COHOMOLOGY OF SPALTENSTEIN VARIETIES

Similar documents
16.2. Definition. Let N be the set of all nilpotent elements in g. Define N

SYMMETRIC FUNCTIONS, PARABOLIC CATEGORY O AND THE SPRINGER FIBER

Notes on Partial Resolutions of Nilpotent Varieties by Borho and Macpherson

THE S 1 -EQUIVARIANT COHOMOLOGY RINGS OF (n k, k) SPRINGER VARIETIES

8 Perverse Sheaves. 8.1 Theory of perverse sheaves

Character sheaves and modular generalized Springer correspondence Part 2: The generalized Springer correspondence

AFFINE PUSHFORWARD AND SMOOTH PULLBACK FOR PERVERSE SHEAVES

H(G(Q p )//G(Z p )) = C c (SL n (Z p )\ SL n (Q p )/ SL n (Z p )).

Algebraic Geometry Spring 2009

Combinatorics for algebraic geometers

GEOMETRIC SATAKE, SPRINGER CORRESPONDENCE, AND SMALL REPRESENTATIONS

Character sheaves and modular generalized Springer correspondence Part 2: The generalized Springer correspondence

Chern classes à la Grothendieck

BLOCKS IN THE CATEGORY OF FINITE-DIMENSIONAL REPRESENTATIONS OF gl(m n)

Representations and Linear Actions

Vector bundles in Algebraic Geometry Enrique Arrondo. 1. The notion of vector bundle

An overview of D-modules: holonomic D-modules, b-functions, and V -filtrations

Groups of Prime Power Order with Derived Subgroup of Prime Order

ELEMENTARY SUBALGEBRAS OF RESTRICTED LIE ALGEBRAS

Parameterizing orbits in flag varieties

Equivariant Algebraic K-Theory

EKT of Some Wonderful Compactifications

PERVERSE SHEAVES ON A TRIANGULATED SPACE

GEOMETRIC STRUCTURES OF SEMISIMPLE LIE ALGEBRAS

Math 249B. Geometric Bruhat decomposition

PERVERSE SHEAVES. Contents

PART I: GEOMETRY OF SEMISIMPLE LIE ALGEBRAS

Exercises on chapter 0

Math 249B. Nilpotence of connected solvable groups

Citation Osaka Journal of Mathematics. 43(2)

where m is the maximal ideal of O X,p. Note that m/m 2 is a vector space. Suppose that we are given a morphism

LOCAL VS GLOBAL DEFINITION OF THE FUSION TENSOR PRODUCT

HARTSHORNE EXERCISES

Three Descriptions of the Cohomology of Bun G (X) (Lecture 4)

Construction of M B, M Dol, M DR

Introduction to Chiral Algebras

On the geometric Langlands duality

Defining equations for some nilpotent varieties

LECTURE 11: SOERGEL BIMODULES

Math 248B. Applications of base change for coherent cohomology

Tilting exotic sheaves, parity sheaves on affine Grassmannians, and the Mirković Vilonen conjecture

APPENDIX 3: AN OVERVIEW OF CHOW GROUPS

On the singular elements of a semisimple Lie algebra and the generalized Amitsur-Levitski Theorem

CHAPTER 1. AFFINE ALGEBRAIC VARIETIES

The Lusztig-Vogan Bijection in the Case of the Trivial Representation

The Affine Grassmannian

THE QUANTUM CONNECTION

Algebraic varieties. Chapter A ne varieties

Exercises of the Algebraic Geometry course held by Prof. Ugo Bruzzo. Alex Massarenti

k k would be reducible. But the zero locus of f in A n+1

ALGEBRAIC GROUPS JEROEN SIJSLING

Introduction to Arithmetic Geometry Fall 2013 Lecture #18 11/07/2013

3.1. Derivations. Let A be a commutative k-algebra. Let M be a left A-module. A derivation of A in M is a linear map D : A M such that

Formal power series rings, inverse limits, and I-adic completions of rings

3. The Sheaf of Regular Functions

Representation Theory and Orbital Varieties. Thomas Pietraho Bowdoin College

FILTERED RINGS AND MODULES. GRADINGS AND COMPLETIONS.

L(C G (x) 0 ) c g (x). Proof. Recall C G (x) = {g G xgx 1 = g} and c g (x) = {X g Ad xx = X}. In general, it is obvious that

1. Algebraic vector bundles. Affine Varieties

MIXED HODGE MODULES PAVEL SAFRONOV

Noetherian property of infinite EI categories

Fourier Mukai transforms II Orlov s criterion

LECTURE 10: KAZHDAN-LUSZTIG BASIS AND CATEGORIES O

Math 121 Homework 5: Notes on Selected Problems

THE p-smooth LOCUS OF SCHUBERT VARIETIES. Let k be a ring and X be an n-dimensional variety over C equipped with the classical topology.

FOUNDATIONS OF ALGEBRAIC GEOMETRY CLASS 25

Schubert Varieties. P. Littelmann. May 21, 2012

Characteristic classes in the Chow ring

ARITHMETICALLY COHEN-MACAULAY BUNDLES ON HYPERSURFACES

REPRESENTATIONS OF FINITE GROUPS OF LIE TYPE: EXERCISES. Notation. 1. GL n

This is a closed subset of X Y, by Proposition 6.5(b), since it is equal to the inverse image of the diagonal under the regular map:

Systems of linear equations. We start with some linear algebra. Let K be a field. We consider a system of linear homogeneous equations over K,

Representations of algebraic groups and their Lie algebras Jens Carsten Jantzen Lecture III

SPACES OF RATIONAL CURVES IN COMPLETE INTERSECTIONS

NOTES ON RESTRICTED INVERSE LIMITS OF CATEGORIES

A CATEGORIFICATION OF INTEGRAL SPECHT MODULES. 1. Introduction

SYMMETRIC SUBGROUP ACTIONS ON ISOTROPIC GRASSMANNIANS

Notes on nilpotent orbits Computational Theory of Real Reductive Groups Workshop. Eric Sommers

3. Lecture 3. Y Z[1/p]Hom (Sch/k) (Y, X).

Row-strict quasisymmetric Schur functions

9. Birational Maps and Blowing Up

mult V f, where the sum ranges over prime divisor V X. We say that two divisors D 1 and D 2 are linearly equivalent, denoted by sending

NOTES ON PROCESI BUNDLES AND THE SYMPLECTIC MCKAY EQUIVALENCE

Algebraic Varieties. Notes by Mateusz Micha lek for the lecture on April 17, 2018, in the IMPRS Ringvorlesung Introduction to Nonlinear Algebra

π X : X Y X and π Y : X Y Y

Proof of Langlands for GL(2), II

Braid group actions on categories of coherent sheaves

The Hitchin map, local to global

1 Replete topoi. X = Shv proét (X) X is locally weakly contractible (next lecture) X is replete. D(X ) is left complete. K D(X ) we have R lim

Coherent sheaves on elliptic curves.

ALGEBRAIC GEOMETRY I - FINAL PROJECT

Combinatorial models for the variety of complete quadrics

Synopsis of material from EGA Chapter II, 4. Proposition (4.1.6). The canonical homomorphism ( ) is surjective [(3.2.4)].

12. Projective modules The blanket assumptions about the base ring k, the k-algebra A, and A-modules enumerated at the start of 11 continue to hold.

Constructible Derived Category

Toshiaki Shoji (Nagoya University) Character sheaves on a symmetric space and Kostka polynomials July 27, 2012, Osaka 1 / 1

Definition 9.1. The scheme µ 1 (O)/G is called the Hamiltonian reduction of M with respect to G along O. We will denote by R(M, G, O).

