Topology of the Affine Springer Fiber in Type A

Similar documents
Background on Chevalley Groups Constructed from a Root System

ON A CONNECTION BETWEEN AFFINE SPRINGER FIBERS AND L U DECOMPOSABLE MATRICES

Parameterizing orbits in flag varieties

Fundamental Lemma and Hitchin Fibration

ELEMENTARY SUBALGEBRAS OF RESTRICTED LIE ALGEBRAS

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

Intersection of stable and unstable manifolds for invariant Morse functions

Math 249B. Geometric Bruhat decomposition

EQUIVARIANT COHOMOLOGY. p : E B such that there exist a countable open covering {U i } i I of B and homeomorphisms

The Affine Grassmannian

arxiv:math/ v1 [math.ag] 17 Mar 2005

Lecture on Equivariant Cohomology

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

EQUIVARIANT COHOMOLOGY IN ALGEBRAIC GEOMETRY LECTURE FOUR: LOCALIZATION 1

EKT of Some Wonderful Compactifications

Fundamental Lemma and Hitchin Fibration

Lemma 1.3. The element [X, X] is nonzero.

9. Integral Ring Extensions

IRREDUCIBLE REPRESENTATIONS OF SEMISIMPLE LIE ALGEBRAS. Contents

LECTURE 4: REPRESENTATION THEORY OF SL 2 (F) AND sl 2 (F)

MATH SOLUTIONS TO PRACTICE MIDTERM LECTURE 1, SUMMER Given vector spaces V and W, V W is the vector space given by

SPHERICAL UNITARY REPRESENTATIONS FOR REDUCTIVE GROUPS

Chow rings of Complex Algebraic Groups

SYMMETRIC SUBGROUP ACTIONS ON ISOTROPIC GRASSMANNIANS

Spherical varieties and arc spaces

Combinatorics for algebraic geometers

Review of Linear Algebra

Parabolic subgroups Montreal-Toronto 2018

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

REPRESENTATION THEORY OF S n

Geometry of Schubert Varieties RepNet Workshop

Notation. For any Lie group G, we set G 0 to be the connected component of the identity.

1 Fields and vector spaces

Topics in linear algebra

Counting matrices over finite fields

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

Equivariant cohomology of infinite-dimensional Grassmannian and shifted Schur functions

GEOMETRIC STRUCTURES OF SEMISIMPLE LIE ALGEBRAS

chapter 12 MORE MATRIX ALGEBRA 12.1 Systems of Linear Equations GOALS

Classification of root systems

EXERCISE SET 5.1. = (kx + kx + k, ky + ky + k ) = (kx + kx + 1, ky + ky + 1) = ((k + )x + 1, (k + )y + 1)

Math 210C. The representation ring

LECTURE 11: SOERGEL BIMODULES

THE COHOMOLOGY RINGS OF REGULAR NILPOTENT HESSENBERG VARIETIES IN LIE TYPE A

ALGEBRAIC GEOMETRY I - FINAL PROJECT

arxiv: v4 [math.rt] 9 Jun 2017

Schubert Varieties. P. Littelmann. May 21, 2012

On the geometric Langlands duality

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

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).

FILTERED RINGS AND MODULES. GRADINGS AND COMPLETIONS.

Root systems and optimal block designs

10. Smooth Varieties. 82 Andreas Gathmann

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

Infinite-Dimensional Triangularization

ON THE CLASSIFICATION OF RANK 1 GROUPS OVER NON ARCHIMEDEAN LOCAL FIELDS

EQUIVARIANT COHOMOLOGY IN ALGEBRAIC GEOMETRY LECTURE TWO: DEFINITIONS AND BASIC PROPERTIES

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.

1. Affine Grassmannian for G a. Gr Ga = lim A n. Intuition. First some intuition. We always have to rst approximation

Citation Osaka Journal of Mathematics. 43(2)

Definitions. Notations. Injective, Surjective and Bijective. Divides. Cartesian Product. Relations. Equivalence Relations

Category O and its basic properties

PERMUTATION REPRESENTATIONS ON SCHUBERT VARIETIES arxiv:math/ v2 [math.rt] 4 Jun 2007

Combinatorial models for the variety of complete quadrics

On the Splitting of MO(2) over the Steenrod Algebra. Maurice Shih

MODEL ANSWERS TO THE FIRST HOMEWORK

Lecture 1. Toric Varieties: Basics

Margulis Superrigidity I & II

The Hurewicz Theorem

Codimensions of Root Valuation Strata

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

YOUNG TABLEAUX AND THE REPRESENTATIONS OF THE SYMMETRIC GROUP

The tangent space to an enumerative problem

DIVISORS ON NONSINGULAR CURVES

Rigid monomial ideals

ON NEARLY SEMIFREE CIRCLE ACTIONS

KAZHDAN LUSZTIG CELLS IN INFINITE COXETER GROUPS. 1. Introduction

Vector Space Basics. 1 Abstract Vector Spaces. 1. (commutativity of vector addition) u + v = v + u. 2. (associativity of vector addition)

Equality of P-partition Generating Functions

Notes 10: Consequences of Eli Cartan s theorem.

Noetherian property of infinite EI categories

REPRESENTATIONS OF S n AND GL(n, C)

Tropical Constructions and Lifts

RELATION SPACES OF HYPERPLANE ARRANGEMENTS AND MODULES DEFINED BY GRAPHS OF FIBER ZONOTOPES

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

LECTURE 11: SYMPLECTIC TORIC MANIFOLDS. Contents 1. Symplectic toric manifolds 1 2. Delzant s theorem 4 3. Symplectic cut 8

Notes on p-divisible Groups

arxiv: v1 [math.rt] 14 Nov 2007

ALGEBRA EXERCISES, PhD EXAMINATION LEVEL

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

FINITE GROUP THEORY: SOLUTIONS FALL MORNING 5. Stab G (l) =.

Equivalence Relations

ALGEBRAIC GROUPS J. WARNER

Math 249B. Nilpotence of connected solvable groups

Introduction to affine Springer fibers and the Hitchin fibration

Radon Transforms and the Finite General Linear Groups

NOTES ON POINCARÉ SERIES OF FINITE AND AFFINE COXETER GROUPS

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

First we introduce the sets that are going to serve as the generalizations of the scalars.

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.

Transcription:

University of Massachusetts Amherst ScholarWorks@UMass Amherst Doctoral Dissertations Dissertations and Theses 26 Topology of the Affine Springer Fiber in Type A Tobias Wilson University of Massachusetts - Amherst Follow this and additional works at: https://scholarworks.umass.edu/dissertations_2 Part of the Algebraic Geometry Commons Recommended Citation Wilson, Tobias, "Topology of the Affine Springer Fiber in Type A" (26). Doctoral Dissertations. 6. https://scholarworks.umass.edu/dissertations_2/6 This Open Access Dissertation is brought to you for free and open access by the Dissertations and Theses at ScholarWorks@UMass Amherst. It has been accepted for inclusion in Doctoral Dissertations by an authorized administrator of ScholarWorks@UMass Amherst. For more information, please contact scholarworks@library.umass.edu.

