum in the reverse dominance
ordering. By induction hypothesis, the left-hand-sides of (9.27) and (9.28) coincide,
and all the terms on the right-hand-sides of (9.27) and (9.28) involving μ not equal
= mi (λ) = 0.
to λ coincide. Thus (9.22) follows since aλ,λ
is represented precisely by the tautological rank-n vector bundle (Lα )[n] defined in
(1.25). Therefore, we obtain from Lemma 9.6 that
(−1)i−(μ) a−λ (1X )a−μ (α)|0
(9.32) ci (p1 )! (Vα ) = .
zλ zμ
|λ|+|μ|=n
(λ)=n−i
9.4. UNIMODULARITY 185
Finally, we conclude from (9.31) and Lemma 4.13 that (9.32) holds for a general
integral class α ∈ H 2 (X). Setting i = n in (9.32) and using (9.16), we obtain the
integral class
(−1)n−(μ) a−μ (α)|0 n
cn (p1 )! (Vα ) = = [L(1 ) α].
zμ
|μ|=n
Theorem 9.14. Let the class α ∈ H 2 (X) be integral. Then for every partition
λ, the class [Lλ α] and the operator mλ,α are integral.
Proof. Recall from Section 1.2 that the monomial symmetric functions
m(1i ) , i≥1
generate the integral ring Λ. So the monomial symmetric function mλ can be
written as an integral polynomial P of the functions m(1i ) , i ≥ 1, i.e.,
E
mλ = P m(1i ) E i ≥ 1 .
Combining this with the two formulas in (9.15), we see that
E E
(9.33) [Lλ α] = Φα (mλ ) = Φα P m(1i ) E i ≥ 1 = P m(1i ),α E i ≥ 1 |0 .
Hence, the class [Lλ α] is integral by Lemma 9.13. Finally, we conclude from Propo-
sition 9.5 and (9.13) that the operator mλ,α is integral.
9.4. Unimodularity
Now Lemma 9.4, Lemma 9.7 and Theorem 9.14 have provided us with three
types of integral operators:
1
a−λ (1X ), a−μ (x), mν,α
zλ
where α ∈ H 2 (X) is integral. These operators enable us to write down integral
classes for the Hilbert scheme X [n] . To ensure that a certain set of these integral
classes is an integral linear basis, we must show that their intersection matrix is
unimodular. In this section, we will prove that if the intersection matrix of the
cohomology classes α1 , · · · , αk ∈ H 2 (X) has determinant ±1, then so does the
intersection matrix of the following cohomology classes in H 2n (X [n] ):
mλ1 ,α1 · · · mλk ,αk |0 , |λ1 | + . . . + |λk | = n.
First of all, we begin with some linear algebra preparations. Let V be a k-
dimensional complex vector space with a symmetric bilinear form
·, · : V × V → C.
Fix a linear basis v = {v1 , . . . , vk } of V , and let
Mv = vi , vj 1≤i,j≤k
i
{v1i1 · · · vk−1 ṽk | i1 + . . . + ik−1 + ik = n}.
k−1 ik
Since the transition matrix from the second basis to the first basis is lower triangular
with all diagonal entries being 1, we conclude that
(9.38) det Mn,v = det Mn,ṽ .
ik−1 ik j
Now note that v1i1 · · · vk−1 ṽk , v1j1 · · · vk−1
k−1 jk
ṽk is equal to
i j
v1i1 · · · vk−1
k−1
, v1j1 · · · vk−1
k−1
· δik ,jk · ik ! · ṽk , ṽk ik
.
It follows that
Mn,ṽ = diag · · · , Mn−m,vk−1 · m! · ṽk , ṽk m
, ···
n−m+k−2
where m runs from 0 to n. The matrix Mn−m,vk−1 has k−2 rows. So
n "
#
m·(n−m+k−2 )
det Mn,ṽ = c1 (n, k) · det Mn−m,vk−1 · ṽk , ṽk k−2
m=0
9.4. UNIMODULARITY 187
for some constant c1 (n, k). Applying induction to det Mn−m,vk−1 yields
(9.39) det Mn,ṽ
n
(n−m+k−2 ) m·(n−m+k−2 )
= c(n, k) · det Mvk−1 k−1
· ṽk , ṽk k−2
m=0
n n
(n−m+k−2 ) m·(n−m+k−2 )
= c(n, k) · det Mvk−1 m=0 k−1
· ṽk , ṽk m=0 k−2
(n+k−1 )
= c(n, k) · det Mvk−1 · ṽk , ṽk k
for some constant c(n, k), where we have used the combinatorial identities:
n n
n−m+k−2 n−m+k−2 n+k−1
(9.40) = m· = .
m=0
k−1 m=0
k−2 k
Indeed, all the three terms in (9.40) compute the dimension of the space of degree-
(n − 1) homogeneous polynomials in (k + 1)-variables.
Combining (9.38), (9.39) and (9.37), we conclude that
(n+k−1 )
det Mn,v = c(n, k) · det Mvk−1 · ṽk , ṽk k
(n+k−1 )
= c(n, k) · det Mv k
.
Finally, we come to the general case. Assume that the bilinear form ·, · on V
is not identically zero (otherwise the lemma trivially holds). Set zij = vi , vj for
1 ≤ i ≤ j ≤ k. Then both sides of (9.35) are easily seen to be polynomials in the
variables zij , 1 ≤ i ≤ j ≤ k. Under the assumption (9.36), the proof above implies
that the (polynomial) identity (9.35) holds for a Zariski open subset (zij )1≤i≤j≤k
of Ck(k+1)/2 . Thus, (9.35) holds for an arbitrary (zij )1≤i≤j≤k .
for uIr , vJr ∈ S mr (V [r]). Denote by Mμ,v the matrix of the pairings of the induced
monomial basis for S μ V (see (9.34) for the monomial basis of S n (V )).
Lemma 9.17. For some constant c(μ, k) and some integer d(μ, k) ≥ 1, we have
d(μ,k)
det Mμ,v = c(μ, k) · det Mv .
Proof. Follows immediately from our definitions and Lemma 9.16.
188 9. INTEGRAL COHOMOLOGY OF HILBERT SCHEMES
This orthogonality together with (9.45) implies by standard linear algebra that
k
dim Hnj ,Cβj
det Mn,β = (det Mni ,βi ) 1≤j≤k,j=i .
n i=1
Recall that the transition matrix from the basis an,β to mn,β is B −1 . So we conclude
from Lemma 9.15 and (9.46) that
Mn,β = B −1 · diag(· · · , ±zλ , · · · ) · (B −1 )t
where λ runs over all partitions of n. Hence we have
(9.47) det Mn,β = ±(det B)−2 · zλ .
λn
By (9.12), Lemma 9.9, and (1.5), we see that the transition matrix from the
basis {pλ | λ n} of ΛnC ⊂ ΛC to the basis {mλ | λ n} is B −1 as well. Introduce
the standard bilinear form ·, · on ΛnC by letting
(9.48) pλ , pμ = δλ,μ · zλ .
The matrix Mn formed by the pairings of the monomial symmetric functions
mλ , λ n is unimodular, and thus det Mn = ±1. By Lemma 9.15 and (9.48),
Mn = B −1 · diag(· · · , zλ , · · · ) · (B −1 )t
where λ runs over all partitions of n. It follows immediately that
(9.49) (det B)−2 · zλ = det Mn = ±1.
λn
For μ n, let Hμ,α ⊂ Hn,V be the span of a−μ1 (α1 ) · · · a−μk (αk )|0 where
μ , . . . , μk are partitions such that μ1 ∪ · · · ∪ μk = μ. Note that
1
Proof. By (3.9), a−μ1 (∗) · · · a−μk (∗) and a−ν 1 (∗) · · · a−ν (∗)|0 are orthogonal
unless μ1 ∪ · · · ∪ μk = ν 1 ∪ · · · ∪ ν . Here, ∗’s denote unspecified classes in H 2 (X).
Thus, we have an orthogonal direct sum
Hn,V = Hμ,α .
μn
By Lemma 9.20, the intersection matrix Mn,α of the basis an,α for Hn,V is given
by the diagonal block matrix whose diagonal consists of Mμ,α , μ n. So
(9.50) n,α =
det M det Mμ,α .
μn
Similarly, by repeating the above with the α’s replaced by the β’s, we obtain
(9.51) n,β =
det M det Mμ,β .
μn
n,α = A M
By Lemma 9.15 and the definition of the matrix A, we have M n,β At .
Combining this with (9.52) yields (det A)2 = ±1.
1
k
(9.53) a−λ (1X )a−μ (x)mν 1 ,α1 · · · mν k ,αk |0 , |λ| + |μ| + |ν i | = n
zλ i=1
∗
are integral, and furthermore, they form an integral basis for H (X [n] ; Z)/Tor.
Proof. By Lemma 9.4, Lemma 9.7 and Theorem 9.14, we have three types of
integral operators:
1
a−λ (1X ), a−μ (x), mνi ,αi
zλ
where λ, μ, νi stand for partitions. Therefore, we obtain the integral classes (9.53).
Since H 1 (X) = H 3 (X) = 0, we see from (3.10) that the number of classes in (9.53)
9.6. COMPARISON OF TWO INTEGRAL BASES OF H ∗ ((P2 )[n] ; Z) 191
is equal to the rank of H ∗ (X [n] ; Z)/Tor. Now consider the intersection number of
two classes from (9.53):
1
a−λ (1X )a−μ (x)mν 1 ,α1 · · · mν k ,αk |0 ,
zλ
1
a (1X )a−μ̃ (x)mν̃ 1 ,α1 · · · mν̃ k ,αk |0 .
zλ̃ −λ̃
By the Heisenberg algebra commutation relation (3.9), the intersection number is
±δλ,μ̃ · δλ̃,μ · mν 1 ,α1 · · · mν k ,αk |0 , mν̃ 1 ,α1 · · · mν̃ k ,αk |0 .
By Theorem 9.22, the intersection matrix formed by the pairings of the integral
classes in (9.53) is unimodular. Therefore, the classes in (9.53) form an integral
basis of H ∗ (X [n] ; Z)/Tor.
Remark 9.24. Under the same assumptions as in Theorem 9.23, one can fur-
ther prove that the integral ring H ∗ (X [n] ; Z)/Tor is generated over Z by all the
integral classes:
[n]
ci OX , 1−(n−j) m(1j ),αs |0 , 1−(n−j) a−j (x)|0
where 1 ≤ i ≤ n − 1, 1 ≤ j ≤ n, and 1 ≤ s ≤ k. We refer to [LQ3, Section 4] for
the details.
Remark 9.25. When X is a K3 surface, the ring generators for the integral
cohomology ring of the Hilbert scheme X [n] are also obtained by Markman [Mar2].
where |λ| + |μ| + |ν| = n. It follows that the integral cohomology H ∗ ((P2 )[n] ; Z) is
a Z-module freely generated by the fundamental classes of the closures:
(9.60) [C λ,μ,ν ], |λ| + |μ| + |ν| = n.
Moreover, it is known that
(9.61) dim(C0,μ,0 ) = |μ|;
if (w2 − w0 ) (w2 − w1 ), then
(9.62) dim(C0,0,ν ) = |ν| + (ν);
if (w2 − w0 ) (w1 − w0 ), then
(9.63) dim(Cλ,0,0 ) = |λ| − (λ).
Since the integral cohomology H ∗ ((P2 )[n] ; Z) is torsion free, we conclude from
Theorem 9.23 that another integral basis of H ∗ ((P2 )[n] ; Z) consists of the classes
1
(9.64) a−λ (1)mμ,H a−ν (x)|0 , |λ| + |μ| + |ν| = n
zλ
where 1 denotes 1P2 and x ∈ H 4 (P2 ; Z) is the class corresponding to a point in P2 .
We will study the transition matrix between the two integral bases (9.60) and
(9.64) of H ∗ ((P2 )[n] ; Z). To begin with, we will express the three special types of
classes [C 0,μ,0 ], [C 0,0,ν ] and [C λ,0,0 ] in terms of the basis (9.64) . For the cohomology
classes [C 0,μ,0 ], we put
T0 T2
u= , v= .
T1 T1
Then u, v form a system of affine coordinates at P1 .
9.6. COMPARISON OF TWO INTEGRAL BASES OF H ∗ ((P2 )[n] ; Z) 193
Note from (9.56) that φ(t)u = tw0 −w1 u with w1 − w0 > 0, and φ(t)v = tw2 −w1 v
with w2 − w1 > 0. Choosing (w2 − w1 ) (w1 − w0 ), we see from (9.68) that
uj v μj+1 ∈ lim φ(t)I
t→0
for with 0 ≤ j < k. Similar, we have u ∈ limt→0 φ(t)I. Since μ1 +. . .+μk = |μ| = n
k
and the flat limit limt→0 φ(t)I is of colength-n in C[u, v], we must have
lim φ(t)I = v μ1 , uv μ2 , . . . , uk−1 v μk , uk .
t→0
where I denotes the ideal in C[u, v] defining the element (9.69). Note that
I
ν1
νk
= u − a1 , v − b1,i · · · u − ak , v − bk,i
i=1 i=1
= u − a1 , (v − b1,1 ) · · · (v − b1,ν1 ) · · · u − ak , (v − bk,1 ) · · · (v − bk,νk ) .
Since (u − a1 ) · · · (u − ak ) ∈ I, we get uk ∈ limt→0 φ(t)I. Next, fix j with 0 ≤ j ≤
(k − 1). Then the ideal I contains the (k − j) polynomials:
j
νs
(u − ai ) · (v − bs,i ) · (u − ai )
i=1 i=1 j+1≤i≤k, i=s
is equal to 1. It follows from (9.71) that there exists a polynomial f (u, v) with
j
(9.72) (u − ai ) · (v νj+1 + f (u, v)) ∈ I
i=1
9.6. COMPARISON OF TWO INTEGRAL BASES OF H ∗ ((P2 )[n] ; Z) 195
and with the degree of f (u, v) in v being less than νj+1 . Since (w2 −w0 ) (w2 −w1 ),
we conclude that uj v νj+1 ∈ limt→0 φ(t)I for 0 ≤ j ≤ (k − 1). In summary,
v ν1 , uv ν2 , . . . , uk−1 v νk , uk ⊂ lim φ(t)I.
t→0
ν1 ν2 k−1 νk k
Since both the ideals v , uv , . . . , u v , u and limt→0 φ(t)I have co-length
n in the polynomial ring C[u, v], we see that (9.70) holds.
The next lemma studies the intersection between the integral classes a−λ (x)|0
and [C 0,0,ν ] where λ, ν n.
Lemma 9.28. (i) Let λ, ν n. If the intersection pairing between the cohomol-
ogy classes a−λ (x)|0 and [C 0,0,ν ] is nonzero, then λ = ν;
(ii) The intersection pairing between a−ν (x)|0 and [C 0,0,ν ] is positive.
Proof. (i) Let λ = (λ1 ≥ . . . ≥ λ ) and ν = (ν1 ≥ . . . ≥ νk ). Let
a1 , b1 , . . . , a , b ∈ C be different numbers. Put
x1 = (a1 , b1 ), . . . , x = (a , b ) ∈ C2 .
Then, the points x1 , . . . , x are distinct. Now a−λ (x)|0 is represented by the sub-
variety
Mλ1 (x1 ) × · · · × Mλ (x )
whose dimension is equal to (n − ). Since dim(C0,0,ν ) = n + k, we must have
= k. Since the pairing between a−λ (x)|0 and [C 0,0,ν ] is nonzero, the intersection
between Mλ1 (x1 ) × · · · × Mλ (x ) and C 0,0,ν is nonempty. Let ξ be a point in
the intersection. Then in view of (9.69), ξ = ξ1 + . . . + ξk where for each i with
1 ≤ i ≤ k,
ξi ∈ Mi (xi ) and λi = (ξi ) = νji
for some ji such that {j1 , . . . , jk } = {1, . . . , k}. Therefore, λ = ν.
(ii) Let notations be the same as in the previous paragraph. Then the only
point in the intersection between Mν1 (x1 ) × · · · × Mνk (xk ) and C 0,0,ν is
ξ = ξ1 + . . . + ξk
where each element ξi ∈ Mνi (xi ) is defined by the ideal u − ai , (v − bi )νi in the
polynomial ring C[u, v]. In particular, a−ν (x)|0 , [C 0,0,ν ] > 0.
So the divisor [C 0,0,ν ] is an integral linear combination of these two classes. Using
the proof of Proposition 9.30 (ii), we see that f˜ν = 1. So it remains to show that
(9.75) [C 0,0,ν ], a−1 (H)a−1 (x)n−1 |0 = (n − 1).
Fix a line H in P2 such that H , viewed as a line in C2 , is not parallel to the
v-axis. Choose distinct numbers a1 , b1 , . . . , an−1 , bn−1 ∈ C such that none of the
(n − 1) points xi = (ai , bi ), 1 ≤ i ≤ (n − 1) in C2 is contained in the line H . Then
the cohomology class a−1 (H)a−1 (x)n−1 |0 is represented by the curve:
H + x1 + . . . + xn−1 ⊂ (P2 )[n] .
For 1 ≤ i ≤ (n − 1), let Hi be the line in P2 defined by u = ai when Hi is viewed
as a line in C2 . Using the description (9.69) of generic elements in C 0,0,ν , we see
that the intersection C 0,0,ν ∩ (H + x1 + . . . + xn−1 ) consists of (n − 1) elements:
(H ∩ Hi ) + x1 + . . . + xn−1 , 1 ≤ i ≤ (n − 1).
Moreover, the intersection is transversal at these points. Therefore, the pairing
between the classes [C 0,0,ν ] and a−1 (H)a−1 (x)n−1 |0 is equal to (n − 1).
9.6. COMPARISON OF TWO INTEGRAL BASES OF H ∗ ((P2 )[n] ; Z) 197
where ∈ Z. Recall from Definition 1.1 the total ordering ≺ on the set of all
eλν
partitions.
Proposition 9.32. For every partition λ, there exist integers eλν ∈ Z for par-
titions ν with |λ| = |ν|, (λ) = (ν), λ ≺ ν such that
(9.77) [C λ,0,0 ] = eλλ a−λ (x)|0 + eλν a−ν (x)|0 .
|λ|=|ν|, (λ)=(ν)
λ≺ν
for some λ , μ , ν .
The cell Cλ,μ,ν ⊂ (P2 )[n] consists of elements of the form
(9.80) ξ1 + ξ2 + ξ3
with ξ1 ∈ Cλ,0,0 , ξ2 ∈ C0,μ,0 and ξ3 ∈ C0,0,ν . In particular, Supp(ξ1 ) = F0 = {P0 },
Supp(ξ2 ) ⊂ F1 = H − {P0 }, and Supp(ξ2 ) ⊂ F2 = P2 − H.
Choose a partition μ̃ with |μ̃| = |μ |. Let H̃ ⊂ P2 be a line different from H.
Fix distinct points x1 , . . . , x(ν ) ∈ P2 not lying on the lines H and H̃. Let
Zλ ,μ̃,ν ⊂ (P2 )[n]
be the subset consisting of the elements of the form
(λ ) (ν )
(9.81) η1,i + η2 + η3,i
i=1 i=1
198 9. INTEGRAL COHOMOLOGY OF HILBERT SCHEMES
where η3,i ∈ Mνi (xi ), η2 ∈ Lμ̃ H̃, and η1,i ∈ Mλi (yi ) such that y1 , . . . , y(λ ) ∈ P2
are distinct and not in the subset {x1 , . . . , x(ν ) } ∪ H̃. Let Z λ ,μ̃,ν be the closure
of Zλ ,μ̃,ν . Let λ = (1m1 (λ ) 2m2 (λ ) · · · ). As in Proposition 3.16, we have
(λ )
1
(9.82) [Z λ ,μ̃,ν ] = imi (λ ) ·
a−λ (1)mμ̃,H̃ a−ν (x)|0 .
i=1
zλ
By (9.78) and (9.82), the pairing [C λ,μ,ν ], [Z λ ,μ̃,ν ] is equal to
(λ )
|−(ν ))
(9.83) imi (λ ) · (−1)(|λ |−(λ ))+(|ν
i=1
< =
· eλ,μ,ν
λ ,μ ,ν mμ ,H |0 , mμ̃,H̃ |0 .
μ
The next lemma determines the leading coefficient in the expression (9.78). The
main idea in its proof is to exploit the non-emptiness of the intersecton C λ,μ,ν ∩
Z λ ,μ̃,ν for a suitable μ̃, whenever eλ,μ,ν
λ ,μ ,ν = 0.
where eλ,μ,ν∈ Z, |λ | + |μ | + |ν | = |λ| + |μ| + |ν|, and either |ν | < |ν|, or ν = ν
λ ,μ ,ν
and |μ | < |μ|, or ν = ν, μ = μ, (λ ) = (λ) and λ ≺ λ .
Proof. Recall the expression (9.78) and other notations introduced in the
previous paragraphs. As in (9.79), assume eλ,μ,ν λ ,μ ,ν = 0 for some λ , μ , ν . By
Theorem 9.22, there exists a partition μ̃ such that |μ̃| = |μ | and
< =
λ,μ,ν
(9.84) eλ ,μ ,ν mμ ,H |0 , mμ̃,H̃ |0 = 0.
μ
By (9.83), [C λ,μ,ν ], [Z λ ,μ̃,ν ] = 0. So C λ,μ,ν ∩ Z λ ,μ̃,ν = ∅. Let
ξ ∈ C λ,μ,ν ∩ Z λ ,μ̃,ν .
Since {x1 , . . . , x(ν ) } ∩ H = ∅, we see from (9.80) and (9.81) that
ξ = ξ + ξ3
where Supp(ξ ) ⊂ H, Supp(ξ3 ) ∩ H = ∅, (ξ ) ≥ |λ| + |μ|, and |ν| ≥ (ξ3 ) ≥ |ν |. If
|ν| > |ν |, we are done. In the following, we assume |ν| = |ν |.
Since |ν| = |ν |, we obtain (ξ3 ) = |ν| = |ν |, and (ξ ) = |λ| + |μ|. Moreover,
(ν )
ξ3 ∈ C 0,0,ν ∩ Mνi (xi ).
i=1
9.6. COMPARISON OF TWO INTEGRAL BASES OF H ∗ ((P2 )[n] ; Z) 199
From the proof of Lemma 9.28 (i), we see that ν = ν . Also, by the proof of
(ν)
Lemma 9.28 (ii), C 0,0,ν and i=1 Mνi (xi ) intersect at a unique point which will be
denoted by ξ3 for simplicity. Now, ξ ∈ C λ,μ,ν ∩ Z λ ,μ̃,ν if and only if
ξ = ξ + ξ3
with ξ ∈ C λ,μ,0 ∩ Z λ ,μ̃,0 . This decomposition allows us to split (P2 )[n] locally into
the product (P2 )[|λ|+|μ|] × (P2 )[|ν|] , i.e., we obtain
(9.85) [C λ,μ,ν ], [Z λ ,μ̃,ν ] = [C λ,μ,0 ], [Z λ ,μ̃,0 ] · [C 0,0,ν ], a−ν (x)|0 .
By Proposition 9.30, [C 0,0,ν ], a−ν (x)|0 = f˜ν . Therefore,
(9.86) [C λ,μ,ν ], [Z λ ,μ̃,ν ] = f˜ν [C λ,μ,0 ], [Z λ ,μ̃,0 ] .
Note that (9.86) is independent of the nonvanishing condition (9.84), i.e., it holds
as long as |μ̃| = |μ |. Combining (9.86) and (9.83), we see that for every partition
μ̃ with |μ̃| = |μ |, the number f˜ν [C λ,μ,0 ], [Z λ ,μ̃,0 ] is equal to
(λ )
(9.87) imi (λ ) · (−1)(|λ |−(λ ))+(|ν|−(ν))
i=1
< =
· eλ,μ,ν
λ ,μ ,ν mμ ,H |0 , mμ̃,H̃ |0 .
μ
In the rest of the proof, we work with [C λ,μ,0 ], [Z λ ,μ̃,0 ] , i.e., both ν and ν
are empty partitions. Now (9.80) and (9.81) are simplified to:
(λ )
ξ1 + ξ2 , η1,i + η2
i=1
where ξ1 ∈ Cλ,0,0 , ξ2 ∈ C0,μ,0 , η2 ∈ Lμ̃ H̃, and η1,i ∈ Mλi (yi ) such that the points
y1 , . . . , y(λ ) ∈ P2 are distinct and not on the line H̃. Assuming (9.84), we see from
(9.87) that
[C λ,μ,0 ], [Z λ ,μ̃,0 ] = 0.
ξ = ξ1 + ξ2
where Supp(ξ1 ) = {P0 }, P0 ∈ Supp(ξ2 ), (ξ1 ) ≥ |λ|, and |μ| ≥ (ξ2 ) ≥ |μ̃|. Since
|μ̃| = |μ |, we get |μ| ≥ |μ |. If |μ| > |μ |, we are done.
Assume that |μ| = |μ |. Then, we have |μ| = |μ̃| = (ξ2 ) and |λ| = |λ | = (ξ1 ).
Moreover, we see that ξ ∈ C λ,μ,0 ∩ Z λ ,μ̃,0 if and only if
ξ = ξ1 + ξ2
where ξ1 ∈ C λ,0,0 ∩ Z λ ,0,0 and ξ2 ∈ C 0,μ,0 ∩ Lμ̃ H̃. As in (9.85), we obtain
(9.88) [C λ,μ,0 ], [Z λ ,μ̃,0 ] = [C λ,0,0 ], [Z λ ,0,0 ] · [C 0,μ,0 ], [Lμ̃ H̃]
200 9. INTEGRAL COHOMOLOGY OF HILBERT SCHEMES
for every μ̃ with |μ̃| = |μ|. Note that [C 0,μ,0 ] = mμ,H |0 by Proposition 9.26, and
[Lμ̃ H̃] = mμ̃,H̃ |0 = mμ̃,H |0 . Combining with (9.87) and (9.88) yields
(λ )
< =
mi (λ ) (|λ |−(λ ))+(|ν|−(ν))
i · (−1) · eλ,μ,ν
λ ,μ ,ν mμ ,H |0
, mμ̃,H̃ |0
i=1 μ
= f˜ν [C λ,0,0 ], [Z λ ,0,0 ] · mμ,H |0 , mμ̃,H |0 .
In view of (9.82), we conclude for all the partitions μ̃ with |μ̃| = |μ| that
< =
λ,μ,ν
(|λ |−(λ ))+(|ν|−(ν))
(9.89) (−1) · eλ ,μ ,ν mμ ,H |0 , mμ̃,H̃ |0
μ
F G
1
= f˜ν [C λ,0,0 ], a−λ (1)|0 · mμ,H |0 , mμ̃,H |0 .
zλ
Let L ⊂ H 2n ((P2 )[n] ; Z) be the Z-span of all the classes mμ̃,H |0 , μ̃ |μ|. Since
H, H = 1, we see from Theorem 9.22 that L, together with the bilinear form ·, · ,
is an integral lattice, and the classes mμ̃,H |0 , μ̃ |μ| form a basis of L. It follows
from (9.89) that
λ,μ,ν
(−1)(|λ |−(λ ))+(|ν|−(ν)) · eλ ,μ ,ν mμ ,H |0
μ
F G
1
= f˜ν [C λ,0,0 ], a−λ (1)|0 · mμ,H |0 .
zλ
Therefore, we must have μ = μ and
F G
1
(−1)(|λ |−(λ ))+(|ν|−(ν)) · eλ,μ,ν ˜
λ ,μ,ν = fν [C λ,0,0 ], a−λ (1)|0 .
zλ
In particular, the right-hand-side is nonzero. By Proposition 9.32, (λ) = (λ ) and
λ # λ . If λ ≺ λ , then we are done. If λ = λ , then
F G
(−1)(|λ|−(λ))+(|ν|−(ν)) · eλ,μ,ν = ˜ν [C λ,0,0 ], 1 a−λ (1)|0 .
f
λ,μ,ν
zλ
By Proposition 9.32 again, the right-hand-side is equal to (−1)|λ|−(λ) eλλ f˜ν . Hence
|ν|−(ν) λ ˜
we obtain eλ,μ,ν
λ,μ,ν = (−1) eλ fν .