We can choose generators of this k-algebra: s i H 0 (X, L r i. H 0 (X, L mr )

Notes on Green functions

1 Notations and Statement of the Main Results

Transcription:

Transformation Groups, Vol.?, No.?,??, pp. 1?? c Birkhäuser Boston (??) COHOMOLOGY OF SPALTENSTEIN VARIETIES JONATHAN BRUNDAN Department of Mathematics University of Oregon Eugene, Oregon, USA brundan@uoregon.edu VICTOR OSTRIK Department of Mathematics University of Oregon Eugene, Oregon, USA vostrik@uoregon.edu To Professor Tonny Springer on the occasion of his eighty-fifth birthday Abstract. We give a presentation for the cohomology algebra of the Spaltenstein variety of all partial flags annihilated by a fixed nilpotent matrix, generalizing the description of the cohomology algebra of the Springer fiber found by De Concini, Procesi and Tanisaki. 1. Introduction Throughout the article, we fix integers n, d 0 and write Λ(n, d) for the set of all n-part compositions µ = (µ 1,..., µ n ) of d, so the µ i s are non-negative integers summing to d. Let Λ + (n, d) Λ(n, d) denote the n-part partitions of d, i.e. the λ Λ(n, d) satisfying λ 1 λ n. For λ Λ + (n, d), we write λ T for the transpose partition (which may have more than n non-zero parts). Let X be the complex projective variety of flags (U 0,..., U d ) in C d, so {0} = U 0 < U 1 < < U d = C d, dim U i /U i 1 = 1. By a classical result of Borel, the cohomology algebra H (X, C) is isomorphic to the coinvariant algebra, which is the quotient P/I of the polynomial algebra P := C[x 1,..., x d ], graded so that each x i is in degree 2, by the ideal I generated by the homogeneous symmetric polynomials of positive degree. To fix a specific isomorphism, let Ũi be the sub-bundle of the trivial vector bundle C d X X having fiber U i over the point (U 0,..., U d ) X. Then there is a unique graded algebra isomorphism ϕ : C H (X, C) with ϕ(x i ) = c 1 (Ũi/Ũi 1) H 2 (X, C) for each i = 1,..., d; see e.g. [F, 10.2, Proposition 3]. Associated to λ Λ + (n, d), we have the Springer fiber X λ, which is the closed subvariety of X consisting of all flags annihilated by the nilpotent matrix x λ of Jordan type λ T, so X λ = {(U 0,..., U d ) X x λ U i U i 1 for each i = 1,..., d}. Supported by NSF grant DMS-0654147 Supported by NSF grant DMS-0602263 Received December 16, 2010. Accepted April 25, 2011.

2 JONATHAN BRUNDAN AND VICTOR OSTRIK The pull-back homomorphism i : H (X, C) H (X λ, C) arising from the inclusion i : X λ X is surjective. In [DCP], De Concini and Procesi computed generators for the ideal of C that maps to ker i under the isomorphism ϕ, thus obtaining an explicit presentation for H (X λ, C). Soon after that, a slightly simplified description was given by Tanisaki in [T], as follows. Let C λ := P/I λ where I λ is the ideal generated by the elementary symmetric functions { e r (x i1,..., x im ) } m 1, 1 i 1 < < i m d,. r > m λ d m+1 λ n Then there is a unique isomorphism ϕ : C λ H (X λ, C) making the diagram ϕ C H (X, C) i C λ ϕ H (X λ, C) (1.1) commute, where the left hand map is the canonical quotient map which makes sense because I I λ. The symmetric group S d acts on P, hence also on the quotients C and C λ, by algebra automorphisms so that w x i = x w(i) for w S d. Using the isomorphism ϕ, we get induced an S d -action on H (X, C), which has (at least) two geometric definitions: the classical one as in [J, 13.1], and another more sophisticated construction originating in the remarkable work of Springer [S1, S2] (see also [J, 13.6]). Springer s construction yields also an action of S d on H (X λ, C), uniquely determined by the property that the surjection i is S d -equivariant (see [HS, Theorem 1.1] or [J, 13.13]). This plays an essential role in the derivation of the De Concini-Procesi-Tanisaki presentation. It is also one of the reasons the cohomology algebras H (X λ, C) are so interesting: the top degree cohomology is isomorphic as an S d -module to the irreducible representation S(λ T ) indexed by the transpose partition λ T (see [HS, Proposition 2.7] or [J, 13.16]). The main goal of this article is to prove a parabolic analogue of (1.1). Fix λ Λ + (n, d) as before and also µ Λ(n, d). We are going to replace the flag variety X by the partial flag variety X µ consisting of flags (V 0,..., V n ) of type µ in C d, so {0} = V 0 V 1 V n = C d, dim V i /V i 1 = µ i, and to replace the Springer fiber X λ by the Spaltenstein variety X λ µ = {(V 0,..., V n ) X µ x λ V i V i 1 for each i = 1,..., n}. Our main result gives an explicit presentation for the cohomology algebra of the Spaltenstein variety. To formulate this, we must first recall the classical description of the cohomology algebra H (X µ, C). Let P µ := P Sµ-inv be the subalgebra of P consisting of all S µ - invariants, where S µ denotes the parabolic subgroup S µ1 S µn of S d. For 1 i 1 < < i m n and r 1, we let e r (µ; i 1,..., i m ) and h r (µ; i 1,..., i m )

SPALTENSTEIN VARIETIES 3 denote the rth elementary and complete symmetric functions in the variables X i1 X im, where X j = {x k µ 1 + + µ j 1 + 1 k µ 1 + + µ j }. We interpret e r (µ; i 1,..., i m ) and h r (µ; i 1,..., i m ) as 1 if r = 0 and as 0 if r < 0. Note also that e r (µ; i 1,..., i m ) = h r (µ; i 1,..., i m ) = r 1+ +r m=r r 1+ +r m=r e r1 (µ; i 1 ) e rm (µ; i m ), (1.2) h r1 (µ; i 1 ) h rm (µ; i m ). (1.3) The algebra P µ is freely generated either by {e r (µ; i) 1 i n, 1 r µ i } or by {h r (µ; i) 1 i n, 1 r µ i }. Let I µ be the ideal of P µ generated by homogeneous symmetric polynomials of positive degree. Set C µ := P µ /I µ. Then there is a unique isomorphism ψ : C µ H (X µ, C) with ψ(e r (µ; i)) = ( 1) r c r (Ṽi/Ṽi 1) for each i = 1,..., n and r > 0, where Ṽi denotes the sub-bundle of the trivial vector bundle C d X µ X µ with fiber V i over (V 0,..., V n ) X µ. Let C λ µ := P µ /I λ µ where I λ µ is the ideal of P µ generated by the elements { h r (µ; i 1,..., i m ) m 1, 1 i } 1 < < i m n,. (1.4) r > λ 1 + + λ m µ i1 µ im Equivalently (see [B, Lemma 2.2]) I λ µ is generated by e r(µ; i 1,..., i m ) m 1, 1 i 1 < < i m n, r > µ i1 + + µ im λ l+1 λ n where l := #{i µ i > 0, i i 1,..., i m }, (1.5) from which it is easy to see that I λ µ = I λ, hence C λ µ = C λ, if µ is regular (all parts 1). Finally let j : X λ µ X µ be the inclusion, and note that the pull-back j : H (X µ, C) H (X λ µ, C) is surjective; see 2. Theorem 1.1. There is a unique isomorphism ψ : Cµ λ diagram C µ ψ H (X µ, C) j H (X λ µ, C) making the (1.6) C λ µ ψ H (X λ µ, C)

4 JONATHAN BRUNDAN AND VICTOR OSTRIK commute, where the left hand map is the canonical quotient map which makes sense because I µ I λ µ. The precise form of the relations (1.4) (1.5) was originally worked out in [B], which is concerned with the centers of integral blocks of parabolic category O for the general linear Lie algebra. These centers provide another natural occurrence of the algebras Cµ λ = H (Xµ, λ C); see also [St, Theorem 4.1.1] (which treats regular µ) and [BLPPW, Remark 9.10] (for all µ). The latter work shows moreover that the equivariant cohomology of Xµ λ is isomorphic to the center of the universal deformation of the corresponding block of parabolic category O, and extends the duality between equivariant cohomology rings of partial flag varieties and Springer fibers discovered by Goresky and MacPherson [GM] to Spaltenstein varieties. We next formulate a result which plays an essential role in the proof of Theorem 1.1. Let p : X X µ be the projection sending (U 0,..., U d ) X to (V 0,..., V n ) X µ where V i = U µ1+ +µ i for each i. It is classical that the pull-back p defines a graded algebra isomorphism between H (X µ, C) and H (X, C) Sµ-inv. Moreover the diagram C µ ψ H (X µ, C) C Sµ-inv ϕ p H (X, C) Sµ-inv (1.7) commutes, where the top map is induced by the inclusion P µ P. Using Poincaré duality, it also makes sense to consider the push-forward p as a map in cohomology. Let n d µ := dim X dim X µ = 1 2 µ i (µ i 1). Then p restricts to give an H (X µ, C)-module isomorphism between the space H (X, C) Sµ-anti of S µ -anti-invariants in H (X, C) and the module H (X µ, C). Thus, there is a commuting diagram C Sµ-anti ϕ H (X, C) Sµ-anti i=1 C µ [ 2d µ ] ψ p H (X µ, C)[ 2d µ ], (1.8) where for a graded vector space M we write M[i] for the graded vector space obtained by shifting the grading down by i, i.e. M[i] j = M i+j. Viewed as a C µ - module via the isomorphism C µ = C S µ-inv from (1.7), C Sµ-anti is free of rank one generated by the element ε µ := 1 S µ 1 i<j d, same S µ-orbit (x i x j ). The isomorphism at the top of (1.8) sends xε µ x for each x C µ ; see e.g. [B, Lemma 3.2].