TOPOLOGY OF THE AFFINE SPRINGER FIBER IN TYPE A A Dissertation Presented by TOBIAS WILSON Submitted to the Graduate School of the University of Massachusetts Amherst in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY February 26 Department of Mathematics and Statistics

c Copyright by Tobias Wilson 26 All Rights Reserved

TOPOLOGY OF THE AFFINE SPRINGER FIBER IN TYPE A A Dissertation Presented by TOBIAS WILSON Approved as to style and content by: Alexei Oblomkov, Chair Tom Braden, Member Julianna Tymoczko, Member Andrew McGregor Computer Science, Outside Member Farshid Hajir, Department Head Mathematics and Statistics

ACKNOWLEDGEMENTS I would like to acknowledge my indebtedness to my advisor, Alexei Oblomkov, for his patience, guidance, and advice throughout the past 5 years. His assistance, suggestions, explanations, and re-explanations made this work possible. I am extremely grateful to Tom Braden, Julianna Tymoczko, and Andrew McGregor for many helpful conversations about my work. I have received a very great deal of support from family and friends while in graduate school. Fellow mathematicians Nico, Luke, Jeff, Steve, Jenn, and Tom provided much needed commiseration, advice, and occasional distraction. Laura and Nicky, always integral parts of my educational career, offered regular sanity checks. Cat and Rosie have given me near-daily encouragement and, towards the end, a second home. My parents have been a constant source of support and advice. I am indebted to all of them, and others. And finally, I owe endless thanks to Andy, who has been my favorite part of the last 6 years. iv

ABSTRACT TOPOLOGY OF THE AFFINE SPRINGER FIBER IN TYPE A FEBRUARY 26 TOBIAS WILSON M.S., UNIVERSITY OF MASSACHUSETTS AMHERST Ph.D., UNIVERSITY OF MASSACHUSETTS AMHERST Directed by: Professor Alexei Oblomkov We develop algorithms for describing elements of the affine Springer fiber in type A for certain γ g(c[[t]]). For these γ, which are equivalued, integral, and regular, it is known that the affine Springer fiber, X γ, has a paving by affines resulting from the intersection of Schubert cells with X γ. Our description of the elements of X γ allow us to understand these affine spaces and write down explicit dimension formulae. We also explore some closure relations between the affine spaces and begin to describe the moment map for the both the regular and extended torus action. v

TABLE OF CONTENTS Page ACKNOWLEDGEMENTS............................... iv ABSTRACT........................................ v LIST OF TABLES.................................... viii LIST OF FIGURES.................................... CHAPTER. INTRODUCTION................................ Affine Springer Fibers...........................2 Main Results and Organization.................... 2 2. AFFINE GRASSMANNIANS AND AFFINE SPRINGER FIBERS..... 3 2. The Affine Grassmannian........................ 3 2.. Basic Definitions......................... 3 2..2 Lattices.............................. 4 2..3 Torus Action........................... 5 2.2 Affine Springer Fiber.......................... 5 2.3 Equivariant Cohomology........................ 7 2.4 Equivariant Cohomology for the Affine Springer Fiber....... 9 3. DESCRIPTION OF THE AFFINE SPACES.................. 3. Lattice Notation............................. 3.2 Two-Dimensional Case......................... 3.3 Three-Dimensional Case........................ 3 3.4 General Case............................... 8 3.5 Dimension................................ 2 3.6 Lattices and Schubert Cells....................... 22 4. CLOSURE RELATIONSHIPS......................... 24 4. Basic Definitions............................. 24 4.2 Two Dimensional Case......................... 25 4.2. Lattice Decompositions..................... 25 4.2.2 Motivation............................ 26 4.2.3 Forming K............................ 27 4.2.4 Closure Picture.......................... 29 4.3 Three Dimensional Case........................ 3 ix vi

4.3. Constructing Lattices From Gr(K, m) t,γ............ 3 4.3.2 Summarizing Lattices...................... 34 4.4 Summarizing Dimensions and Closure Relations.......... 35 4.5 General Case............................... 37 5. ONE-DIMENSIONAL ORBITS AND MOMENT GRAPHS........ 42 5. Zero and One-Dimensional Orbits.................. 42 5.2 Moment Graph for n = 2 and n = 3.................. 44 5.2. Index Lattices......................... 44 5.2.2 Index and 2 Lattices...................... 48 5.2.3 One-Dimensional Orbits.................... 5 6. DIRECTIONS FOR FUTURE WORK..................... 54 APPENDICES A. POSSIBLE LATTICE TYPES WHEN n = 3.................. 56 B. CLOSURE IN 3 DIMENSIONAL CASE................... 66 BIBLIOGRAPHY..................................... 74 vii

LIST OF TABLES Table Page. Lattice Subsets In Generic Closure....................... 37 viii

LIST OF FIGURES Figure Page. Two-dimensional lattices and closures..................... 3 2. Sub-varieties in the closure of a generic lattice variety............ 37 3. Vertices of lattices with index labeled by minimum degree......... 45 4. Arrangement of the vertices, colored by type.................. 46 5. Arrangement of the vertices, colored by dimension.............. 47 6. Index lattices, with vertices colored by type.................. 49 7. Index lattices, with vertices colored by dimension.............. 49 8. Index 2 vertices, colored by type......................... 5 9. Index 2 vertices, colored by dimension..................... 5. One-dimensional orbits for widely spaced degree tuples........... 52. All one-dimensional orbits in index...................... 53 ix

C H A P T E R INTRODUCTION. Affine Springer Fibers The study of affine Springer fibers, originally defined by Kazhdan and Lusztig in [5] was motivated by the theory of Springer fibers, which enjoy some particularly useful properties. Let G be a connected semisimple algebraic group over an algebraically closed field and fix T, a maximal torus and B, a Borel subgroup containing T. For γ g, the Lie algebra of G, the Springer fiber is Y γ = {g G/B Ad(g (γ) b}, the fiber of γ in the Grothendiek simultaneous resolution. The connected components of Y γ are indexed by a quotient of the Weyl group; when γ is regular and semisimple, the connected components are indexed by the Weyl group itself. The Weyl group also acts on the cohomology of Y γ Kazhdan and Lusztig extended the study of Springer fibers by considering their analogue over local function fields F = k((t)). Working within the loop group, G(F ), they defined the affine Springer fiber X γ as a sub-ind scheme of the affine Grassmannian. Lusztig showed in [7] that there is an action of the affine Weyl group W = Hom(k, t) W on the cohomology of X γ. The theory of affine Springer fibers was motivated in part by attempts to prove the fundamental lemma, because orbital integrals have a connection to the number of points of affine Springer fibers. The method used by Laumon and Ngô in [6] relied on embedding affine Springer fibers into Hitchin fibers. Despite this, the geometry of the affine

