You are on page 1of 122

On category O for the rational

Cherednik algebra of the complex


reflection group (Z/!Z) ! Sn

by
Richard Thomas Robert Vale

A thesis submitted to the


Faculty of Information and Mathematical Sciences
at the University of Glasgow
for the degree of
Doctor of Philosophy

December 2006

c Richard Vale 2006


"
2
Acknowledgements

I would like to thank my supervisors Prof. K. A. Brown and Prof. I. Gordon for giving me a vast
amount of help and suggesting many problems and their solutions. I would also like to thank Prof.
T. Lenegan of the University of Edinburgh for helping me to obtain funding to visit Chicago in
April 2006, and for arranging for me to visit Edinburgh in July 2006. I would like to thank Q.
Gashi, V. Ginzburg and I. Gordon for their hospitality during my visit to Chicago. I would like
to thank those mathematicians with whom I have had valuable conversations or correspondence,
in particular S. Ariki, O. Chalykh, and C. Stroppel. I would like to thank many of my colleagues
for their friendship. I acknowledge that my studies were funded by the Engineering and Physical
Sciences Research Council.

i
Statement

This thesis is submitted in accordance with the regulations for the degree of Doctor of Philosophy
in the University of Glasgow.
Chapters 1 and 2 cover notation, definitions and known results (apart from Theorem 2.1).
Chapters 3, 4 and 5 are the author’s original work except where stated otherwise. Some of the
results therein will appear in [70] and [69].

ii
Summary

The topic of this thesis is the rational Cherednik algebra of the complex reflection group (Z/!Z)!Sn ,
and in particular the category O of modules over the Cherednik algebra. The structure of the thesis
is as follows.
In Chapter 1, we begin by recalling some standard results which will be used in the text.
These include some basic notions of algebraic geometry, the representation spaces of quivers, and
quotients and blocks of abelian categories. We also recall some very basic results of invariant
theory of finite groups and the corresponding skew group algebras. We then introduce symplectic
reflection algebras in Section 1.2, and define Cherednik algebras of complex reflection groups as
a special case. After recalling the most important ring-theoretic properties of Cherednik algebras
and their spherical subalgebras in Proposition 1.18 and Proposition 1.19, we introduce category
O, which will be our main object of study. We then explain how to construct the KZ functor from
category O to the category of finite-dimensional modules over the Hecke algebra of W . Next, in
Section 1.4, we specialise to the case of the group W = (Z/!Z) ! Sn = G(!, 1, n), and describe some
of the main facts about the Hecke algebra of this group (which is also known as the Ariki-Koike
algebra). We end Chapter 1 by defining an isomorphism which is an analogue in the G(!, 1, n) case
of the Fourier automorphism of the Cherednik algebra in the Coxeter case (Theorem 1.40).
In Chapter 2, we give a proof that for any complex reflection group W , simplicity of the
Cherednik algebra of W is equivalent to semisimplicity of category O (Theorem 2.1). In the case
of W = G(!, 1, n), we recall that semisimplicity of category O is equivalent to semisimplicity of the
Ariki-Koike algebra (Theorem 2.4).
We begin Chapter 3 by recalling in Section 3.1 some of the known results on finite-dimensional
representations of the rational Cherednik algebra of W = G(!, 1, n). We then prove a new result
which says that when the KZ functor satisfies a condition called separating simples, we can com-

iii
pletely describe the structure of category O (Theorem 3.7). We prove that separating simples is
equivalent to the Hecke algebra having one less simple module than the group algebra CW (Theo-
rem 3.23). We then prove in Section 3.4 that this property determines the Ariki-Koike algebra up
to isomorphism. This chapter has been submitted for publication in [69].
In Chapter 4, we study shift functors for the Cherednik algebra of W = G(!, 1, n). First we prove
a shift relation, Theorem 4.1, for the spherical subalgebra, and use it to construct the Heckman-
Opdam shift functors. In Section 4.1, we give some conditions under which the Heckman-Opdam
shift functors are equivalences, and we prove that when they are equivalences, the Heckman-Opdam
shift functors commute. We then turn in Section 4.2 to another notion of shift functor, which we
call the Boyarchenko-Gordon shift functor. This functor is only well-defined when a hypothesis
(Hypothesis 4.48) holds. We prove that the hypothesis holds in the case n = 1, and conjecture that
it holds in general. Under Hypothesis 4.48, we construct the Boyarchenko-Gordon shift functor,
and prove that it coincides with a product of Heckman-Opdam shift functors, provided that all of
the functors in the product are equivalences (Theorem 4.62).
In Chapter 5, we give an application of the results of the previous chapters, by proving an
analogue for G(!, 1, n) of a result of Gordon [34] on the diagonal coinvariants of Coxeter groups.
This chapter has been accepted for publication in [70].