This defines a total ordering for all 3-tuples (λ, μ, ν) of partitions. Combining
with Lemma 9.33 yields the main result in this section.
9.6. COMPARISON OF TWO INTEGRAL BASES OF H ∗ ((P2 )[n] ; Z) 201
where eλ,μ,ν
λ ,μ ,ν ∈ Z, |λ | + |μ | + |ν | = |λ| + |μ| + |ν|, and (λ , μ , ν ) ≺ (λ, μ, ν).
We refer to [Mur, Section 5] for the transition matrices between the two integral
bases (9.60) and (9.64) of H ∗ ((P2 )[n] ; Z) in the special cases of n = 2, 3.
conjecture 9.36. [C λ,0,0 ] = a−λ (x)|0 for every partition λ.
By Theorem 9.35, Conjecture 9.36 holds when λ = (1n ), (1n−2 2), or (n).
Corollary 9.37. Let C∗ act on the affine coordinates u and v of C2 by
t(u, v) = (tw1 −w2 u, tw0 −w2 v), t ∈ C∗ and w1 w0 .
For ν n, let Cν be the cell in (C2 )[n] corresponding to the fixed point ξν . Then,
1
(9.90) a−ν (1C2 )|0 = (−1)|ν|−(ν) [C ν ]
zν
in the integral cohomology H ∗ ((C2 )[n] ; Z) of the Hilbert scheme (C2 )[n] .
Proof. We see from (8.2) and (9.73) that in H ∗ ((C2 )[n] ; Z), we have
1
a−ν (1C2 )|0 = (−1)|ν|−(ν) ι∗n [C 0,0,ν ] .
zν
Now (9.90) follows from ι∗n [C 0,0,ν ] = [ι−1
n (C 0,0,ν )] = [C ν ].
To analyze the coefficients eλν (x), put n = |λ|. For ν n with (ν) = (λ), let
Z ν (X) ⊂ X [n] be the closure of the subset consisting of the elements of the form
(ν)
ξi
i=1
where ξi ∈ Mνi (xi ), and x1 , . . . , x(ν) ∈ X are distinct. Then,
(ν)
1
(9.93) [Z ν (X)] = imi · a−ν (1)|0
i=1
zν
where ν = (1m1 2m2 · · · ). Note that the pairing [M λ (x)], [Z ν (X)] depends only
on an analytic neighborhood of x, which can be chosen to be independent of X and
x. Setting X = P2 and x = P0 , we obtain from Theorem 9.35 that
(ν)
(9.94) [M λ (x)], [Z ν (X)] = (−1)n−(ν) imi · eλν
i=1
where eλλ = 1 and eλν = 0 for ν ≺ λ. On the other hand, by (9.92) and (9.93),
(ν)
[M λ (x)], [Z ν (X)] = (−1) n−(ν)
imi · eλν (x).
i=1
∗
The ring structure of Horb (X (n) )
10.1. Generalities
∗
In this section, we will review the orbifold cohomology Horb (M/G) of the global
orbifold M/G where M is a complex manifold of complex dimension d with a finite
group G action.
First of all, we introduce the space
&
M $ G = {(g, x) ∈ G × M | gx = x} = M g,
g∈G
203
∗
204 10. THE RING STRUCTURE OF Horb (X (n) )
where G∗ denotes the set of conjugacy classes of G and Z(g) = ZG (g) denotes the
centralizer of g in G.
For g ∈ G and x ∈ M g , write the eigenvalues of the action of g on the complex
tangent space Tx,M to be μk = e2πirk , where 0 ≤ rk < 1 and k = 1, . . . , d. The
degree shift number (or age in the terminology of Ito and Reid) is the rational
number
d
Fxg = rk .
k=1
It depends only on the connected component Z which contains x, so we can de-
note it by FZg . Then associated to a cohomology class in H r (Z), we assign the
corresponding element in H ∗ (M, G) (and thus in Horb ∗
(M/G)) the degree r + 2FZg .
∗
A ring structure on Horb (M/G) was introduced in [CR2, Section 4]. This was
subsequently clarified in [FG1] by introducing a ring structure on H ∗ (M, G) first
∗
and then passing to Horb (M/G) by restriction. We will use • to denote this product.
The ring structure on H ∗ (M, G) is degree-preserving, and has the property that
α • β lies in H ∗ (M gh ) for α ∈ H ∗ (M g ) and β ∈ H ∗ (M h ).
For 1 ∈ G,
H ∗ (M 1 /Z(1)) ∼
= H ∗ (M/G),
and thus we can regard α ∈ H ∗ (M/G) to be α ∈ Horb ∗
(M/G) by this isomorphism.
Also, given
a= ag g
g∈G
by sending α ∈ H ∗ (M h ), where h ∈ K, to
1
IndG
K (α) = adg(α).
|K|
g∈G
Note that IndGK (α) is clearly G-invariant. When restricted to the invariant part,
we obtain a degree-preserving linear map
∗ ∗
K : Horb (M/K) → Horb (M/G).
IndG
We often write the restriction and induction maps as ResK , Res and IndG , Ind,
when the groups involved are clear from the context. In particular, when M is a
∗
point, Horb (pt/G) reduces to the Grothendieck ring RC (G) of G, and we recover
the induction and restriction functors in the theory of finite groups.
+∞
∗
FX = Horb (X (n) ).
n=0
ωn : H ∗ (X) → Horb
∗
(X (n) )
as follows. Let σn be any permutation in the conjugacy class [n] ∈ (Sn )∗ which
consists of the n-cycles, and let τn : H ∗ (X) → H ∗ ((X n )σn ) be the obvious isomor-
phism. Given α ∈ H ∗ (X), we define
ωn (α) = τn (nα).
r+d(n−1)
Note that if α ∈ H r (X), then ωn (α) ∈ Horb (X (n) ). We also define
∗
chn : Horb (X (n) ) → H ∗ (X)
and the annihilation operator pn (α) ∈ End(FX ) given by the composition (k ≥ 0):
∗ Res ∗
Horb (X n+k /Sn+k ) −→ Horb (X n+k /(Sn × Sk ))
∼ C
= ∗ ∗
−→ Horb (X n /Sn ) Horb (X k /Sk )
C
H ∗ (X) ∗
ch
−→
n
Horb (X k /Sk )
(α,·) ∗
−→ Horb (X k /Sk ).
Theorem 10.1. The operators pn (α) ∈ End(FX ) (n ∈ Z, α ∈ H ∗ (X)) generate
a Heisenberg (super)algebra with commutation relations given by
[pm (α), pn (β)] = mδm,−n (α, β) · IdFX
where n, m ∈ Z, α, β ∈ H ∗ (X). Furthermore, FX is an irreducible representation
of the Heisenberg algebra with the vacuum vector |0 = 1 ∈ H ∗ (pt) ∼
= C.
This theorem can be proved in the same way as an analogous theorem formu-
lated by using the equivariant K-group KSn (X n ) ⊗ C. This analogous theorem
was established in [Seg] (we refer to [Wan1, Theorem 4] and its proof for the de-
tails). Note that there is a fundamental sign difference in the two commutators of
Theorems 3.8 and Theorems 10.1.
|y|
In particular, for a given y ∈ Horb (X (n−1) ), by the definition of p−1 (α) (where
α ∈ H |α| (X)) and the induction map, we can write that
1
p−1 (α)(y) = adg (α ⊗ y)
(n − 1)!
g∈Sn
(−1)|α|·|y|
= adg (y ⊗ α).
(n − 1)!
g∈Sn
∗
For 0 ≤ i < n, we introduce the following cohomology class in Horb (X (n) ):
1
(10.1) Pi (α, n) = · p−i−1 (α)p−1 (1X )n−i−1 |0 .
(n − i − 1)!
This is the analogue of the class Bi (α, n) defined in Definition 7.3. The analogue
of the class Gi (α, n) (in the setting of Hilbert schemes) will be studied in the next
section.
L denotes a line bundle over X. The classes Ok (α, n) play the role of the classes
Gk (α, n) for the Hilbert scheme X [n] which are defined in Definition 4.1.
Let
+∞
ch : C(Sn ) −→ ΛC = Λ ⊗Z C
n=0
be the Frobenius characteristic map from the direct sum of class algebras of the
symmetric group Sn to the ring ΛC of symmetric functions in infinitely many vari-
ables (with complex coefficients). Denote by ηn and εn the trivial and alternating
characters of Sn . Then ch sends ηn and εn to the n-th complete and elementary
10.3. THE COHOMOLOGY CLASSES ηn (γ) AND Ok (α, n) 207
When it is clear from the text, we may simply write ξj;n as ξj . Put
Ξn = {ξ1 , . . . , ξn }.
Then the k-th elementary symmetric function ek (Ξn ) of Ξn is equal to the sum of
all permutations in Sn having exactly (n − k) cycles. Therefore, we obtain
n
n
(10.2) εn = (−1)d(σ) σ = (−1)k ek (Ξn ) = (1 − ξi )
σ∈Sn k=0 i=1
Noting that εn (1) (respectively, εn (−1)) coincides with the alternating character εn
(respectively, the trivial character ηn ), we obtain two classical identities involving
ηn , εn , and pr by setting = ±1 in (10.3). See also (8.13).
In the rest of this chapter, we will assume that X is a closed complex manifold
of even complex dimension d. Given γ ∈ H ∗ (X), we denote
γ (i) = 1⊗i−1 ⊗ γ ⊗ 1⊗n−i ∈ H ∗ (X n ),
and regard it to be a cohomology class in H ∗ (X n , Sn ) associated to the identity
conjugacy class. We define
ξi (γ) = ξi + γ (i) ∈ H ∗ (X n , Sn ).
We sometimes write ξi (γ) as ξi;n (γ) to specify its dependence on n when necessary.
Definition 10.2. Given γ ∈ H ∗ (X), we define ηn (γ) (respectively, εn (γ)) to
∗
be the cohomology class in Horb (X (n) ) whose component associated to an element
σ in the conjugacy class of partition λ of n is given by
γ ⊗(λ) ∈ H ∗ ((X n )σ ) ∼
= H ⊗(λ)
(respectively, by (−1)d(λ) γ ⊗(λ) ). We further define an operator η(γ) (respectively,
∗
ε(γ)) in End(FX ) by letting it act on Horb (X (n) ) by the orbifold product with
ηn (γ) (respectively, εn (γ)) for every n.
This definition is motivated by its counterpart in terms of equivariant K-groups
[Seg, Wan1]. Parallel to the Proposition 4 in [Wan1], we have
⎛ ⎞
+∞ 1
ηn (γ)z n = exp ⎝ p−r (γ)z r ⎠ · |0 .
n=0
r
r≥1
∗
208 10. THE RING STRUCTURE OF Horb (X (n) )
Proposition 10.3. Given γ ∈ H ∗ (X), the orbifold cup product of the n ele-
ments ξi (γ) (i = 1, . . . , n) in H ∗ (X n , Sn ) lies in Horb
∗
(X (n) ), and furthermore the
following identity holds:
n
(10.4) ηn (γ) = ξi (γ) = ξ1 (γ) • ξ2 (γ) • . . . • ξn (γ).
i=1
Proof. It suffices to prove (10.4), since the first claim follows from (10.4) and
the fact that ηn (γ) is Sn -invariant.
A typical monomial on the right-hand side of (10.4) is of the form
(ξi1 · · · ξik ) • (γ (j1 ) · · · γ (jn−k ) )
where i1 < . . . < ik , j1 < . . . < jn−k , and
{i1 , . . . , ik , j1 , . . . , jn−k } = {1, . . . , n}.
Here we have used the observation that ξi1 • · · · • ξik is just the usual multiplication
of permutations ξi1 · · · ξik and γ (j1 ) • · · · • γ (jn−k ) is just the ordinary cup product
γ (j1 ) · · · γ (jn−k ) in H ∗ (X n ) ∼
= H ∗ (X)⊗n . Note that every cycle of each permutation
σ appearing in ξi1 · · · ξik has exactly one number which does not belong to i1 , . . . , ik ,
and in addition, (σ) = n − k and d(σ) = k. Using the definition of the orbifold cup
product, we see that the product σ•γ (j1 ) · · · γ (jn−k ) does not involve the obstruction
bundles (or the group defects are trivial in the terminology of Lehn-Sorger) and is
equal to
∼ H ∗ ((X n )σ ).
γ ⊗(σ) ∈ (H ∗ (X))⊗(σ) =
This proves (10.4).
Put
n
εn (γ, ) = (γ (i) − ξi ).
i=1
Then we have
⎛ ⎞
+∞ p−r (γ)
(10.5) εn (γ, )z n = exp ⎝ (−)r−1 zr ⎠ .
n=0
r
r≥1
∗
Regarding ξi = ξi (0) ∈ H (X , Sn ), we denote
n
ξi•k = ξi • . . . • ξi ∈ H ∗ (X n , Sn ),
/ 01 2
k times
and define
1
e−ξi = (−ξi )•k ∈ H ∗ (X n , Sn ).
k!
k≥0
Definition 10.4. For a homogeneous class α ∈ H |α| (X), we define the class
∗
Ok (α, n) ∈ Horb (X (n) ) to be
1
n
(−ξi )•k • α(i) ∈ Horb (X (n) ).
dk+|α|
(10.6) Ok (α, n) =
k! i=1
10.4. INTERACTIONS BETWEEN HEISENBERG ALGEBRA AND Ok (γ) 209
k≥0
1
p−1 (α)(y) = adg (y ⊗ α).
(n − 1)!
g∈Sn
ι : H ∗ (X n−1 , Sn−1 ) → H ∗ (X n , Sn )
1
Ok (γ, n) − ι(Ok (γ, n − 1)) = · (−ξn;n )•k • γ (n) .
k!
Thus, we obtain
1
(n − 1)! [Ok (γ), p−1 (α)](y) = · adg [(−ξn;n )•k • γ (n) • (y ⊗ α)]
k! g
1
= · adg [(−ξn;n )•k • (y ⊗ γα)].
k! g
Under the assumption that the formula (10.7) is true for k, we have
adg [(−ξn;n )•(k+1) • (y ⊗ γα)]
g
= adg [(O1 (1X , n) − ι(O1 (1X , n − 1))) • (−ξn;n )•k • (y ⊗ γα)]
g
= O1 (1X , n) • adg [(−ξn;n )•k • (y ⊗ γα)]
g
− adg [ι(O1 (1X , n − 1)) • (−ξn;n )•k • (y ⊗ γα)],
g
Proof. The proof of the second formula is similar to that of the first, so we
will prove the first one only. For simplicity of signs in the proof, we assume that
the cohomology class α is of even degree.
By definition,
n
ηn (γ) = (ξi;n + γ (i) ).
i=1
It follows that
∗
Given y ∈ Horb (X (n−1) ), we have
(n − 1)! [η(γ)·p−1 (α)(y) − p−1 (γα) · η(γ)(y)]
= ηn (γ) • adg (y ⊗ α) − adg [(ηn−1 (γ) • y) ⊗ γα]
g∈Sn g∈Sn
= adg [(ηn (γ) − ηn−1 (γ) ⊗ γ) • (y ⊗ α)]
g
= adg [ξn;n • ι(ηn−1 (γ)) • (y ⊗ α)]
g
= adg [ξn;n • ((ηn−1 (γ) • y) ⊗ α)]
g
= −O1 (1X , n) • adg [(ηn−1 (γ) • y) ⊗ α]
g
+ adg [(O1 (1X , n − 1) • ηn−1 (γ) • y) ⊗ α]
g
= (n − 1)! [−O1 (1X ) · p−1 (α) · η(γ)(y) + p−1 (α) · O1 (1X ) · η(γ)(y)]
= −(n − 1)! p−1 (α) · η(γ)(y).
This finishes the proof.
∗
10.5. The ring structure of Horb (X (n) )
Using Theorem 10.1, Theorem 10.6 and Theorem 10.8, we can follow the ap-
proaches in Chapter 3, Chapter 4 and Chapter 7 to study the ring structure of
∗
Horb (X (n) ). Note that the terms involving the canonical class KX of X in various
formulas for the Hilbert schemes X [n] will disappear because there is no KX -term
in Theorem 10.6. Also, the fact e2X = 0 was used in previous chapters (where X
is a surface and eX is the Euler class of X). Thus, when dealing with the ring
∗
Horb (X (n) ) in this section, we sometimes need to treat separately and carefully the
case when X is a point (i.e. when e2X = 0).
Recall the classes Pi (α, n) and Oi (α, n) from (10.1) and (10.6) respectively.
The next result is parallel to Theorem 7.5.
Theorem 10.10. Let X be a closed complex manifold of even dimension. Then,
∗
the ring Horb (X (n) ) is generated by the cohomology classes Oi (α, n) (respectively,
the cohomology classes Pi (α, n)) where 0 ≤ i < n and α runs over a fixed linear
basis of H ∗ (X).
Theorem 10.11. Let X be a closed complex manifold of even dimension. Then
∗
the ring Horb (X (n) ) is determined uniquely by the ring H ∗ (X).
We refer to Theorem 10.11, and Proposition 10.12 and Proposition 10.13 be-
∗
low as the universality of the ring Horb (X (n) ). Theorem 10.11 follows from the
more quantitative descriptions of the orbifold cup product of ring generators of
∗
Horb (X (n) ) in Proposition 10.12 and Proposition 10.13. It also follows from the
results in [FG1]. Let 1−k = 0 if k < 0, and
1
1−k =p−1 (1X )k
k!
if k ≥ 0. The following proposition is parallel to Lemma 4.13.
∗
10.5. THE RING STRUCTURE OF Horb (X (n) ) 213
m
i −ri
0 < ni,1 ≤ . . . ≤ ni,mi −ri , ni,j ≤ (kj + 1) for every i, and
j=1 j∈πi
⎛ ⎞
(π) i −ri
m
s
⎝mi − 2 + ni,j ⎠ = ki .
i=1 j=1 i=1
Moreover, all the coefficients in this linear combination are independent of the man-
ifold X, the cohomology classes α1 , . . . , αs , and the integer n.
For the case d = 0 (i.e., X is a point), we adopt the simplified notations pm
and Ok (n) for pm (1X ) and Ok (1X , n) respectively. We have the following analog of
Proposition 10.12.
Proposition 10.13. Let n, s ≥ 1, k1 , . . . , ks ≥ 0. Then, Ok1 (n) • · · · • Oks (n)
is a finite linear combination of expressions of the form
⎛ ⎞
(π) mi −2ri
1 (π) −2ri
⎝ p−ni,j ⎠ · |0
mi
− n− ni,j i=1 j=1
i=1 j=1
mi
−2ri
0 < ni,1 ≤ . . . ≤ ni,mi −2ri , ni,j ≤ (kj + 1) for every i, and
j=1 j∈πi
⎛ ⎞
(π) i −2ri
m
s
⎝mi − 2 + ni,j ⎠ = ki .
i=1 j=1 i=1
ki
fix n with n ≥ ni,j for all 1 ≤ i ≤ s. Then the orbifold cup product
j=1
⎛ ⎞
s
ki
⎝1 ki p−ni,j (αi,j ) · |0 ⎠
−(n− j=1 ni,j )
i=1 j=1
∗
in Horb (X (n) ) is equal to a finite linear combination of monomials of the form
N
1−(n− N
a=1 ma ) p−ma (γa ) · |0
a=1
where
N
s
ki
ma ≤ ni,j ,
a=1 i=1 j=1
and γ1 , . . . , γN depend only on eX , αi,j , 1 ≤ i ≤ s, 1 ≤ j ≤ ki . Moreover, the coeffi-
cients in this linear combination are independent of αi,j and n. These coefficients
are also independent of X provided d > 0.
Parallel to Section 7.2, we can use Theorem 10.14 to define a ring RX , called
∗
the stable ring, which completely encodes the ring structure of Horb (X (n) ) for each
∗
n. The stable ring RX depends only on the cohomology ring H (X). We refer to
[QW1, Subsection 4.4] for details.
where the λ’s are generalized partitions. Note that J0m (α) = pm (α). We define
WX to be the linear span of the identity operator IdFX and the operators
Jpm (α)
∗
in End(FX ), where p ≥ 0, m ∈ Z and α ∈ H (X).
The following theorem describes the operator Ok (α) in terms of the Heisenberg
generators explicitly. It is a counterpart of Theorem 4.7.
Theorem 10.15. Let k ≥ 0, and α ∈ H ∗ (X). Then, Ok (α) is equal to
⎛ ⎞
1 s(λ) − 2
(−1)k · ⎝ pλ (τ∗ α) + pλ (τ∗ (eX α))⎠ .
λ! 24λ!
(λ)=k+2,|λ|=0 (λ)=k,|λ|=0
Since Ok (α, n) = Ok (α)p−1 (1X )n |0 /n!, we conclude from Theorem 10.15 that
Ok (α, n) is equal to
⎛
⎜ 1
(10.8) (−1)k · ⎝ · 1−(n−j−1) p−λ (τ∗ α)|0
0≤j≤k,λ(j+1)
λ! · |λ|!
(λ)=k−j+1
⎞
1 |λ| + s(λ) − 2 ⎟
+ · · 1−(n−j−1) p−λ (τ∗ (eX α))|0 ⎠ .
λ! · |λ|! 24
0≤j≤k λ(j+1)
(λ)=k−j−1
1 1
(10.9) : pp+1 :m (τ∗ α) + p(m2 − 3m − 2p) : pp−1 :m (τ∗ (eX α))
(p + 1) 24
p(p − 1)
+ : (∂ 2 p) pp−2 :m (τ∗ (eX α)).
24
1 p
Jpm (α) = : pp+1 :m (τ∗ α) + (∂ 2 : pp−1 :)m (τ∗ (eX α))
(p + 1) 24
(p + 1)p
+ (∂ : pp−1 :)m (τ∗ (eX α))
12
p(p2 − p − 2) p−1
+ :p :m (τ∗ (eX α))
24
p(p − 1)
+ : (∂ 2 p)pp−2 :m (τ∗ (eX α)).
24
(−1)p p
[Op (α), pn (β)] = −n · · Jn (αβ).
p!
Ωp,q
[
Jpm (α),
Jqn (β)] = (qm − pn) · p+q−1 m,n p+q−3
Jm+n (αβ) + · Jm+n (eX αβ)
12
∗
216 10. THE RING STRUCTURE OF Horb (X (n) )
where (p, q) ∈ Z2+ except for the unordered pairs (0, 0), (1, 0), (2, 0) and (1, 1). In
addition, for these four exceptional cases, we have
,
J0m (α),
[ J0n (β)] = mδm,−n (αβ) · IdFX ,
X
J1m (α),
[ J0n (β)] = −n ·
J0m+n (αβ),
,
m3 − m
[Jm (α), Jn (β)] = −2n · Jm+n (αβ) +
2 0 1
δm,−n (eX αβ) · IdFX ,
6 X
,
m3 − m
[Jm (α), Jn (β)] = (m − n) · Jm+n (αβ) +
1 1 1
δm,−n (eX αβ) · IdFX .
12 X
There exist fundamental and deep connections between the geometry of the
Hilbert schemes X [n] of points on a (quasi-)projective surface X and combinatorics
of symmetric functions. The monomial symmetric functions can be realized as cer-
tain ordinary cohomology classes of the Hilbert schemes associated to an embedded
curve in a surface ((9.12) and [Nak5, (9.13)]). Nakajima [Nak7] further showed
that the Jack polynomials whose Jack parameter is a positive integer γ can be re-
alized as certain T-equivariant cohomology classes of the Hilbert schemes of points
on the surface X(γ) which is the total space of the line bundle OP1 (−γ) over the
complex projective line P1 . Here T stands for the one-dimensional complex torus,
and the Jack parameter is interpreted as minus the self-intersection number of the
zero-section in X(γ). With very different motivations, Haiman [Hai1, Hai2, Hai3]
developed connections between the Macdonald polynomials and the geometry of
Hilbert schemes, and in particular realized the Macdonald polynomials as certain
T-equivariant K-homology classes of the Hilbert schemes of points on the affine
plane C2 . Moreover, the equivariant cohomology of Hilbert schemes are related to
Toda hierarchies, Hurwitz theory, equivariant (local) Gromov-Witten theory, and
equivariant (local) Donaldson-Thomas theory. These connections will be covered
in Chapters 12 and 14.
In this chapter, we will deal with the equivariant cohomology of the Hilbert
schemes X [n] where X is either the complex affine plane C2 or the total space of
a line bundle over P1 , and is equipped with a torus action. As in Section 3.2,
Heisenberg algebra actions on the equivariant cohomology of the Hilbert schemes
X [n] will be constructed. These actions enable the interpretation of the equivariant
cohomology of X [n] in terms of the ring of symmetric functions. In addition, the
equivariant cohomology classes corresponding to the Jack symmetric functions will
be identified. This chapter follows the presentation of [LQW7] which generalizes
the results in [Nak6, Vas]. We refer to [Nak7] for an excellent expository on
further topics such as geometric proofs of the norm formula and Pieri formula
of Jack symmetric functions, and a representation of the Virasoro algebra on the
equivariant cohomology of the Hilbert schemes (C2 )[n] (which is a special case of
Lehn’s result, Theorem 3.24).
bundle over P1 . The T-actions on these surfaces are specified in the following two
examples.
Example 11.2. Fix two nonzero integers α and β with the same signs. Let
u, v be the standard coordinate functions on C2 . We define the action of T on C2
by
(11.1) s · (u, v) = (sα u, s−β v), s ∈ T.
The origin of C2 is the only fixed point, which will be denoted by x. Let Σ and
Σ be the u-axis and v-axis respectively in C2 . As T-modules, we have Tx,Σ = θ −α
and Tx,Σ = θ β . By the localization theorem, we get
(11.2) [Σ] = −α−1 t−1 [x], [Σ ] = β −1 t−1 [x].
Example 11.3. Fix an integer γ > 1. Let X(γ) be the total space of the line
bundle OP1 (−γ) over P1 . The quasi-projective surface X(γ) can be regarded as the
quotient space of C × (C2 − {0}) by the C∗ -action defined by
(11.3) s · (b, b1 , b2 ) = (s−γ b, sb1 , sb2 ), s ∈ C∗ .
We use [(b, b1 , b2 )] to denote the equivalence class. Define a T-action on X(γ) by
(11.4) s · [(b, b1 , b2 )] = [(sb, s−1 b1 , b2 )], s ∈ T.
For i = 1, 2, let Xi be the open subset of X(γ) given by
(11.5) X1 = {[(b, b1 , b2 )] | b2 = 1}, X2 = {[(b, b1 , b2 )] | b1 = 1}.
Then X1 and X2 form an affine open cover of X(γ). Moreover, each Xi is T-
invariant. For simplicity, denote the point [(b, b1 , 1)] ∈ X1 by (b, b1 ). Similarly,
denote [(b, 1, b2 )] ∈ X2 by (b, b2 ). Then T acts on the points of X1 by
s · (b, b1 ) = (sb, s−1 b1 ),
i.e., T acts on the coordinate functions u1 and v1 of X1 by
(11.6) s · (u1 , v1 ) = (s−1 u1 , sv1 ), s ∈ T.
Similarly, T acts on the coordinate functions u2 and v2 of X2 by
(11.7) s · (u2 , v2 ) = (sγ−1 u2 , s−1 v2 ), s ∈ T.
Let xi be the origin of Xi . Then X(γ)T = {x1 , x2 }. Let
ρ : X(γ) → P1
be the projection sending [(b, b1 , b2 )] to [b1 , b2 ]. Let Σ0 ∼
= P1 be the zero section of
ρ, and
(11.8) Σ1 = ρ−1 ([0, 1]), Σ2 = ρ−1 ([1, 0]).
Then as T-modules, we have
Tx1 ,Σ1 = θ, Tx1 ,Σ0 = θ −1 , Tx2 ,Σ0 = θ, Tx2 ,Σ2 = θ 1−γ .
By the localization theorem, we get
(11.9) [Σ1 ] = t−1 [x1 ], [Σ0 ] = −t−1 [x1 ] + t−1 [x2 ], [Σ2 ] = (1 − γ)−1 t−1 [x2 ].
222 11. EQUIVARIANT COHOMOLOGY OF HILBERT SCHEMES
Remark 11.4. (i) Let X(1) be the total space of OP1 (−1). Using [(b, b1 , b2 )] ∈
X(1) to denote the equivalence class defined by (11.3), we define a T-action on
X(1) by
s · [(b, b1 , b2 )] = [(sb, s−2 b1 , b2 )], s ∈ T.