Springer fiber remains less understood than the geometry of the finite Springer fibers..2 Main Results and Organization We will take F = C((t)) with ring of integers o = C[[t]] and consider the affine Grassmannian X = G(F )/G(o). For γ = diag{γ t, γ 2 t,..., γ n t} g(o), γ i distinct, we will consider the affine Springer fiber X γ. This choice of γ was originally motivated by the study of the Hilbert scheme of points for curves of the form y m = x k. In this case, it was proved in [2] that X γ has a paving by affine spaces, gotten by intersecting the Schubert cells of X with X γ. The closure relations of the Schubert cells in X is well known and are given by the Bruhat order, but similar relations have not been found for affine Springer fibers. In this thesis, I attempt to understand this paving by developing methods to compute explicit bases for elements of the affine paving and then computing the closure relationships between the spaces. In Chapter 2, I present background material on the affine Grassmannian and affine Springer fiber. In Chapter 3, I construct a decomposition of X γ into affine varieties consisting of sets of lattices, indexed by elements of the lattice of translations. These varieties are easy to describe and have a straightforward dimension formula. In Chapter 4, I develop a method that, given one of the affine varieties, uses a Grassmannian to identify a subset of lattices containing the closure. Conjecturally, this subset of lattices is precisely the closure of the variety. In Chapter 5, using the bases from Chapter 3, I identify all the and dimensional orbits of the extended torus action on X γ and consider the moment graph of X γ for n = 2 and n = 3. Directions for future work appear in Chapter 6 and a few longer computations appear in the appendices for reference. 2

C H A P T E R 2 AFFINE GRASSMANNIANS AND AFFINE SPRINGER FIBERS In this chapter, we recall the definition and basic properties of affine Grassmannians and affine Springer fibers. We also present the pertinent results about the affine Springer fiber that will be used in the sequel, as well as the definition of equivariant cohomology. 2. The Affine Grassmannian 2.. Basic Definitions Let F = C((t)), with ring of integers o = C[[t]]. Let G be a reductive group over C with Lie algebra g. In particular, we will work in type A and take G = SL n or G = GL n. Let T be a maximal torus in G, with Lie algebra t and B be a Borel subgroup containing T. Define the Weyl group W of G to be N(T )/T, where N(T ) is the normalizer of T. Since we are considering GL n and SL n, we will always take T to be the diagonal matrices and B to be the upper triangular matrices. The Weyl group in both cases is S n, the symmetric group on n elements. Denote G(F ) as G and define g(f ) = g C F and g(o) = g C o. The affine Grassmannian X is the quotient G/G(o). Let L = T (F )/T (o) be the lattice of translations. In the case of G = GL n, this is just the set of matrices of the form diag{t a,..., t an }, a i Z. For G = SL n, we have the additional restriction that a i =. 3

X is an ind-scheme, and can thus be written as n X n, where X n is a scheme. There are multiple characterizations of the affine Grassmannian; in the case of GL n and SL n, we can express the elements of X n in terms of lattices. 2..2 Lattices We define Λ st = o n to be the standard lattice. It is trivially an o module; we will view it as a submodule of F n. Definition 2.. ([4]) A lattice Λ F n is an o n submodule so that. There exists N Z so that t N Λ st Λ t N Λ st 2. t N Λ st /Λ is locally free of finite rank over C. More concretely, a full rank lattice Λ is the o span of n vectors {v,..., v n } in F n. For G = GL n, the affine Grassmannian is precisely the set of full rank lattices. In the case of G = SL n, the affine Grassmannian is the set of full rank lattices such that k Λ = Λ st. Given some x X, we can associate x to the lattice xλ st. For the other direction, given a lattice Λ, choose any basis {v,..., v n } and let x X be the matrix with columns {v,..., v n }. Any other choice of basis amounts to right multiplying x by some g G(o), so this is well defined. These lattices also allow us to see the ind-scheme structure of X. For N, let X N be the set of lattices such that t N Λ st Λ t N Λ st. It is shown [4] that X N is a projective scheme, with X = X N. Given a lattice Λ, one particularly useful piece of data is the minimum degree tuple. Definition 2.2. For a lattice Λ X, let d i be the minimum degree for which e i t d i +higher degree terms appears in Λ. For Λ with minimum degree tuple (d,..., d n ), we say the index of Λ is N = i d i. 4

2..3 Torus Action There is an evaluation map from G(o) G(C) sending g to g(). The Iwahori subgroup, I, corresponding to the Borel subgroup is the preimage of B under this map. We have the Bruhat decomposition of X into Schubert cells: X = l L IlG(o) Let l L be the diagonal matrix with entries {t d,..., t dn }. In terms of lattices, the Schubert cell corresponding to l is the set of all lattices with minimum degrees (d,..., d n ). We can define a torus action on X by T (C), acting on X by left multiplication. We will also consider the action of C on F given by scaling t: λ t m = λ m t m. On G(F ), this action preserves G(o), so we can define the action of this torus on X. Additionally, the scaling or rotation action commutes with the action of T (C), so we can define the extended torus T = T (C) C, which acts on X by (g, λ) xg(o) = λ (gxg )G(o). Lemma 2.3. ([2]) The fixed point set of T (C) is Lx, where x is the identity in X. 2.2 Affine Springer Fiber For any element γ g(o), we can define the affine Springer fiber X γ = {g X : Ad(g )(γ) g(o)}. We wish to use lattices to study the affine Springer fiber; to do so, we redefine X γ in terms of lattices. Proposition 2.4. It is equivalent to define X γ = {Λ X : γλ Λ} Proof. Suppose g X γ, so g γg = x g(o). Define Λ = gλ st and consider γλ = γgλ st. Since γg = gx, x g(o), the columns of γg are in the o span of the columns of g, so 5

γλ Λ. Therefore every element of X γ corresponds to lattice Λ such that γλ Λ. Similarly, if γλ Λ, choose any basis for Λ and form gg(o) X. Since γλ = Λx for some x g(o), g γg = x g(o). For arbitrary γ, X γ may be empty, a scheme, or an ind-scheme. We will restrict our consideration to γ that satisfy a number of definitions. We say that γ g(o) is semisimple if ad γ : g(o) g(o) is diagonalizable. A semisimple element is regular if its centralizer is a maximal torus. In the case of G = GL n and SL n, an element will be regular if it has distinct eigenvalues. Due to Kazhdan and Lusztig, we have the following result in the regular semisimple case: Proposition 2.5. ([5]) For γ g(o), X γ will be finite dimensional if and only if γ is regular and semisimple. Given γ, define T γ to be the centralizer of γ in G and let Hom F (C, T γ ) be the cocharacter lattice. Theorem 2.6. ([5]) Let γ g(o) be regular and semisimple. Then Hom F (C, T γ ) acts freely on X γ and Hom F (C, T γ )\X γ is a projective scheme. We say that γ g(o) is integral if it is diagonalizable over F and equivalued if, in addition to being diagonalizable, all of the eigenvalues have the same valuation. In [CITE GKM purity], Goresky, Kottwitz, and MacPherson show that for regular integral equivalued γ, X γ admits a paving by Hessenberg varieties. In the case where T is weakly Coxeter, however, they proved a stronger result. The G conjugacy classes of maximal tori are indexed by the conjugacy classes in W ; if the conjugacy class associated to T contains only Coxeter elements, we say that T is Coxeter in G. We say that T is weakly Coxeter in G if it is Coxeter in M, the centralizer of the maximal split torus in T. Since all split maximal tori - and all maximal tori in GL n are weakly Coxeter, we will always be considering weakly Coxeter T. 6