SPALTENSTEIN VARIETIES 5 Theorem 1.2. There is a unique homogeneous linear map p of degree zero making the following diagram commute: H (X, C) i p H (X µ, C)[ 2d µ ] j (1.9) H (X λ, C) p H (X λ µ, C)[ 2d µ ]. Moreover the restriction of p is an isomorphism of graded vector spaces p : H (X λ, C) Sµ-anti H (X λ µ, C)[ 2d µ ]. The existence of an isomorphism H (X λ, C) Sµ-anti = H (Xµ, λ C)[ 2d µ ] was established already by Borho and Macpherson [BM2, Corollary 3.6(b)]. The point of Theorem 1.2 is to make this isomorphism canonical; see 4. Granted Theorem 1.2, one possible approach to the proof of Theorem 1.1 is sketched in [B, Remark 4.6]. The proof of Theorem 1.1 given in 5 below is quite different and gives a more natural explanation of the relations; it is a generalization of Tanisaki s original argument in [T]. In the course of the proof, we also obtain a homogeneous algebraic basis for H (Xµ, λ C) indexed by certain λ-tableaux of type µ, which is of independent interest; see 2 3. We point out finally that the algebras H (Xµ, λ C) arise in Ginzburg s construction of analogues of the Springer representations for the general linear Lie algebra; see [G1, BG]. In particular, there is a natural way to define an action of gl n (C) on the direct sum of the H (Xµ, λ C) s for all µ Λ(n, d), so that the top cohomology H 2d λ 2d µ (Xµ, λ C) µ Λ(n,d) is irreducible of highest weight λ. Following the reformulation by Braverman and Gaitsgory [BG] (see also [G2, 7]), this action can be constructed by applying the signed Schur functor Vsgn d CSd? to the S d -module H (X λ, C). In more detail, let Vsgn d denote the dth tensor power of the natural gl n (C)-module viewed as a right CS d -module so that w S d acts by (v 1 v d )w = sgn(w)v w(1) v w(d). For µ Λ(n, d), the µ-weight space of Vsgn d CSd H (X λ, C) is canonically isomorphic to H (X λ, C) Sµ-anti ; see [B, Lemma 3.1]. Using also Theorem 1.2, we get a canonical vector space isomorphism V d sgn CSd H (X λ, C) = µ Λ(n,d) H (X λ µ, C), (1.10) hence can transport the gl n (C)-module structure from the left to the right hand space. This yields the desired action. Using the presentation for the algebras H (X λ µ, C) from Theorem 1.1, it is possible to give a purely algebraic construction (no cohomology) of these representations, in the same spirit as the approach of Garsia and Procesi to the type A Springer representations in [GP]. This is pursued further in [B, 4], where one can find explicit algebraic formulae for the actions of the Chevalley generators of gl n (C) on the space on the right hand side of (1.10).

6 JONATHAN BRUNDAN AND VICTOR OSTRIK 2. Affine paving By an affine paving of a variety M, we mean a partition of M into disjoint subsets M i indexed by some finite poset (I, ), such that the following hold for all i I: j i M j is closed in M; M i (which is automatically locally closed) is isomorphic as a variety to A di for some d i 0. Suppose we are given λ Λ + (n, d) and µ Λ(n, d). The Spaltenstein variety Xµ λ from the introduction was defined and studied originally in [Spa]. In that paper, Spaltenstein constructed an explicit affine paving of the Springer fiber X λ, and deduced from that a parametrization of the irreducible components of Xµ. λ The goal in this section is to extend Spaltenstein s argument slightly to produce an affine paving of Xµ λ itself; actually we explain the dual construction which produces a more convenient parametrization from a combinatorial point of view. This is a widely known piece of folk-lore, but still we could not find it explicitly written in the literature. We begin with a few more conventions regarding partitions. We draw the Young diagram of a partition λ in the usual English way as in [M]. The sum of the parts of λ is λ, and h(λ) denotes the height of λ, that is, the number of non-zero parts. We write P k for the set of all partitions of height at most k, and P k,l for the set of all partitions fitting into a k l-rectangle, i.e. h(λ) k and λ 1 l. For partitions λ and µ, we write λ µ if λ i µ i for all i. Also denotes the usual dominance ordering. For µ Λ(n, d), we let µ + Λ + (n, d) be the unique partition obtained from µ by rearranging the parts in weakly decreasing order. Given 0 k d, let Gr k,d be the Grassmannian of k-dimensional subspaces of C d. To a partition γ P k, we associate its column sequence (c 1,..., c k ) defined from γ k+1 i = c i i. (2.1) This gives a bijection between P k,d k and the set of all sequences (c 1,..., c k ) with 1 c 1 < < c k d. Given a fixed basis f 1,..., f d for C d and γ P k,d k, we have the Schubert variety Y γ := {U Gr k,d dim U f 1,..., f ci i for i = 1,..., k}, (2.2) where (c 1,..., c k ) is the column sequence associated to γ; see e.g. [F, 9.4]. This is an irreducible closed subvariety of dimension γ. Also Y γ Y γ if and only if γ γ. The Schubert cells Yγ := Y γ \ γ γ Y γ give an affine paving of Gr k,d indexed by the poset (P k,d k, ). In terms of coordinates, every U Yγ can be represented as the span of the vectors f ci + for unique a i,j C. In particular, Y γ = A γ. a i,jf j (i = 1,..., k) (2.3) 1 j c i j c 1,...,c i

SPALTENSTEIN VARIETIES 7 In this section, we fix the basis f 1,..., f d as follows. Let e 1,..., e d be the standard basis for C d. If we identify the basis vectors e 1,..., e d with the boxes of the Young diagram of λ working up columns starting from the first (leftmost) column, then the nilpotent matrix x λ of Jordan type λ T is the endomorphism of C d sending a basis vector to the one immediately below it in the diagram, or to zero if it is at the bottom of its column. Then let f 1,..., f d be the basis for C d obtained by reading the boxes of the Young diagram in order along rows starting from the first (top) row. For example: λ = (4, 3, 2) e 3 e 6 e 8 e 9 e 2 e 5 e 7 f 1 f 2 f 3 f 4 f 5 f 6 f 7 x λ = (2.4) e 1 e 4 f 8 f 9 Also let (.,.) be the symmetric bilinear form on C d such that (f i, f j ) = δ i,j. The following proposition provides the key induction step. Proposition 2.1. Suppose n 1 and that k := µ n, s := λ 1 and t := λ n satisfy s k t. Let π : X µ Gr k,d be the morphism (V 0,..., V n ) V n 1. Let µ Λ(n 1, d k) be obtained from µ by forgetting the last part µ n, let β := ((s k) k t ) P k,d k, and take any γ P k,d k with column sequence (c 1,..., c k ). (i) The restriction of π defines a morphism π : Xµ λ Y β. (ii) There is an isomorphism of varieties f γ : π 1 (Yγ ) Yγ X µ. Assume in addition that γ β. Let λ Λ + (n 1, d k) be the partition whose Young diagram is obtained by removing one box from the bottom of each of the columns numbered c 1 = 1,..., c t = t, c t+1,..., c k in the Young diagram of λ, then re-ordering the columns to get a proper partition shape. (iii) For (V 0,..., V n ) π 1 (Yγ ), we have x λ (V n 1 ) x λ (V n ) V n 1, and the restriction of x λ to V n 1 is of Jordan type (λ) T. (iv) The isomorphism f γ can be chosen so that it restricts to an isomorphism f γ : π 1 (Y γ ) Y γ X λ µ. Proof. Note that Y β = {U Gr k,d f 1,..., f t U f 1,..., f s }. Take (V 0,..., V n ) Xµ. λ Then x λ (V n ) V n 1 and (x λ ) n 1 (V n 1 ) = {0}, or equivalently, f 1,..., f t = (ker(x λ ) n 1 ) Vn 1 (imx λ ) = f 1,..., f s. This shows π(xµ) λ Y β, establishing (i). Now we ll prove (ii), (iii) and (iv) all under the assumption that γ β, noting that (ii) for more general γ is actually a particular case of (iv) on replacing λ by the partition of d with just one non-zero part. The subspace f c1,..., f ck belongs to Yγ. Take U Yγ represented by the coordinates (a i,j ) according to (2.3). The non-zero vectors of the form (x λ ) i (f j ) for i 0 and 1 j s form a basis for C d. Hence there exists a unique matrix g(u) GL d (C) such that g(u)(f ci ) = f ci + j {1,...,c i}\{c 1,...,c i} a i,jf j for each i = 1,..., k; g(u)(f j ) = f j for each j {1,..., s} \ {c 1,..., c k }; g(u)(x λ (f j )) = x λ (g(u)(f j )) for any 1 j d such that x λ (f j ) 0.