Then the methods and results below apply to X(1) as well.
(ii) Other T-actions on the surfaces X(γ), γ ≥ 1, with isolated fixed points can
be treated similarly.
In the rest of this chapter, let X be a surface in Example 11.2 or Exam-
ple 11.3. Next we will define some distinguished equivariant cohomology classes
for the Hilbert scheme X [n] . The T-action on X induces a T-action on X [n] . The
support of a T-fixed point in X [n] is contained in X T . As in Section 9.6, the T-fixed
points of X [n] are isolated and parametrized in terms of (multi-)partitions.
First of all, let X = C2 as in Example 11.2. The T-fixed points of X [n] are
supported in X T = {x} and indexed by partitions λ of n. We use ξλ to denote
the fixed point in (X [n] )T corresponding to a partition λ of n, and use Tξλ ,X [n] to
denote the T-equivariant tangent space of X [n] at the fixed point ξλ . As in (9.58),
we adopt the convention that ξλ is defined by the ideal
(11.10) Iξλ = v λ1 , uv λ2 , . . . , uk−1 v λk , uk
when λ = (λ1 ≥ λ2 ≥ · · · ≥ λk ). Then by [Nak6, Lemma 6.2] (also [ES1, Got2]),
for each λ n, we have
" #
(11.11) Tξλ ,X [n] = θ α(()+1)+βa() ⊕ θ −α()−β(a()+1)
∈Dλ
Note that [ξλ ] ∈ HT4n (X [n] ). We define the following distinguished class:
(−1)n −n
(11.15) [λ] = t [ξλ ].
cλ (α, β)
Now let X = X(γ) as in Example 11.3. In view of (11.1) and (11.6), there is
a T-equivariant identification between X1 and the complex plane in Example 11.2
with α = β = −1. Similarly, there is a T-equivariant identification between X2
and the complex plane in Example 11.2 with α = γ − 1 and β = 1. Then the
T-fixed points of X [n] are of the form ξλ1 + ξλ2 where λ1 and λ2 are partitions with
|λ1 | + |λ2 | = n, and ξλ1 and ξλ2 are defined in the previous paragraph as we identify
X1 and X2 with C2 respectively. For simplicity, put
ξλ1 ,λ2 = ξλ1 + ξλ2 .
11.1. EQUIVARIANT COHOMOLOGY RINGS OF HILBERT SCHEMES 223
By (11.12),
eT (Tξλ1 ,λ2 ,X [n] ) = (−1)n cλ1 (−1, −1)cλ1 (−1, −1)cλ2 (γ − 1, 1)cλ2 (γ − 1, 1)t2n .
Also, as in (11.15), we introduce the distinguished class
(−1)n
(11.16) [λ1 , λ2 ] = t−n [ξλ1 ,λ2 ].
cλ1 (−1, −1)cλ2 (γ − 1, 1)
In the following, we study bilinear forms on the equivariant cohomology of the
Hilbert schemes. Define
(11.17) Hn = HT2n (X [n] ).
Since X [n] admits a cell decomposition, its odd Betti numbers are equal to zero. By
calculating the dimensions of the cells, we conclude that H k (X [n] ) = 0 for k > 2n.
Therefore, the spectral sequence associated with the fibration X [n] ×T ET → BT
degenerates at the E2 -term. We now have
HT2k (X [n] ) = tk−n ∪ HT2n (X [n] )
for k ≥ n, which shows that the classes defined in (11.15) and (11.16) are contained
in Hn . Furthermore, following [Vas, Section B], we define a product structure on
Hn by
(11.18) tn ∪ (A B) = A ∪ B ∈ HT4n (X [n] )
for A, B ∈ HT2n (X [n] ). We see that (Hn , ) is a ring.
Let
HT∗ (·) = HT∗ (·) ⊗C[t] C(t)
be the localization, and let
+∞
(11.19) HX = HT∗ (X [n] ) .
n=0
Let
ι : (X [n] )T → X [n]
be the inclusion map. By abusing notations, we also use ι! to denote the induced
Gysin map on the localized equivariant cohomology groups:
(11.20) ι! : HT∗ ((X [n] )T ) → HT∗ (X [n] ) ,
which is an isomorphism by the localization theorem.
Define a bilinear form ·, · : HT∗ (X [n] ) × HT∗ (X [n] ) → C(t) by putting
(11.21) A, B = (−1)n p! ι−1
! (A ∪ B)
where p is the projection (X [n] )T → pt. This induces a bilinear form ·, · on HX .
Remark 11.5. Passing from HT∗ (X [n] ) to Hn does not lead to loss of infor-
mation since we can recover the cup product and bilinear form on HT∗ (X [n] ) from
those on Hn . Thus, understanding the ring Hn is the same as understanding the
equivariant intersection theory on X [n] .
224 11. EQUIVARIANT COHOMOLOGY OF HILBERT SCHEMES
where Iη and Iξ are the sheaves of ideals corresponding to η and ξ respectively. Let
p1 and p2 be the projections of X [n+i] × X [n] to the two factors. As in [Vas, B.4],
we define a linear operator
p−i ([Y ]) ∈ End(HX )
by
" #
(11.26) p−i ([Y ])(A) = p1! p∗2 A ∪ [Yn,i ]
for A ∈ HT∗ (X [n] ) . Note that the restriction of p1 to Yn,i is proper. We define
pi ([Y ]) ∈ End(HX )
11.3. EQUIVARIANT COHOMOLOGY AND JACK POLYNOMIALS 225
to be the adjoint operator of p−i ([Y ]) with respect to the bilinear form ·, · on HX .
For A ∈ HT∗ (X [n] ) , we have
" #
(11.27) pi ([Y ])(A) = (−1)i p2! ι × Id)−1 ∗
! (p1 A ∪ [Yn−i,i ]
where cλ,μ ∈ C and “<” denotes the dominance partial ordering of partitions.
Recall from Section 1.2 the ring Λ of symmetric functions in infinitely many
variables, the space Λn of degree-n symmetric functions, the monomial symmetric
function mλ associated to a partition λ, the power sum symmetric function pλ
(α)
associated to λ, and the Jack symmetric function Pλ . Put
ΛC = Λ ⊗Z C,
ΛnC = Λn ⊗Z C,
1
(11.32) p̃λ = pλ .
zλ
The symmetric functions p̃λ form a linear basis of ΛC since {pλ }λ is a linear basis
of ΛC . So we can define a bilinear form ·, · on ΛC by
1
(11.33) p̃λ , p̃μ = δλ,μ (β/α)(λ) .
zλ
We introduce a ring structure ◦ on ΛnC defined by
(β/α) (β/α) (β/α)
Pλ Pμ P
(11.34) ◦ = δλ,μ λ .
cλ (α, β) cμ (α, β) cλ (α, β)
Theorem 11.7. There exists a linear isomorphism φ : HX → ΛC preserving
bilinear forms such that
φ(p̃−λ |0 ) = p̃λ ,
φ([Lλ Σ]) = mλ ,
(β/α)
φ([λ]) = Pλ .
Furthermore, the restriction φn of φ to Hn is an isomorphism of rings, i.e.,
φn : (Hn , ) ∼
= (ΛnC , ◦).
11.3. EQUIVARIANT COHOMOLOGY AND JACK POLYNOMIALS 227
where μ and aλ,μ are the same as defined in (1.5). We conclude from an induction,
(11.36) and (11.35) that
φ([Lλ Σ]) = mλ .
By (11.30) and (11.33), φ preserves the bilinear forms. So we see from (11.23)
that
c (α, β) c (1, β/α)
φ([λ]), φ([μ]) = δλ,μ λ = δλ,μ λ .
cλ (α, β) cλ (1, β/α)
By (11.31), we have
[λ] = [Lλ Σ] + dλ,μ [Lμ Σ].
μ<λ
It follows that
φ([λ]) = mλ + dλ,μ mμ .
μ<λ
11.3.2. The case of the surface X(γ). Let X = X(γ) be the surface from
Example 11.3. Recall that there is a T-equivariant identification between the affine
open subset X1 of X and the complex plane in Example 11.2 with α = β = −1;
similarly for X2 and the complex plane in Example 11.2 with α = γ − 1 and β = 1.
By Proposition 11.6, HX is the Fock space of the Heisenberg algebra generated by
pi ([Σ1 ]) and pi ([Σ2 ]) with i ∈ Z, where Σ1 and Σ2 are the two T-equivariant fibers
in X defined by (11.8). Note also that HX1 is the Fock space of the Heisenberg
algebra generated by pi ([Σ1 ]) for i ∈ Z, where Σ1 is considered as a T-equivariant
closed curve in the affine open subset X1 ⊂ X. Similarly, HX2 is the Fock space of
the Heisenberg algebra generated by pi ([Σ2 ]) for i ∈ Z. To avoid confusion, we use
pX
i
1
to denote the operators pi ([Σ1 ]) acting on HX1 , pX i
2
to denote the operators
pi ([Σ2 ]) acting on HX2 , and pi ([Σ1 ]) and pi ([Σ2 ]) for the Heisenberg operators acting
on HX . As in (11.29), for a partition λ = (1m1 2m2 . . .) and for j = 1 or 2, we define
1
p̃−λ ([Σj ]) = p−i ([Σj ])mi .
zλ
i≥1
228 11. EQUIVARIANT COHOMOLOGY OF HILBERT SCHEMES
−i ⊗ Id
Ψ ◦ pX 1
= p−i ([Σ1 ]) ◦ Ψ,
Ψ ◦ Id ⊗ pX
−i
2
= p−i ([Σ2 ]) ◦ Ψ.
Proof. By the symmetry between Σ1 and Σ2 , we need only to prove the first
identity. Also, since Ψ preserves the bilinear forms and pi is the adjoint operator
of p−i , it suffices to prove the first identity for i > 0.
Since the two fibers Σ1 and Σ2 do not intersect, we see that for partitions λ1
and λ2 with |λ1 | + |λ2 | = n, the T-equivariant closed subvariety
1 2
Lλ Σ1 × Lλ Σ2 ⊂ X [n]
" 1 #
|λ | |λ2 |
is the closure of ρ−1
n Sλ1 Σ1 × Sλ2 Σ2 in X [n] , where ρn : X [n] → X (n) denotes
the Hilbert-Chow morphism. We conclude that
" 1 2
# 1 2
Ψ [Lλ Σ1 ] ⊗ [Lλ Σ2 ] = [Lλ Σ1 × Lλ Σ2 ]
1 2
by writing [Lλ Σ1 × Lλ Σ2 ] in terms of [λ1 , λ2 ] and [μ1 , μ2 ] where μ1 < λ1 or
1 2
μ2 < λ2 , similar to (11.31). It implies that the classes [Lλ Σ1 ×Lλ Σ2 ] form a linear
1 2
basis of HX . By (11.36) and a similar computation for p−i ([Σ1 ])[Lλ Σ1 × Lλ Σ2 ],
we have
" #
λ1 λ2
Ψ pX −i ⊗ Id [L Σ1 ] ⊗ [L Σ2 ]
1
" #
λ1 λ2
= Ψ pX 1
−i [L Σ 1 ] ⊗ [L Σ 2 ]
1 2
= p−i ([Σ1 ])[Lλ Σ1 × Lλ Σ2 ]
1 2
= p−i ([Σ1 ]) Ψ [Lλ Σ1 ] ⊗ [Lλ Σ2 ] .
It follows that Ψ ◦ pX
−i ⊗ Id = p−i ([Σ1 ]) ◦ Ψ for i > 0.
1
Let ΛC,1 be the ring ΛC of symmetric functions with the bilinear form
1
p̃λ , p̃μ = δλ,μ ,
zλ
ΛC,2 be the same ring ΛC equipped with a different bilinear form
1 (λ)
p̃λ , p̃μ = δλ,μ 1/(γ − 1) ,
zλ
and ΛnC,i
i
be the space of degree-ni symmetric polynomials in ΛC,i . The tensor
product ΛC,1 ⊗C ΛC,2 has an induced bilinear form. Let
(ΛC,1 ⊗C ΛC,2 )n = ΛnC,1
1
⊗C ΛnC,2
2
.
n1 +n2 =n
11.3. EQUIVARIANT COHOMOLOGY AND JACK POLYNOMIALS 229
Hilbert/Gromov-Witten correspondence
It is evident from Part 2 and Part 3 that the Chern character operators intro-
duced in Definition 4.1 (ii) are very essential to the understanding of the cohomology
ring structure and intersection theory of the Hilbert scheme of points on a surface.
In this chapter, we will discuss their analogues in the equivariant setting. More
precisely, using the setups and results from Chapter 11, we will define equivariant
Chern character operators on the equivariant cohomology of the Hilbert schemes
(C2 )[n] of points on the complex affine plane C2 . These operators give rise to a mas-
ter operator H(z) acting on certain Fock space. In the study of the class algebras
of symmetric groups, a counterpart of this operator has also played a distinguished
role [LT, Wan4]. Via the standard boson-fermion correspondence, the operator
H(z) turns out to be related to another operator ε0 (z), which has played a key role
in the study of stationary Gromov-Witten theory of curves [OP1, OP2], by the
following formula:
1 1
(12.1) H(z) = ε0 (z) − I .
e − e−z/2
z/2 ez/2 − e−z/2
2 dimC (Y )
α ∈ H ∗ (Y, C) = H j (Y, C)
j=0
= α1 α3 , α2 , α4 0,β + α1 , α3 , α2 α4 0,β
+ α1 , α3 , Δa 0,β1 Δa , α2 , α4 0,β2
β1 +β2 =β, β1 ,β2 =0 a
where c() is the content of the cell in the Young diagram Dλ , and is defined in
Definition 1.4. Denote by
[n]
chk,T
the k-th T-equivariant Chern character of O[n] . In particular the zero-th Chern
[n]
character ch0,T equals the rank of the vector bundle O[n] , which is n. Then
[n] 1
chk,T |ξλ = (c()t)k .
k!
∈Dλ
[n] , 0 ≤ k < n,
Remark 12.2. The Theorem 5.10 in [Wan4] implies that chk
form a set of ring generators of Hn = HT2n (X [n] ).
Recall from (11.18) the ring product on Hn . We define an operator G (re-
spectively, Gk ) in End(HX ) by sending a ∈ Hn to
[n]
a
chk
k≥0
12.2. THE HILBERT/GROMOV-WITTEN CORRESPONDENCE 235
1 z(λi −i+1/2)
+∞
Gz ([λ]) = (e − ez(−i+1/2) ) · [λ]
ς(z) i=1
where we have used λi = 0 for i > (λ). In view of the identity
+∞
1
ez(−i+1/2) = ,
i=1
ς(z)
our lemma follows.
Next, denote by Om the T-equivariant line bundle over X = C2 associated to
the T-character tm , where m ∈ Z. Let
M(m, n)
be the moduli space which parameterizes all rank-1 subsheaves of Om such that
the quotients are supported at finitely many points of X and have length n. Given
I ∈ X [n] , then Om ⊗ I is an element in M(m, n). Define
H(m)
n = HT2n (M(m, n)).
As before, the study of the equivariant cohomology ring HT∗ (M(m, n)) leads to the
(m)
ring Hn whose product is denoted by . The natural identification
M(m, n) ∼
= X [n]
236 12. HILBERT/GROMOV-WITTEN CORRESPONDENCE
where the notation p̃−λ is introduced in (11.29). Moreover, define the q-trace Trq (f)
of an operator f ∈ End(HX ) by
(12.21) Trq (f) = zλ p̃−λ , f(p̃−λ ) q |λ| ,
λ
and let
FE (z1 , . . . , zn ; q)
be the N -point disconnected series (with an additional variable q inserted) of the
stationary Gromov-Witten invariants of an elliptic curve E. Since Gromov-Witten
invariants are deformation invariant, FE (z1 , . . . , zn ; q) is independent of the elliptic
curve E. We have
Gz = H(z).
Proof. Follows from Lemma 12.3, formula (12.25) and Lemma 12.4.
In light of the interpretation of H(z) in Theorem 12.5 and the role of ε0 (z)
in the study of the stationary Gromov-Witten invariants (12.20) and (12.22), the
identity (12.26) defines the Hilbert/Gromov-Witten correspondence.
238 12. HILBERT/GROMOV-WITTEN CORRESPONDENCE
Ifλ−
U ⊂ μ, or equivalently if λ ⊂ λU + μ, then μ − λ− U = (λU + μ) − λ, and we can
show by induction that
zμ
(12.33) pλi · p−μ1 . . . p−μs |0 = p−μa1 . . . p−μat |0
zλU +μ−λ
i∈r\U
where < =
•
Fλ,μ (zU ) = p̃−λ , ε0 (zi )p̃−μ .
i∈U n
where η(q) = q 1/24 (q; q)∞ is the Dedekind eta function. By the Jacobi triple prod-
uct identity, we obtain
Θ(z) = (ez/2 − e−z/2 )(qez ; q)∞ (qe−z ; q)∞ /(q; q)2∞ .
We further define
dk
Θ(k) (z) = Θ(z), k ≥ 0.
dz k
We agree that Θ(k) (z) = 0 for k < 0. Given a positive integer N , we denote
N = {1, 2, . . . , N }. Given a finite set U , we denote by SU the symmetric group
on U . In particular, SN = SN . Given U = {u1 , . . . , uk } ⊂ N with u1 < . . . < uk
12.4. τ -FUNCTIONS OF 2-TODA HIERARCHIES 241
Define the τ -function τ (x, t, s) to be the following generating function for the
equivariant intersection numbers λ, ch [n] . . . ch
[n] , μ n from (12.28):
k1 kN
< +∞ =
[n]
τ (x, t, s) = tλ sμ λ, exp
xk ch ,μ
k
n |λ|=|μ|=n k=0 n
Note that
Γ− (s) = tλ p̃−λ ,
n≥0 |λ|=n
Γ+ (t) = Γ− (t)† .
From the definition of Hk and Theorem 12.5, we see that the τ -function affords an
operator formulation:
< +∞ =
τ (x, t, s) = Γ+ (t) exp xk Hk Γ− (s) .
k=0
Next, we define the Chern character operators from the space M(m, n). We
have a universal exact sequence:
0 → Jm → π2∗ Om → Qm → 0
[n]
where π1 , π2 are the projections of M(m, n) × X to the two factors. Denote by Om
the T-equivariant rank-n vector bundle over M(m, n) given by the push-forward
π1∗ (Qm ), whose fiber over a point ξλ ∈ M(m, n) ∼ = X [n] is given by
C
(12.38) Om[n]
|ξλ = Om O[n] |ξλ .
Equivalently, if we define
(m) = G(m) + e − 1 Id.
mz
(12.39) G z z
ς(z)2
(m) (m) z k , then
and further write G z = mz −1 Id + k≥0 G k
we denote by [λ](m) the image of [λ]. By the identification of toric action (12.38),
the same proof as in Lemma 12.3 implies that
+∞
emz 1 emz
− 1
([λ] ) =
G(m) (m) z(λ
e i −i+1/2)
− + · [λ](m) .
z
ς(z) i=1 ς(z) ς(z)2
Theorem 12.12.
(i) The function τ (x, t, s, m) can be reformulated as:
< +∞ =
−m m
τ (x, t, s, m) = S Γ+ (t) exp xk Hk Γ− (s)S .
k=0
(ii) The function τ (x, t, s, m), m ∈ Z, satisfies the 2-Toda hierarchy of Ueno-
Takasaki [UT, Section 1]. The lowest equation among the hierarchy reads:
∂2 τ (t, s, x, m + 1) τ (t, s, x, m − 1)
(12.40) ln τ (t, s, x, m) = .
∂t1 ∂s1 τ (t, s, x, m)2
Proof. (i) follows from Lemma 12.11 and the definition of τ (x, t, s, m), while
I ∞.
(ii) is standard since the operator Hk lies in sl
thanks to the fact that G0 ([λ]) = |λ| · [λ]. Setting u = x0 + ln(t1 s1 ), we denote
the above generating function by τ (u, x1 ). A simple computation reduces the Toda
equation (12.40) to the following:
∂2 τ (u + x1 , x1 ) τ (u − x1 , x1 )
e−u 2
ln τ (u, x1 ) = .
∂u τ (u, x1 )2
It is interesting to observe that this τ function can also be interpreted [Ok1, Sub-
section 2.9] as generating functions of certain Hurwitz numbers.
Recall the notations I [n] and a−λ (α) from Definition 8.5 and (9.2) respectively.
For a partition λ = (1m1 2m2 · · · ) and a cohomology class α ∈ H ∗ (X), define
1 1
(12.41) ã−λ (α) = a−λ (α) = a−r (α)mr .
zλ zλ
r≥1
The following theorem relates the construct constant cλk,μ to the 1-point discon-
•
nected series Fλ,μ (z).
4
Theorem 12.13. Let 2 ≤ k ≤ n, μ n, and I = H i (X). Then,
i=1
is related to the cup product in HT∗ ((C2 )[n] ). Then, we will apply (12.30) which
is a consequence of Theorem 12.5 where an algebraic relation between the equi-
variant cohomology ring of (C2 )[n] and the Gromov-Witten theory of P1 have been
established.
To convert (12.43) to a cup product in H ∗ ((C2 )[n] ), define ã−μ (1C2 ) as in
(12.41). By Theorem 8.12, the quotient ring H ∗ (X [n] )/I [n] is isomorphic to the co-
homology ring H ∗ ((C2 )[n] ). Moreover, we see from (8.18), (8.19), (8.4) that via this
isomorphism, the cohomology classes ã−(1n−k k) (1X )|0 (mod I [n] ) and ã−μ (1X )|0
[n]
(mod I [n] ) correspond to the classes (−1)k−1 (k − 1)! chk−1 and ã−μ (1C2 )|0 in
H ∗ ((C2 )[n] ) respectively. Here, chk−1 stands for the (k − 1)-th component of the
[n]
Chern character of the rank-n tautological vector bundle (OC2 )[n] . Thus, proving
(12.43) is equivalent to showing that
[n]
(12.44) (−1)k−1 (k − 1)! chk−1 ∪ã−μ (1C2 )|0
= c̃λk,μ ã−λ (1C2 )|0 .
λn
(λ)=(μ)+1−k
Note that 1C2 = [C2 ] = −t−1 [{0} × C] in HT∗ ((C2 )[n] ). It follows that
(12.46) p̃−μ (1C2 ) = (−t−1 )(μ) p̃−μ ([{0} × C]) = (−t)−(μ) p̃−μ
where the last p̃−μ denotes p̃−μ ([{0} × C]) for simplicity. By (12.45),
[n]
(12.47) (−1)k−1 (k − 1)! chk−1 ∪ã−μ (1C2 )|0
" #
= (−1)k−1−(μ) (k − 1)! Ψ t−(μ) chk−1,T ∪p̃−μ |0 .
[n]
Recall from Section 11.2 that {p̃−λ |0 | λ n} is a linear basis of HT2n ((C2 )[n] ).
Put
(12.48) [n] p̃−μ |0 =
ch dλk,μ p̃−λ |0
k−1
λn
where all the coefficients dλk,μ are independent of t. So by (12.16) and (11.18),
[n]
[n]
chk−1,T ∪p̃−μ |0
= t−n+(k−1) ch
(12.49) k−1,T ∪ p̃−μ |0
[n]
= tk−1 ch k−1,T p̃−μ |0
= t k−1
dλk,μ p̃−λ |0 .
λn
Since Ψ(p̃−λ (1C2 )|0 ) = ã−λ (1C2 )|0 = 0 for each λ n, we obtain from (12.50)
that
[n]
(12.51) (−1)k−1 (k − 1)! chk−1 ∪ã−μ (1C2 )|0
= (k − 1)! dλk,μ ã−λ (1C2 )|0 .
λn
(λ)=(μ)+1−k
By (12.51), to finish the proof of (12.44), it remains to show that the structure
constant dλk,μ in (12.48) is the coefficient of z k−1 in the expansion of the series
zλ •
(12.52) F (z).
ς(z) λ,μ
By (12.29) and (12.30), we have
[n] p̃−μ |0 1 • δλ,μ
z i p̃−λ |0 , chi n = Fλ,μ (z) − .
i
ς(z) zλ ς(z)
Combining these observations with (12.48) and (11.30) with α = β = 1, we get
[n]
= zλ p̃−λ |0 , ch
dλk,μ k−1 p̃−μ |0 n
zλ • δλ,μ
= Coeffzk−1 Fλ,μ (z) − .
ς(z) zλ ς(z)
12.6. RELATION TO THE HURWITZ NUMBERS OF P1 247
4
Remark 12.14. Let n ≥ 2, μ n, and I = H i (X). Theorem 12.13 handles
i=1
the cup product a−(1n−k k) (1X )|0 ∪a−μ (1X )|0 (mod I [n] ) for the partition (1n−k k)
of n. By Lemma 8.11, the quotient ring H ∗ (X [n] )/I [n] is generated by the (n − 1)
classes a−(1n−k k) (1X )|0 (mod I [n] ), 2 ≤ k ≤ n. It follows that the cup product
a−ν (1X )|0 ∪ a−μ (1X )|0 (mod I [n] )
for an arbitrary partition ν n can be obtained by repeatedly applying Theo-
rem 12.13. In other words, the structure constants of a−ν (1X )|0 ∪ a−μ (1X )|0
(mod I [n] ) should also be related to the Gromov-Witten theory of P1 .
which is valid for both connected and disconnected Hurwitz covers. In disconnected
theory, the domain genus may be negative. Two Hurwitz covers π : D → C and
π : D → C are defined to be isomorphic if there exists an isomorphism of curves
f : D → D satisfying π ◦ f = π. Up to isomorphism, there are only finitely many
Hurwitz covers of C of genus g, degree n and monodromy λi at ci . Each cover π
has a finite group of automorphisms Aut(π). The Hurwitz number
(12.54) HnC (λ1 , . . . , λk )
is defined to be the weighted count of the distinct, possibly disconnected Hurwitz
covers π with the prescribed data. Each such cover is weighted by 1/|Aut(π)|. We
refer to [GJV] for more references on (single) Hurwitz numbers and some interesting
development which relates double Hurwitz numbers to Gromov-Witten theory.
The enumeration of Hurwitz covers of the projective line P1 is classically known
to be equivalent to multiplication in the class algebra of the symmetric group. Fix
(k + 1) partitions λ0 , λ1 , . . . , λk of n. Hurwitz covers with profile λi over ci ∈ P1
canonically yield (k + 1)-tuples of permutations (s0 , . . . , sk ), in the n-th symmetric
group Sn , defined up to conjugation satisfying:
(i) si is of cycle type λi for each i;
(ii) s0 s1 . . . sk = 1.
The elements si are determined by the monodromies of π around the points ci .
This leads to the following well-known lemma (e.g. [OP1, Subsection 0.3.3]).
Lemma 12.15. The Hurwitz number HnP (λ0 , . . . , λk ) is equal to the number of
1
i=1
zλ0 λ ,...,λ
and (12.56) is equivalent to
c̃λλ1 ,...,λk = zλ0 · HnP (λ0 , λ1 , . . . , λk ).
0 1
(12.57)
Let Ψ1 , Ψ2 and Cλ be from (8.4), (8.11) and (8.9) respectively. By (8.5) and
(8.12), we obtain
Ψ1 Ψ2 (Cλ ) = ã−λ (1C2 )|0 .
0
By the assumption (12.55), we see from Theorem 8.3 that c̃λλ1 ,...,λk is equal to
the number of conjugacy class Cλ0 in the cup product of the conjugacy classes
Cλ1 , . . . , Cλk in the class algebra C(Sn ), denoted by
k
[Cλ0 ] C λi .
i=1
Since the conjugacy class Cλ0 consists of n!/zλ0 elements, we have
k
n! k
n! λ0
[C(1n ) ]Cλ0 C λi = · [Cλ0 ] C λi = · c̃ 1 k.
i=1
zλ 0 i=1
zλ0 λ ,...,λ
On the other hand, we conclude from Lemma 12.15 that
1 k
HnP (λ0 , λ1 , . . . , λk ) =
1
· [C(1n ) ]Cλ0 C λi .
n! i=1
Now (12.57) follows from putting the above together.