Theorem 2.7. ([2]) Suppose that T (F ) is weakly Coxeter and let γ be a regular integral equivalued element of t(o). Then X γ admits a paving by affine spaces. These affine spaces are the intersection of the Schubert cells from the Bruhat decomposition with X γ. In the following chapter, we will construct the affine spaces by describing bases of the lattices in each space. 2.3 Equivariant Cohomology Because X γ has a torus and extended torus action, we are interested in studying the equivariant cohomology of X γ. We will briefly recall the definition of equivariant cohomology here, following [], as well as the relevant results in the case of X γ. The equivariant cohomology of any T -variety Y is defined by considering a contractible space ET with a free T action and defining BT = ET/T to be the classifying space of T. Let Y T ET = (Y ET )/T, where T acts on Y ET diagonally. This creates a fiber bundle over BT with fiber Y. We define the T equivariant cohomology H T (Y ) := H (Y T ET ). We will use two facts about equivariant cohomology in particular; first, that as in usual cohomology, HT (Y ) is a module over H T (pt) and second, that H T (pt) = S(t ), the symmetric ring of the dual lie algebra of T. This is especially helpful for certain T actions on Y. Definition 2.8. ([]) A variety Y is equivariantly formal with respect to the action of T if E 2 = E in the Leray spectral sequence associated to Y Y T ET BT. If Y is equivariantly formal with respect to T, HT (Y ) is in fact a free H T (pt) module. In [], Goresky, Kottwitz, and MacPherson give an algorithm for computing H T (Y ) when Y is equivariantly formal and satisfies two additional conditions. Proposition 2.9. ([]) If Y is equivariantly formal with respect to T and Y has finitely many T fixed points and finitely many one-dimensional T orbits, then the closure of each one-dimensional 7

orbit contains exactly two fixed points and is isomorphic to CP, with the fixed points appearing at the poles. The torus acts on each one-dimensional orbit by rotation. For each one-dimensional orbit, there is a codimension one subtorus that fixes the orbit pointwise. In this case, the fixed points and one dimensional orbits give combinatorial data that allow for direct computation of HT (Y ). When Y is equivariantly formal and has finitely many fixed points, the inclusion map Y T Y of the fixed points into Y induces a map H T (Y ) T fixed points S(t ). This map is actually an injection []; GKM give an explicit presentation for the image. Theorem 2.. [] Given Y and T so that Y is equivariantly formal with respect to the T action and has finitely many T fixed points and one-dimensional orbits, denote the one-dimensional orbits O,..., O n, each with poles N i and S i and stabilizer T i in T. Then H T (Y ) = {(f, f 2,..., f m ) T fixed points S(t ) f Ni ti = f Si ti for all i n}, as rings. Thus, in the equivariantly formal case, the equivariant cohomology can be determined via the T fixed points and one-dimensional orbits of T in Y. This data can be represented by the moment graph, with the T fixed points as vertices and one-dimensional orbits as edges between their poles. This moment graph is a linear graph in the image of the moment map, µ : Y t. Under the moment map, the fixed points will be sent to points in t. Each one dimensional orbit O i will be sent to the line segment between the images of its poles. The direction of this line segment is the annihilator of t i, the lie algebra of the codimension one subtorus that stabilizes O i. 8

2.4 Equivariant Cohomology for the Affine Springer Fiber Thus far, we have described computing equivariant cohomology for varieties Y that, among other properties, have finitely many T fixed points. In the case of the affine Grassmannian and affine Springer fiber, however, that assumption immediately fails. Taking T and T as defined above, the T and T fixed points of X and X γ are the elements corresponding to l L, the lattice of translations, or diagonal matrices of the form diag(t d, t d 2,..., t dn ). Following [], we can extend the definition of equivariant formality to an ind-scheme with the following assumptions. Let X = X N be an indscheme with an action of T so that X N is preserved under T and the action of T on X N is equivariantly formal and further assume that each X N has finitely many T fixed points and H 2 (X) is finite dimensional. It is shown in [2] that for our choice of γ, these assumptions are met, and so X γ is equivariantly formal. 9

C H A P T E R 3 DESCRIPTION OF THE AFFINE SPACES We will first write down explicit bases for elements in the affine spaces, which will be the intersection of the Schubert cells with X γ. We will always take γ g(o) to be the diagonal matrix with entries {γ t, γ 2 t,..., γ n t}, with all γ i distinct. Since γ is diagonal and has distinct eigenvalues, it is regular and semisimple. In this chapter, we will construct valid lattices in X γ, then identify them with the Schubert cells of the affine Grassmannian. We will do all computations for G = GL n, since the lattices for G = SL n will appear as the subset of lattices with index. 3. Lattice Notation Let Λ = span{v, v 2,, v n } be a C[[t]] lattice, with the v i ordered by non-decreasing initial degree, d i. Elements of Λ are n-dimensional vectors with entries in F or, equivalently, Laurent series with n-dimensional complex vectors as coefficients. We begin by performing row reduction on the v i until the following two conditions hold:. The first non-zero entry in the lowest degree coefficient vector of v i is. Let a i be the position of this entry. Thus we can write v i = e ai t d i + higher degree terms. 2. The a th i equal to d i. position is in all the other coefficient vectors with degree greater than or

We can collect the data above into two tuples: let d = ( d, d 2,, d n ) be the initial degrees of the v i and let σ = (a, a 2,, a n ) be the positions of the first non-zero entry in the v i. We choose the notation σ to reflect that it is an element of the symmetric group S n, a fact which we will exploit later. In order for σ to be well defined, we need the convention that we order the v i so that if v i and v i+ have the same initial degree, then a i < a i+. For a given lattice Λ, we will refer to d as the ordered degree tuple and σ as the initial arrangement. We will first consider GL 2, which is a short computation. We will then consider the case of GL 3, which is more complex. Based on the computation for GL 3, we can describe the general case. 3.2 Two-Dimensional Case Consider the lattice Λ = span{v, v 2 }. There are two possible cases to consider, σ = (, 2) and σ = (2, ). First, suppose that, after performing row reduction, we can write the basis as v = t d + t d+ + + c c v 2 = t e, c e d t e, where e d. In order to have γλ Λ, we must have γv span{tv, v 2 }. We have γv = γ t d+ + t d+2 + + t e, γ 2 c γ 2 c γ 2 c e d so we can subtract γ tv yielding γv γ tv = t d+ + t d+2 + + t e. (γ 2 γ )c (γ 2 γ )c (γ 2 γ )c e d