8 JONATHAN BRUNDAN AND VICTOR OSTRIK The transpose matrix (x λ ) T annihilates f 1,..., f s, and (x λ ) T (x λ (f j )) = f j whenever x λ (f j ) 0. Using this we see that g(u) commutes with (x λ ) T, or equivalently, g(u) T commutes with x λ. Moreover g(u)( f c1,..., f ck ) = U, hence g(u) T (U ) = f c1,..., f ck. The restriction of x λ to the space f c1,..., f ck is of Jordan type (λ) T. Hence we can pick an isomorphism θ : f c1,..., f ck C d k such that θ x λ = x λ θ. Define f γ : π 1 (Yγ ) Yγ X µ to be the map sending (V 0,..., V n ) π 1 (Yγ ) to (U, (V 0,..., V n 1 )) Y γ X µ, where U = Vn 1 and V i = θ(g(u) T (V i )). This makes sense because g(u) T (V n 1 ) = f c1,..., f ck. It is clear that the map f γ is invertible, hence it is an isomorphism of varieties as in (ii). To deduce (iii), take (V 0,..., V n ) π 1 (Yγ ). We have that Vn 1 Yγ Y β, hence Vn 1 f 1,..., f s. This shows that x λ (V n ) = f 1,..., f s V n 1, so in particular x λ leaves V n 1 invariant. Moreover multiplication by g(vn 1) T is an x λ -equivariant isomorphism between V n 1 and f c1,..., f ck. Hence the Jordan type of x λ on V n 1 is the same as its Jordan type on f c1,..., f ck, namely (λ) T. Finally, for (iv), suppose we are given (V 0,..., V n ) π 1 (Y γ ) mapping to (U, (V 0,..., V n 1 )) under the isomorphism f γ. We need to show that (V 0,..., V n ) belongs to Xµ λ if and only if (V 0,..., V n 1 ) belongs to Xµ λ. By (iii), we know automatically that x λ (V n ) V n 1. Hence we just need to check for each i = 1,..., n 1 that x λ (V i ) V i 1 if and only if x λ (V i ) V i 1. This is clear as x λ θ g(u) T = θ g(u) T x λ on V n 1. By a column-strict (resp. semi-standard) λ-tableau we mean a filling of the boxes of the Young diagram of λ by integers so that entries are strictly increasing down columns (resp. strictly increasing down columns and weakly increasing along rows). Let Col λ µ (resp. Std λ µ) denote the set of all column-strict (resp. semi-standard) λ- tableaux that have exactly µ i entries equal to i for each i = 1,..., n. The following definition will also be needed in the next section. Definition 2.2. Suppose that n 1 and we are given T Col λ µ. The existence of T implies that the integers k, s, t defined as in Proposition 2.1 satisfy s k t. We define (c 1,..., c k ), γ, µ, T, and λ as follows: 1 c 1 < < c k s index the columns of T containing entry n; γ P k,d k is the partition with column sequence (c 1,..., c k ); µ Λ(n 1, d k) is defined by forgetting the last part of µ; T is obtained by removing all boxes labelled by n from T, then repeatedly interchanging columns i and (i+1) whenever the ith column contains fewer boxes than the (i + 1)th column for some i; λ Λ + (n 1, d k) is the shape of T, so T Col λ µ. Example 2.3. (i) If n = 6 and T = 2 1 2 2 3 2 4 4 4 6 6 5 then (c 1, c 2 ) = (1, 3), γ = (1) and T = 1 2 2 2 2 3 4. 4 4 5

(ii) If n = 4 and T = 3 1 1 2 1 4 4 2 3 4 2 T = 1 1 1 3 2. 2 3 2 SPALTENSTEIN VARIETIES 9 then (c 1, c 2, c 3 ) = (1, 4, 6), γ = (3, 2) and Now we make several definitions by induction on n. Suppose first that n = d = 0. We define the degree deg(t) of the unique tableau T Col λ µ (the empty tableau) to be zero, let YT := Xµ λ (which is a single point), and let be the trivial partial order on Col λ µ. Now suppose that n 1, take T Col λ µ and define γ and T as in Definition 2.2. Then set deg(t) := γ + deg(t). (2.5) For example, the tableau T from Example 2.3(i) is of degree 2 and the one from Example 2.3(ii) is of degree 8. Also define a subset YT Xµ λ by setting Y T := 1 f γ (Yγ Y ), (2.6) T where f γ is the map chosen in Proposition 2.1(iv). Finally define a partial order on Col λ µ by declaring that T T if γ γ, or γ = γ and T T, where γ and T are defined using Definition 2.2 again but starting from T. Theorem 2.4. The sets YT give an affine paving of Xµ λ indexed by the poset (Col λ µ, ), such that YT = A deg(t). Moreover the complement X µ \ Xµ λ also has an affine paving. Proof. Consider the map π : Xµ λ Y β from Proposition 2.1(i). Since Y β is paved by the affine spaces Yγ for γ β, it suffices to show that the inverse images π 1 (Yγ ) all have affine pavings. This is follows by induction on n using Proposition 2.1(iv) and the definition (2.6). The same induction gives also that YT = A deg(t). Finally, to see that the complement X µ \ Xµ λ has an affine paving, we use also Proposition 2.1(ii) to see that the complement is the disjoint union of the spaces π 1 (Yγ ) \ π 1 (Yγ ) = Yγ (X µ \ Xµ λ ) for each γ β, each of which has an affine paving by induction, together with π 1 (Yγ ) = Yγ X µ for γ β, which have affine pavings too. Corollary 2.5. The cohomology H (X λ µ, C) vanishes in all odd degrees, and in even degrees dim H 2r (X λ µ, C) is equal to the number of T Col λ µ with deg(t) = r. Moreover the pull-back j : H (X µ, C) H (X λ µ, C) is surjective. Proof. This is a standard consequence of the existence of an affine paving as in Theorem 2.4; see e.g. [HS, Corollary 2.3] or the discussion at the bottom of [J, p.163]. Corollary 2.6. The variety X λ µ is non-empty if and only if λ µ +. Assuming that is the case, X λ µ is connected. Proof. The first statement follows because Col λ µ is non-empty if and only if λ µ +. For the second statement, we observe that there is a unique T Col λ µ of degree zero, hence dim H 0 (X λ µ, C) = 1 by Corollary 2.5.