Part 5
The Heisenberg algebra of Grojnowski and Nakajima offers a basic tool for
writing down the cohomology classes of the Hilbert schemes of points on a sur-
face. The geometric representation (Proposition 3.16) of the Heisenberg mono-
mial classes provides a convenient way to analyze the Gromov-Witten invariants of
these Hilbert schemes. The study of the Gromov-Witten theory of the Hilbert
schemes of points on a surface began with the work [LQ1] where the 1-point
genus-0 extremal Gromov-Witten invariants are computed. The results in [LQ1]
give the first evidence to Ruan’s Cohomological Crepant Resolution Conjecture
[Rua2, Section 2]. In [ELQ], the genus-0 extremal Gromov-Witten invariants
of the Hilbert scheme of 3 points on the complex projective plane P2 are calcu-
lated. Motivated by the Gromov-Witten and Donaldson-Thomas correspondence
conjectured in [MNOP1, MNOP2], Okounkov and Pandharipande [OP3, OP4]
investigated the equivariant Gromov-Witten theory of the Hilbert schemes of points
on the affine plane C2 , and proved its equivalence with the local Gromov-Witten
theory of the 3-fold C2 × P1 . Using the technique of cosection localization and
the equivariant quantum corrected boundary operator of Okounkov and Pandhari-
pande [OP3], Jun Li and W.-P. Li [LL] determined the 2-point genus-0 extremal
Gromov-Witten invariants of the Hilbert scheme X [n] of points on an arbitrary
surface X.
In this chapter, we will address the technique of cosection localization developed
in [KL1,KL2], which provides a powerful tool for studying the virtual fundamental
cycles. We will apply it to the Gromov-Witten theory of the Hilbert schemes of
points on surfaces, and review the reduction results from [LL] when the images
of the stable maps to the Hilbert schemes are contracted by the Hilbert-Chow
morphism. When the smooth projective surface X has a non-trivial holomorphic 2-
form, we will prove some vanishing results regarding the Gromov-Witten invariants
of the Hilbert scheme X [n] . Moreover, we will apply the results to the Hilbert
scheme X [2] , and calculate its genus-1 Gromov-Witten invariant. Proposition 13.3
is from [LL], while Proposition 13.15, Theorem 13.19 and their proofs are from
[HLQ].
X (n) . We will present the reduction results from [LL] regarding the corresponding
virtual fundamental cycles.
First of all, we review the technique of cosection localization. Following the con-
cept of a localized Gysin map, a localized virtual fundamental cycle is constructed
by Kiem and J. Li [KL2, Theorem 1.1].
Theorem 13.1. Let M be a Deligne-Mumford stack endowed with a perfect
obstruction theory. Suppose the obstruction sheaf ObM admits a surjective homo-
morphism σ : ObM |U → OU over an open U ⊂ M. Let
M(σ) = M − U.
be the degeneracy locus. Then (M, σ) has a localized virtual fundamental cycle
loc ∈ A∗ M(σ) .
[M]vir
This cycle has the usual properties of the virtual fundamental cycles, and relates to
the usual virtual fundamental cycle [M]vir via
loc ∈ A∗ M
[M]vir = ι∗ [M]vir
where ι : M(σ) → M is the inclusion map.
To apply the technique of cosection localization to Gromov-Witten theory, let
Y be a smooth quasi-projective variety with a non-trivial holomorphic 2-form θ̃ ∈
H 0 (Y, Ω2Y ). Let M = Mg,r (Y, β) be the moduli space of r-pointed genus-g stable
maps to Y with homology class β. The 2-form θ̃ on Y induces a cosection σ of the
obstruction sheaf on M as follows. Let f : C → Y and π : C → M be the universal
family of M. Let S be the Artin stack of genus-g connected nodal curves. Then
the relative obstruction sheaf of the standard relative obstruction theory of M/S is
ObM/S = R1 π∗ f ∗ TY .
Now regard the holomorphic 2-form θ̃ ∈ H 0 (Y, Ω2Y ) as an anti-symmetric homo-
morphism
(13.1) θ̂ : TY −→ ΩY , θ̂(v), v = 0.
Then θ̂ defines the first arrow in the following sequence of homomorphisms
(13.2) R1 π∗ f ∗ TY −→ R1 π∗ f ∗ ΩY −→ R1 π∗ ΩC/M −→ R1 π∗ ωC/M
where the second is induced by f ∗ ΩY → ΩC/M , and the third is induced by the
tautological map ΩC/M → ωC/M . Since R1 π∗ ωC/M ∼= OM , the composite of this
sequence provides
(13.3) σ rel : R1 π∗ f ∗ TY = ObM/S −→ OM .
The obstruction sheaf of M is the cokernel of p∗ TS → ObM/S where p : M → S is
the projection. Using the universal family f and R1 π∗ f ∗ TY = Ext1π (f ∗ ΩY , OC ), we
obtain the exact sequence
(13.4) Ext1π (ΩC/M , OC ) −→ Ext1π (f ∗ ΩY , OC ) −→ ObM −→ 0
where the first arrow is induced by f ∗ ΩY → ΩC/M . It can be proved that the
composition
σ rel
Ext1π (ΩC/M , OC ) −→ Ext1π (f ∗ ΩY , OC ) −→ OM
13.1. COSECTION LOCALIZATION OF KIEM AND J. LI 255
Thus, θ [n] is the direct sum of the forms θ [ni ] , and ϕ∗ (TCreg ) is contained in the null
space of θ [n] if and only if ϕi∗ (TCreg ) is contained in the null space of θ [ni ] for each i.
According to the work of Beauville, the form θ [ni ] is non-degenerate along Mni (x)
if x ∈ θ −1 (0). Applying this to the support xi of ϕi , we see that ϕ ∈ Λθ .
Lemma 13.2 is sufficient if we have a regular section θ ∈ H 0 (X, OX (KX )). For
a general surface, we take a meromorphic section θ of OX (KX ). Let D0 and D∞ be
the vanishing and pole divisors of θ respectively. Viewed as a meromorphic section
of Ω2X , θ induces a meromorphic section θ [n] of Ω2X [n] and hence a meromorphic
homomorphism
σ : E → OM .
Here, E is a certain locally free sheaf admitting a surjection onto R1 π∗ f ∗ TX [n] ,
and f : C → X [n] and π : C → M are the universal family of M. Define the
degeneracy locus of σ to be the subset Deg(σ) ⊂ M consisting of all the points
ϕ ∈ M such that either σ is undefined or vanishes at ϕ. Adopting the proof of
13.2. VANISHING OF GROMOV-WITTEN INVARIANTS 257
Lemma 13.2, one sees that Deg(σ) is contained in the subset of M consisting of
all those ϕ = (ϕ1 , . . . , ϕ ) ∈ M such that either for some i the support of ϕi is
contained in D0 ∪ D∞ , or for each i the map ϕi is constant.
vir
The following reduction result states that the virtual fundamental cycle M
is supported in a much smaller subset than Deg(σ).
Proposition 13.3. Let Λθ ⊂ M = Mg,0 (X [n] , dβn ) be the subset consisting of
the points
(13.14) ϕ = (ϕ1 , . . . , ϕ ) ∈ M
such that for each i, either ϕi is a constant map or the support xi = Spt(ϕi ) lies
vir
in D0 ∪ D∞ . Then M is supported in Λθ .
The proof of Proposition 13.3 is based on the observation that when ϕ ∈ M is
decomposed into maps into smaller Hilbert schemes as in (13.13), the obstruction
sheaf can be decomposed into the direct sum of factors. We refer to [LL, Section 3]
for the details.
Proposition 13.3 deals with stable maps whose images are contracted by the
Hilbert-Chow morphisms, and only requires the presence of a meromorphic section
θ of OX (KX ). Therefore, it is very useful in studying extremal Gromov-Witten
invariants of the Hilbert scheme of points on an arbitrary surface, as we will see
in Chapter 15. However, to apply the technique of cosection localization to handle
stable maps whose images are not necessarily contracted by the Hilbert-Chow mor-
phisms, we will have to assume the existence of a non-trivial holomorphic section
of OX (KX ). This leads us to the next section.
is trivial over the regular locus Γreg of Γ. Moreover, the cosection σ is surjective
away from the degeneracy locus M(σ), and there exists a localized virtual cycle
loc ∈ A∗ (M(σ)) such that
[M]vir
(13.18) loc ∈ A∗ (M)
[M]vir = ι∗ [M]vir
where ι : M(σ) → M stands for the inclusion map.
Our next lemma is parallel to Lemma 13.2. On one hand, it applies to an
arbitrary β ∈ H2 (X [n] ). On the other hand, it provides less information than
Lemma 13.2 does.
Lemma 13.4. Let C0 be the zero divisor of θ. Let u : Γ → X [n] be a stable map
in the degeneracy locus M(σ), and let Γ0 be an irreducible component of Γ with
non-constant restriction u|Γ0 . Then there exists ξ1 ∈ X [n0 ] for some n0 such that
Supp(ξ1 ) ∩ C0 = ∅ and
(13.19) u(Γ0 ) ⊂ ξ1 + ξ2 | Supp(ξ2 ) ⊂ C0 .
Proof. For notational convenience, we assume that Γ = Γ0 is irreducible.
Then there exist a nonempty open subset U ⊂ Γ and an integer n0 ≥ 0 such that
U is smooth and for every element p ∈ U , the image u(p) is of the form
(13.20) u(p) = ξ1 (p) + ξ2 (p)
where ξ1 (p) ∈ X with Supp(ξ1 (p)) ∩ C0 = ∅ and Supp(ξ2 (p)) ⊂ C0 . This
[n0 ]
for some integers d1 , . . . , ds ≥ 0 and d . Now our lemma follows immediately from
(13.24).
Proof. By Lemma 13.5, the degeneracy locus M(σ) from (13.16) is empty. It
follows from (13.18) that
[Mg,r (X [n] , β)]vir = 0.
Therefore, all the Gromov-Witten invariants of X [n] defined via the moduli space
Mg,r (X [n] , β) vanish.
Remark 13.8. From the proof of Corollary 13.7, we see that if m is from
(13.25), then the rational number d0 in Corollary 13.7 may be required to be of the
form d0 /m for some integer d0 ≥ 0.
260 13. COSECTION LOCALIZATION FOR THE HILBERT SCHEMES
Recall from Theorem 1.24 (ii) that KX [n] = DKX . Thus, if β = d0 βKX − dβn
for some rational number d0 ≥ 0 and integer d, then the expected dimension of
Mg,r (X [n] , β) is
(13.26) d = −KX [n] · β + (dim X [n] − 3)(1 − g) + r
= −d0 KX
2
+ (2n − 3)(1 − g) + r.
Our first corollary deals with the case when X is an elliptic surface.
Corollary 13.9. Let X be a simply connected (minimal) elliptic surface with-
out multiple fibers and with positive geometric genus. Let n ≥ 2 and β = 0. Then all
the Gromov-Witten invariants without descendant insertions defined via the moduli
space Mg,r (X [n] , β) vanish, except possibly when 0 ≤ g ≤ 1 and β = d0 βKX − dβn
for some integer d and rational number d0 ≥ 0.
Proof. Since X is a simply connected elliptic surface without multiple fibers,
KX = (pg − 1)f where pg ≥ 1 is the geometric genus of X and f denotes a smooth
fiber of the elliptic fibration. By Corollary 13.7, it remains to consider the case
when β = d0 βKX − dβn for some integer d and rational number d0 ≥ 0. By (13.26)
2
and KX = 0, the expected dimension of the moduli space Mg,r (X [n] , β) is equal to
d = (2n − 3)(1 − g) + r.
By the Fundamental Class Axiom, all the Gromov-Witten invariants without de-
scendant insertions are equal to zero if g ≥ 2.
Our second corollary concentrates on the case when X is of general type.
Corollary 13.10. Let X be a simply connected minimal surface of general type
admitting a holomorphic 2-form with irreducible zero divisor. Let n ≥ 2 and β = 0.
Then all the Gromov-Witten invariants without descendant insertions defined via
Mg,r (X [n] , β) vanish, except possibly in the following cases
(i) g = 0 and β = dβn for some integer d > 0;
(ii) g = 1 and β = dβn for some integer d > 0;
(iii) g = 0 and β = d0 βKX − dβn for some integer d and rational number
d0 > 0.
Proof. In view of Corollary 13.7, it remains to consider the case when β =
d0 βKX − dβn for some integer d and rational number d0 ≥ 0.
When d0 = 0 and β = dβn with d > 0, we see from (13.26) that the expected
dimension of the moduli space Mg,r (X [n] , β) is equal to
d = (2n − 3)(1 − g) + r.
If g ≥ 2, then all the Gromov-Witten invariants without descendant insertions
defined via Mg,r (X [n] , β) vanish by the Fundamental Class Axiom.
Next, assume that d0 > 0. Since KX 2
≥ 1, we see from (13.26) that
d < (2n − 3)(1 − g) + r.
By the Fundamental Class Axiom, all the Gromov-Witten invariants without de-
scendant insertions vanish except possibly in the case when g = 0.
In Section 13.4 below, for the Hilbert scheme X [2] where X is a simply con-
nected minimal surface of general type admitting a non-trivial holomorphic 2-form
13.3. INTERSECTIONS ON MODULI SPACE OF GENUS-1 STABLE MAPS 261
with irreducible zero divisor, we will study the exceptional cases listed in Corol-
lary 13.10 (i), (ii) and (iii). The computation will involve certain integrals over
some moduli space of genus-1 stable maps. These integrals will be calculated in the
next section.
Note that R1 (f1,0 )∗ ev∗1 TS is a rank-2d bundle on the 2d-dimensional moduli space
M1,0 (S, d[C]) whose virtual dimension is 0. By Proposition 12.1,
,
c2d R1 (f1,0 )∗ ev∗1 OP1 (−2) = deg [M1,0 (S, d[C])]vir .
[M1,0 (P1 ,d[P1 ])]
The following lemma deals with the case when V is a rank-2 bundle over P1 .
The main idea in its proof is to reduce to the special cases
OP1 ⊕ OP1 , OP1 (2) ⊕ OP1 (−1)
depending on the parity of the degree of V .
Lemma 13.14. Let d ≥ 1, and V be a rank-2 bundle over P1 . Then,
,
deg(V )
(13.34) λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2) = .
[M1,0 (P(V ),df )] 12d
Proof. First of all, assume that deg(V ) = 2k for some integer k. Then V can
be deformed to
OP1 (k) ⊕ OP1 (k) = OP1 ⊕ OP1 ⊗ OP1 (k).
By Lemma 13.12,
,
λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2)
[M1,0 (P(V ),df )]
, " #
= λ · c2d R1 (f1,0 )∗ ev∗1 OP1 ×P1 (−2) ⊗ OP1 ×P1 (−2kf ) .
[M1,0 (P1 ×P1 ,df )]
Note that
M1,0 (P1 × P1 , df ) ∼
= P1 × M1,0 (P1 , d[P1 ]),
Thus, we obtain
,
λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2)
[M1,0 (P(V ),df )]
, " #
= π2∗ λ · c2d π2∗ R1 (f1,0 )∗ ev∗1 OP1 (−2) ⊗ π1∗ OP1 (−2k)
[P1 ×M1,0 (P1 ,d[P1 ])]
where π1 and π2 are the projection on P1 × M1,0 (P1 , d[P1 ]). Hence,
,
(13.35) λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2)
[M1,0 (P(V ),df )]
,
= π2∗ λ · π2∗ c2d−1 R1 (f1,0 )∗ ev∗1 OP1 (−2) · π1∗ c1 OP1 (−2k)
[P1 ×M1,0 (P1 ,d[P1 ])]
deg(V )
=
12d
where we have used formula (13.31) in the last step.
Next, assume that deg(V ) = 2k+1 for some integer k. Then V can be deformed
to OP1 (2) ⊕ OP1 (−1) ⊗ OP1 (k). As in the previous paragraph, we have
,
λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2)
[M1,0 (P(V ),df )]
, " #
= λ · c2d R1 (f1,0 )∗ ev∗1 OS (−2) ⊗ OS (−2kf )
[M1,0 (S,df )]
,
2k
= λ · c2d R1 (f1,0 )∗ ev∗1 OS (−2) +
[M1,0 (S,df )] 12d
264 13. COSECTION LOCALIZATION FOR THE HILBERT SCHEMES
where S = P OP1 (2) ⊕ OP1 (−1) . Let F1 be the blown-up of P2 at a point p, and
σ be the exceptional curve. Then, TF1 |σ ∼= OP1 (2) ⊕ OP1 (−1). So
,
(13.36) λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2)
[M1,0 (P(V ),df )]
,
2k
= λ · c2d R1 (f1,0 )∗ ev∗1 OP(TF1 |σ ) (−2) +
[M1,0 (P(TF1 |σ ),df )] 12d
,
2k
= φ∗ [σ] · λ · c2d R1 (f1,0 )∗ ev∗1 OP(TF1 ) (−2) +
[M1,0 (P(TF1 ),df )] 12d
where the morphism φ : M1,0 P(TF1 ), df → F1 is from (13.27). Let f0 be a fiber
of the ruling F1 → P1 , and C be a smooth conic in P2 such that p ∈ C. We use C
to denote its strict transform in F1 . Then, [σ] = [C]/2 − [f0 ]. By (13.36),
,
λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2)
[M (P(V ),df )]
,1,0
1
= · φ∗ [C] · λ · c2d R1 (f1,0 )∗ ev∗1 OP(TF1 ) (−2)
2 [M1,0 (P(TF1 ),df )]
,
2k
− φ∗ [f0 ] · λ · c2d R1 (f1,0 )∗ ev∗1 OP(TF1 ) (−2) +
[M1,0 (P(TF1 ),df )] 12d
,
1
= · λ · c2d R1 (f1,0 )∗ ev∗1 OP(TF1 |C ) (−2)
2 [M1,0 (P(TF1 |C ),df )]
,
2k
− λ · c2d R1 (f1,0 )∗ ev∗1 OP(TF1 |f0 ) (−2) + .
[M1,0 (P(TF1 |f0 ),df )] 12d
Note that deg TF1 |C = 6 and deg TF1 |f0 = 2. By (13.35),
,
λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2)
[M1,0 (P(V ),df )]
1 6 2 2k
= · − +
2 12d 12d 12d
deg(V )
= .
12d
This completes the proof of the lemma.
The following is the main result in this section. The ideas in its proof are
to deform the rank-2 vector bundle V over a curve C to a direct sum of two line
bundles over C, reduce the situation to a rank-2 bundle over P1 , and then apply
Lemma 13.14.
Proposition 13.15. Let d be a positive integer. Assume that V is a rank-2
vector bundle over a smooth projective curve C. Then, we have
,
deg(V )
(13.37) λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2) = .
[M1,0 (P(V ),df )] 12d
Proof. It is well-known that there exist a rank-2 bundle V over P1 × C and
two points b1 , b2 ∈ P1 such that
V|{b1 }×C = V,
V|{b2 }×C = (OC ⊕ M ) ⊗ N
13.4. GROMOV-WITTEN INVARIANTS OF THE HILBERT SCHEME X [2] 265
where M and N are line bundles on C with M being very ample. As in the proof
of Lemma 13.14, we conclude from Lemma 13.12 that
,
(13.38) λ · c2d R1 (f1,0 )∗ ev∗1 OP(V ) (−2)
[M1,0 (P(V ),df )]
,
2 deg(N )
= λ · c2d R1 (f1,0 )∗ ev∗1 OP(OC ⊕M ) (−2) + .
[M1,0 (P(OC ⊕M ),df )] 12d
Since M is very ample, there exists a morphism α : C → P1 such that M =
α∗ OP1 (1). Then,
OC ⊕ M = α∗ OP1 ⊕ OP1 (1) .
This induces an isomorphism
M1,0 (P(OC ⊕ M ), df ) ∼
= C ×P1 M1,0 P(OP1 ⊕ OP1 (1)), df .
Let
: M1,0 (P(OC ⊕ M ), df ) → M1,0 P(OP1 ⊕ OP1 (1)), df
α
be the projection. Then
,
λ · c2d R1 (f1,0 )∗ ev∗1 OP(OC ⊕M ) (−2)
[M1,0 (P(OC ⊕M ),df )]
, " #
= ∗ λ · c2d α
α ∗ R1 (f1,0 )∗ ev∗1 O (−2)
P OP1 ⊕OP1 (1)
[M1,0 (P(OC ⊕M ),df )]
,
α) ·
= deg( λ · c2d R1 (f1,0 )∗ ev∗1 OP(OP1 ⊕OP1 (1)) (−2) .
[M1,0 (P(OP1 ⊕OP1 (1)),df )]
rational number d0 > 0 and some integer d. We see from (13.26) that the expected
dimension of the moduli space M0,r (X [2] , β) is equal to
d = −d0 KX
2
+ 1 + r.
2
Since d0 KX must be a positive integer, we conclude from the Fundamental Class
Axiom that all the Gromov-Witten invariants without descendant insertions defined
via M0,r (X [2] , β) vanish except possibly when d0 KX
2
= 1. Now write KX = mC0
where C0 is an irreducible and reduced curve, and m ≥ 1 is an integer. By Re-
mark 13.8, d0 = d0 /m for some integer d0 ≥ 1. Therefore, we obtain
2
1 = d0 KX = d0 m(C0 )2 .
It follows that d0 = m = (C0 )2 = 1. Hence KX
2
= 1 and d0 = 1.
Case (i) in Proposition 13.16 will be dealt with in Example 15.10. By the
Divisor Axiom, Case (ii) in Proposition 13.16 can be reduced to the invariant
X [2]
(13.40) 1 1,dβ2 .
The invariants (13.40) and (13.41) have been computed in [HLQ, Section 6].
In the rest of this section, we will present the calculation of (13.40) for an
arbitrary smooth projective surface X. Let d ≥ 2 be an integer. For simplicity, we
put
(13.42) Mg,r,d = Mg,r (X [2] , dβ2 ).
The next lemma determines the obstruction sheaf over the moduli space M1,0,d and
the corresponding virtual fundamental class.
Lemma 13.17. (i) The obstruction sheaf Ob = R1 (f1,0 )∗ (ev∗1 TX [2] ) over
the moduli space M1,0,d is locally free of rank 2d + 2;
(ii) [M1,0,d ]vir = c2d+2 (Ob) ∩ [M1,0,d ].
Proof. (i) Let ev1 : M1,1,d → X [2] and f1,0 : M1,1,d → M1,0,d be the evalu-
ation map and the forgetful map respectively. Let u = [μ : D → X [2] ] ∈ M1,0,d .
Then
H 1 f −1 (u), (ev∗ TX [2] )| −1 ∼ H 1 (D, μ∗ TX [2] ) = H 1 D, μ∗ (TX [2] |μ(D) ) .
=
1,0 1 f1,0 (u)
By (13.45),
∗1 TB2 ∼
R1 (f1,0 )∗ ev ∗1 (ρ∗ TX ).
= R1 (f1,0 )∗ ev
1 = φ ◦ f1,0 , we get
Since ρ ◦ ev
∗ TB ∼
R1 (f1,0 )∗ ev 1 2 = R1 (f1,0 )∗ f ∗ (φ∗ TX )
1,0
∼
= R1 (f1,0 )∗ OM1,1,d ⊗ φ∗ TX
∼
= H ∗ ⊗ φ ∗ TX .
1 , we see from Lemma 13.17 (i) that
(ii) Since ev1 factors through ev
∗1 (TX [2] |B2 ) .
Ob = R1 (f1,0 )∗ (ev∗1 TX [2] ) = R1 (f1,0 )∗ ev
268 13. COSECTION LOCALIZATION FOR THE HILBERT SCHEMES
Since B2 is a smooth divisor in X [2] and OB2 (B2 ) = OB2 (−2), we have
∗1 OB2 (−2) → 0
→ R1 (f1,0 )∗ ev
u = [μ : D → X [2] ] ∈ M1,0,d ,
and assume that μ(D) = M2 (x). Then, since OB2 (−2)|M2 (x) = OM2 (x) (−2),
−1
H 0 f1,0 ∗1 OB2 (−2)|f −1 (u) ∼
(u), ev = H 0 (D, μ∗ OM2 (x) (−2)) = 0.
1,0
X [2]
1 1,dβ2 = deg[M1,0,d ]vir = deg c2d+2 (Ob) ∩ [M1,0,d ] .
1 X [2]
1,dβ2 ∗1 OB2 (−2)
= c2 H∗ ⊗ φ∗ TX · c2d R1 (f1,0 )∗ ev
= ∗1 OB2 (−2)
λ2 + φ∗ KX · λ + φ∗ c2 (TX ) · c2d R1 (f1,0 )∗ ev
= ∗1 OB2 (−2) .
φ∗ KX · λ + φ∗ c2 (TX ) · c2d R1 (f1,0 )∗ ev
∗1 OB2 (−2)
φ∗ c2 (TX ) · c2d R1 (f1,0 )∗ ev
,
= χ(X) · c2d R1 (f1,0 )∗ ev∗1 OP1 (−2)
[M1,0 (P1 ,d[P1 ])]
= 0.
Hence,
1 X [2]
1,dβ2 ∗1 OB2 (−2) .
= φ∗ KX · λ · c2d R1 (f1,0 )∗ ev
13.4. GROMOV-WITTEN INVARIANTS OF THE HILBERT SCHEME X [2] 269
Choose two smooth irreducible curves C1 and C2 satisfying [C1 ]−[C2 ] = KX . Since
B2 ∼
= P(TX∗
), we see that
1 X [2]
1,dβ2 ∗1 OB2 (−2)
= φ∗ ([C1 ] − [C2 ]) · λ · c2d R1 (f1,0 )∗ ev
,
= λ · c2d R1 (f1,0 )∗ ev∗1 OP(TX∗ |C1 ) (−2)
∗|
[M1,0 (P(TX C1 ),df )]
,
− λ · c2d R1 (f1,0 )∗ ev∗1 OP(TX∗ |C2 ) (−2)
∗|
[M1,0 (P(TX C2 ),df )]
∗ ∗
deg(TX |C1 ) deg(TX |C2 )
= −
12d 12d
2
KX
=
12d
where we have used Proposition 13.15 in the third step.
Problem 13.20. Let n ≥ 3 and d ≥ 2. Let X be a smooth projective complex
[n]
surface. What is the genus-1 extremal Gromov-Witten invariant 1 X
1,dβn ?
CHAPTER 14
and HT∗ (pt) = C[t1 , t2 ]. Similar to Example 11.2, the 2-dimensional algebraic torus
T acts on C2 by
(14.2) (s1 , s2 ) · (u, v) = (s1 u, s2 v), s 1 , s 2 ∈ C∗
where u and v denote the coordinate functions of C2 . This action lifts to a T-action
on the Hilbert schemes (C2 )[n] . The study of the equivariant cohomology of the
Hilbert schemes (C2 )[n] in terms of the Heisenberg algebra action can be carried
out as in Section 11.1 and Section 11.2. The Heisenberg algebra is generated by the
operators pk ([Y ]) defined in (11.26) and (11.27), where Y denotes a T-invariant
closed subscheme of C2 . The analogue of Proposition 11.6 gives the following
commutation relations:
(14.3) [pk ([C2 ]), p ([x])] = −kδk,− Id.
It allows us to define a linear isomorphism of Fock spaces:
+∞
(14.4) Ψ: F ⊗C C[t1 , t2 ] → HT∗ (C2 )[n]
n=0
by putting
(14.5) Ψ α−k1 · · · α−kr |0 = p−k1 ([C2 ]) · · · p−kr ([C2 ])|0
for k1 , . . . , kr > 0.
Lemma 14.1. Let k > 0, and x denote the origin of C2 . Then via the iso-
morphism Ψ, the operators α−k and αk correspond to the operators p−k ([C2 ]) =
1/(t1 t2 ) · p−k ([x]) and −pk ([x]) respectively.
Proof. Note that [C2 ] = 1/(t1 t2 )·[x]. By (14.5), the operator α−k corresponds
to the operator p−k ([C2 ]) = 1/(t1 t2 )·p−k ([x]). In view of the commutation relations
(14.1) and (14.3), αk corresponds to −pk ([x]).