Since this must now be in the span of v 2, we see that all of the c i must be except for c e d, giving us a general form for the lattice: v = t d + t e, v 2 = t e. c The only exception to this form occurs when d = e. In this case, Λ = t d Λ st and there is no constant c. The second possibility is that after row reduction, the basis can be written as v = t d + c t d+ + + c e d t e, v 2 = t e where e > d. This case proceeds exactly as in case, except when e = d +. In the general case, we have lattices of the form v = t d + c t e, v 2 = t e. When e = d +, however, c would occupy the first position in degree d - which has already been assumed to be. Therefore, when e = d +, there is only one possible lattice, with basis v = t d, v 2 = t d+. From this case, we see that in general, the structure of the lattices does not depend on the initial arrangement. The lattices with σ = (2, ) are simply the lattices with σ = (, 2) after permutation by (2) S 2. When the initial degrees only differ by, however, we lose the free variable. To distinguish these cases, we will call a lattice in which at least two of the initial degrees differ by exactly a close case and refer to other lattices as widely spaced. Note that a widely spaced lattice can have the same initial degree occur 2

more than once; e.g, in n = 3, we would say a lattice with initial degrees ( 3, 3, 6) was widely spaced, while (,, ) was a close case. 3.3 Three-Dimensional Case We will first consider the case where all initial degrees are distinct and widely spaced. For simplicity, we will assume that σ = (, 2, 3). Lattices with other initial arrangements can be formed via the action of S n. Let Λ = span{v, v 2, v 3 } and perform row reduction so that v = a td + a td+ + a 2 td+2 + + b + b te + b 2 te+ + + a e d 2 b e d 2 te 2 + a e d tf 2 + b e d b e d b e d+ b f d 2 b f d c c c f e 2 c f e te v 2 = te + te+ + + tf 2 + tf, v 3 = tf (3.) tf Let γ = γ γ 2 t, γ, γ 2, γ 3 distinct and non-zero γ 3 Claim 3.. We must have γv i span C[t] {v i, v i+,..., v n } because by performing row reduction, we know that v i has a in all positions where v,..., v i has a pivot. Given that, we can work backwards: 3

γv 2 = γ 2 γ 3 c te+ + γ 3 c te+2 + + γ 3 c f e 2 tf + γ 3 c f e tf γv 2 γ 2 tv 2 c(γ 3 γ 2 )v 3 = γ 3 c γ 2 c te+ + γ 3 c γ 2 c te+2 + + γ 3 c f e 2 γ 2 c f e 2 tf We can no longer use any multiples of v 2 without reintroducing a non-zero second coordinate and we can t cancel any of the 3rd coordinates with v 3 because the degrees are too low. So if γv 2 span{v 2, v 3 } then we must have that c (γ 3 γ 2 ) = c (γ 3 γ 2 ) = = c f e 2 (γ 3 γ 2 ) = and since γ 3 and γ 2 are distinct, that gives that c = = c f e 2 =. Therefore, v 2 = te + c f e tf Now consider γv : γv = γ 2 a γ 3 b td+ + γ 2 a γ 3 b td+2 + γ 2 a 2 γ 3 b 2 td+3 + + γ 2 a e d 2 γ 3 b e d 2 te + γ 2 a e d γ 3 b e d te + γ 3 b e d te+ + γ 3 b e d+ te+2 + + γ 3 b f d 2 tf + γ 3 b f d tf (3.2) 4

γv γ tv = a (γ 2 γ ) td+ + a (γ 2 γ ) td+2 + a 2 (γ 2 γ ) td+3 + + a e d 2 (γ 2 γ ) b (γ 3 γ ) b (γ 3 γ ) b 2 (γ 3 γ ) b e d 2 (γ 3 γ ) + a e d (γ 2 γ ) te + te+ b e d (γ 3 γ ) b e d (γ 3 γ ) + te+2 + + tf + tf (3.3) b e d+ (γ 3 γ ) b f d 2 (γ 3 γ ) b f d (γ 3 γ ) There are no other terms in degrees e +,..., f 2 that we can use to cancel without reintroducing the first or second coordinate, so b e d = = b f d 3 =. te γv γ tv a e d (γ 2 γ )v 2 b f d (γ 3 γ )v 3 = a (γ 2 γ ) td+ + a (γ 2 γ ) td+2 + a 2 (γ 2 γ ) td+3 + + a e d 2 (γ 2 γ ) b (γ 3 γ ) b (γ 3 γ ) b 2 (γ 3 γ ) b e d 2 (γ 3 γ ) + te + tf (3.4) b e d (γ 3 γ ) b f d 2 (γ 3 γ ) a e d c f e (γ 2 γ ) te Now there are no vectors we can use in the remaining degrees without reintroducing a first or second coordinate, so all the remaining coordinates must be. Thus a =... a e d 2 = b = = b e d =, b f d 2 (γ 3 γ ) a e d c f e (γ 2 γ ) = This last condition can be rewritten as: b f d 2 = a e d c f e γ 2 γ γ 3 γ 5

Letting x = γ 2 γ γ 3 γ and removing unnecessary subscripts, the basis can now be written: v = td + a te + acx tf 2 + b tf v 2 = te + c tf, v 3 = tf By acting on this basis by σ S 3, we can write down bases for the lattices with the same initial degrees, but other initial arrangements. We need to consider, however, the action of σ on x. We defined x above to be γ 2 γ γ 3 γ, but for other σ S 3, performing the same row reduction will create a different constant. Therefore, we define an action of σ on x by permuting the subscripts; in general, we will write the constant γ j γ i γ k γ i as x ijk or x σ. Using this notation, we can easily write down bases for lattices with other initial arrangements. For example, if σ = (2,, 3), Λ has basis v = td + a te + acx 23 tf 2 + b tf v 2 = te + c tf, v 3 = tf The bases for all initial arrangements appear in Appendix. Although the close cases are essentially similar, as in n = 2, some of the constants may be forced to be zero by assumption. For example, consider the case with initial 6

degrees (d, d +, f) and initial arrangement (3, 2, ). As before, we can find v = td + b td + b 2 td 2 + + b f d tf v 2 = td+ + c tf, v 3 = tf Note that by assumption, the first term of v is. Thus we have lost a constant; there is no a in degree (d + ) and therefore also no cax σ. The final simplified form is: v = td + b tf v 2 = td+ + c tf, v 3 = tf The cases that can cause this constant loss are (d, d+, f), (d, e, e+), and (d, d+, d+ 2). The simplified forms of lattices of those types for all initial arrangements appear in Appendix. So far, we have only discussed lattices with distinct initial degrees. We must also consider initial degrees of the form (d, d, e), (d, e, e), and (d, d, d). Note that in these cases, the possible initial arrangements will be only a subset of S 3. The bases in for these cases are listed in Appendix, but we will also summarize some of them here. Clearly if Λ has degree tuple (d, d, d), Λ = t d Λ st. In the other two cases, Λ will have two free variables when the degrees are widely spaced. For example, the widely spaced case with d = (d, d, e) and σ = (, 3, 2) has basis 7

v = td + a te, v 2 = td + b te, v 3 = te. In the close cases, we may lose none, one or both of the free variables, depending on σ. If Λ has d = (d, d, d + ) with σ = (, 3, 2), the basis is v = a td, v 2 = td, v 3 = td+. In general, a free variable is lost any time the higher degree vector also has a higher initial position: if d j = d i + and a j < a i, v i will lose a free variable. 3.4 General Case that Recall that for any Λ, we choose a basis {v,..., v n } and perform row reduction so. The first non-zero entry in the lowest degree coefficient vector of v i is. We denote this position by a i. 2. The a th i equal to d i. position is in all the other coefficient vectors with degree greater than or With these conventions, the condition that γλ Λ requires that for each v i, γv i is in the C[[t]] span of {v i, v i+,, v n }. Since v i is the only basis element with a non-zero entry in the a th i position, we can rewrite this condition as γv i γ ai tv i span C[[t]] {v i+, v i+2,, v n }, or equivalently, (γ γ ai t)v i = j>i c j v j, c j C[[t]]. 8