10 JONATHAN BRUNDAN AND VICTOR OSTRIK Although not really needed in the rest of the article, we end the section by explaining for completeness how to recover Spaltenstein s classification of the irreducible components of Xµ λ from Theorem 2.4. Suppose we are given (V 0,..., V n ) Xµ. λ Assuming n 1, Proposition 2.1(iii) shows that the restriction of x λ to V n 1 is of Jordan type (λ) T for some λ Λ + (n 1, d µ n ) such that λ i+1 λ i λ i for all 1 i n 1. By induction we deduce for j = 0, 1,..., n that the restriction of x λ to V j is of Jordan type (λ (j) ) T for λ (j) Λ + (j, µ 1 + + µ j ) satisfying λ (j+1) i+1 λ (j) i λ (j+1) i for all 1 i j. It follows that there is a well-defined map J : X λ µ Std λ µ, (V 0,..., V n ) S, (2.7) where S is the semi-standard tableau having entry j in all boxes belonging to the Young diagram of λ (j) but not of λ (j 1) for j = 1,..., n. Theorem 2.7. For S Std λ µ, we have that J 1 (S) is a locally closed, smooth, irreducible subvariety of Xµ λ of dimension d λ d µ. Moreover YS is a dense open subset of J 1 (S). Proof. For T Col λ µ, define T + Std λ µ as follows. Let T Col λ µ be as in Definition 2.2. Then let T + be obtained from the recursively defined (T) + Std λ µ by adding the entry n to all boxes belonging to the Young diagram of λ but not of λ. For example if T is as in Example 2.3(ii) then T + = 1 1 1 2 3 4. All points of Y 2 2 3 4 4 T map to T + under the map J. Hence for S Std λ µ we have that J 1 (S) = T Ω(S) Y T where Ω(S) := {T Col λ µ T + = S}. (2.8) In particular, J 1 (S) is locally closed as each YT is so by Theorem 2.4. Note further that S belongs to Ω(S), and all other T Ω(S) are strictly smaller than S in the partial order, so Theorem 2.4 implies YS is open in J 1 (S). It remains to prove that J 1 (S) is smooth and irreducible of the given dimension. Define γ and S Std λ µ as in Definition 2.2, using S in place of T. Let (c 1,..., c k ) be the column sequence of γ. In other words, the c i s index the columns of S containing the entry n. Also define π as in Proposition 2.1(i). Let Ω 0 (S) := {T Ω(S) T has entry n in columns c 1,..., c k }, so that M := J 1 (S) π 1 (Yγ ) = YT. (2.9) T Ω 0(S) Observe that every element of Ω(S) \ Ω 0 (S) is smaller than every element of Ω 0 (S) in the partial order. Hence using Theorem 2.4 again, we deduce that M is an open subset of J 1 (S). The map T T is a bijection between Ω 0 (S) and Ω(S). So, comparing (2.9) with (2.8) for S, we see that the isomorphism f γ from Proposition 2.1(iv) restricts to an isomorphism M Yγ J 1 (S). By induction, we deduce that M is a smooth irreducible variety of dimension γ + d λ d µ = d λ d µ.

SPALTENSTEIN VARIETIES 11 Finally let Z be the centralizer of x λ in GL d (C). This is a connected algebraic group which acts naturally on X λ µ. Moreover Z leaves the subset J 1 (S) invariant, so the action restricts to a morphism m : Z M J 1 (S). To complete the proof of the theorem, we just need to show that this map is surjective. Take (V 0,..., V n ) J 1 (S). By (2.8), it lies in YT for some T Ω(S). Suppose that the entries of T equal to n are in columns 1 b 1 < < b k s. There is a permutation of the columns of equal height in the Young diagram of λ sending columns b 1,..., b k to columns c 1,..., c k. Let x GL d (C) be the matrix inducing the associated permutation of the basis f 1,..., f d, labelling boxes as in (2.4). So we have that x( f b1,..., f bk ) = f c1,..., f ck, and obviously x Z. Also, as in the proof of Proposition 2.1, there is an element y := g(vn 1) T Z such that y(v n 1 ) = f b1,..., f bk. Hence xy(v n 1 ) = f c1,..., f ck, so xy Z maps (V 0,..., V n ) to a point of M. This implies that m is surjective. Corollary 2.8 (Spaltenstein). Assuming λ µ +, Xµ λ is equidimensional of dimension d λ d µ, with irreducible components Y S := YS = J 1 (S) for S Std λ µ (closure in the Zariski topology). Proof. By Theorem 2.7, the subvarieties J 1 (S) for S Std λ µ are irreducible of dimension d λ d µ, and they partition Xµ λ into disjoint subsets. Hence their closures give all the irreducible components. Finally, YS has the same closure as J 1 (S) as it is a dense subset. Corollary 2.9. For T Col λ µ, we have that deg(t) d λ d µ, with equality if and only if T is semi-standard. Proof. Let S := T +, notation as in the proof of Theorem 2.7. By Theorem 2.7 and the decomposition (2.8), YT is an irreducible subset of the irreducible variety J 1 (S), and it is dense in J 1 (S) if and only if T is semi-standard (equivalently, T = S). The corollary follows from this since dim J 1 (S) = d λ d µ and dim YT = deg(t). (It is not hard to supply a purely combinatorial proof of this corollary.) 3. Algebraic basis Continue with fixed λ Λ + (n, d) and µ Λ(n, d). Recall the definition of the elements h r (µ; i 1,..., i m ) P µ and the algebra C λ µ := P µ /I λ µ from the introduction. Lemma 3.1. The algebra C λ µ is non-zero if and only if λ µ +. Proof. Since everything is graded, Cµ λ is non-zero if and only if all the generators from (1.4) are of positive degree, i.e. λ 1 + + λ m µ i1 + + µ im for all m 1 and 1 i 1 < < i m n. By the definition of the dominance ordering on partitions, this is the statement that λ µ +.

12 JONATHAN BRUNDAN AND VICTOR OSTRIK For T Col λ µ, we inductively define an element h(t) P µ as follows. If n = d = 0 then T is the empty tableau and we simply set h(t) := 1. If n 1, we let γ P k and T Col λ µ be as in Definition 2.2. The natural embedding C[x 1,..., x d k ] C[x 1,..., x d ] induces an embedding P µ P µ. This allows us to view the recursively defined element h(t) P µ as an element of P µ. Then we set k h(t) := h(t)h γ (µ; n) where h γ (µ; n) := h γi (µ; n), (3.1) where each h γi (µ; n) is as defined in the introduction. Recalling (2.5), h(t) is homogeneous of degree 2 deg(t). For example, if T is as in Example 2.3(i) then h(t) = h 1 (µ; 3)h 1 (µ; 6) = x 6 (x 11 + x 12 ), while for Example 2.3(ii) we get that h(t) = h 1 (µ; 2)h 2 (µ; 3)h 1 (µ; 3)h 3 (µ; 4)h 2 (µ; 4). We use the same notation for the canonical image of h(t) in the quotient Cµ. λ Theorem 3.2. The elements {h(t) T Col λ µ} give a basis for C λ µ. In the remainder of the section, we will prove the spanning part of Theorem 3.2, postponing the proof of linear independence to 5. The approach is similar to [T, Lemmas 2 3] where the case of regular µ was treated, but the general case turns out to be considerably more delicate. The argument goes by induction on n. The theorem is trivial in the case n = 0, so assume for the rest of the section that n 1 and that we have proved the spanning part of Theorem 3.2 for all smaller n. Let k := µ n, s := λ 1 and t := λ n. In view of Lemma 3.1, we may as well assume that λ µ +, hence we have that s k t. Set β := ((s k) k t ) P k, and let µ Λ(n 1, d k) be obtained from µ by forgetting the last part (cf. Proposition 2.1). Let be the partial order on the set P k of partitions of height at most k such such that γ κ if either γ > κ, or γ = κ and γ κ in the dominance ordering. For γ P k, let J γ (resp. J γ ) denote the ideal of P µ generated by I λ µ and all h κ (µ; n) for κ γ (resp. κ γ). Lemma 3.3. Fix m 0, 1 i 1 < < i m n 1, r 0 and γ P k. Suppose we are given 1 c k such that r + c > λ 1 + + λ m µ i1 µ im, (3.2) r + γ c > λ 1 + + λ m + λ m+1 µ i1 µ im µ n. (3.3) Then h r (µ; i 1,..., i m )h γ (µ; n) J γ. Proof. We first formulate and prove two technical claims. By a marked partition we mean a pair (π; p) consisting of a partition π and a non-zero part p of π. We write π {q} for the partition obtained by adding one extra part equal to q to the partition π, so that (π {q}; q) is a marked partition. Claim 1. Suppose we are given vectors v(π; p) for each marked partition (π; p) with 1 π c, where c is fixed as in the statement of the lemma. For any partition π with π c, let v(π) denote p v(π; p) summing over the set of all non-zero parts p of π; in particular v( ) := 0. Assume for 1 b c that b v(π) + v(π {q}; q) = 0 (3.4) q=1 i=1