For a partition λ with parts λi , put
1
|λ = α−λi · |0 ,
zλ i
|λ([x]) = (t1 t2 )(λ) · |λ
where zλ is defined by (1.1). Under the identification (14.4), the elements |λ , λ
n form a linear basis of the equivariant cohomology HT∗ (C2 )[n] of (C2 )[n] . By
Lemma 14.1, |λ([x]) corresponds to the class
1
(14.6) p−λi ([x]) · |0 .
zλ i
For simplicity, we will denote the classes |λ and |λ([x]) by λ and λ([x]) respectively
if no confusion arises.
The standard inner product on the equivariant cohomology induces the follow-
ing nonstandard inner product on Fock space after an extension of scalars:
(−1)|μ|−(μ) · δμ,ν
μ, ν = |μ , |ν = .
(t1 t2 )(μ) · zμ
The equivariant Gromov-Witten theory can be defined as in Section 12.1. Let
βn be the homology class of (C2 )[n] defined in (1.34). The T-fixed points ξλ of
(C2 )[n] are in one-to-one correspondence with the partitions λ n. Throughout this
14.1. EQUIVARIANT QUANTUM COHOMOLOGY OF (C2 )[n] 273
chapter, as in [OP3, OP4], we adopt the convention that ξλ is defined by the ideal
(12.12). By the localization theorem, the equivariant cohomology classes [ξλ ] form
a linear basis of the localized equivariant cohomology HT∗ (C2 )[n] ⊗C C(t1 , t2 ). The
locus of the stable maps in Mg,k ((C2 )[n] , dβn ) meeting a T-fixed point is compact.
The T-equivariant Gromov-Witten invariants of (C2 )[n] are defined by putting
@ A ,
[ξλ1 ], . . . , [ξλk ] = ev∗1 ([ξλ1 ]) ∪ · · · ∪ ev∗k ([ξλk ])
g,dβn [Mg,k ((C2 )[n] ,dβn )]vir
where ev1 , . . . , evk are the evaluation maps on Mg,k ((C2 )[n] , dβn ), and λ1 , . . . , λk
are partitions of n.
The following is [OP3, Theorem 1].
Theorem 14.2. Let w1 , w2 ∈ HT∗ (C2 )[n] ⊗C C(t1 , t2 ), and let M(q, t1 , t2 ) be
k (−q)k + 1 1 (−q) + 1
(14.7) (t1 + t2 ) − α−k αk
2 (−q)k − 1 2 (−q) − 1
k>0
1
+ (t1 t2 αk+ α−k α− − α−k− αk α ).
2
k,>0
Then,
[n]
w1 , c1 OC2 , w2 0,dβn q d = w1 , M(q, t1 , t2 ) w2 .
d≥0
[n]
Note that c1 OC2 = −Bn /2 where Bn ⊂ (C2 )[n] is the boundary divisor
defined by (1.26). The operator M(q, t1 , t2 ) is known as the equivariant quantum
[n]
corrected boundary operator. Since βn · c1 OC2 = 1, Theorem 14.2 implies that
@ A
d w1 , w2 0,dβ q d = w1 , M(q, t1 , t2 ) − M(0, t1 , t2 ) w2
n
d≥1
The structure of the small equivariant quantum cohomology ring QHT∗ (C2 )[n]
of (C2 )[n] is given by the following [OP3, Corollary 1].
[n]
Corollary 14.3. The divisor class c1 OC2 generates the small equivariant
quantum cohomology ring QHT∗ (C2 )[n] of (C2 )[n] over C(q, t1 , t2 ).
Proof. Since the limiting operator
1 k (−q)k + 1 1 (−q) + 1
−1
lim M(q, t, t ) = − α−k αk
t→∞ t 2 (−q)k − 1 2 (−q) − 1
k>0
is diagonal with distinct eigenvalues, M(q, t1 , t2 ) has distinct eigenvalues for generic
values of the parameters. By the Localization Theorem (Theorem 11.1), the equi-
variant cohomology ring HT∗ (C2 )[n] of (C2 )[n] is semisimple after localization.
274 14. EQUIVARIANT QUANTUM OPERATOR
Thus the equivariant quantum cohomology ring QHT∗ (C2 )[n] of (C2 )[n] is also
semisimple after localization. So
QHT∗ (C2 )[n] = C(q, t1 , t2 ) · wi
i
where {wi }i is the (finite) set of idempotents of QHT∗ (C2 )[n] satisfying wi · wj =
δi,j ·wi . Note that the idempotents wi are eigenvectors of quantum multiplication by
[n]
c1 OC2 . So they are eigenvectors under the action of the operator M(q, t1 , t2 ). The
element 1/n! · α−1
n
|0 represents the unit of QHT∗ (C2 )[n] . Since the unit is equal to
i wi , the action of M(q, t1 , t2 ) on 1/n!·α−1 |0 generates the n-eigenvalue subspace
n
of Fock space. Hence, c1 OC2 generates QHT∗ (C2 )[n] over C(q, t1 , t2 ).
[n]
In the rest of this section, we will outline the proof of Theorem 14.2. The first
step is to introduce a T-equivariant reduced obstruction theory for stable maps (of
positive degrees) to (C2 )[n] relative to the moduli space of pointed genus-g curves.
The new obstruction theory differs from the standard theory by a 1-dimensional
obstruction space L with c1 (L) = (t1 + t2 ). Therefore, when d > 0, we have
(14.10) [Mg,k ((C2 )[n] , dβn )]vir = (t1 + t2 ) · [Mg,k ((C2 )[n] , dβn )]vir
red
for which the degree splitting has at least two non-zero terms corresponding to
two different points xj and xk . The component M[j, k] has a standard obstruction
theory obtained from the standard obstruction theory of M0,3 ((C2 )[n] , dβn ). The
standard obstruction theory of M[j, k] has a 2-dimensional quotient obtained from
the 2-dimensional family of holomorphic symplectic forms (at xj and xk ). So we
obtain a doubly reduced obstruction theory by reducing the obstruction space by
the 2-dimensional quotient. The nonequivariant integral of the (singly) reduced
theory over such a component vanishes since the singly reduced theory contains an
additional 1-dimensional trivial factor. Therefore, the only components of (14.11)
[n]
which contribute to the integral μ([x]), c1 OC2 , μ d are those for which the degree
d is distributed entirely to a single factor of the product.
In summary, the following properties have been verified:
[n]
(i) μ, c1 OC2 , ν d = 0 if d > 0 and μ = ν;
(ii) μ, c1 OC2 , μ d = γn,d · (t1 t2 )−(μ) · (t1 + t2 ) for some γn,d ∈ Q;
[n]
points on the domain are unmarked, and the maps are required to be nonconstant
on all connected components.
The partition function ZGW (C2 × P1 )n[P1 ],λ1 ,...,λr of the local Gromov-Witten
theory of C2 × P1 is defined to be
,
u 2h−2
· •
e − R• π∗ f ∗ (OP1 ⊕ OP1 )
h∈Z [Mh (P1 ,λ1 ,...,λr )]vir
•
where π and f are from the universal diagram for Mh (P1 , λ1 , . . . , λr ):
f
U
⏐ −→ P1
⏐
π%
•
Mh (P1 , λ1 , . . . , λr ).
The shifted generating function GW∗n (C2 × P1 )λ1 ,...,λr is defined to be
r
(λj )
(−iu)n(2−r)+ j=1 · ZGW (C2 × P1 )n[P1 ],λ1 ,...,λr .
(C2 )[n]
Fix a point ξ ∈ M0,r . Define the multipoint invariant λ1 , . . . , λr ξ to be
+∞ ,
qd · ev∗1 (λ1 ) ∪ · · · ∪ ev∗r (λr ) ∪ π̃ ∗ (ξ)
d=0 [M0,r ((C2 )[n] ,dβn )]vir
where π̃ : M0,r ((C2 )[n] , dβn ) → M0,r is the T-equivariant forgetful map, and
ev1 , . . . , evr are the evaluation maps on M0,r ((C2 )[n] , dβn ). The following corre-
spondence theorem [OP3, Theorem 2] relates the local Gromov-Witten theory of
(C2 )[n]
C2 × P1 to the genus-0 invariants λ1 , . . . , λr ξ of the Hilbert scheme (C2 )[n] .
Theorem 14.4. After the variable change eiu = −q,
(C2 )[n]
(14.14) GW∗n (C2 × P1 )λ1 ,...,λr = (−1)n · λ1 , . . . , λr ξ .
Proof. A direct comparison of the formulas of Theorem 14.2 and [BP, Theo-
rem 6.5] yields the result in case r = 3 and λ1 = (1n−2 2). Moreover, a verification
shows that the degeneration formula of local Gromov-Witten theory is compatible
via the correspondence (14.14) with the splitting formula for genus-0 fixed mod-
uli invariants of (C2 )[n] . So the correspondence (14.14) follows from the observa-
tion that by Corollary 14.3 and the reconstruction result in [BP, Appendix], both
sides of (14.14) are canonically determined by the 3-point case with one 2-cycle
(1n−2 2).
It would be interesting to compare the present Hilbert/Gromov-Witten cor-
respondence Theorem 14.4 with the Hilbert/Gromov-Witten correspondence dis-
cussed in Chapter 12.
The Gromov-Witten and Donaldson-Thomas theories of C2 × P1 are related
by the correspondence conjectured in [MNOP1, MNOP2] and refined for the
equivariant context in [BP]. The proof of this equivalence has been completed in
[OP5].
In the context of the Crepant Resolution Conjecture, Bryan and Graber [BG]
proved that the equivariant quantum cohomology of the Hilbert scheme (C2 )[n]
is equivalent to the equivariant quantum orbifold cohomology of the symmetric
product (C2 )(n) . Related material will be presented in Chapter 16.
14.2. EQUIVALENCE OF FOUR THEORIES 277
GW theory of C2 × P1 DT theory of C2 × P1
where Cg,r denotes the universal curve over the moduli space Mg,r of genus-g stable
curves with r marked points, and (C2 × Cg,r )/Mg,r is treated as a family of 3-folds
parametrized by Mg,r . In establishing the above vertical equivalence between the
278 14. EQUIVARIANT QUANTUM OPERATOR
genus-g Gromov-Witten theory of (C2 )[n] and the genus-g quantum orbifold theory
of (C2 )(n) , a key role was played by the quantum differential equations for which
the genus-0 case is the subject of the next section.
describes particles moving on the torus |zi | = 1 interacting via the potentials |zi −
zj |−2 . The parameter θ adjusts the strength of the interaction. The function
φ(z) = (zi − zj )θ
i<j
is an eigenfunction of HCS , and the operator φHCS φ−1 preserves the space of sym-
metric polynomials in the variables z1 , z2 , z3 , . . .. Identify the space of symmetric
functions in the variables z1 , z2 , z3 , . . . with the Fock space F via the linear map
pμ → zμ |μ
14.3. QUANTUM DIFFERENTIAL EQUATIONS 279
where pμ is the power symmetric function defined in Section 1.2. Via this identifi-
cation, the operator φHCS φ−1 acts on the Fock space F.
A direct computation shows that the operator φHCS φ−1 is equal to
1−θ 1
ΔCS = kα−k αk + (θαk+ α−k α− + α−k− αk α )
2 2
k>0 k,>0
∂
modulo scalars and a multiple of the momentum operator i zi ∂z i
. We have
(·)+1 −(·)
(14.18) M(0) = −t1 ΔCS |θ=−t2 /t1 t1
where (·) is the function mapping the basis |μ to (μ). This implies that the
behavior of (14.16) near the regular singularities q = 0, ∞ is described by the
Schrödinger equation for (14.17).
For Ψ in the n-th component of F, the equation (14.15) is a linear first order
ODE in P2 (n) unknowns, where P2 (n) is the number of partitions of n. In view of
(14.7), it has regular singularities. These are q = 0, ∞, and the solutions ξ of
(−ξ)k = 1, k = 2, . . . , n
excluding q = −1. Using L’Hospital’s rule, we see that for k > 0,
k (−q)k + 1 1 (−q) + 1
(14.19) lim − = 0.
q→−1 2 (−q)k − 1 2 (−q) − 1
Therefore, q = −1 is a nonsingular point of (14.15).
Two special values of q play a special role. These are q = −1 and q = 0, which
may be called the Gromov-Witten and Donaldson-Thomas points, respectively. On
one hand, the point q = 0 is a regular singularity of (14.15), and (14.18) relates
the residue of (14.15) at q = 0 to the Calogero-Sutherland operator. In particular,
the eigenfunctions of M(0) are, up to normalization, the Jack symmetric functions.
(α)
More precisely, recall the Jack symmetric function Pλ ∈ F defined in Section 1.2
with α = 1/θ. Then the eigenfunctions of M(0) are
|λ| (·) (α)
J λ = t2 t1 Pλ |α=−t1 /t2 .
Note that geometrically, M(0) is the equivariant boundary operator
[n]
c1 OC2 ∪,
n
(2πi)(μ)
Γ(t1 , t2 )|μ = GGW (t1 , t2 )|μ .
j μj
Define the connection
d
%Γ = Γ q − M(q, t1 , t2 ) Γ−1 .
dq
The monodromy at q = −1 is given by [OP4, Theorem 3].
Theorem 14.5. The monodromy of %Γ based at q = −1 is a Laurent polyno-
mial in
T1 = e2πit1 , T2 = e2πit2 .
Moreover, it is unitary with respect to the Hermitian form defined by
δμ,ν μ /2 −μ /2 μ /2 −μ /2
μ, ν = (T1 i − T1 i )(T2 i − T2 i ) .
z(μ) i
Next, the connection problem for the QDE (14.15) between the special points
q = −1 and q = 0 is solved. The connection problem is to find the value of the
matrix Y(q) at q = −1. It is proved in [OP4] that up to normalization, the value
of the matrix Y(q) at q = −1 is given in terms of Macdonald polynomials with
parameters e2πit1 and e2πit2 . More precisely, let Pλ ∈ F ⊗ Q(q, t) be the monic
Macdonald polynomial as defined in [Mac2]. In studying the K-theory of the
Hilbert scheme, Haiman [Hai3] used the polynomial
Hλ (q, t) = tn(λ) 1 − q a() t−()−1 · ΥPλ (q, t−1 )
∈Dλ
where n(λ) = i (i − 1)λi , and Υ is the diagonal operator defined by
Υ|μ = (1 − t−μi )−1 · |μ .
i
Haiman’s parameters q and t are identified with the parameters T1 and T2 in The-
orem 14.5 respectively. Define the operator O(a) in K-theory by
O(a)Hλ = e−2πiac(λ) Hλ .
Let GDT = GDT (t1 , t2 ) be the diagonal operator defined by
1
GDT (t1 , t2 )|λ = q −c(λ) · |λ
w
Γ(w + 1)
14.3. QUANTUM DIFFERENTIAL EQUATIONS 281
k
k
Note from (3.11) that if n = ni and m = (2ni − 2 + |αi |), then
i=1 i=1
η ) are
Note that the (n − 1) exceptional curves in the surface gn (X
(15.9) M2 (xi ) + (η\{xi }), i = 1, . . . , n − 1.
Finally, choose η ∈ X such that Supp(η) ∩ (C ∪ C)
[n−2] = ∅. Then according to
Proposition 3.16, sα,
α is represented by the closure of the subset
(15.10) and x = x̃} ⊂ X [n] .
{x + x̃ + η| x ∈ C, x̃ ∈ C,
Via the above notations, the following elementary lemma lists the basis elements
of the cohomology groups H 2 (X [n] ), H 4n−2 (X [n] ) and H 4n−4 (X [n] ).
Lemma 15.2. Let n ≥ 2 and X be simply connected. Let {α1 , . . . , αs } be a
basis of H 2 (X) represented by real-2-dimensional cycles {C1 , . . . , Cs } respectively.
Then,
(i) a basis of H 2 (X [n] ) consists of the cohomology classes
Bn , a−1 (αi )1−(n−1) |0 (i = 1, . . . , s).
(ii) a basis of H 4n−2 (X [n] ) consists of the cohomology classes
a−2 (x)a−1 (x)n−2 |0 , a−1 (αi )a−1 (x)n−1 |0 (i = 1, . . . , s).
(iii) a basis of H 4n−4 (X [n] ) consists of the cohomology classes
sn,1 , sn,2 , sn,3 , sαi ,1 (i = 1, . . . , s),
sαi ,2 (i = 1, . . . , s), sαi ,αj (i, j = 1, . . . , s).
Proof. Note that (i) and (ii) follow from Lemma 2.1. To prove (iii), fix a
point x ∈ X. Expand the basis {α1 , . . . , αs } of H 2 (X) to the basis
{α0 = 1X , α1 , . . . , αs , αs+1 = x}
∗
of H (X) = H (X) ⊕ H (X) ⊕ H 4 (X). By (15.1), a basis of H 4n−4 (X [n] ) consists
0 2
of the classes
(15.11) a−n1 (αm1 ) . . . a−nk (αmk )|0
k
satisfying ni ≥ 1, ni = n, and
i=1
k
(2ni − 2 + |αmi |) = 4n − 4.
i=1
Thus,
k
|αmi | ≥ 4k − 4.
i=1
Since X is simply connected, |αmi | ∈ {0, 2, 4} for every i. It follows that |αmi | = 4
for at least (k − 2) many i’s, i.e., αm3 = . . . = αmk = x. Then,
|αm1 | + |αm2 | = 2(n − k) + 4 ≥ 4.
286 15. THE GENUS-0 EXTREMAL GROMOV-WITTEN INVARIANTS
Assume that |αm2 | = 4, that is, αm2 = x. Then, |αm1 | = 2(n−k). If |αm1 | = 4,
n−2
then αm1 = x, k = n − 2, and ni = n. In this case, (15.11) is equal to either
i=1
sn,2 or sn,3 . If |αm1 | = 2, i.e., αm1 = αi for some 1 ≤ i ≤ s, then k = n − 1 and
n−1
ni = n. In this case, (15.11) is equal to either sαi ,1 or sαi ,2 . If |αm1 | = 0, then
i=1
n
αm1 = 1X , k = n, and ni = n. In this case, (15.11) is equal to sn,1 .
i=1
Finally, assume that |αm2 | ≤ 2. Since |αm1 | ≤ |αm2 | and
|αm1 | + |αm2 | = 2(n − k) + 4 ≥ 4,
n
we must have |αm1 | = |αm2 | = 2. So k = n and ni = n. In this case, (15.11) is
i=1
equal to sαi ,αj for some i and j satisfying 1 ≤ i, j ≤ s.
where Δi,j
n∗ = {(x1 , . . . , xi , . . . , xj , . . . , xn ) ∈ X∗ | xi = xj } for 1 ≤ i < j ≤ n.
n
'
X [n] ∗ = X n
∗ /Sn .
'
Let σ̃ : X ∗ → X
n [n] i,j '
∗ be the quotient map. Let E∗ ⊂ X∗ be the exceptional locus
n
i,j
over Δn∗ . Consider the following morphisms:
(15.13) p1,2 n∗ −→ X,
: Δ1,2 (x, x, x3 , . . . , xn ) → x,
(n)
(15.14) j2 : Xs∗ −→ X, 2x + x3 + . . . + xn → x.
15.1. 1-POINT GENUS-0 EXTREMAL GROMOV-WITTEN INVARIANTS 287
∗
Since the normal bundle of Δ1,2 n
n∗ in X∗ is isomorphic to p1,2 TX , we have
E∗1,2 ∼
= P(p∗1,2 TX
∗
).
The subgroup S2 × Sn−2 ⊂ Sn acts on Δ1,2 n∗ with the S2 -factor acting trivially on
Δ1,2 1,2 1,2
n∗ . The action of S2 × Sn−2 on Δn∗ lifts to an action on E∗ . It is easy to see
that
(n)
Xs∗ = Δ1,2n∗ /(S2 × Sn−2 )
and
B∗ = E∗1,2 /(S2 × Sn−2 ).
Regard p1,2 : Δ1,2n∗ → X as an S2 × Sn−2 -equivariant morphism where S2 × Sn−2
acts on X trivially. Then, S2 × Sn−2 acts on p∗1,2 TX ∗
, and the isomorphism E∗1,2 ∼
=
P(p∗1,2 TX
∗
) is S2 × Sn−2 -equivariant. So we get an isomorphism
j1 : B∗ = E∗1,2 /(S2 × Sn−2 ) ∼
= P(p∗1,2 TX
∗
)/(S2 × Sn−2 ) ∼
= P(j2∗ TX
∗
).
where the last isomorphism is due to the fact that the S2 -factor acts trivially on
p∗1,2 TX and the Sn−2 -factor commutes with the morphism p1,2 .
Next, we study OB∗ (B∗ ). Since
⎛ ⎞
σ̃ ∗ OX [n] ∗ (B∗ ) ∼
= OX [n] ∗ ⎝2 E∗i,j ⎠
1≤i<j≤n
Consider the open subset U0 of M0,0 (X [n] , dβn ) consisting of stable maps
[μ : D → X [n] ] such that μ(D) ⊂ X [n] ∗ . Similarly, take the open subset U1 of
M0,1 (X [n] , dβn ) consisting of stable maps [μ : (D; p) → X [n] ] such that μ(D) ⊂
−1
X [n] ∗ . Clearly U1 = f1,0 (U0 ). Let [μ : (D; p) → X [n] ] ∈ U1 . Since μ∗ (D) ∼ dβn , we
must have
μ(D) = M2 (x2 ) + x3 + . . . + xn
for some distinct points x2 , . . . , xn ∈ X. Hence μ(D) ⊂ B∗ . Moreover, the compos-
ite ρn ◦ ev1 sends the stable map [μ : (D; p) → X [n] ] to the point 2x2 + x3 + . . . + xn ,
which is independent of the marked point p on D. Hence ev1 induces a morphism φ
288 15. THE GENUS-0 EXTREMAL GROMOV-WITTEN INVARIANTS
where π : P(j2∗ TX
∗ (n)
) → Xs∗ is the natural projection of the P1 -bundle.
Note that the fiber φ−1 (2x2 +x3 +. . .+xn ) over a fixed point 2x2 +x3 +. . .+xn ∈
ρn (B∗ ) is simply
M0,0 M2 (x2 ) + x3 + . . . + xn , d[M2 (x2 ) + x3 + . . . + xn ]
which is isomorphic to the moduli space M0,0 (P1 , d[P1 ]) via the isomorphism
M2 (x2 ) + x3 + . . . + xn ∼
= P1 .
Hence the complex dimension of U0 is equal to
dim M0,0 (P1 , d[P1 ]) + 2(n − 1) = 2n − 3 + 2d − 1.
Since KX [n] ·dβn = 0, the expected dimension of M0,0 (X [n] , dβn ) is 2n−3 according
to (12.3) . Hence the excess dimension of U0 is e = (2d − 1).
Lemma 15.4. With notations as above, the restriction of the direct image sheaf
R1 (f1,0 )∗ (ev∗1 TX [n] ) to U0 is a locally free sheaf of rank (2d − 1).
Proof. Take a stable map u = [μ : D → X [n] ] in U0 , and consider
H 1 (f −1 (u), (ev∗ TX [n] )| −1 ) ∼
1,0 1 = H 1 (D, μ∗ TX [n] ).
f1,0 (u)
and
V = (R1 (f1,0 )∗ (ev1 )∗ TX [n] )|U0
= R1 (f˜1,0 )∗ ((ev1 )∗ TX [n] )|U 1
1 )∗ TX [n] |B∗ .
= R1 (f˜1,0 )∗ (ev
Since B∗ is a smooth codimension-1 subvariety of X [n] , we obtain the exact sequence
(15.20) 0 → TB∗ → TX [n] |B∗ → OB∗ (B∗ ) → 0.
∗
1 ) and (f˜1,0 )∗ to the exact sequence (15.20), we get
Applying (ev
1 )∗ TB∗ → V → R1 (f˜1,0 )∗ (ev
R1 (f˜1,0 )∗ (ev 1 )∗ OB∗ (B∗ ) → 0
1 )∗ TB∗ = 0 since f˜1,0 is of relative dimension 1.
where we have used R2 (f˜1,0 )∗ (ev
If [μ : D → X ] is a stable map in U0 , then
[n]
μ(D) = M2 (x2 ) + x3 + . . . + xn .
Hence the normal bundle of μ(D) in B∗ is trivial since μ(D) is a fiber of the P1 -
bundle P(j2∗ TX
∗ (n)
) over Xs∗ . Thus
TB∗ |μ(D) ∼
⊕(2n−2)
= Oμ(D) (2) ⊕ Oμ(D) .
Therefore,
H 1 (D, μ∗ TB∗ ) ∼
⊕(2n−2)
= H 1 (D, μ∗ (Oμ(D) (2) ⊕ Oμ(D) )) = 0,
By Lemma 15.5 (ii), we get c2d−1 (V) = −(j2 ◦ φ)∗ KX · c2d−2 (E) where
1 )∗ ((j2 ◦ π)∗ TX ⊗ OP(j2∗ TX∗ ) (−1)).
E = R1 (f˜1,0 )∗ (j1 ◦ ev
In view of (12.5) and (15.18), we conclude that
,
γ 0,dβn = (ev1 )∗ γ
[M0,1 (Y,β)]vir
Thus, γ = PD(ρ−1
n (2C1 + η̃)). So we see from (15.18) and Lemma 15.5 (ii) that
(15.25) γ 0,dβn
,
= (ev1 )∗ γ
[M0,1 (Y,β)]vir
,
= (ev1 )∗ γ
−(f˜1,0 )∗ (j2 ◦φ)∗ KX ·(f˜1,0 )∗ c2d−2 (E)
,
= − (ev1 )∗ γ
1 )∗ (ρ∗
(ev ∗ ˜ ∗
n j2 KX )·(f1,0 ) c2d−2 (E)
,
= − 1 )∗ PD(ρ−1
(ev n (2C1 + η̃) · c1 (OB∗ (B∗ )))
1 )∗ (ρ∗
(ev ∗ ˜ ∗
n j2 KX )·(f1,0 ) c2d−2 (E)
, " " # #
= − 1 )∗ PD ρ−1
(ev n (j ∗
2 XK ) · (2C 1 + η̃) · c 1 (OB∗ (B ∗ ))
(f˜1,0 )∗ c2d−2 (E)
,
= 2 KX , C 1 · 1 )∗ PD(ξ)
(ev
(f˜1,0 )∗ c2d−2 (E)
where ξ ∈ ρ−1 −1
n (2x + η̃) = ρn (2x + x1 + . . . + xn−2 ) is a fixed point for some fixed
point x ∈ C1 . Also, we have used the isomorphism (15.16) in the last step.
Let M1 ⊂ M0,1 (X [n] , dβn ) be the subset consisting of all the stable maps
[μ : (D; p) → X [n] ] with μ(p) = ξ. If [μ : (D; p) → X [n] ] ∈ M1 , then
ρn (μ(D)) = ρn (μ(p)) = 2x + x1 + . . . + xn−2 .
So μ(D) = ρ−1 ˜
n (2x + x1 + . . . + xn−2 ). Thus the restriction of the forgetful map f1,0
−1
to M1 gives a degree-d morphism from M1 to φ (2x + x1 + . . . + xn−2 ). Put
M0 = φ−1 (2x + x1 + . . . + xn−2 ).
Now, as algebraic cycles, we have
(f˜1,0 )∗ [M1 ] = d[M0 ] = d · φ∗ [2x + x1 + . . . + xn−2 ].
By (15.25), we obtain
(15.26) γ 0,dβ2 = 2 KX , C1 · [M1 ] · (f˜1,0 )∗ c2d−2 (E)
= 2 KX , C1 · (f˜1,0 )∗ [M ] · c2d−1 (V)
1
∗
= 2d KX , C1 · φ [2x + x1 + . . . + xn−2 ] · c2d−2 (E)
= 2d KX , C1 · c2d−2 (E|φ−1 (2x+x1 +...+xn−2 ) ).
By Remark 15.6,
E|φ−1 (2x+x1 +...+xn−2 ) ∼
= R1 (f1,0 )∗ (ev1 )∗ (OP1 (−1) ⊕ OP1 (−1))
where f1,0 and ev1 denote the forgetful map and the evaluation map from the moduli
space M0,1 (P1 , d[P1 ]) to M0,0 (P1 , d[P1 ]) and P1 respectively. By the Theorem 9.2.3
in [CK],
c2d−2 (R1 (f1,0 )∗ (ev1 )∗ (OP1 (−1) ⊕ OP1 (−1))) = 1/d3 .