In fact, once we have performed the above row reduction, we can make a much stronger claim: in order to have γλ Λ, we must have γv i γ ai tv i span C {v i+, v i+2,, v n }. To verify this, note that γ γ ai t = diag{γ γ ai,, γ ai γ ai,, γ ai + γ ai,, γ n γ ai } t acts by scaling each term and increasing the degree by. If γv i γ ai tv i contains a nonzero multiple of t k v j for some j > i, k >, then v i contains a non-zero entry in the a th j position of its d j + k coefficient vector. This, however, contradicts our second row-reduction condition. Therefore, we can reduce γλ Λ to the condition that (γ γ ai t)v i = j>i c j v j, c j C. The row reduction conditions give one further constraint on the v i : in close cases, v i may be required to be linearly independent of some of the v j. Recall that we say ( d, d 2,, d n ) is a close case if for some i, d j = d i +, j > i. Definition 3.2. We say that v j is below v i if:. j > i, and 2. dj > d i, and 3. If d j = d i +, then a i < a j Lemma 3.3. To guarantee that γλ Λ, it is enough to require that for each i, (γ γ ai t)v i = v j below v i c j v j, c j C. Proof. We have already shown that γλ Λ implies that (γ γ ai t)v i span C {v i+,, v n }. Now suppose that v j is not below v i. If γv i γ ai tv i contains a multiple of v j, then v i has a non-zero entry in the a th j position of the degree d j coefficient vector. If v j fails to be below v i because d j = d i, then by assumption, v i cannot have non-zero entries below degree d j. The only other possibility is that d j = d i + and a i > a j. We have reduced v i so that the first non-zero entry is in the a th i position of the degree d i coefficient vector. 9

When a i > a j, the a th j position of the d i coefficient vector is by definition, so γv i γ ai tv i cannot contain a multiple of v j. Let e ai denote the a th i standard basis vector. Note that the kernel of γ γ ai t is span F {e ai }. We can decompose v i into e ai t d i +v i where v i span F {e a,..., e ai, e ai+,..., e an } = e a i, using the standard pairing. While the diagonal matrix γ γ ai t is obviously not invertible, it is invertible on e a i. Since (γ γ ai t)v i = (γ γ ai t)v i, we can consider the action of (γ γ ai t) on e a i to write v i = (γ γ ai t) c j v j This leads to the final v j below v i expression for the basis of Λ. Theorem 3.4. Every element Λ X γ can be expressed by a basis {v,..., v n }, v i = e ai t d i + (γ γ ai t) c j v j v j below v i where σ = (a,..., a n ) is an element of S n and d = ( d,..., d n ) is a non-decreasing sequence of integers. 3.5 Dimension We showed in the previous section that the form of a lattice Λ X γ is determined by two tuples: the initial degrees d = ( d, d 2,, d n ) and the first non-zero positions σ = (a, a 2,, a n ). Let Λ d,σ be the subvariety of all lattices in X γ with degree d and initial arrangement σ. We wish to develop combinatorial formulas for the dimension of Λ d,σ. Although these formulas are trivial given the decomposition in the last section, they introduce notation that will be useful in later sections and make dimension computations easier and more explicit. Definition 3.5. We say a lattice in Λ d,σ is of type (k,..., k l ), where k i is the frequency of the i th distinct d j. Example 3.6. A lattice with degrees (,, 3, 5, 5, 5, 6) is of type (2,, 3, ) 2

l Note that k i = n. The type does not give us any new information, but it will i= assist in computing the dimension of Λ d,σ. When we need to recover the degree of the k i vectors, we will write d(k i ). Proposition 3.7. For Λ d,σ, define the following:. s i = #{v k dk = d i + and a i > a k } 2. S = n s i. i= 3. b i = l j=i+ k j Then the dimension of Λ d,σ is b i k i ) S l ( i= Proof. The proof of this follows directly from the decomposition. We showed above that for each i, v i = e ai + v i, where v i is any element of the vector space spanned by all v j below v i. From this, it is clear that the dimension of Λ d,σ is #{v j below v i } = i i #{v j d j > d i } i #{v j d j = d i +, a j < a i }. Recall that v j is below v i if either d i + < d j or d i + = d j and a i < a j. S is exactly the count of all v j that fail to be below v i for some i < j because of the second condition, so all that remains is to check that #{v j d j > d l i } = b i k i. To verify, note that for each i i= of the k i vectors of degree d(k i ), there are b i vectors of higher degree. Corollary 3.8. The subvariety Λ d,σ is isomorphic to A l, where l = ( l i= b ik i ) S. One benefit of this formula is that it allows us to quickly compute the dimension of any Λ d,σ without writing out the basis. Although clumsy to define, the formula written this way is somewhat more intuitive than counting the number of v j below each v i. The dimension formula is particularly useful in the widely spaced case, where S will be. 2

In this case which is essentially the generic case the dimension of Λ d,σ will depend only on the type of d, not on the specific degrees or on σ. In later chapters, we will be able to visualize how the different types and dimensions interact. Corollary 3.9. The largest cells in X γ correspond to d in which all d i are distinct and widely spaced. In this case, the dimension of Λ d,σ is n(n )/2. Proof. In this case, v j will be below v i for all j > i, so Λ d,σ will have the maximal number of free variables. Notice that for d of this type, σ does not matter. In the notation of the above formula, S =, k i = and b i = n i for all i. Thus, the dimension is n n i = i= n(n ). 2 3.6 Lattices and Schubert Cells Thus far, we have described types of lattices by using two tuples, d = ( d,..., d n ) listing the initial degrees in increasing order, and σ = (a,..., a n ) listing the first nonzero position for each of the v i. This notation is most convenient for writing down explicit bases and computing dimensions of Λ d,σ. In order to associate Λ d,σ to a Schubert cell, we need to develop slightly different notation. Definition 3.. For a lattice Λ X γ with basis {v, v 2,..., v n }, let d i be the initial degree of the basis vector whose first term is e i t d i. Definition 3.. Where appropriate, we will refer to Λ d,σ as Λ d. Example 3.2. For the tuples d = (, 4, 8,,, 2, 8), a = (4, 5,, 3, 7, 6, 2), we have d = (8, 8,,, 4, 2, ). Note that d is just a permutation of d - in fact, d = σ d. In many ways, d is the more natural notation; it contains all the information encoded in d and σ and, as we 22

will see, makes the relationships between the Λ d more clear. It also allows us to see the relationship between Λ d and the Schubert cells: Λ d = IlG(o) X γ, where l L is the diagonal matrix with entries {t d,..., tdn n }. In the following chapters, we will use the d notation almost exclusively, except when writing down concrete bases. It will be helpful to have the following recast definition: Definition 3.3. For d = {d,..., d n }, with corresponding basis {v,..., v n }, v i = e i t d i + v i, where v i consists of higher degree terms, we say that v j is below v i if d j > d i + or if d j = d i + and j > i. 23