SPALTENSTEIN VARIETIES 13 for each partition π of (c b). Then we have that π =c ( 1)h(π) v(π) = 0. To see this, we note for i = 1,..., c that π <c h(π)=i 1 ( 1) h(π) v(π) = π <c h(π)=i 1 = π <c h(π)=i c π ( 1) h(π)+1 v(π {q}; q) q=1 ( 1) h(π) v(π) + π =c h(π)=i ( 1) h(π) v(π), using (3.4) for the first equality and the definition of v(π) for the second. Using this identity repeatedly, we get that v( ) = π <c h(π)=0 = π <c h(π)=1 = π <c h(π)=c ( 1) h(π) v(π) + ( 1) h(π) v(π) + ( 1) h(π) v(π) + π =c h(π) 0 π =c h(π) 1 π =c h(π) c Since v( ) = 0 this establishes Claim 1. ( 1) h(π) v(π) ( 1) h(π) v(π). ( 1) h(π) v(π) = ( 1) h(π) v(π). π =c Call a function f : {1,..., c} {1,..., c} a quasi-permutation of descent b 0 if there exist distinct integers 1 j 1,..., j b+1 c and a bijection f : {1,..., c} \ {j 1,..., j b } {1,..., c} \ {j 1,..., j b } such that j 1 = c; f(j 1 ) = j 2, f(j 2 ) = j 3,..., f(j b ) = j b+1 ; f(j) = f(j) for each j {1,..., c} \ {j 1,..., j b }. Note b, j 1,..., j b+1 and f are uniquely determined by the quasi-permutation f, and quasi-permutations of descent 0 are ordinary permutations belonging to the symmetric group S c. The marked cycle type of the quasi-permutation f is the marked partition (π; p) defined by letting π be the partition of (c b) recording the usual cycle type of the permutation f of {1,..., c} \ {j 1,..., j b } and p be the length of the cycle that involves j b+1 when f is written as a product of disjoint cycles. For example, the quasi-permutation ( 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 ) 2 1 5 6 7 10 8 9 7 4 12 13 14 11 3 of c = 15 has marked cycle type ((4, 3 2, 2); 3). Let S c (π; p) denote the set of all quasi-permutations of marked cycle type (π; p). In particular, S c ((1 c ); 1) = {id}.

14 JONATHAN BRUNDAN AND VICTOR OSTRIK Given f S c (π; p), let ω(f) := (1 f(1), 2 f(2),..., c f(c)) Z c, ω(f) := (1 f(1)) + + (c f(c)) Z 0, c k h γ ω(f) (µ; n) := h γa a+f(a)(µ; n) h γa (µ; n). a=1 a=c+1 Claim 2. For each marked partition (π; p) with 1 π c, we set v(π; p) := h r+ ω(f) (µ; i 1,..., i m )h γ ω(f) (µ; n). f S c(π;p) Also define v(π) as in Claim 1. Then, for 1 b c and each partition π of (c b), we have that v(π) + b q=1 v(π {q}; q) J γ. To establish this, fix 1 b c and a partition π of (c b), and let j = (j 1,..., j b ) a tuple of distinct integers with Ω := (j; g) j 1 = c and 1 j 2,..., j b < c; g a permutation. of {1,..., c} \ {j 1,, j b } of cycle type π Given (j; g) Ω and 1 i c, we let f i (j; g) : {1,..., c} {1,..., c} be the function mapping j 1 j 2,..., j b 1 j b, j b i and j g(j) for all j {1,..., c} \ {j 1,..., j b }. If i {1,..., c} \ {j 1,..., j b } then f i (j; g) is a quasipermutation of descent b and marked cycle type (π; p) where p is the length of the cycle of g containing i. Otherwise, we have that i = j b q+1 for some 1 q b, and f i (j; g) is a quasi-permutation of descent (b q) and marked cycle type (π {q}; q). It follows that ω(f i (j; g)) = c i and v(π) + b v(π {q}; q) = q=1 (j;g) Ω i=1 Therefore it suffices to show for each (j; g) Ω that x := belongs to J γ. Let y := c h r+c i (µ; i 1,..., i m )h γ ω(fi(j;g))(µ; n). c h r+c i (µ; i 1,..., i m )h γ ω(fi(j;g))(µ; n) i=1 c h γa a+fi(j;g)(a)(µ; n) a=1 a j b k a=c+1 h γa (µ; n) so that c x = h r+c i (µ; i 1,..., i m )h γjb j b +i(µ; n)y. i=1

Using (1.3), we can expand h r+c+γjb j b (µ; i 1,..., i m, n) = SPALTENSTEIN VARIETIES 15 r+c i=j b γ jb h r+c i (µ; i 1,..., i m )h γjb j b +i(µ; n). As j b c, we have that c + γ jb j b γ c, hence using (3.3) and (1.4) we get that h r+c+γjb j b (µ; i 1,..., i m, n) I λ µ. Using (3.2) we have that h r+c i (µ; i 1,..., i m ) I λ µ for i 0. We deduce that r+c h r+c i (µ; i 1,..., i m )h (µ; n) γjb j b +i Iλ µ. i=1 As I λ µ J γ, this implies that x c+r i=c+1 h r+c i (µ; i 1,..., i m )h γjb j b +i(µ; n)y (mod J γ ). But for i > c the term h γjb j b +i(µ; n)y is of the form h κ (µ; n) for some κ P k with κ > γ, so it belongs to J γ. This proves Claim 2. Now we can complete the proof of the lemma. Let v(π; p) be as in Claim 2 and then define v(π) as in Claim 1. From Claims 1 and 2, we get that ( 1) h(π) c v(π) J γ. π =c For π = c, the set S c (π; p) appearing in the definition of v(π; p) consists of quasipermutations of descent zero, i.e. ordinary permutations. So the above sum is equal to v((1 c )) = h r (µ; i 1,..., i m )h γ (µ; n) plus a linear combination of terms h r+ ω(f) (µ; i 1,..., i m )h γ ω(f) (µ; n) for ordinary permutations f id. It remains to observe for all such f that ω(f) < 0 in the dominance ordering on Z c, hence h γ ω(f) (µ; n) J γ. Lemma 3.4. For γ P k with γ β, we have that J γ = J γ. Moreover J γ = Iµ λ if γ > k(s k). Proof. If γ 1 > s k = λ 1 µ n then h γ1 (µ; n) I λ µ by (1.4), hence h γ (µ; n) I λ µ J γ. In particular, this shows J γ = I λ µ when γ > k(s k), as in that case all γ κ P k satisfy κ 1 > s k. It remains to show that h γ (µ; n) J γ for γ β with γ 1 s k. In that case, we have that t 1 and γ k t+1 > 0. Now apply Lemma 3.3, taking r = 0, m = n 1 and c = k t + 1 and noting the right hand sides of (3.2) and (3.3) equal k t and 0, respectively. Lemma 3.5. Take γ P k such that γ β. Define (c 1,..., c k ) and λ Λ + (n 1, d k) as in the statement of Proposition 2.1. The quotient J γ /J γ is spanned by the images of the elements h(t) for T Col λ µ such that T has entry n in each of columns c 1,..., c k.