15.1. 1-POINT GENUS-0 EXTREMAL GROMOV-WITTEN INVARIANTS 293
So we have
c2d−2 (E|φ−1 (2x+x1 +...+xn−2 ) )
= c2d−2 (R1 (f1,0 )∗ (ev1 )∗ (OP1 (−1) ⊕ OP1 (−1))
= 1/d3 .
Combining this with (15.26), we conclude that γ 0,dβn = 2 KX , C1 /d2 .
Lemma 15.8. Let d ≥ 1. If γ = sn,2 or sn,3 , then γ 0,dβn = 0.
Proof. Since similar argument works for sn,2 , we will only prove the lemma
for sn,3 . So assume γ = sn,3 . Let x1 , . . . , xn−2 ∈ X be fixed distinct points on
X contained in a small analytic open subset U of X. We may assume that U is
independent of the smooth surface X. Let U1 ⊂ M0,1 (X [n] , dβn ) be the analytic
open subset consisting of all stable maps [μ : (D; p) → X [n] ] with μ(p) ∈ U [n] . Since
μ∗ (D) ∼ dβn , we see that Supp(μ(D)) = Supp(μ(p)) for
[μ : (D; p) → X [n] ] ∈ M0,1 (X [n] , dβn ).
So μ(D) ⊂ U [n] , and U1 is independent of X.
Next, recall from (15.3) that sn,3 is represented by M3 (x1 ) + x2 + . . . + xn−2 .
Let M1 ⊂ M0,1 (X [n] , dβn ) be the closed subset consisting of all stable maps
[μ : (D; p) → X [n] ] with μ(p) ∈ M3 (x1 ) + x2 + . . . + xn−2 . Then, M1 ⊂ U1 since
M3 (x1 ) + x2 + . . . + xn−2 ⊂ U [n] . In addition, since Supp(μ(D)) = Supp(μ(p)), we
must have
μ(D) ⊂ M3 (x1 ) + x2 + . . . + xn−2
for every [μ : (D; p) → X [n] ] ∈ M1 . So M1 is independent of X. Thus the pull-back
ev∗1 (γ) is also independent of X.
In summary, M1 ⊂ U1 , U1 is analytic open in M0,1 (X [n] , dβn ), and M1 and U1
are independent of X. It follows from the constructions of the virtual fundamental
cycle in [LT2, LT3] that the restriction [M0,1 (X [n] , dβn )]vir |M1 is independent of
the smooth surface X. So the 1-point Gromov-Witten invariant γ 0,dβn , which is
defined to be [M0,1 (X [n] , dβn )]vir · ev∗1 (γ) with ev∗1 (γ) being independent of X, is
independent of X as well. Since all the Gromov-Witten invariants γ1 , . . . , γk 0,β
with β = 0 for a K3 surface are zero, we conclude that γ 0,dβn = 0 for d ≥ 1.
The following is the main result in this section.
Theorem 15.9. Let X be a simply connected smooth projective surface. Let
n ≥ 2, d ≥ 1, and α1 , α2 ∈ H 2 (X).
(i) If γ = sn,1 , sα1 ,α2 , sα1 ,1 , sn,2 or sn,3 , then
γ 0,dβn = 0.
(ii) If γ = sα1 ,2 , then
2
γ =
KX , α1 .
0,dβn
d2
Proof. Follows immediately from Lemma 15.7 and Lemma 15.8.
Example 15.10. Let X be a simply connected smooth projective surface. Let
us consider all the genus-0 extremal Gromov-Witten invariants of X [2] :
(15.27) γ1 , . . . , γk 0,dβ2
294 15. THE GENUS-0 EXTREMAL GROMOV-WITTEN INVARIANTS
= 2(n + (μ ) − (μ )).1 3
(15.32) Λθ ∩ ev−1 λ −1
1 (We ) ∩ ev2 (Wc ) = ∅
μ
where evi is the i-th evaluation map from (12.2). A detailed analysis involving
(15.32), (15.30) and the definition of Λθ in Proposition 13.3 leads to the following.
Proposition 15.11. Let d ≥ 1. Assume that Aλe , Aμc 0,dβ = 0. Then,
n
3 1
(15.33) (λ ) = (μ ) + δ
(15.34) 3
(μ ) = (λ1 ) + (1 − δ)
where δ = 0 or 1. If δ = 0, then λ3 = μ1 , and there exists an integer = μ3i = λ2j
for some i and j such that the partition λ1 is obtained from μ3 with deleted, and
the partition μ2 is obtained from λ2 with deleted.
λ−λ2j 2
Let Ae be the cohomology class in H ∗ (X [n−λj ] ) obtained from Aλe with the
λ−λ1j λ−λ2i −λ2j
factor a−λ2j (ej ) deleted. We similarly define Ae , Ae , etc.
By the Lemma 4.3 in [LL], there exists a universal function c,d depending only on
and d such that for every e ∈ H 2 (X), we have
(15.37) a− (e)|0 , a− (1X )|0 0,dβ = KX , e · c,d .
= − Bn /2 · dβn · Ae , Ac 0,dβn
= d · Aλe , Aμc 0,dβn
if d ≥ 1. Hence we obtain the following corollary from Lemma 15.12.
Corollary 15.13. Let λ and μ be as in Proposition 15.11 with δ = 0. Then
[n]
Aλe , c1 OX , Aμc 0,dβn q d
d≥0
[n] λ−λ2j μ−μ3i
= Aλe , c1 OX , Aμc + Ae , Ac · KX , e j · d c,d q d .
1≤i≤t d≥1
1≤j≤b
μ3 =λ2 =
i j
Next, recall the notation τk∗ from Definition 3.9. Consider the operator
(15.38) M(q)
k (−q)k + 1 1 (−q) + 1
= − a−k ak (τ2∗ KX )
2 (−q)k − 1 2 (−q) − 1
k>0
1
− a−k a− ak+ + a−k− ak a (τ3∗ [X])
2
k,>0
k(−q)k q
k−1
= − a−k ak (τ 2∗ K X ) − a−k ak (τ2∗ KX )
(−q)k − 1 1 + q 2
k>0 k>0
1
− a−k a− ak+ + a−k− ak a (τ3∗ [X]).
2
k,>0
By (4.19),
M(0) = d.
In view of this and Theorem 15.17 below, M(q) is defined to be the quantum
corrected boundary operator. The operator M(q) is closely related to the operator
M(q, t1 , t2 ) introduced in Theorem 14.2.
To compute M(q)(Aμc ), it suffices to calculate
(15.39) a−k ak (τ2∗ KX )(Aμc ),
(15.40) a−k a− ak+ (τ3∗ 1X )(Aμc ),
(15.41) a−k− ak a (τ3∗ 1X )(Aμc )
where k, > 0. Using Lemma 3.12, we see that a−k ak (τ2∗ KX )(Aμc ) is equal to
μ−μ2
μ−μ3
(15.42) (−μ2i )a−k (KX ci )Ac i + (−μ3i )a−k (KX )Ac i .
1≤i≤s 1≤i≤t
μ2 =k μ3 =k
i i
Proposition 15.14. If Aλe , M(q)(Aμc ) is not a constant function of q, then
(15.43) (λ3 ) = (μ1 ) + δ
(15.44) (μ3 ) = (λ1 ) + (1 − δ)
15.2. THE 2-POINT GENUS-0 EXTREMAL GROMOV-WITTEN INVARIANTS 297
Corollary 15.13 and Corollary 15.15 will be connected via the complex projec-
tive plane P2 . To achieve this, let X = P2 , T = (C∗ )3 and HT∗ (pt) = C[t1 , t2 , t3 ].
The group T acts on a point [z1 , z2 , z3 ] ∈ X by
N1 ∼
= T1−1 T2 ⊕ T1−1 T3
where Ti is the 1-dimensional representation given by (s1 , s2 , s3 ) → si . Similarly,
N2 ∼
= T2−1 T1 ⊕ T2−1 T3 ,
N3 ∼
= T3−1 T1 ⊕ T3−1 T1 .
in HT∗ (X) = HT∗ (X) ⊗C[t1 ,t2 ,t3 ] C(t1 , t2 , t3 ). Similarly, we obtain
[q2 ] [q3 ]
(15.47) [L1 ] = + ,
t1 − t2 t1 − t3
[q1 ] [q3 ]
(15.48) [L2 ] = + ,
t2 − t1 t2 − t3
[q1 ] [q2 ] [q3 ]
(15.49) [X] = + + .
(t2 − t1 )(t3 − t1 ) (t1 − t2 )(t3 − t2 ) (t1 − t3 )(t2 − t3 )
Using −KX = [L1 ] + [L2 ] + [L3 ] and the above formulas, we see that −KX is equal
to
(t2 + t3 − 2t1 )[q1 ] (t1 + t3 − 2t2 )[q2 ] (t1 + t2 − 2t3 )[q3 ]
(15.50) + + .
(t2 − t1 )(t3 − t1 ) (t1 − t2 )(t3 − t2 ) (t1 − t3 )(t2 − t3 )
Let Y1 , . . . , Yk be T-equivariant subvarities of X. Put
an1 · · · ank ([Y1 × . . . × Yk ]) = an1 ([Y1 ]) · · · ank ([Yk ]).
Proposition 15.16. Let L be a line in the projective plane X = P2 . Then,
[n]
(15.51) a−n ([L])|0 , c1 OX , a−n ([X])|0 0,dβ q d
n
d≥0
= a−n ([L])|0 , M(q)(a−n ([X])|0 ) .
[n]
Proof. Note that M(0) = d = c1 OX ∪. Thus, we have
[n]
a−n ([L])|0 , c1 OX , a−n ([X])|0 = a−n ([L])|0 , M(0)(a−n ([X])|0 ) .
Therefore (15.51) is equivalent to the following:
[n]
(15.52) a−n ([L])|0 , c1 OX , a−n ([X])|0 0,dβn q d
d≥1
= a−n ([L])|0 , (M(q) − M(0))(a−n ([X])|0 ) .
Recall from Section 11.2 that the equivariant Heisenberg operators are denoted by
pi (α). Take L = L3 . For d ≥ 1, we have
[n]
a−n ([L3 ])|0 , c1 OX , a−n ([X])|0 0,dβ
n
i = j. Thus for d > 0, a−n ([L3 ])|0 , c1 OX , a−n ([X])|0 0,dβn is equal to
d
p−n ([q1 ])|0 , p−n ([q1 ])|0 0,dβ
(t2 − t1 )(t3 − t1 ) 2 n
15.2. THE 2-POINT GENUS-0 EXTREMAL GROMOV-WITTEN INVARIANTS 299
d
(15.53) + p−n ([q2 ])|0 , p−n ([q2 ])|0 0,dβn .
(t1 − t2 )(t3 − t2 ) 2
The term p−n ([qi ])|0 , p−n ([qi ])|0 0,dβ has been computed by Theorem 14.2,
n
(14.8) and (14.9):
(15.54) d p−n ([qi ])|0 , p−n ([qi ])|0 0,dβn q d
d≥1
[n]
= p−n ([qi ])|0 , c1 OC2 , p−n ([qi ])|0 0,dβn q d
d≥1
= p−n ([qi ])|0 , M i (0) p−n ([qi ])|0
i (q) − M
k (−q)k + 1 1 (−q) + 1
1 (q) = − (t2 + t3 − 2t1 )
M − p−k ([q1 ])pk ([q1 ]),
(t2 − t1 )(t3 − t1 ) 2 (−q)k − 1 2 (−q) − 1
k>0
k (−q)k + 1 1 (−q) + 1
2 (q) = − (t1 + t3 − 2t2 )
M − p−k ([q2 ])pk ([q2 ]),
(t1 − t2 )(t3 − t2 ) 2 (−q)k − 1 2 (−q) − 1
k>0
(t1 + t2 − 2t3 ) k (−q)k + 1 1 (−q) + 1
M3 (q) = − − p−k ([q3 ])pk ([q3 ]).
(t1 − t3 )(t2 − t3 ) 2 (−q)k − 1 2 (−q) − 1
k>0
1
+ 2 (q) − M
p−n ([q2 ])|0 , M 2 (0) p−n ([q2 ])|0 .
(t1 − t2 )(t3 − t2 ) 2
(t2 + t3 − 2t1 )
(15.56) − p−k ([q1 ])pk ([q1 ])
(t2 − t1 )(t3 − t1 )
1
+ 2 (q) − M
p−n ([q2 ])|0 , M 2 (0) p−n ([q2 ])|0 .
(t1 − t2 )(t3 − t2 )2
In view of (15.55), we have proved the desired formula (15.52).
The following determines all the 2-point genus-0 extremal Gromov-Witten in-
variants of X [n] .
Theorem 15.17. Let X be a simply connected smooth projective complex sur-
face. Then
[n]
(15.57) Aλe , c1 OX , Aμc 0,dβn q d = Aλe , M(q)Aμc .
d≥0
We now sketch the proof of Theorem 15.17, and refer to [LL, Section 7] for the
details. Since [n]
M(0) = d = c1 OX ∪,
n
Proposition 15.11 and Proposition 15.14 reduce (15.57) to the case when λ3 = μ1
and (μ3 ) = (λ1 ) + 1. By Corollary 15.13 and Corollary 15.15, the proof of (15.57)
is further reduced to the proof of the following identity:
(−q) q
(15.58) d c,d q d = (−1) 2 − .
(−q) − 1 1 + q
d≥1
To prove (15.58), consider the special case when X = P2 , Aλe = a− ([L]) and
Aμc = a− ([X]) where L denotes a line in X. By Corollary 15.13,
[n] [n]
Aλe , c1 OX , Aμc 0,dβ q d = Aλe , c1 OX , Aμc − 3 d c,d q d .
n
d≥0 d≥1
β
For α = β ∈ PΛ , define Ξα β = Ξα∧β ∪ Ξα∧β . The discrepancy between X
α [[α]]
and
[[β]] [[α]]
X (in X ) and its complement are defined to be
[[α]]
Ξβ = X [[α]] ×X Λ Ξα
β,
[[α]] [[α]]
Xβ = X [[α]] − Ξβ
respectively. More precisely, by Lemma 1.2 in [LiJ], there exists a functorial open
embedding
[[α]]
ζαβ : Xβ → X [[β]]
induced by the universal property of the respective moduli spaces such that
Im(ζαβ ) = Xα[[β]] .
[[α]] ∼
= [[β]]
Thus we obtain an isomorphism (equivalence) ζαβ : Xβ −→ Xα . We define
"& #
(15.61) X [[≤α]] = X [[β]] / ∼,
β≤α
[[β]] [[γ]]
where the equivalence is by identifying Xγ ⊂ X [[β]] and Xβ ⊂ X [[γ]] via ζβγ for all
β, γ ≤ α. Note that X [[≤α]] is non-separated (except when α = 1Λ ), and contains
the spaces X [[≤β]] , β ≤ α, as open subschemes.
Next, let
X [n,d] = M0,3 (X [n] , dβn )
be the moduli space of 3-pointed genus-0 degree-dβn stable maps to X [n] . Recall
from (13.12) that every stable map (ϕ, C) ∈ X [n,d] has a standard decomposition
ϕ = (ϕ1 , . . . , ϕl ) ∈ X [n,d]
where the stable reduction ϕst
i is contained in X
[ni ,di ]
for some ni and di ,
ρni (Im(ϕi )) = ni xi ,
the points x1 , . . . , xl are distinct, and ϕ(p) = i ϕi (p) for all p ∈ C. We use the
ideas from [LiJ, Section 1] to index the support of
ρn (Im(ϕ)) = ni xi ∈ X (n) .
i
This is done by introducing the notion of 3-pointed genus-0 degree-δ α-stable maps
to X [n] , where α = (α1 , · · · , αl ) denotes a partition of the set [n] = {1, . . . , n} and
δ = (δ1 , · · · , δl ) with δi ’s being nonnegative integers. The set of such pairs (α, δ)
with i δi = d is denoted by P[n],d . The techniques in [LiJ, Section 1] and the
product formula [Beh2, (1)] for Gromov-Witten invariants enable us to express the
virtual fundamental cycle [X [n,d] ]vir in terms of certain discrepancy cycles
[Θ[[α,δ]] ], (α, δ) ∈ P[n],d .
In fact, one of the key points in [LQ5] is to study such decomposition of [X [n,d] ]vir
as a sum of cycles indexed by the partition type of ρn (Im(ϕ)) ∈ X (n) . However, this
cannot be done on the moduli space X [n,d] . The technique to overcome this impasse
is to use the Hilbert scheme X [[α]] of α-points defined in (15.60) and the space
X [[≤α]] defined in (15.61). Then the cycle ev∗ ([X [n,d] ]vir ) is a sum of various [Θ[[α,δ]] ]
in (X [[≤[n]]] )3 . Combining with the cosection localization theory in Section 13.1,
15.3. THE STRUCTURE OF THE GENUS-0 EXTREMAL INVARIANTS 303
pairings with [Θ[[α,δ]] ] can be studied via C ∞ -maps from X to the Grassmannians.
For d ≥ 1, we assemble those
[Θ[[α,δ]] ], (α, δ) ∈ P[n],d
with δi > 0 for every i into a homology class
Zn,d ∈ H∗ ((X [n] )3 ).
Note that we are back to the original Hilbert scheme X [n] . Now the structure
of the 3-point genus-0 extremal Gromov-Witten invariants of X [n] is given by the
following two theorems from [LQ5, Section 1].
Theorem 15.19. Let X be a simply connected smooth projective surface. Let
A1 , A2 , A3 ∈ H ∗ (X [n] ) be Heisenberg monomial classes, and πm,i be the i-th pro-
jection on (X [m] )3 . Then, A1 , A2 , A3 0,dβn is equal to
< =
3
∗
(15.62) A1,1 , A2,1 , A3,1 · Zm,d , πm,i Ai,2 .
m≤n A1,1 ◦A1,2 =A1 i=1
A2,1 ◦A2,2 =A2
A3,1 ◦A3,2 =A3
and whose coefficients depend only on d, n, λ(i) , ni,j (and hence are inde-
pendent of the surface X and the classes αi,j ).
We refer to Definition 3.13 for the operation ◦ appearing in (15.62), and to
Definition 3.28 (i) for the notation a−λ(i) (1X ) appearing in Theorem 15.20 (ii).
Geometrically, we may think of the pairing
< =
3
∗
Zm,d , πm,i Ai,2
i=1
if 2 ≤ m ≤ n − 1. It follows that
< =
3
∗
Zn,d , πn,i Ai = 0.
i=1
(ii) We compute A1 , A2 , A3 0,dβn by using (15.62). Note that the class A1,2
in (15.62) is equal to a−1 (1X )m |0 , or is equal to a−1 (1X )m−1 a−1 (α)|0 , or con-
tains a factor a−i (x) for some i > 0. By (i) and Theorem 15.20 (i), we get
A1 , A2 , A3 0,dβn = 0.
Gk (α) defined in Definition 4.1. From previous chapters, it is clear that the coho-
mology classes Gk (α, n) ∈ H ∗ (X [n] ) and the operators Gk (α) play significant roles
307
308 16. RUAN’S COHOMOLOGICAL CREPANT RESOLUTION CONJECTURE
the quantum corrected cohomology ring Hρ∗n (X [n] ). We will prove the identity
1 {k}
k (α), a−1 (β)] =
[G a (αβ).
k! −1
Following [LQ5, Section 6], an axiomatization originated from [Leh1, LQW1] and
formulated in [QW1, Subsection 2.6] completes the proof of Ruan’s conjecture.
Theorem 16.5 and its proof are from [LL], while Theorem 16.20, Theorem 16.22
and their proofs are from [LQ5].
Remark 16.2. A priori, it is not clear why the power series (16.1) should be
convergent at q = −1 so that (16.2) makes sense. The convergence of (16.1) at
q = −1 will follow from the proofs in Subsection 16.2 which reduce the convergence
for an arbitrary surface X to the case of smooth toric surfaces. The convergence of
(16.1) at q = −1 for smooth toric surfaces is a consequence of [Cheo2, Theorem 0.2]
where a stronger version of Ruan’s Cohomological Crepant Resolution Conjecture
was proved for smooth toric surfaces.
Note that as complex vector spaces, Hρ∗n (X [n] ) and H ∗ (X [n] ) are the same.
Thus,
+∞
+∞
HX = H ∗ (X [n] ) = Hρ∗n (X [n] )
n=0 n=0
as complex vector spaces. Moreover, we see from (16.2), (16.1) and the Fundamental
Class Axiom (12.6) that
w ·ρn 1X [n] = w
for every w ∈ Hρ∗n (X [n] ).
16.1. THE QUANTUM CORRECTED COHOMOLOGY RING Hρ∗n (X [n] ) 309
In the setting of the quantum corrected product, the Chern character opera-
tor Gi (α) ∈ End(HX ) and the cohomology class Gi (α, n) ∈ H ∗ (X [n] ) defined in
Definition 4.1 (iii) must be modified as follows.
Definition 16.3. Let k ≥ 0 and α ∈ H ∗ (X).
k (α, n) ∈ Hρ∗ (X [n] ) to be
(i) Define the cohomology class G n
(−1)|λ|−1
(16.3) · 1−(n−j−1) a−λ (τ∗ α)|0
λ! · |λ|!
0≤j≤k λ(j+1)
(λ)=k−j+1
which acts on the n-th component Hρ∗n (X [n] ) by the quantum corrected
k (α, n)·ρ .
product G n
(16.4) G k (α)w1 , w2 = G
k (α, n) ·ρ w1 , w2 = G k (α, n), w1 , w2 ρn (−1)
n
From now on, let the surface X be simply connected. By Remark 16.4,
1 (1X , n) = G1 (1X , n) = c1 O[n] .
G X
The following was proved in [LL, Subsection 7.2].
Theorem 16.5. We have
1
1 (1X ) =
G
[n]
c1 OX ·ρn = − : a3 :0 (τ3∗ 1X ).
n
6
Proof. First of all, recall from (14.19) that
k (−q)k + 1 1 (−q) + 1
(16.6) lim − = 0.
q→−1 2 (−q)k − 1 2 (−q) − 1
Next, by (16.2) and Theorem 15.17,
[n]
c1 OX ·ρn = M(−1).
n
310 16. RUAN’S COHOMOLOGICAL CREPANT RESOLUTION CONJECTURE
{k}
Comparing with Theorem 3.31, we see that the operator am (α) consists of
(k)
those terms in am (α) which do not contain the canonical divisor KX .
∗,T
Now let aTm (α), Hρ∗,T n
(X [n] ) and pTm (α), Horb (X (n) ) be the equivariant versions
of am (α), Hρ∗n (X [n] ) and pm (α), Horb∗
(X (n) ) respectively. Following [Cheo2, Sub-
section 3.1], there exists a ring isomorphism
∗,T
ΨTn : Horb (X (n) ) → Hρ∗,T
n
(X [n] )
defined by
√ n1 +...+ns −s T
ΨTn −1 p−n1 (α1 ) · · · pT−ns (αs )|0
= aT−n1 (α1 ) · · · aT−ns (αs )|0 .
Note that up to a scalar factor which depends only on the partition λ = (n1 , . . . , ns )
and the tuple −→
α = (α1 , . . . , αs ), our notation pT−n1 (α1 ) · · · pT−ns (αs )|0 coincides
with the notation λ −→
α used in [Cheo2]. Also, our notation aTm (α) coincides with
the notation pm (α) used in [Cheo2]. The integer n1 + . . . + ns − s is the age.
Passing the map ΨTn to the ordinary cohomology, we obtain a ring isomorphism
∗
Ψn : Horb (X (n) ) → Hρ∗n (X [n] )
with
√ n1 +...+ns −s
Ψn −1 p−n1 (α1 ) · · · p−ns (αs )|0
= a−n1 (α1 ) · · · a−ns (αs )|0 .
√ k
Using (10.8) and (16.3), we see that Ψn −1 Ok (α, n) = G k (α, n).
∗
Next, let A = a−n1 (α1 ) · · · a−ns (αs )|0 ∈ H (X [n−1]
). By definition,
[G k (α)a−1 (β)A − a−1 (β)G
k (α), a−1 (β)]A = G k (α)A
k (α, n) ·ρ a−1 (β)A − a−1 (β) G
= G k (α, n − 1) ·ρ A .
n n−1
Put P = p−n1 (α1 ) · · · p−ns (αs )|0 and a = n1 + . . . + ns − s. Let • denote the
orbifold ring product. Then,
" √ k √ a #
Ψn p−1 (β) −1 Ok (α, n − 1) • −1 P
"√ k √ a #
= a−1 (β)Ψn −1 Ok (α, n − 1) • −1 P
= k (α, n − 1) ·ρ
a−1 (β) G A ,
n−1
√ a
and Ψn k (α), a−1 (β)]A is equal to
−1 p−1 (β)P = a−1 (β)A. So [G
√ k √ a
Ψn −1 Ok (α, n) • −1 p−1 (β)P
√ k √ a
− Ψn p−1 (β) −1 Ok (α, n − 1) • −1 P .
Since Ok (α, n) • p−1 (β)P = Ok (α)p−1 (β)P , we obtain
√
k (α), a−1 (β)]A = −1k+a · Ψn [Ok (α), p−1 (β)]P .
[G
By Theorem 10.1 (iii), we conclude that [G k (α), a−1 (β)]A is equal to
⎛ ⎞
√ k+a ⎜ 1 s(λ) − 1 ⎟
−1 ·(−1)k ·Ψn ⎝ p (τ (αβ))P +
! λ ∗ !
pλ (τ∗ (eX αβ))P ⎠ .
(λ)=k+1
λ (λ)=k−1
24λ
|λ|=−1 |λ|=−1
312 16. RUAN’S COHOMOLOGICAL CREPANT RESOLUTION CONJECTURE
Therefore,
To prove (16.8) for every simply connected smooth projective surface X, for
simplicity, we put
w1 , w2 , w3 d = w1 , w2 , w3 0,dβn .
for all cohomology classes w1 ∈ Hρ∗n (X [n−1] ) = H ∗ (X [n−1] ) and w2 ∈ Hρ∗n (X [n] ) =
H ∗ (X [n] ). Put
(16.11) k (α; q), a−1 (β)]w1 , w2 − 1 a{k} (αβ)w1 , w2
Dβα (w1 , w2 ; q) = [G −1
k!
where we have omitted k in Dβα (w1 , w2 ; q) since it will be clear from the context.
The following lemma gives an explicit expression to Dβα (w1 , w2 ; q), which en-
ables us to analyze Dβα (w1 , w2 ; q).
(−1)|λ|−1 "
1−(n−j−1) a−λ (τ∗ α)|0 , a−1 (β)w1 , w2 d
λ! · |λ|!
0≤j≤k λ(j+1) d≥1
(λ)=k−j+1
#
− 1−(n−j−2) a−λ (τ∗ α)|0 , w1 , a−1 (β)† w2 d q d
+ f˜| | (λ) · aλ (τ∗ (αβ))w1 , w2
∈{KX ,KX
2 } (λ)=k+1−||/2
|λ|=−1
"
− g̃| | (λ) · 1−(n−j−1) a−λ (τ∗ (α))|0 , a−1 (β)w1 , w2
∈{KX ,K 2 } λ(j+1)
X (λ)=k−j+1−||/2
0≤j≤k
#
− 1−(n−j−2) a−λ (τ∗ (α))|0 , w1 , a−1 (β)† w2
where a−1 (β)† = −a1 (β) is the adjoint operator of a−1 (β), and the functions f˜| | (λ)
and g̃| | (λ) depend only on k, || and λ.
16.2. THE COMMUTATOR [G̃k (α), a−1 (β)] 313
Proof. By (16.5), [G k (α; q), a−1 (β)]w1 , w2 is equal to
(16.12) G k (α; q) a−1 (β)w1 , w2 − a−1 (β)G k (α; q)(w1 ), w2
= G k (α; q) a−1 (β)w1 , w2 − G k (α; q)(w1 ), a−1 (β)† w2
"
= G k (α, n), a−1 (β)w1 , w2
d
d≥0
# d
− G k (α, n − 1), w1 , a−1 (β)† w2 q .
d
Next, we study the two terms with d = 0 in (16.12). By (16.3) and Theo-
k (α, n) is equal to
rem 4.10, G
Gk (α, n) − g̃| | (λ) · 1−(n−j−1) a−λ (τ∗ (α))|0
∈{KX ,K 2 } λ(j+1)
X (λ)=k−j+1−||/2
0≤j≤k
Combining this with (16.12), (16.13) and (16.14), we prove the lemma.