C H A P T E R 4 CLOSURE RELATIONSHIPS Given the description of the Λ d from the last chapter, we can begin to look for closure relationships between the affine varieties. For n = 2, we can compute the closure of Λ d by hand and show that the elements of the closure also arise as elements of the t and γ stable points of a Grassmannian, Gr(K, m) t,γ, where K and m depend on d. Based on this, we compute Gr(K, m) t,γ for generic d in n = 3 and prove that for all n, the closure of Λ d will be contained in the image of Gr(K, m) t,γ in X γ. The proof of this is very simple, but involves developing a tool, degree tables, which makes it simple to describe the possible Λ d in the closure of Λ d. 4. Basic Definitions Let X N γ be the lattices of index N in X γ. Consider the variety, Λ d of lattices in X N γ described by the degree tuple d = (d,..., d n ). Recall that we must have d i = N. Let Λ Λ d be a generic element. To determine the closure of Λ d, we define the following two objects: Definition 4.. Let Λ min be the C[[t]] lattice with basis {e i t m i }, where m i is the smallest degree for which a non-zero multiple of e i appears in one of the basis vectors of the general form of Λ. Definition 4.2. We define Λ max to be the C[[t]] lattice with basis {e i t k i }, where k i is the smallest degree for which e i t k i Λ. 24

Note that in general, neither Λ min nor Λ max will be valid lattices in Xγ N. Since Λ max Λ min, we can form the quotient space, K = Λ min /Λ max, which we will consider as a C vector space. Furthermore, we can consider the subspace Λ/Λ max K. Multiplication by t and γ are linear maps on K which preserve Λ/Λ max. Let m = dim C (Λ/Λ max ). We prove in 4.5 that the subvarieties in the closure of Λ d will appear as at least a subset of the elements of Gr(K, m) preserved by t and γ, and conjecturally, all the elements of Gr(K, m) t,γ will correspond to subvarieties in the closure. 4.2 Two Dimensional Case 4.2. Lattice Decompositions Recall that in X γ we have the following cases:. d = ( x, x), σ = (, 2): t x + t x, t x a 2. d = ( x, x), σ = (2, ): t x + a t x, t x 3. d = (, ), σ = (, 2):, In the case of X γ, we have the following varieties:. d = ( x, x + ), σ = (, 2), x : t x + t x, t x+ a 25

2. d = ( x, x + ), σ = (2, ), x : t x + a t x, t x+ 3. d = (, ), σ = (, 2):, t a 4. d = (, ), σ = (2, ):, t For any other values of N, X N γ = X N mod 2 γ so all other possible varieties are isomorphic to those above. 4.2.2 Motivation To compute the closure, we can consider the limit points of curves lying in Λ d. To do so, we introduce an additional variable, s, and consider taking the limit as s. That is, we take F = C[[t]][s] and consider lattices generated by bases of the form v = t x + t x, v 2 = t x a(s) where a(s) is a non-constant polynomial in s. By factoring out the highest power of s, we can write this basis as s k t x + t x, t x. b + b( s ) Now, upon taking the limit as s, we see the basis becomes: t x, t x, b 26

which is spanned by t x when b. This, however, is not a complete lattice. Note that in the original lattice, we also had the element tv a(s)v 2 = t x+. This vector is unaffected by the limit and is the lowest degree element in the lattice. Thus, a complete basis for the lattice is t x+, t x, indicating that the origin of the variety given by d = ( x +, x ), σ = (, 2) is in the closure of d = ( x, x), σ = (, 2). This type of computation informs our definition of Λ min and Λ max. Upon setting up and taking the limit, the smallest degree that can occur in each e i will be the first time a non-zero multiple of e i appears. 4.2.3 Forming K Using the enumeration from above:. d = ( x, x), σ = (, 2): Λ min = { t x, t x } Λ max = { t x+, t x } K = {v = t x, v 2 t x } Then Λ/Λ max = v + av 2, a one dimensional subset of K. Multiplying elements of K by t or γ is equivalent to multiplying by, so Gr(K, ) t,γ = Gr(K, ). The 27

elements of Gr(K, ) can be enumerated as {v +av 2 a C} {v 2 }. Any subspace of the first type, when lifted back to X γ, will form an element of Λ d : { t x + t x } Λ max = { t x + t x, t x } Λ d. a a The final element of Gr(K, ) lifts to the lattice { t x+, t x } Thus, we have recovered the results of the computation in the previous section: Λ d = Λ d { t x+, t x } 2. d = ( x, x), σ = (2, ): This case proceeds essentially identically to the first. We find that K = {v = t x, v 2 t x } and by computing that Gr(K, ) t,γ = {v + av 2 a C} {v 2 }, find that Λ d = Λ d { t x+, t x }. 3. d = (, ), σ = (, 2): Since in this case Λ d is a single point, note that it is already closed. We have that Λ min = Λ = Λ max, so K is the trivial group. For lattices in X γ, the computations are largely identical. The only interesting case is d = (, ), where the closure depends on the arrangement, σ.. d = ( x, x + ), σ = (, 2), x : Λ d = Λ d { t x+, t x } 28

2. d = ( x, x + ), σ = (2, ), x : Λ d = Λ d { t x+, t x } 3. d = (, ), σ = (, 2): Λ min = { t, t } Λ max = { t, t } K = {v =, v 2 } Since Λ/Λ max = v + av 2, we want to consider elements of Gr(K, ) and this case proceeds as above, giving Λ d = Λ d { t, t } 4. d = (, ), σ = (2, ): As for d = (, ) in X γ, Λ d is a single point and K is the trivial group. 4.2.4 Closure Picture In order to summarize and visualize the data above, we can draw a diagram of the Λ d and their closures. We represent lattices of the form { t a, t b } by nodes labeled with d. If Λ d is one-dimensional, we then represent those lattices with an arrow emanating from the appropriate node, terminating at the additional lattice in 29

( 4, 4) ( 3, 3) ( 2, 2) (, ) (, ) (, ) (2, 2) (3, 3) (4, 4) ( 3, 4) ( 2, 3) (, 2) (, ) (, ) (2, ) (3, 2) (4, 3) (5, 4) Figure. Two-dimensional lattices and closures the closure. These nodes and lines are precisely the and dimensional orbits of the torus action discussed in the next chapter. 4.3 Three Dimensional Case 4.3. Constructing Lattices From Gr(K, m) t,γ We will first provide a completely worked example of computing the elements of Gr(K, m) t,γ for generic d. For our example case, let d = (, 4, 8) and σ = (, 2, 3). This choice is arbitrary as long as d is widely spaced and all the d i are distinct. In this case, we have Λ = span{v, v 2, v 3 } with v = + b t3 + bgx 23 t6 + c v 2 = t4 + t7, v 3 = g where x 23 = γ 2 γ γ 3 γ. For simplicity, since it is unambiguous, we will omit the subscript for the duration of this section. t8 t7 3

Then Λ min = span, t3, t6 Λ max = span t2, t5, t8 and K = span u =, u 2 = t, u 3 = t3, u 4 = t4, u 5 = t6, u 6 = t7 Since Λ/Λ max = span{u +bu 3 +bgxu 5 +cu 6, u 2 +bg(x )u 6, u 4 +gu 6 }, we have m = 3. The actions of t and γ on K are given by t =, γ = γ γ 2 γ 3 Instead of working with t and γ, it is equivalent and computationally simpler to consider N = (γ 2 γ ) (γ 2 t γ) and N 2 = (γ 2 γ ) (γ γ t): N = x, N 2 = x 3