16 JONATHAN BRUNDAN AND VICTOR OSTRIK Proof. Consider the homomorphism C[x 1,..., x d ] C[x 1,..., x d k ] which maps x i x i for 1 i d k and x i 0 for d k + 1 i d. It restricts to a homomorphism P µ P µ such that h r (µ; i) h r (µ; i) for 1 i n 1, and h r (µ; n) 0 for r 1. Let Φ : P µ Cµ λ be the composite of this homomorphism with the natural quotient map P µ Cµ λ. The P µ -module J γ /J γ is cyclic, generated by the image of h γ (µ; n). We claim that the action of P µ on J γ /J γ factors through Φ to make J γ /J γ into a welldefined cyclic Cµ λ -module. To prove this, we need to show that (ker Φ)J γ J γ. By definition, ker Φ is generated by the elements {h r (µ; n) r > 0} together with the elements { h r (µ; i 1,..., i m ) m 1, 1 i 1 < < i m n 1, r > λ 1 + + λ m µ i1 µ im So we need to show that multiplication by either of these families of elements sends h γ (µ; n) into J γ. The first family is easy to deal with: for r > 0 the element h r (µ; n)h γ (µ; n) is a symmetric polynomial in the variables x d k+1,..., x d of degree strictly greater than γ. Hence it can be expressed as a linear combination of h κ (µ; n) s with κ > γ, as the h κ (µ; n) s for κ P k give a basis for the space of all symmetric polynomials in x d k+1,..., x d. To deal with the second family, we need to show that h r (µ; i 1,..., i m )h γ (µ; n) J γ for m 1, 1 i 1 < < i m n 1 and r > λ 1 + + λ m µ i1 µ im. Let c := λ 1 + + λ m λ 1 λ m. If c = 0 then h r (µ; i 1,..., i m ) I λ µ already, so there is nothing to do. So we may assume that c 1, and are in the situation of Lemma 3.3. The hypothesis (3.2) is immediate, so we are left with checking (3.3). To see that, note that λ has c fewer boxes than λ on the first m rows. So by the definition of λ, we must have that k + 1 c + γ c > λ m+1. Hence r + γ c > λ 1 + + λ m µ i1 µ im + λ m+1 k λ 1 + + λ m + λ m+1 µ i1 µ im µ n, and the claim is proved. Now to prove the lemma, we have shown that J γ /J γ is a cyclic Cµ λ -module generated by the image of h γ (µ; n). By the induction hypothesis which we have been assuming since the paragraph before Lemma 3.3, we know that Cµ λ is spanned by the elements h(t) for T Col λ µ. Hence J γ is spanned by the images of the elements h(t) = h(t)h γ (µ; n) as described in the statement of the lemma. Now we can complete the proof of the spanning part of Theorem 3.2. Enumerate the partitions γ β as γ (1) =, γ (2),..., γ (N 1), γ (N) = β so that γ (i) γ (j) implies i j. Let J i be the ideal of C λ µ generated by {h γ (j)(µ; n) i j N}. Then C λ µ = J 1 J N J N+1 := {0} }.

SPALTENSTEIN VARIETIES 17 is a filtration of C λ µ. Lemma 3.4 implies for i = 1,..., N that the canonical map J γ (i) J i induces a surjection J γ (i)/j γ (i) J i /J i+1. Hence by Lemma 3.5 we get that J i /J i+1 is spanned by the images of the h(t) for T Col λ µ with entry n in each of columns c 1,..., c k, where (c 1,..., c k ) is the column sequence of γ (i). Hence the h(t) for all T Col λ µ span C λ µ itself. 4. Proof of Theorem 1.2 In order to prove Theorem 1.2, we need to exploit the construction of the Springer representations via perverse sheaves. As is recalled in detail below, there are two basic approaches, one (by restriction) due to Lusztig, Borho and Macpherson [BM1], and the other (by Fourier transform) due to Kashiwara and Brylinski [Bry]. We refer to [J, ch.13] and [G2, 6] for more recent accounts of these two approaches, and also [KW, VI.15] which explains the relationship between them. (Some of these references work in terms of étale cohomology so require some translation before they can be used in our complex setting.) For any complex variety Y, we let D b (Y ) be the bounded derived category of constructible sheaves of C-vector spaces on Y ; see e.g. [G2, 3]. Let D : D b (Y ) D b (Y ) be the Verdier duality functor. Let Perv(Y ) be the (abelian) full subcategory of D b (Y ) consisting of perverse sheaves; see e.g. [G2, 4]. The constant sheaf on Y is denoted C Y, which we often view as an object in D b (Y ) concentrated in degree zero. For M D b (Y ), M[i] denotes the object obtained from M by translating down by i. If V = j Z V j is a graded vector space, we ll write M V for j Z M[ j] V j, so M (V [i]) = (M[i]) V = (M V )[i]. Assume from now on that Y is irreducible and smooth. In that case, C Y [dim Y ] is a perverse sheaf. For a holomorphic vector bundle E Y, let D b mon(e) be the derived category of the category of bounded complexes of sheaves of C-vector spaces on E whose cohomology sheaves are monodromic, i.e. locally constant over orbits of the natural C -action on E. Let Perv mon (E) be the full subcategory of Perv(E) consisting of the monodromic perverse sheaves. We will need the Fourier transform F : D b mon(e) D b mon(e ), where E Y is the dual bundle; see [G2, 8] or [Bry, 6] for its definition (our F is the normalized Fourier transform denoted F on [Bry, p.69]). The Fourier transform induces an equivalence of categories F : Perv mon (E) Perv mon (E ) (see [Bry, Corollaire 7.23]), which corresponds under the Riemann-Hilbert correspondence to the formal Fourier transform on holonomic D-modules with regular singularities (see [Bry, Théorème 7.24]). Lemma 4.1. Let E Y be a vector bundle on the smooth irreducible variety Y as above. Let ι : V E be a sub-bundle with annihilator ῑ : V E. Also

18 JONATHAN BRUNDAN AND VICTOR OSTRIK let ˆι : 0 E be the zero sub-bundle. The unit of adjunction Id ῑ ῑ 1 defines a canonical map res : C E ῑ C V. Similarly there is a canonical map res : ι C V ˆι C 0 defined by applying ι to the unit of adjunction for the inclusion 0 V. There are unique (up to scalars) horizontal isomorphisms in the following diagram: DF(ˆι C 0 [dim V ]) C E [dim V ] DF(res) res DF(ι C V [dim V ]) ῑ C V [dim V ] Moreover the scalars can be chosen so the diagram commutes. Proof. As ῑ C V [dim V ] is self-dual, the existence of the bottom isomorphism amounts to the assertion that F(ι C V [dim V ]) = ῑ C V [dim V ], which is a basic property of Fourier transform; see [G2, Proposition 8.3(4)] or [KW, Corollary III.13.4] in the étale setting. The uniqueness of the bottom isomorphism follows as End(ῑ C V ) = C. Existence and uniqueness of the top isomorphism is proved in a similar way. Finally the commutativity of the diagram is justified in [KW, Remark III.13.6] (we have applied D to the statement there). Now let g := gl d (C) and b be the Borel subalgebra of upper triangular matrices. As always X and X µ are the varieties of full flags and partial flags of type µ in C d. We abbreviate r := 2 dim X = d(d 1), r µ := 2 dim X µ = d(d 1) 2d µ. We will always identify g with g via the trace form, noting that b is the nilpotent radical n of b. We can view g as a self-dual vector bundle over a point, so that the Fourier transform for g gives a self-equivalence F : Perv mon (g) Perv mon (g). We also work with the trivial vector bundles g X X and g X µ X µ. These are again identified with their duals, so Fourier transform gives two more self-equivalences F : Perv mon (g X) Perv mon (g X), F : Perv mon (g X µ ) Perv mon (g X µ ). Let N be the nilpotent cone in g, and set Ñ := {(x, (U 0,..., U d )) N X xu i U i 1 for each i = 1,..., d}, g := {(x, (U 0,..., U d )) g X xu i U i for each i = 1,..., d}. These are sub-bundles of g X X, with fibers isomorphic to n and b, respectively. In fact, g is the annihilator of Ñ in the self-dual bundle g X, while Ñ is canonically isomorphic to the cotangent bundle T X of the flag variety; see e.g. [J,

SPALTENSTEIN VARIETIES 19 6.5]. Hence both varieties are smooth and irreducible, with dim Ñ = 2 dim X = r and dim g = dim g = d 2. Analogously in the parabolic case, we consider Ñ µ := {(x, (V 0,..., V n )) N X µ xv i V i 1 for each i = 1,..., n}, g µ := {(x, (V 0,..., V n )) g X µ xv i V i for each i = 1,..., n}, which are sub-bundles of g X µ X µ such that g µ is the annihilator of ѵ, and ѵ is isomorphic to the cotangent bundle T X µ. So ѵ and g µ are smooth, irreducible varieties with dim ѵ = 2 dim X µ = r µ and dim g µ = dim g = d 2, respectively. We will often exploit the following inclusions of sub-bundles: g X ῑ ι g Ñ {0} X ˆι ῑ ι g X µ g µ Ñ µ {0} X µ ˆι Composing the named maps with the first projections pr : g X g and pr : g X µ g gives the named maps in the following diagrams: π g π g Ñ {0} X ˆπ π g π g µ Ñ µ {0} X µ ˆπ The morphism π : g g here is part of Grothendieck s simultaneous resolution of the quotient of g by the adjoint action of G := GL d (C). It is a projective, hence proper, morphism which is small in the sense of Goresky-Macpherson (see e.g. [J, Lemma 13.2]). Similarly the map π : g µ g in the parabolic case is small. Hence Rπ C g [d 2 ] and Rπ C gµ [d 2 ] are (monodromic) perverse sheaves on g, which we call the Grothendieck sheaf and the partial Grothendieck sheaf, respectively. There is a natural action of the Weyl group S d on the Grothendieck sheaf. To recall some of the details of the construction of this action, let g rs denote the set of regular semisimple elements in g, and set g rs := π 1 (g rs ). Let π rs be the restriction of π to g rs. Then π rs : g rs g rs is a regular covering whose automorphism group is canonically identified with the Weyl group S d ; see e.g. [G2, Proposition 9.3]. So S d acts freely on g rs by deck transformations, and there is a unique isomorphism between g rs and the quotient S d \ g rs making the following diagram commute: g rs π rs g rs can S d \ g rs. (4.1)