In the following, the proof of (16.8) will be divided into three cases:
• α = x is the cohomology class of a point,
• |α| = 2,
• α = 1X .
We start with the simplest case when α = x and β is arbitrary.
Proposition 16.10. If α = x is the cohomology class of a point, then ( 16.8)
is true.
Proof. By Corollary 15.21 (ii), every term in Lemma 16.9 is equal to zero.
So Dβx (w1 , w2 ; q) = 0. In particular, Dβx (w1 , w2 ; −1) = 0. By (16.11),
Next, we deal with the case |α| = 2. We will study Dβα (w1 , w2 ; q) when |α| = 2.
We begin with two lemmas about the structures of the intersections in H ∗ (X [n] ).
Lemma 16.11. Let λ be a partition with |λ| ≤ n. For i = 1 and 2, let
(16.15) wi = a−λ(i) (x)a−μ(i) (1X )a−ni,1 (αi,1 ) · · · a−ni,ui (αi,ui )|0
where p(σ) depends only on σ, n, λ and all the λ(i) , μ(i) , ni,j .
Proof. By Remark 8.10 (ii-a), a−1 (1X )n−|λ| a−λ (x)|0 is a polynomial of the
classes Gk (x, n), k ≥ 0 whose coefficients are independent of X. In addition, the
integers k involved depend only on λ. Note that
Gk1 (x, n) · · · Gkl (x, n), w1 , w2
= Gk1 (x) · · · Gkl (x)w1 , w2
= a−λ(1) (x)a−n1,1 (α1,1 ) · · · a−n1,u1 (α1,u1 )Gk1 (x) · · · Gkl (x)a−μ(1) (1X )|0 , w2 .
So by Theorem 4.7 and Theorem 3.8, Gk1 (x, n) · · · Gkl (x, n), w1 , w2 is equal to
u1
(16.17) δu1 ,u2 · α1,i , α2,σ(i) · p̃(σ)
σ∈Perm{1,...,u1 } i=1
where p̃(σ) depends only on σ, n, k1 , . . . , kl and all the λ(i) , μ(i) , ni,j .
16.2. THE COMMUTATOR [G̃k (α), a−1 (β)] 315
u2 u1
+ α, α2,j · δu1 ,u2 −1 · α1,i , α2,σ3 (i) · p3 (σ3 )
j=1 σ3 i=1
where σ2 runs over all bijections {1, . . . , u1 } − {j} → {1, . . . , u2 }, σ3 runs over all
bijections {1, . . . , u1 } → {1, . . . , u2 } − {j}, and p1 (σ1 ) (respectively, p2 (σ2 ), p3 (σ3 ))
depend only on σ1 (respectively, σ2 , σ3 ), n, n0 , λ and all the λ(i) , μ(i) , ni,j .
Proof. By Remark 8.10 (ii-b), 1−(n−|λ|−n0 ) a−λ (x)a−n0 (α)|0 can be written
as
KX , α · F1 (n) + Gki (α, n) · F2,i (n)
i
where F1 (n) and F2,i (n) are polynomials of Gk (x, n), k ≥ 0 whose coefficients are
independent of n and α. Moreover, the integers k and ki depend only on λ and n0 .
Thus,
(16.19) 1−(n−|λ|−n0 ) a−λ (x)a−n0 (α)|0 , w1 , w2
= KX , α · F1 (n), w1 , w2 + Gki (α, n) · F2,i (n), w1 , w2 .
i
As in the proof of Lemma 16.11, F1 (n), w1 , w2 is of the form
u1
(16.20) δu1 ,u2 · α1,i , α2,σ1 (i) · p̃1,1 (σ1 )
σ1 ∈Perm{1,...,u1 } i=1
where p̃1,1 (σ1 ) depends only on σ1 , n, n0 , λ and all the λ(i) , μ(i) , ni,j . Also,
Gki (α, n)Gs1 (x, n) · · · Gsl (x, n), w1 , w2
= Gs1 (x) · · · Gsl (x)Gki (α)w1 , w2 .
By Theorem 4.7 and Lemma 3.12, Gki (α, n)Gs1 (x, n) · · · Gsl (x, n), w1 , w2 is equal
to
u1
(16.21) KX , α · δu1 ,u2 · α1,i , α2,σ1 (i) · p̃1,2 (σ1 )
σ1 ∈Perm{1,...,u1 } i=1
u1
+ α, α1,j · δu1 −1,u2 · α1,i , α2,σ2 (i) · p̃2 (σ2 )
j=1 σ2 i=j
u2 u1
+ α, α2,j · δu1 ,u2 −1 · α1,i , α2,σ3 (i) · p̃3 (σ3 )
j=1 σ3 i=1
316 16. RUAN’S COHOMOLOGICAL CREPANT RESOLUTION CONJECTURE
where σ2 runs over all the bijections {1, . . . , u1 } − {j} → {1, . . . , u2 }, and σ3 runs
over all the bijections {1, . . . , u1 } → {1, . . . , u2 } − {j}. Hence
Gki (α, n) · F2,i (n), w1 , w2
i
is of the form (16.21) as well. Combining with (16.19) and (16.20), we obtain
formula (16.18).
Now we introduce the notion of universal polynomials
P (KX , S1 , S2 )
in KX , KX of degree at most m and of type (u1 , u2 ), and prove a vanishing lemma.
Definition 16.13. Fix three integers m, u1 , u2 ≥ 0. Then a universal polyno-
mial P (KX , S1 , S2 ) in KX , KX of degree at most m and of type (u1 , u2 ) is of the
form
(16.22) KX , α1,i · KX , α2,i
1≤j1 <...<js ≤u1 i∈{j1 ,...,js } i∈{l1 ,...,ls }
1≤l1 <...<ls ≤u2
s
· α1,ji , α2,σ(li ) · p(j1 , . . . , js ; l1 , . . . , ls ; σ)
σ∈Perm{l1 ,...,ls } i=1
Our next lemma is about the structure of certain 3-pointed extremal Gromov-
Witten invariants, and provides the motivation for Definition 16.13.
Lemma 16.15. Let d, n0 ≥ 1 and |α| = 2. Let w1 and w2 be given by ( 16.15).
Then,
(16.24) 1−(n−|λ|−n0 ) a−λ (x)a−n0 (α)|0 , w1 , w2 d
= KX , α · P (KX , S1 , S2 )
where S1 = {α1,1 , . . . , α1,u1 }, S2 = {α2,1 , . . . , α2,u2 }, and P (KX , S1 , S2 ) is a uni-
versal polynomial in KX , KX of degree at most (n − n0 )/2 and of type (u1 , u2 ).
Proof. For simplicity, let
w0 = 1−(n−|λ|−n0 ) a−λ (x)a−n0 (α)|0 .
Also, for i = 1 and 2, let
i = a−μ(i) (1X )a−ni,1 (αi,1 ) · · · a−ni,ui (αi,ui )|0 .
w
We compute w0 , w1 , w2 d by using (15.62). Consider the following term from
(15.62):
F G
∗ w0 ∗ w1 ∗ w2
(16.25) B0 , B1 , B2 · Zm,d , πm,1 · πm,2 · πm,3
B0 B1 B2
where m ≤ n, B0 , B1 , B2 ∈ H ∗ (X [n−m] ), B0 ⊂ w0 , B1 ⊂ w1 , and B2 ⊂ w2 . By
Theorem 15.20 (i) and Corollary 15.21 (i), such a term is nonzero only if
B0 = a−1 (1X )j a−λ (x)|0 ,
B1 = a−λ(1) (x)B1 ,
B2 = 2
a−λ(2) (x)B
for some 1 ≤ j1 < . . . < js ≤ u1 and some sub-partition ν (1) of μ(1) (i.e., every part
of ν (1) is a part of μ(1) ). Similarly,
2 = a−ν (2) (1X )a−n (α2,l ) · · · a−n (α2,l )|0
B 2,l1 1 2,lt t
where p1 is a number independent of the surface X and the classes αi,j . By Theo-
rem 15.20, we see that the factor (16.27) is equal to
KX , α · KX , α1,i · KX , α2,i · p2 (j1 , . . . , js ; l1 , . . . , ls ; σ)
i∈{j1 ,...,js } i∈{l1 ,...,ls }
a−1 (1X )s a−λ̃ (x)a−n2 (γj,2 )|0 where s ≤ ñ. In the following, we assume that (16.34)
is nonzero. By symmetry, we need only to consider two cases for B0 :
B0 = a−1 (1X )s a−λ̃ (x)|0
or
B0 = a−1 (1X )s a−λ̃ (x)a−n1 (γj,1 )|0 .
We begin with the case B0 = a−1 (1X )s a−λ̃ (x)|0 . Then (16.34) becomes
1 , a−λ(2) (x)B
a−1 (1X )s a−λ̃ (x)|0 , a−λ(1) (x)B 2
j
F G
∗ 1−ñ a−n1 (γj,1 )a−n2 (γj,2 )|0 ∗ 1
w ∗ 2
w
· Zm,d , πm,1 · πm,2 · πm,3 .
a−1 (1X ) |0
s 1
B 2
B
Applying the same arguments as in the computations of (16.26) and (16.27), we
conclude that the term (16.34) is equal to
KX , γj,1 · KX , γj,2 · P3 (KX , S1 , S2 )
j
is defined by requiring
τk∗ (a), b1 ⊗ · · · ⊗ bk = T (ab1 · · · bk ).
Now the axiomatization in [QW1, Subsection 2.6] states that the algebra structure
on each A[n] in a sequence of graded Frobenius algebras A[n] (n ≥ 0) is determined
if
(A1) the direct sum
+∞
A[n]
n=0
affords the structure of the Fock space of a Heisenberg algebra modeled
on A which denotes A[1] .
(A2) There exists a sequence of elements G k (α, n) ∈ A[n] depending on α ∈ A
(linearly) and a non-negative integer k. Define the operators G k (α) on
+∞
A[n]
n=0
1 (1A ) = 1 3
(16.38) G − : a :0 (τ3∗ 1A ),
6
k (α), a−1 (β)] = 1 {k}
(16.39) [G a (αβ)
k! −1
where : a3 :0 is the zero mode in the normally ordered product : a3 :, and
{k} {0}
a−1 (α) denotes the k-th derivative with a−1 (α) = a−1 (α) and
{k} 1 (1A ), a{k−1} (α)]
a−1 (α) = [G −1
for k ≥ 1.
When (A1) and (A2) are satisfied, the algebra A[n] is generated by the elements
k (α, n) ∈ A[n] ,
G α ∈ A, k ≥ 0.
In addition, the product is determined by (16.38) and (16.39).
On one hand, with
∗
A[n] = Horb (X (n) )
∗
(viewed as an algebra over C) and the sequence of classes Ok (α, n) ∈ Horb (X (n) ), we
see from Theorem 10.1, Theorem 10.6 and Theorem 10.8 that (A1) and (A2) hold
∗
for the orbifold cohomology rings Horb (X (n) ). On the other hand, by Theorem 3.8,
Theorem 16.5 and Theorem 16.20, the quantum corrected cohomology rings
A[n] = Hρ∗n (X [n] )
also satisfy (A1) and (A2) with A = A[1] = H ∗ (X) and the sequence of classes
k (α, n) ∈ Hρ∗ (X [n] ). This leads to the following result.
G n
Proof. Recall the operators pn (α) from Section 10.2. Note that the shift
number (or the age) of p−n1 (α1 ) · · · p−ns (αs )|0 is equal to n1 + . . . + ns − s. Define
a linear isomorphism
+∞
+∞
∗
(16.40) Ψ: Horb (X (n) ) −→ Hρ∗n (X [n] )
n=0 n=0
by
"√ n +...+n −s #
−1 p−n1 (α1 ) · · · p−ns (αs )|0 = a−n1 (α1 ) · · · a−ns (αs )|0 .
1 s
Ψn
This induces a linear isomorphism
∗
Ψn : Horb (X (n) ) −→ Hρ∗n (X [n] ) = H ∗ (X [n] )
for each n. Moreover, Ψ1 is simply the identity map on the cohomology group of
the surface X.
By Theorem 16.5 and Theorem 16.20, the two formulas (16.38) and (16.39)
hold for A[n] = Hρ∗n (X [n] ). By the proof of Theorem 4.7, we have
1 s(λ) − 2
k (α) = −
G aλ (τ∗ α) + aλ (τ∗ (eX α)).
λ ! 24λ!
(λ)=k+2,|λ|=0 (λ)=k,|λ|=0
[AGV] D. Abramovich, T. Graber, and A. Vistoli, Algebraic orbifold quantum products, Orb-
ifolds in mathematics and physics (Madison, WI, 2001), Contemp. Math., vol. 310,
Amer. Math. Soc., Providence, RI, 2002, pp. 1–24, DOI 10.1090/conm/310/05397.
MR1950940
[AGT] L. F. Alday, D. Gaiotto, and Y. Tachikawa, Liouville correlation functions from
four-dimensional gauge theories, Lett. Math. Phys. 91 (2010), no. 2, 167–197, DOI
10.1007/s11005-010-0369-5. MR2586871
[AIK] A. B. Altman, A. Iarrobino, and S. L. Kleiman, Irreducibility of the compactified Jaco-
bian, Real and complex singularities (Proc. Ninth Nordic Summer School/NAVF Sym-
pos. Math., Oslo, 1976), Sijthoff and Noordhoff, Alphen aan den Rijn, 1977, pp. 1–12.
MR0498546
[And] G. E. Andrews, The theory of partitions, Addison-Wesley Publishing Co., Reading,
Mass.-London-Amsterdam, 1976. Encyclopedia of Mathematics and its Applications,
Vol. 2. MR0557013
[ACGH] E. Arbarello, M. Cornalba, P. A. Griffiths, and J. Harris, Geometry of algebraic curves.
Vol. I, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of
Mathematical Sciences], vol. 267, Springer-Verlag, New York, 1985. MR770932
[ABCH] D. Arcara, A. Bertram, I. Coskun, and J. Huizenga, The minimal model program for
the Hilbert scheme of points on P2 and Bridgeland stability, Adv. Math. 235 (2013),
580–626, DOI 10.1016/j.aim.2012.11.018. MR3010070
[AH] M. F. Atiyah and F. Hirzebruch, Vector bundles and homogeneous spaces, Proc. Sym-
pos. Pure Math., Vol. III, American Mathematical Society, Providence, R.I., 1961,
pp. 7–38. MR0139181
[BBC] D. H. Bailey, J. M. Borwein, and R. E. Crandall, Computation and theory of extended
Mordell-Tornheim-Witten sums, Math. Comp. 83 (2014), no. 288, 1795–1821, DOI
10.1090/S0025-5718-2014-02768-3. MR3194130
[BB] V. V. Batyrev and L. A. Borisov, Mirror duality and string-theoretic Hodge numbers,
Invent. Math. 126 (1996), no. 1, 183–203, DOI 10.1007/s002220050093. MR1408560
[BBM] P. Baum, J.-L. Brylinski, and R. MacPherson, Cohomologie équivariante délocalisée
(French, with English summary), C. R. Acad. Sci. Paris Sér. I Math. 300 (1985),
no. 17, 605–608. MR791098
[BC] P. Baum and A. Connes, Chern character for discrete groups, A fête of topology, Aca-
demic Press, Boston, MA, 1988, pp. 163–232, DOI 10.1016/B978-0-12-480440-1.50015-
0. MR928402
[Bea1] A. Beauville, Variétés Kähleriennes dont la première classe de Chern est nulle
(French), J. Differential Geom. 18 (1983), no. 4, 755–782 (1984). MR730926
[Bea2] , Variétés kählériennes compactes avec c1 = 0 (French), Astérisque 126
(1985), 181–192. Geometry of K3 surfaces: moduli and periods (Palaiseau, 1981/1982).
MR785234
[Beh1] K. Behrend, Gromov-Witten invariants in algebraic geometry, Invent. Math. 127
(1997), no. 3, 601–617, DOI 10.1007/s002220050132. MR1431140
[Beh2] , The product formula for Gromov-Witten invariants, J. Algebraic Geom. 8
(1999), no. 3, 529–541. MR1689355
[BF1] K. Behrend and B. Fantechi, The intrinsic normal cone, Invent. Math. 128 (1997),
no. 1, 45–88, DOI 10.1007/s002220050136. MR1437495
325
326 BIBLIOGRAPHY
[BF2] K. Behrend and B. Fantechi, Symmetric obstruction theories and Hilbert schemes
of points on threefolds, Algebra Number Theory 2 (2008), no. 3, 313–345, DOI
10.2140/ant.2008.2.313. MR2407118
[BSG] M. Beltrametti and A. J. Sommese, Zero cycles and kth order embeddings of smooth
projective surfaces, Problems in the theory of surfaces and their classification (Cortona,
1988), Sympos. Math., XXXII, Academic Press, London, 1991, pp. 33–48. With an
appendix by Lothar Göttsche. MR1273371
[BT] A. Bertram and M. Thaddeus, On the quantum cohomology of a symmetric product
of an algebraic curve, Duke Math. J. 108 (2001), no. 2, 329–362, DOI 10.1215/S0012-
7094-01-10825-9. MR1833394
[BO] S. Bloch and A. Okounkov, The character of the infinite wedge representation, Adv.
Math. 149 (2000), no. 1, 1–60, DOI 10.1006/aima.1999.1845. MR1742353
[Boi1] S. Boissière, Chern classes of the tangent bundle on the Hilbert scheme of points on
the affine plane, J. Algebraic Geom. 14 (2005), no. 4, 761–787, DOI 10.1090/S1056-
3911-05-00412-1. MR2147349
[Boi2] , On the McKay correspondences for the Hilbert scheme of points on the
affine plane, Math. Ann. 334 (2006), no. 2, 419–438, DOI 10.1007/s00208-005-0738-z.
MR2207706
[BN] S. Boissière and M. A. Nieper-Wisskirchen, Generating series in the cohomology of
Hilbert schemes of points on surfaces, LMS J. Comput. Math. 10 (2007), 254–270,
DOI 10.1112/S146115700000139X. MR2320831
[Bor] R. E. Borcherds, Vertex algebras, Kac-Moody algebras, and the Monster, Proc.
Nat. Acad. Sci. U.S.A. 83 (1986), no. 10, 3068–3071, DOI 10.1073/pnas.83.10.3068.
MR843307
[Bra1] D. M. Bradley, Multiple q-zeta values, J. Algebra 283 (2005), no. 2, 752–798, DOI
10.1016/j.jalgebra.2004.09.017. MR2111222
[Bra2] , On the sum formula for multiple q-zeta values, Rocky Mountain J. Math. 37
(2007), no. 5, 1427–1434, DOI 10.1216/rmjm/1194275927. MR2382894
[Bri] J. Briançon, Description de Hilbn C{x, y}, Invent. Math. 41 (1977), no. 1, 45–89, DOI
10.1007/BF01390164. MR0457432
[BG] J. Bryan and T. Graber, The crepant resolution conjecture, Algebraic geometry—
Seattle 2005. Part 1, Proc. Sympos. Pure Math., vol. 80, Amer. Math. Soc., Providence,
RI, 2009, pp. 23–42, DOI 10.1090/pspum/080.1/2483931. MR2483931
[BP] J. Bryan and R. Pandharipande, The local Gromov-Witten theory of curves, J. Amer.
Math. Soc. 21 (2008), no. 1, 101–136, DOI 10.1090/S0894-0347-06-00545-5. With an
appendix by Bryan, C. Faber, A. Okounkov and Pandharipande. MR2350052
[Can] M. B. Can, Nested Hilbert schemes and the nested q, t-Catalan series (English, with
English and French summaries), 20th Annual International Conference on Formal Power
Series and Algebraic Combinatorics (FPSAC 2008), Discrete Math. Theor. Comput.
Sci. Proc., AJ, Assoc. Discrete Math. Theor. Comput. Sci., Nancy, 2008, pp. 61–69.
MR2721442
[Car1] E. Carlsson, Vertex operators and moduli spaces of sheaves, ProQuest LLC, Ann Arbor,
MI, 2008. Thesis (Ph.D.)–Princeton University. MR2711527
[Car2] , Vertex operators and quasimodularity of Chern numbers on the Hilbert scheme,
Adv. Math. 229 (2012), no. 5, 2888–2907, DOI 10.1016/j.aim.2011.10.003. MR2889150
[CO] E. Carlsson and A. Okounkov, Exts and vertex operators, Duke Math. J. 161 (2012),
no. 9, 1797–1815, DOI 10.1215/00127094-1593380. MR2942794
[dCM] M. A. A. de Cataldo and L. Migliorini, The Douady space of a complex surface, Adv.
Math. 151 (2000), no. 2, 283–312, DOI 10.1006/aima.1999.1896. MR1758249
[CG] F. Catanese and L. Gœttsche, d-very-ample line bundles and embeddings of
Hilbert schemes of 0-cycles, Manuscripta Math. 68 (1990), no. 3, 337–341, DOI
10.1007/BF02568768. MR1065935
[CL] H.-l. Chang and J. Li, Semi-perfect obstruction theory and Donaldson-Thomas in-
variants of derived objects, Comm. Anal. Geom. 19 (2011), no. 4, 807–830, DOI
10.4310/CAG.2011.v19.n4.a6. MR2880216
[Che1] J. Cheah, On the cohomology of Hilbert schemes of points, J. Algebraic Geom. 5 (1996),
no. 3, 479–511. MR1382733
BIBLIOGRAPHY 327
[Che2] , Cellular decompositions for nested Hilbert schemes of points, Pacific J. Math.
183 (1998), no. 1, 39–90, DOI 10.2140/pjm.1998.183.39. MR1616606
[Che3] , The virtual Hodge polynomials of nested Hilbert schemes and related varieties,
Math. Z. 227 (1998), no. 3, 479–504, DOI 10.1007/PL00004387. MR1612677
[CR1] W. Chen and Y. Ruan, Orbifold Gromov-Witten theory, Orbifolds in mathematics and
physics (Madison, WI, 2001), Contemp. Math., vol. 310, Amer. Math. Soc., Providence,
RI, 2002, pp. 25–85, DOI 10.1090/conm/310/05398. MR1950941
[CR2] , A new cohomology theory of orbifold, Comm. Math. Phys. 248 (2004), no. 1,
1–31, DOI 10.1007/s00220-004-1089-4. MR2104605
[Cheo1] Wan Keng Cheong, Orbifold quantum cohomology of the symmetric product of Ar .
Preprint. arXiv:0910.0629
[Cheo2] W. K. Cheong, Strengthening the cohomological crepant resolution conjecture for
Hilbert-Chow morphisms, Math. Ann. 356 (2013), no. 1, 45–72, DOI 10.1007/s00208-
012-0833-x. MR3038121
[ChG] W. K. Cheong and A. Gholampour, Orbifold Gromov-Witten theory of the symmetric
product of Ar , Geom. Topol. 16 (2012), no. 1, 475–527, DOI 10.2140/gt.2012.16.475.
MR2916292
[Coa] T. Coates, On the crepant resolution conjecture in the local case, Comm. Math. Phys.
287 (2009), no. 3, 1071–1108, DOI 10.1007/s00220-008-0715-y. MR2486673
[CCIT] T. Coates, H. Iritani, and H.-H. Tseng, Wall-crossings in toric Gromov-Witten
theory. I. Crepant examples, Geom. Topol. 13 (2009), no. 5, 2675–2744, DOI
10.2140/gt.2009.13.2675. MR2529944
[CR] T. Coates and Y. Ruan, Quantum cohomology and crepant resolutions: a conjecture
(English, with English and French summaries), Ann. Inst. Fourier (Grenoble) 63 (2013),
no. 2, 431–478. MR3112518
[Cos] K. Costello, Higher genus Gromov-Witten invariants as genus zero invariants of sym-
metric products, Ann. of Math. (2) 164 (2006), no. 2, 561–601, DOI 10.4007/an-
nals.2006.164.561. MR2247968
[CoG] K. Costello, I. Grojnowski, Hilbert schemes, Hecke algebras and the Calogero-
Sutherland system. Preprint. math.AG/0310189.
[CK] D. A. Cox and S. Katz, Mirror symmetry and algebraic geometry, Mathematical Sur-
veys and Monographs, vol. 68, American Mathematical Society, Providence, RI, 1999.
MR1677117
[DHVW] L. Dixon, J. A. Harvey, C. Vafa, and E. Witten, Strings on orbifolds, Nuclear Phys. B
261 (1985), no. 4, 678–686, DOI 10.1016/0550-3213(85)90593-0. MR818423
[ELQ] D. Edidin, W.-P. Li, and Z. Qin, Gromov-Witten invariants of the Hilbert scheme of 3-
points on P2 , Asian J. Math. 7 (2003), no. 4, 551–574, DOI 10.4310/AJM.2003.v7.n4.a7.
MR2074891
[EG] G. Ellingsrud and L. Göttsche, Hilbert schemes of points and Heisenberg algebras,
School on Algebraic Geometry (Trieste, 1999), ICTP Lect. Notes, vol. 1, Abdus Salam
Int. Cent. Theoret. Phys., Trieste, 2000, pp. 59–100. MR1795861
[EGL] G. Ellingsrud, L. Göttsche, and M. Lehn, On the cobordism class of the Hilbert scheme
of a surface, J. Algebraic Geom. 10 (2001), no. 1, 81–100. MR1795551
[EL] G. Ellingsrud and M. Lehn, Irreducibility of the punctual quotient scheme of a surface,
Ark. Mat. 37 (1999), no. 2, 245–254, DOI 10.1007/BF02412213. MR1714770
[ES1] G. Ellingsrud and S. A. Strømme, On the homology of the Hilbert scheme of points
in the plane, Invent. Math. 87 (1987), no. 2, 343–352, DOI 10.1007/BF01389419.
MR870732
[ES2] , Towards the Chow ring of the Hilbert scheme of P2 , J. Reine Angew. Math.
441 (1993), 33–44. MR1228610
[ES3] , An intersection number for the punctual Hilbert scheme of a surface, Trans.
Amer. Math. Soc. 350 (1998), no. 6, 2547–2552, DOI 10.1090/S0002-9947-98-01972-2.
MR1432198
[EtG] P. Etingof and V. Ginzburg, Symplectic reflection algebras, Calogero-Moser space, and
deformed Harish-Chandra homomorphism, Invent. Math. 147 (2002), no. 2, 243–348,
DOI 10.1007/s002220100171. MR1881922
[FPa] C. Faber and R. Pandharipande, Hodge integrals and Gromov-Witten theory, Invent.
Math. 139 (2000), no. 1, 173–199, DOI 10.1007/s002229900028. MR1728879
328 BIBLIOGRAPHY
[FG1] B. Fantechi and L. Göttsche, Orbifold cohomology for global quotients, Duke Math. J.
117 (2003), no. 2, 197–227, DOI 10.1215/S0012-7094-03-11721-4. MR1971293
[FG2] , Local properties and Hilbert schemes of points, Fundamental algebraic ge-
ometry, Math. Surveys Monogr., vol. 123, Amer. Math. Soc., Providence, RI, 2005,
pp. 139–178. MR2223408
[Fog1] J. Fogarty, Algebraic families on an algebraic surface, Amer. J. Math 90 (1968), 511–
521, DOI 10.2307/2373541. MR0237496
[Fog2] J. Fogarty, Algebraic families on an algebraic surface. II. The Picard scheme of the
punctual Hilbert scheme, Amer. J. Math. 95 (1973), 660–687, DOI 10.2307/2373734.
MR0335512
[FU] D. S. Freed and K. K. Uhlenbeck, Instantons and four-manifolds, 2nd ed., Mathemat-
ical Sciences Research Institute Publications, vol. 1, Springer-Verlag, New York, 1991.
MR1081321
[FB] E. Frenkel and D. Ben-Zvi, Vertex algebras and algebraic curves, Mathematical Sur-
veys and Monographs, vol. 88, American Mathematical Society, Providence, RI, 2001.