Since Gr(K, m) t,γ = Gr(K, m) N,N 2, we can now proceed to describe all possible Γ Gr(K, m) N,N 2 and the lattices they correspond to in X γ. This is a lengthy computation with many possible cases. It appears in full in the appendix, but to give a sense of the process, we will include the first cases here. We proceed by assuming Γ Gr(K, m) N,N 2 and consider the possible dimensions of N Γ and N 2 Γ.. Suppose that dim(n Γ) =, so Γ Gr(3, {u 2, u 3, u 4, u 6 }). (a) If dim(n 2 Γ) =, Γ Gr(3, {u 2, u 4, u 6 }). Thus Γ = {u 2, u 4, u 6 } By lifting this back to X γ, we see that Γ corresponds to the subvariety, in this case just a point, of lattices of the form span t, t4, (b) If dim(n 2 Γ) =, we can choose a basis element t7 v = a 2 u 2 + a 3 u 3 + a 4 u 4 + a 6 u 6, a 3. Then we must have v 2 = N 2 v = a 3 u 4, which leaves v 3 = b 2 u 2 + b 3 u 3 + b 4 u 4 + b 6 u 6. We can then perform row reduction on this basis; note we will generally allow the values of the arbitrary constants to change without remarking upon it. a 2 a 6 b 2 b 6 32

i. If b 2 : a 6 b 6 This corresponds to lattices of the form span t + b 6 t7, t3 + a 6 t7, t8 Note that as a lattice, v 2 is in the span of v and so doesn t appear in the lattice representation- instead, we introduce the last lattice element from the quotient. ii. If b 2 =, b 6 : a 2 gives lattice: span a 2 t + t3, t4, t7 A. If a 2, this becomes: span t + a 2 t3, t4, t7 B. If a 2 =, this is span t2, t3, t7 where we reintroduce the first basis element from the quotient. 33

4.3.2 Summarizing Lattices It is possible to compute the closure of Λ d by hand, similarly to motivation for the 2 dimensional case, but the computation is much longer and contains many subcases. We have performed the computation and it suggests that in the above variety, the lattices arising from Gr(K, m) t,γ are indeed the closure of Λ d, so we proceed under that assumption to describe the closure. We can group the lattices in the closure by degree tuple. To generalize this example to any lattice, we consider how each type has changed from the original; the net change vector appears next to each type. (, 4, 8): Net change from original type: [,,] span + a t3 + acx t6 + b t7, t4 + c t7, t8 (, 5, 7): Net change from original type: [,, -] span + a t4 + b t6, t5 + c t7, t7 (, 3, 8): Net change from original type: [, -, ] span t + a t7, t3 + b t7, t8 (, 4, 7): Net change from original type: [,, -] span t + a t3 + b t6, t4, t7 34

span t + a t6, t4 + b t6, t7 These two lattices give two planes that intersect in a line. (, 5, 6): Net change from original type: [,, -2] span t + a t4, t5, t6 (2, 3 7): Net change from original type [2, -, -] span t2, t3 + a t6, t7 (2, 4, 6) : Net change from original type [2,, -2] span t2, t4, t6 The net change vectors allow us to generalize this case to other degree tuples. In order to generalize it to lattices with other initial arrangements, we will need to act on the net change vector by σ. 4.4 Summarizing Dimensions and Closure Relations To determine how these spaces fit together, we need to look at the closures of these new lattices as well. We do so by acting on each tuple by the net change vectors. For 35

example, in the closure of (, 5, 7), there are lattices of the following types: (, 5, 6), (, 4, 7), (2, 4, 6), (, 6, 6), (, 6, 5), (2, 5, 5) Of those, the first three are also in the closure of (, 4, 8), the original lattice. Some work needs to be done to verify that the lattices in the closure match up. For example, it needs to be checked that the lattices of type (, 5, 6) in the closure of (, 5, 7) overlap with the lattices of type (, 5, 6) in the closure of (, 4, 8). To do so, we see that (, 5, 6) arises in the closure of (, 5, 7) via the vector [,, ]. We see that the lattices arising from [,, ] are of the form: span t + a t4 + b t6, t5, t7 span t + a t6, t5 + b t6, t7 (note the shift in initial degrees to accommodate a (, 5, 7) as the original lattice). This set of lattices contains span t + a t4, t5, t6, the lattices of type (, 5, 6) that arise in the closure of (, 4, 8). It is straightforward to verify that the closure of the subvariety and the closure of the original variety always overlap; containment can go both ways, however. Sometimes, as above, the closure of the subvariety contains the lattices in the closure of the variety. For other cases, the relationship is reversed. We can summarize the lattices, listed by the net change vector, the dimensions of the subset in the closure and the other lattices contained in their closure as follows: 36

Table. Lattice Subsets In Generic Closure Type Dimension In Closure: [,, ] C 3 [,, ] C 2 [,, ], [2,, 2], [, 2] [,, ] C 2 [,, ], [2,, 2], [2,, ] [,, ] C 2, C 2 [,, 2, ], [2,, ], [2,, 2] [,, 2] C [2,, 2] [2,, ] C [2,, 2] [2,, 2] point We can also visualize the closure relationships in Figure 2, first by net change vector and then by dimension. (,, ) C 3 (,, ) (,, ) C 2 C 2 (,, ) C 2, C 2 (,, 2) (2,, ) C C (2,, 2) pt Figure 2. Sub-varieties in the closure of a generic lattice variety 4.5 General Case Since the computations above relied on using explicit bases for Λ, it was useful to use d and σ. For the general case, however, we will use the d notation. To define the closure in general, consider the lift of the quotient map f : Gr(K, m) X given by 37

f(v) = π (v), where π : F n F n /Λ max. Theorem 4.3. The closure of the variety Λ d is contained in the image of Gr(K, m) t,γ under f Since Gr(K, m) t,γ is closed and Λ d f(gr(k, m) t,γ ), if we can show that f is well defined and that the image lies in Xγ N, we will have that Λ d f(gr(k, m) t,γ ). In order to fully prove this claim, we will need to define some additional machinery. It is simple, however, to show that f is well defined and that the image of Gr(K, m) t,γ is in X γ. Lemma 4.4. The image f(gr(k, m) t,γ ) lies in X γ. Proof. For any v Gr(K, m) t,γ, consider the image f(v) = π (v). Any element u f(v) can be written as u + w, where u v and w Λ max. To verify that f(v) is a valid lattice, we need to check that tu f(v). Clearly, tw Λ max, since Λ max is itself a lattice. It is also clear that tu π (v), since tv v. Similarly, it is clear that since γλ max is fixed by γ, if v Gr(K, m) t,γ, then π (v) will also be preserved by γ. Thus, Im(f) is indeed a subset of X γ. We have verified that f(gr(k, m) t,γ ) X γ, but still need to show that f(gr(k, m) t,γ ) Xγ N. This follows from the condition that the subspaces of K be preserved by t. Let d min and d max be the list of initial degrees of Λ min and Λ max, respectively. Note that for each i, d min,i d i d max,i. For any v Gr(K, m) t,γ, the initial degrees of f(v) must lie between, or possibly equal to, d min and d max. We can organize this data about the possible degrees of f(v) in a grid with n columns, one for each standard basis vector. d min, d min,2... d min,n d min, + d min,2 +... d min,n +... d max, d max,2... d max,n For computational simplicity, we then normalize the table by subtracting d i from each cell in the i th column; thus, the entries will represent possible change in degree 38