20 JONATHAN BRUNDAN AND VICTOR OSTRIK The action of S d on g rs induces an action on the local system π rs C g rs. Let IC be the intersection cohomology functor from local systems on g rs to perverse sheaves on g. Applying it to π rs C g rs, we get an induced action of S d on the perverse sheaf IC(π rs C g rs). Finally, as in [G2, Corollary 9.2] or [J, Theorem 13.5], the smallness of π implies that there is a unique isomorphism such that the composition π rs C g rs κ : IC(π rs C g rs) Rπ C g [d 2 ] (4.2) H d2 (IC(π rs C g rs)) g rs H 0 (Rπ C g ) g rs π rs C g rs is the identity, where the first and last isomorphisms are canonical and the middle one is induced by κ. Using the isomorphism κ, we transport the action of S d to get the desired action on Rπ C g [d 2 ]. It is important to note that this action induces an algebra isomorphism CS d End(Rπ C g [d 2 ]). (4.3) This follows by the perverse continuation principle (see e.g. [G2, Proposition 4.5]) as the action on π rs C g rs obviously defines an isomorphism CS d End(π rs C g rs). The partial Grothendieck sheaf was studied in detail by Borho and Macpherson, who explained how to identify it with S µ -invariants in the Grothendieck sheaf; see especially [BM2, Proposition 2.7(a)]. We need to reformulate this result slightly. Consider the map p : g g µ arising from the restriction of the map id p : g X g X µ. Applying Rπ to the unit of adjunction Id p p 1 evaluated at C gµ, we get a map BM : Rπ C gµ [d 2 ] Rπ C g [d 2 ]. (4.4) Note also as π 1 ({0}) = {0} X = X and π 1 ({0}) = {0} X µ = Xµ that the proper base change theorem gives us canonical isomorphisms H (X, C) = (Rπ C g ) 0 and H (X µ, C) = (Rπ C g ) 0 between H (X, C) and H (X µ, C) (viewed here as a complexes of vector spaces rather than graded vector spaces) and the stalks of the Grothendieck and partial Grothendieck sheaves at the origin. Moreover, under the isomorphism H (X, C) = (Rπ C g ) 0, the classical action of S d on H (X, C) agrees with the action of S d on (Rπ C g ) 0 induced by its action on the Grothendieck sheaf; see e.g. [J, Lemma 13.6]. Theorem 4.2. The morphism BM just defined gives an isomorphism of perverse sheaves BM : Rπ C gµ [d 2 ] Rπ C g [d 2 ] Sµ-inv. Moreover, letting BM 0 denote the induced map between stalks at the origin, the following diagram commutes: H (X µ, C) can (Rπ C gµ ) 0 p H (X, C) can BM 0 (Rπ C g ) 0 (4.5)

SPALTENSTEIN VARIETIES 21 Proof. The second statement is clear from the definition (4.4) and naturality of proper base change. For the first statement, the idea is to compare our morphism BM with the isomorphism Rπ C gµ [d 2 ] Rπ C g [d 2 ] Sµ-inv constructed in [BM2, Proposition 2.7(a)]. Let S µ \ g rs denote the quotient of g rs by the action of the parabolic subgroup S µ. Let g rs µ := π 1 (g rs ) and π rs be the restriction of π to g rs. The key point is that there is a unique isomorphism g µ rs S µ \ g rs making the following diagram g rs µ π rs g rs p rs g rs can S µ \ g rs can S d \ g rs commute, where the bottom isomorphism comes from (4.1) and p rs is the restriction of p. This isomorphism induces an isomorphism of local systems π rs C g rs µ ( π rs Sµ-inv C g rs). Now apply the functor IC to get an isomorphism of perverse sheaves IC(π rs C g rs ) IC(π rs µ C g rs) Sµ-inv. The Borho-Macpherson isomorphism Rπ C gµ [d 2 ] Rπ C g [d 2 ] Sµ-inv is obtained from this by composing on the right with the inverse of the isomorphism κ from (4.2) and on the left with the analogously defined isomorphism κ : IC(π rs C g rs ) µ Rπ C gµ [d 2 ] from [BM2, Proposition 2.6]. To complete the proof it remains to observe that our map BM is equal to the map from the previous paragraph. This follows by the perverse continuation principle, since the two maps agree on restriction to g rs. As in the statement of Lemma 4.1, the units of adjunction associated to the maps ῑ : Ñ g X and ˆι : {0} X g X induce canonical maps res : C g X ῑ CÑ and res : ˆι C {0} X ι C g. Applying the derived functor R pr, we get morphisms α : C g H (X, C) R π CÑ, (4.6) β : Rπ C g δ 0 H (X, C), (4.7) where we are viewing H (X, C) as a graded vector space with H i (X, C) in degree i, δ 0 denotes the constant skyscraper sheaf on g supported at the origin, and we have implicitly composed with the canonical isomorphisms R pr C g X = C g H (X, C), R pr (ῑ CÑ ) = R π CÑ, R pr (ˆι C {0} X ) = Rˆπ C {0} X = δ0 H (X, C) and R pr (ι C g ) = Rπ C g. Also let γ : DF(δ 0 ) H (X, C) DF(Rπ C g ) (4.8) be the morphism obtained from β by applying the composite functor DF, noting that DF(δ 0 H (X, C)) = DF(δ 0 ) H (X, C) if H (X, C) is identified with the

22 JONATHAN BRUNDAN AND VICTOR OSTRIK graded dual of the graded vector space H (X, C) (so H i (X, C) is in degree i). Repeating all this in exactly the same way in the parabolic case, we get analogous morphisms α : C g H (X µ, C) R π Cѵ, (4.9) β : Rπ C gµ δ 0 H (X µ, C), (4.10) γ : DF(δ 0 ) H (X µ, C) DF(Rπ C gµ ). (4.11) Finally note that there exists a unique (up to scalars) isomorphism σ : DF(δ 0 [d 2 ]) C g, (4.12) as follows for example from the top isomorphism in Lemma 4.1 taking V = E = E = g. We fix a choice of such a map. In the regular case the following proposition is essentially [KW, Theorem VI.15.1]. We are repeating some of the details of the proof to make it clear that it extends to the parabolic case. Proposition 4.3. There exist unique isomorphisms τ and τ making the following diagrams commute: DF(δ 0 [d 2 ]) H (X, C) γ DF(Rπ C g [d 2 ]) σ P D τ C g H (X, C)[r] α R π CÑ [r] (4.13) DF(δ 0 [d 2 ]) H (X µ, C) σ P D C g H (X µ, C)[r µ ] γ α DF(Rπ C gµ [d 2 ]) τ R π Cѵ [r µ ] (4.14) Here P D : H i (X, C) H r i (X, C) and P D : H i (X µ, C) H rµ i (X µ, C) are the isomorphisms defined by Poincaré duality. Proof. We just explain the proof for τ, since the argument for τ is similar. For existence, we apply Lemma 4.1 with E = g X = E, V = g and V = Ñ to get a commuting diagram DF(ˆι C {0} X [d 2 ]) C g X [r] DF(res) res DF(ι C g [d 2 ]) ῑ CÑ [r] Now apply R pr, using Verdier duality and the fact that Fourier transform commutes with direct images (see [Bry, Proposition 6.8] or [KW, Theorem III.13.3] in