MR1849359
[FKRW] E. Frenkel, V. Kac, A. Radul, and W. Wang, W1+∞ and W(glN ) with central charge
N , Comm. Math. Phys. 170 (1995), no. 2, 337–357. MR1334399
[FJW] I. B. Frenkel, N. Jing, and W. Wang, Vertex representations via finite groups and
the McKay correspondence, Internat. Math. Res. Notices 4 (2000), 195–222, DOI
10.1155/S107379280000012X. MR1747618
[FW] I. B. Frenkel and W. Wang, Virasoro algebra and wreath product convolution, J. Algebra
242 (2001), no. 2, 656–671, DOI 10.1006/jabr.2001.8860. MR1848963
[Fu] Y. Fu, Quantum cohomology of a Hilbert scheme of a Hirzebruch surface, ProQuest
LLC, Ann Arbor, MI, 2010. Thesis (Ph.D.)–University of Illinois at Urbana-Champaign.
MR2941780
[FN] A. Fujiki and S. Nakano, Supplement to “On the inverse of monoidal transformation”,
Publ. Res. Inst. Math. Sci. 7 (1971/72), 637–644, DOI 10.2977/prims/1195193401.
MR0294712
[FO] K. Fukaya and K. Ono, Arnold conjecture and Gromov-Witten invariant for general
symplectic manifolds, The Arnoldfest (Toronto, ON, 1997), Fields Inst. Commun.,
vol. 24, Amer. Math. Soc., Providence, RI, 1999, pp. 173–190. MR1733575
[Ful] W. Fulton, Introduction to toric varieties, Annals of Mathematics Studies, vol. 131,
Princeton University Press, Princeton, NJ, 1993. The William H. Roever Lectures in
Geometry. MR1234037
[FP] W. Fulton and R. Pandharipande, Notes on stable maps and quantum cohomology,
Algebraic geometry—Santa Cruz 1995, Proc. Sympos. Pure Math., vol. 62, Amer.
Math. Soc., Providence, RI, 1997, pp. 45–96, DOI 10.1090/pspum/062.2/1492534.
MR1492534
[Get] E. Getzler, Intersection theory on M1,4 and elliptic Gromov-Witten invariants, J.
Amer. Math. Soc. 10 (1997), no. 4, 973–998, DOI 10.1090/S0894-0347-97-00246-4.
MR1451505
[GSY1] A. Gholampour, A. Sheshmani, S.-T. Yau, Nested Hilbert schemes on surfaces: virtual
fundamental class. Preprint. arXiv.1701.08899.
[GSY2] A. Gholampour, A. Sheshmani, S.-T. Yau, Localized Donaldson-Thomas theory of sur-
faces. Preprint. arXiv.1701.08902.
[Got1] L. Göttsche, The Betti numbers of the Hilbert scheme of points on a smooth pro-
jective surface, Math. Ann. 286 (1990), no. 1-3, 193–207, DOI 10.1007/BF01453572.
MR1032930
[Got2] , Hilbert schemes of zero-dimensional subschemes of smooth varieties, Lecture
Notes in Mathematics, vol. 1572, Springer-Verlag, Berlin, 1994. MR1312161
[Got3] , On the motive of the Hilbert scheme of points on a surface, Math. Res. Lett.
8 (2001), no. 5-6, 613–627, DOI 10.4310/MRL.2001.v8.n5.a3. MR1879805
[GS] L. Göttsche and W. Soergel, Perverse sheaves and the cohomology of Hilbert
schemes of smooth algebraic surfaces, Math. Ann. 296 (1993), no. 2, 235–245, DOI
10.1007/BF01445104. MR1219901
BIBLIOGRAPHY 329
[Gou] I. P. Goulden, A differential operator for symmetric functions and the combinatorics of
multiplying transpositions, Trans. Amer. Math. Soc. 344 (1994), no. 1, 421–440, DOI
10.2307/2154724. MR1249468
[GJV] I. Goulden, D. Jackson, R. Vakil, Towards the geometry of double Hurwitz numbers,
math.AG/0309440.
[GP] T. Graber and R. Pandharipande, Localization of virtual classes, Invent. Math. 135
(1999), no. 2, 487–518, DOI 10.1007/s002220050293. MR1666787
[Groj] I. Grojnowski, Instantons and affine algebras. I. The Hilbert scheme and vertex oper-
ators, Math. Res. Lett. 3 (1996), no. 2, 275–291, DOI 10.4310/MRL.1996.v3.n2.a12.
MR1386846
[Grot] A. Grothendieck, Techniques de construction et théorèmes d’existence en géométrie
algébrique. IV. Les schémas de Hilbert (French), Séminaire Bourbaki, Vol. 6, Soc.
Math. France, Paris, 1995, pp. Exp. No. 221, 249–276. MR1611822
[Hai1] M. Haiman, Hilbert schemes, polygraphs and the Macdonald positivity conjecture, J.
Amer. Math. Soc. 14 (2001), no. 4, 941–1006, DOI 10.1090/S0894-0347-01-00373-3.
MR1839919
[Hai2] , Vanishing theorems and character formulas for the Hilbert scheme of points
in the plane, Invent. Math. 149 (2002), no. 2, 371–407, DOI 10.1007/s002220200219.
MR1918676
[Hai3] , Combinatorics, symmetric functions, and Hilbert schemes, Current develop-
ments in mathematics, 2002, Int. Press, Somerville, MA, 2003, pp. 39–111. MR2051783
[Harr] J. Harris, Algebraic geometry, Graduate Texts in Mathematics, vol. 133, Springer-
Verlag, New York, 1992. A first course. MR1182558
[Hart] R. Hartshorne, Algebraic geometry, Springer-Verlag, New York-Heidelberg, 1977. Grad-
uate Texts in Mathematics, No. 52. MR0463157
[Hern] D. Hernandez, An introduction to affine Kac-Moody algebras. DEA. Lecture notes from
CTQM Master Class, Aarhus University, Denmark, October 2006.
[HLQ] J. Hu, W.-P. Li, and Z. Qin, The Gromov-Witten invariants of the Hilbert schemes of
points on surfaces with pg > 0, Internat. J. Math. 26 (2015), no. 1, 1550009, 26, DOI
10.1142/S0129167X15500093. MR3313654
[Iar1] A. A. Iarrobino, Punctual Hilbert schemes, Mem. Amer. Math. Soc. 10 (1977), no. 188,
viii+112, DOI 10.1090/memo/0188. MR0485867
[Iar2] A. Iarrobino, Hilbert scheme of points: overview of last ten years, Algebraic geometry,
Bowdoin, 1985 (Brunswick, Maine, 1985), Proc. Sympos. Pure Math., vol. 46, Amer.
Math. Soc., Providence, RI, 1987, pp. 297–320. MR927986
[IN] Y. Ito and I. Nakamura, McKay correspondence and Hilbert schemes, Proc. Japan
Acad. Ser. A Math. Sci. 72 (1996), no. 7, 135–138. MR1420598
[Juc] A.-A. A. Jucys, Symmetric polynomials and the center of the symmetric group ring,
Rep. Mathematical Phys. 5 (1974), no. 1, 107–112. MR0419576
[Kac] V. Kac, Vertex algebras for beginners, 2nd ed., University Lecture Series, vol. 10, Amer-
ican Mathematical Society, Providence, RI, 1998. MR1651389
[Kap] S. Kapfer, Computing cup products in integral cohomology of Hilbert schemes of
points on K3 surfaces, LMS J. Comput. Math. 19 (2016), no. 1, 78–97, DOI
10.1112/S1461157016000012. MR3542518
[KaM] S. Kapfer, G. Menet, Integral cohomology of the generalized Kummer fourfold. Preprint.
[KMM] Y. Kawamata, K. Matsuda, and K. Matsuki, Introduction to the minimal model prob-
lem, Algebraic geometry, Sendai, 1985, Adv. Stud. Pure Math., vol. 10, North-Holland,
Amsterdam, 1987, pp. 283–360. MR946243
[KL1] Y. Kiem, J. Li, Gromov-Witten invariants of varieties with holomorphic 2-forms.
Preprint. arXiv:0707.2986
[KL2] Y.-H. Kiem and J. Li, Localizing virtual cycles by cosections, J. Amer. Math. Soc. 26
(2013), no. 4, 1025–1050, DOI 10.1090/S0894-0347-2013-00768-7. MR3073883
[King] A. D. King, Moduli of representations of finite-dimensional algebras, Quart. J. Math.
Oxford Ser. (2) 45 (1994), no. 180, 515–530, DOI 10.1093/qmath/45.4.515. MR1315461
[KM] M. Kontsevich and Yu. Manin, Gromov-Witten classes, quantum cohomology, and enu-
merative geometry, Comm. Math. Phys. 164 (1994), no. 3, 525–562. MR1291244
330 BIBLIOGRAPHY
[LQW4] , Hilbert schemes and W algebras, Int. Math. Res. Not. 27 (2002), 1427–1456,
DOI 10.1155/S1073792802110129. MR1908477
[LQW5] , Ideals of the cohomology rings of Hilbert schemes and their applications, Trans.
Amer. Math. Soc. 356 (2004), no. 1, 245–265, DOI 10.1090/S0002-9947-03-03422-6.
MR2020031
[LQW6] , Hilbert schemes, integrable hierarchies, and Gromov-Witten theory, Int. Math.
Res. Not. 40 (2004), 2085–2104, DOI 10.1155/S1073792804140452. MR2064317
[LQW7] , The cohomology rings of Hilbert schemes via Jack polynomials, Algebraic
structures and moduli spaces, CRM Proc. Lecture Notes, vol. 38, Amer. Math. Soc.,
Providence, RI, 2004, pp. 249–258. MR2096149
[LQW8] , Hilbert scheme intersection numbers, Hurwitz numbers, and Gromov-Witten
invariants, Infinite-dimensional aspects of representation theory and applications, Con-
temp. Math., vol. 392, Amer. Math. Soc., Providence, RI, 2005, pp. 67–81, DOI
10.1090/conm/392/07354. MR2189871
[LQZ] W.-P. Li, Z. Qin, and Q. Zhang, Curves in the Hilbert schemes of points on
surfaces, Vector bundles and representation theory (Columbia, MO, 2002), Con-
temp. Math., vol. 322, Amer. Math. Soc., Providence, RI, 2003, pp. 89–96, DOI
10.1090/conm/322/05681. MR1987741
[Mac1] I. G. Macdonald, The Poincaré polynomial of a symmetric product, Proc. Cambridge
Philos. Soc. 58 (1962), 563–568. MR0143204
[Mac2] , Symmetric functions and Hall polynomials, 2nd ed., Oxford Mathematical
Monographs, The Clarendon Press, Oxford University Press, New York, 1995. With
contributions by A. Zelevinsky; Oxford Science Publications. MR1354144
[Mar1] E. Markman, Generators of the cohomology ring of moduli spaces of sheaves on sym-
plectic surfaces, J. Reine Angew. Math. 544 (2002), 61–82, DOI 10.1515/crll.2002.028.
MR1887889
[Mar2] , Integral generators for the cohomology ring of moduli spaces of
sheaves over Poisson surfaces, Adv. Math. 208 (2007), no. 2, 622–646, DOI
10.1016/j.aim.2006.03.006. MR2304330
[Mat] Y. Matsuo, Matrix theory, Hilbert scheme and integrable system, Modern Phys. Lett.
A 13 (1998), no. 34, 2731–2742, DOI 10.1142/S0217732398002904. MR1660859
[Mau] D. Maulik, Gromov-Witten theory of An -resolutions, Geom. Topol. 13 (2009), no. 3,
1729–1773, DOI 10.2140/gt.2009.13.1729. MR2496055
[MNOP1] D. Maulik, N. Nekrasov, A. Okounkov, and R. Pandharipande, Gromov-Witten theory
and Donaldson-Thomas theory. I, Compos. Math. 142 (2006), no. 5, 1263–1285, DOI
10.1112/S0010437X06002302. MR2264664
[MNOP2] , Gromov-Witten theory and Donaldson-Thomas theory. II, Compos. Math. 142
(2006), no. 5, 1286–1304, DOI 10.1112/S0010437X06002314. MR2264665
[MO1] D. Maulik and A. Oblomkov, Quantum cohomology of the Hilbert scheme of points on
An -resolutions, J. Amer. Math. Soc. 22 (2009), no. 4, 1055–1091, DOI 10.1090/S0894-
0347-09-00632-8. MR2525779
[MO2] , Donaldson-Thomas theory of An × P 1 , Compos. Math. 145 (2009), no. 5,
1249–1276, DOI 10.1112/S0010437X09003972. MR2551996
[MJD] T. Miwa, M. Jimbo, and E. Date, Solitons, Cambridge Tracts in Mathematics, vol. 135,
Cambridge University Press, Cambridge, 2000. Differential equations, symmetries and
infinite-dimensional algebras; Translated from the 1993 Japanese original by Miles Reid.
MR1736222
[Mod] L. J. Mordell, On the evaluation of some multiple series, J. London Math. Soc. 33
(1958), 368–371, DOI 10.1112/jlms/s1-33.3.368. MR0100181
[Mum] D. Mumford, Towards an enumerative geometry of the moduli space of curves, Arith-
metic and geometry, Vol. II, Progr. Math., vol. 36, Birkhäuser Boston, Boston, MA,
1983, pp. 271–328. MR717614
[MF] D. Mumford and J. Fogarty, Geometric invariant theory, 2nd ed., Ergebnisse der Math-
ematik und ihrer Grenzgebiete [Results in Mathematics and Related Areas], vol. 34,
Springer-Verlag, Berlin, 1982. MR719371
[Mu] G. E. Murphy, A new construction of Young’s seminormal representation of the sym-
metric groups, J. Algebra 69 (1981), no. 2, 287–297, DOI 10.1016/0021-8693(81)90205-
2. MR617079
332 BIBLIOGRAPHY
[Mur] R. Murphy, A transition matrix for two bases of the integral cohomology of the Hilbert
scheme of points in the projective plane, ProQuest LLC, Ann Arbor, MI, 2010. Thesis
(Ph.D.)–University of Missouri - Columbia. MR2982352
[Nag] K. Nagao, Quiver varieties and Frenkel-Kac construction, J. Algebra 321 (2009),
no. 12, 3764–3789, DOI 10.1016/j.jalgebra.2009.03.012. MR2517812
[Nak1] H. Nakajima, Instantons on ALE spaces, quiver varieties, and Kac-Moody alge-
bras, Duke Math. J. 76 (1994), no. 2, 365–416, DOI 10.1215/S0012-7094-94-07613-8.
MR1302318
[Nak2] , Gauge theory on resolutions of simple singularities and simple Lie alge-
bras, Internat. Math. Res. Notices 2 (1994), 61–74, DOI 10.1155/S1073792894000085.
MR1264929
[Nak3] , Heisenberg algebra and Hilbert schemes of points on projective surfaces, Ann.
of Math. (2) 145 (1997), no. 2, 379–388, DOI 10.2307/2951818. MR1441880
[Nak4] , Quiver varieties and Kac-Moody algebras, Duke Math. J. 91 (1998), no. 3,
515–560, DOI 10.1215/S0012-7094-98-09120-7. MR1604167
[Nak5] , Lectures on Hilbert schemes of points on surfaces, University Lecture Series,
vol. 18, American Mathematical Society, Providence, RI, 1999. MR1711344
[Nak6] , The Heisenberg algebra and Hilbert schemes of points on surfaces (Japanese),
Sūgaku 50 (1998), no. 4, 385–398. MR1690690
[Nak7] , More lectures on Hilbert schemes of points on surfaces, Development of moduli
theory—Kyoto 2013, Adv. Stud. Pure Math., vol. 69, Math. Soc. Japan, [Tokyo], 2016,
pp. 173–205. MR3586508
[Naka] S. Nakano, On the inverse of monoidal transformation, Publ. Res. Inst. Math. Sci. 6
(1970/71), 483–502, DOI 10.2977/prims/1195193917. MR0294710
[Neg] A. Neguţ, Exts and the AGT relations, Lett. Math. Phys. 106 (2016), no. 9, 1265–1316,
DOI 10.1007/s11005-016-0865-3. MR3533570
[NO] N. A. Nekrasov and A. Okounkov, Seiberg-Witten theory and random partitions, The
unity of mathematics, Progr. Math., vol. 244, Birkhäuser Boston, Boston, MA, 2006,
pp. 525–596, DOI 10.1007/0-8176-4467-9 15. MR2181816
[Nie1] M. A. Nieper-Wisskirchen, Characteristic classes of the Hilbert schemes of points on
non-compact simply-connected surfaces, JP J. Geom. Topol. 8 (2008), no. 1, 7–21.
MR2444822
[Nie2] , Twisted cohomology of the Hilbert schemes of points on surfaces, Doc. Math.
14 (2009), 749–770. MR2578804
[Obe] G. Oberdieck, Gromov-Witten invariants of the Hilbert scheme of points of a K3 sur-
face. Preprint. arXiv:1406.1139
[Ok1] A. Okounkov, Toda equations for Hurwitz numbers, Math. Res. Lett. 7 (2000), no. 4,
447–453, DOI 10.4310/MRL.2000.v7.n4.a10. MR1783622
[Ok2] A. Yu. Okunkov, Hilbert schemes and multiple q-zeta values (Russian, with Russian
summary), Funktsional. Anal. i Prilozhen. 48 (2014), no. 2, 79–87; English transl.,
Funct. Anal. Appl. 48 (2014), no. 2, 138–144. MR3288178
[OP1] A. Okounkov and R. Pandharipande, Gromov-Witten theory, Hurwitz theory, and
completed cycles, Ann. of Math. (2) 163 (2006), no. 2, 517–560, DOI 10.4007/an-
nals.2006.163.517. MR2199225
[OP2] , The equivariant Gromov-Witten theory of P1 , Ann. of Math. (2) 163 (2006),
no. 2, 561–605, DOI 10.4007/annals.2006.163.561. MR2199226
[OP3] , Quantum cohomology of the Hilbert scheme of points in the plane, Invent.
Math. 179 (2010), no. 3, 523–557, DOI 10.1007/s00222-009-0223-5. MR2587340
[OP4] , The quantum differential equation of the Hilbert scheme of points in the
plane, Transform. Groups 15 (2010), no. 4, 965–982, DOI 10.1007/s00031-010-9116-
3. MR2753265
[OP5] , The local Donaldson-Thomas theory of curves, Geom. Topol. 14 (2010), no. 3,
1503–1567, DOI 10.2140/gt.2010.14.1503. MR2679579
[OT] J.-i. Okuda and Y. Takeyama, On relations for the multiple q-zeta values, Ramanujan
J. 14 (2007), no. 3, 379–387, DOI 10.1007/s11139-007-9053-5. MR2357443
[PTse] R. Pandharipande, H.-H. Tseng, Higher genus Gromov-Witten theory of the Hilbert
scheme of points of the plane and CohFTs associated to local curves. Preprint.
arXiv:1707.01406
BIBLIOGRAPHY 333
[QT] Z. Qin and Y. Tu, The nef cones of and minimal-degree curves in the Hilbert schemes
of points on certain surfaces, Pacific J. Math. 284 (2016), no. 2, 439–453, DOI
10.2140/pjm.2016.284.439. MR3544309
[QW1] Z. Qin and W. Wang, Hilbert schemes and symmetric products: a dictionary, Orb-
ifolds in mathematics and physics (Madison, WI, 2001), Contemp. Math., vol. 310,
Amer. Math. Soc., Providence, RI, 2002, pp. 233–257, DOI 10.1090/conm/310/05407.
MR1950950
[QW2] , Integral operators and integral cohomology classes of Hilbert schemes, Math.
Ann. 331 (2005), no. 3, 669–692, DOI 10.1007/s00208-004-0602-6. MR2122545
[QW3] , Hilbert schemes of points on the minimal resolution and soliton equations, Lie
algebras, vertex operator algebras and their applications, Contemp. Math., vol. 442,
Amer. Math. Soc., Providence, RI, 2007, pp. 435–462, DOI 10.1090/conm/442/08541.
MR2372578
[QY] Z. Qin, F. Yu, On Okounkov’s conjecture connecting Hilbert schemes of points and mul-
tiple q-zeta values. Intern. Math. Res. Notices (to appear). Published online: December
16, 2016.
[Reid] M. Reid, La correspondance de McKay, Astérisque 276 (2002), 53–72. Séminaire Bour-
baki, Vol. 1999/2000. MR1886756
[Rua1] Y. Ruan, Topological sigma model and Donaldson-type invariants in Gromov the-
ory, Duke Math. J. 83 (1996), no. 2, 461–500, DOI 10.1215/S0012-7094-96-08316-7.
MR1390655
[Rua2] , The cohomology ring of crepant resolutions of orbifolds, Gromov-Witten theory
of spin curves and orbifolds, Contemp. Math., vol. 403, Amer. Math. Soc., Providence,
RI, 2006, pp. 117–126, DOI 10.1090/conm/403/07597. MR2234886
[Seg] G. Segal, Equivariant K-theory and symmetric products. Lecture at Cambridge, July
1996.
[Sim] C. T. Simpson, Moduli of representations of the fundamental group of a smooth pro-
jective variety. I, Inst. Hautes Études Sci. Publ. Math. 79 (1994), 47–129. MR1307297
[Tik] A. S. Tikhomirov, Standard bundles on a Hilbert scheme of points on a surface, Al-
gebraic geometry and its applications (Yaroslavl, 1992), Aspects Math., E25, Friedr.
Vieweg, Braunschweig, 1994, pp. 183–203. MR1282029
[Tor] L. Tornheim, Harmonic double series, Amer. J. Math. 72 (1950), 303–314, DOI
10.2307/2372034. MR0034860
[Tse] H.-H. Tseng, A tale of four theories. Talk at the conference “Geometry of Moduli
Spaces”. University of California at San Diego, May 2017.
[UT] K. Ueno and K. Takasaki, Toda lattice hierarchy, Group representations and systems
of differential equations (Tokyo, 1982), Adv. Stud. Pure Math., vol. 4, North-Holland,
Amsterdam, 1984, pp. 1–95. MR810623
[Uri] B. Uribe, Orbifold cohomology of the symmetric product, Comm. Anal. Geom. 13
(2005), no. 1, 113–128. MR2154668
[VW] C. Vafa and E. Witten, A strong coupling test of S-duality, Nuclear Phys. B 431 (1994),
no. 1-2, 3–77, DOI 10.1016/0550-3213(94)90097-3. MR1305096
[Vas] E. Vasserot, Sur l’anneau de cohomologie du schéma de Hilbert de C2 (French, with
English and French summaries), C. R. Acad. Sci. Paris Sér. I Math. 332 (2001), no. 1,
7–12, DOI 10.1016/S0764-4442(00)01766-3. MR1805619
[Wan1] W. Wang, Equivariant K-theory, wreath products, and Heisenberg algebra, Duke Math.
J. 103 (2000), no. 1, 1–23, DOI 10.1215/S0012-7094-00-10311-0. MR1758236
[Wan2] , Algebraic structures behind Hilbert schemes and wreath products, Recent devel-
opments in infinite-dimensional Lie algebras and conformal field theory (Charlottesville,
VA, 2000), Contemp. Math., vol. 297, Amer. Math. Soc., Providence, RI, 2002, pp. 271–
295, DOI 10.1090/conm/297/05102. MR1919822
[Wan3] , The Farahat-Higman ring of wreath products and Hilbert schemes, Adv. Math.
187 (2004), no. 2, 417–446, DOI 10.1016/j.aim.2003.09.003. MR2078343
[Wan4] , Vertex algebras and the class algebras of wreath products, Proc. London Math.
Soc. (3) 88 (2004), no. 2, 381–404, DOI 10.1112/S0024611503014382. MR2032512
[Wit] E. Witten, On quantum gauge theories in two dimensions, Comm. Math. Phys. 141
(1991), no. 1, 153–209. MR1133264
334 BIBLIOGRAPHY
[Zas] E. Zaslow, Topological orbifold models and quantum cohomology rings, Comm. Math.
Phys. 156 (1993), no. 2, 301–331. MR1233848
[Zhao] J. Zhao, Uniform approach to double shuffle and duality relations of various q-analogs
of multiple zeta values via Rota-Baxter algebras. Preprint. arXiv:1412.8044
[Zud] V. V. Zudilin, Algebraic relations for multiple zeta values (Russian, with
Russian summary), Uspekhi Mat. Nauk 58 (2003), no. 1(349), 3–32, DOI
10.1070/RM2003v058n01ABEH000592; English transl., Russian Math. Surveys 58
(2003), no. 1, 1–29. MR1992130
Index
335
336 INDEX
nef, 16
nef and big, 16
normally ordered product, 61, 93
partition, 3
partition, d-dimensional, 3
partition, generalized, 4
Poincaré polynomial, 7
power symmetric function, 6
S-property, 171
Schur function, 6
stable map, 232
standard decomposition, 256
symmetric product, 7
tautological bundle, 15
transfer properties, 77
translation operator, 130
weight, 100
Young diagram, 4
Selected Published Titles in This Series
228 Zhenbo Qin, Hilbert Schemes of Points and Infinite Dimensional Lie Algebras, 2018
227 Roberto Frigerio, Bounded Cohomology of Discrete Groups, 2017
226 Marcelo Aguiar and Swapneel Mahajan, Topics in Hyperplane Arrangements, 2017
225 Mario Bonk and Daniel Meyer, Expanding Thurston Maps, 2017
224 Ruy Exel, Partial Dynamical Systems, Fell Bundles and Applications, 2017
223 Guillaume Aubrun and Stanislaw J. Szarek, Alice and Bob Meet Banach, 2017
222 Alexandru Buium, Foundations of Arithmetic Differential Geometry, 2017
221 Dennis Gaitsgory and Nick Rozenblyum, A Study in Derived Algebraic Geometry,
2017
220 A. Shen, V. A. Uspensky, and N. Vereshchagin, Kolmogorov Complexity and
Algorithmic Randomness, 2017
219 Richard Evan Schwartz, The Projective Heat Map, 2017
218 Tushar Das, David Simmons, and Mariusz Urbański, Geometry and Dynamics in
Gromov Hyperbolic Metric Spaces, 2017
217 Benoit Fresse, Homotopy of Operads and Grothendieck–Teichmüller Groups, 2017
216 Frederick W. Gehring, Gaven J. Martin, and Bruce P. Palka, An Introduction to
the Theory of Higher-Dimensional Quasiconformal Mappings, 2017
215 Robert Bieri and Ralph Strebel, On Groups of PL-homeomorphisms of the Real Line,
2016
214 Jared Speck, Shock Formation in Small-Data Solutions to 3D Quasilinear Wave
Equations, 2016
213 Harold G. Diamond and Wen-Bin Zhang (Cheung Man Ping), Beurling
Generalized Numbers, 2016
212 Pandelis Dodos and Vassilis Kanellopoulos, Ramsey Theory for Product Spaces,
2016
211 Charlotte Hardouin, Jacques Sauloy, and Michael F. Singer, Galois Theories of
Linear Difference Equations: An Introduction, 2016
210 Jason P. Bell, Dragos Ghioca, and Thomas J. Tucker, The Dynamical
Mordell–Lang Conjecture, 2016
209 Steve Y. Oudot, Persistence Theory: From Quiver Representations to Data Analysis,
2015
208 Peter S. Ozsváth, András I. Stipsicz, and Zoltán Szabó, Grid Homology for Knots
and Links, 2015
207 Vladimir I. Bogachev, Nicolai V. Krylov, Michael Röckner, and Stanislav V.
Shaposhnikov, Fokker–Planck–Kolmogorov Equations, 2015
206 Bennett Chow, Sun-Chin Chu, David Glickenstein, Christine Guenther, James
Isenberg, Tom Ivey, Dan Knopf, Peng Lu, Feng Luo, and Lei Ni, The Ricci Flow:
Techniques and Applications: Part IV: Long-Time Solutions and Related Topics, 2015
205 Pavel Etingof, Shlomo Gelaki, Dmitri Nikshych, and Victor Ostrik, Tensor
Categories, 2015
204 Victor M. Buchstaber and Taras E. Panov, Toric Topology, 2015
203 Donald Yau and Mark W. Johnson, A Foundation for PROPs, Algebras, and
Modules, 2015
202 Shiri Artstein-Avidan, Apostolos Giannopoulos, and Vitali D. Milman,
Asymptotic Geometric Analysis, Part I, 2015
201 Christopher L. Douglas, John Francis, André G. Henriques, and Michael A.
Hill, Editors, Topological Modular Forms, 2014
SURV/228
www.ams.org