iv
Introduction

The aim of this thesis is to investigate the category O of modules for the rational Cherednik algebra.
This category has many interesting properties; it is large enough to contain information about the
structure of the algebra itself, and yet small enough to be understood using finite combinatorics.
There is also a conjectural relationship between category O and categories of coherent sheaves on
the resolutions of a certain symplectic singularity, coming from the fact that the Cherednik algebra
can be viewed as a deformation of this singularity. Furthermore, category O is an analogue in the
Cherednik algebra theory of a similar category of modules for the enveloping algebra of a finite-
dimensional Lie algebra, and as such there are many techniques in existence which can be brought
to bear on the problem of trying to find out how it behaves.
Rational Cherednik algebras are a special case of symplectic reflection algebras. Symplectic
reflection algebras were introduced by Etingof and Ginzburg, [26]. Given a symplectic vector
space, V and a finite subgroup U ⊂ Sp(V ), the symplectic reflection algebras Ht,c are a family
of deformations of the skew group algebra C[V ] ∗ U . They were motivated by the representation
theory, geometry and integrable systems which are related to these algebras. There is a dichotomy
in the theory according to whether the parameter t is 0 or 1. In the Cherednik algebra case, t is
taken to be 1. A good reference for symplectic reflection algebras in the t = 0 case is [26].
In the Cherednik algebra case, one takes V = h ⊕ h∗ where h is the reflection representation
of a complex reflection group W , and takes U to be the subgroup of Sp(V ) defined by the natural
action of W on h ⊕ h∗ . The symplectic reflection algebra associated to (h ⊕ h∗ , U ) is called the
rational Cherednik algebra of W (we often refer to it just as the Cherednik algebra of W ). The
Cherednik algebra Hk (W ) depends on a vector of complex parameters k, and when k is taken to be
0, the Cherednik algebra reduces to the algebra An (C) ∗ W , where An (C) denotes the Weyl algebra
with n = dim(h). In general, the Cherednik algebra is still very similar to the Weyl algebra,

v
since it may be identified with the subalgebra of End(C[h]) generated by C[h], W and a family
of commuting differential-difference operators {∇kv : v ∈ h} called Dunkl operators. The Dunkl
operators were introduced in the Coxeter case in [22] and in the general case in [23]. Although the
Dunkl operators commute, they have a difference term which explicitly depends on the action on
W . Thus, the Cherednik algebra is more sensitive to the W –action than the Weyl algebra.
The PBW Theorem of Etingof and Ginzburg [26] implies that the Cherednik algebra has a
vector space decomposition into three parts

Hk (W ) ∼
= C[h] ⊗ CW ⊗ C[h∗ ].

This is reminiscent of the decomposition of the enveloping algebra of a semisimple Lie algebra g
as U(g) = U (n+ ) ⊗ U(h) ⊗ U(n− ), and it is possible to define an analogue of the BGG category O
(see [56]) in the Cherednik algebra situation. This category was first introduced in [6] and then
studied in some detail in [33]. As in the Lie case, category O is a highest-weight category, meaning
that it is combinatorially well-behaved.
An important fact about category O is the existence of a functor

KZ : O → H − mod

where H denotes the Hecke algebra of W , a finite-dimensional algebra which can be viewed as
a deformed version of the group algebra of W , depending on complex parameters which are the
exponentials of the parameters k. The functor KZ exists because, after a suitable localisation, the
Cherednik algebra becomes the skew group algebra of W over the algebra of differential operators on
an open set hreg ⊂ h. Thus, a module for the Cherednik algebra may be regarded as a W –equivariant
D–module on hreg . This D–module gives a vector bundle with a flat connection, which in turn gives
a finite-dimensional representation of the fundamental group of hreg /W . This representation turns
out to factor through the Hecke algebra. The construction is explained in detail in Section 1.29.
We exploit the KZ functor in Chapter 3. Passing back and forth between the Cherednik algebra
and Hecke algebra allows us to prove theorems about both of these objects. The Hecke algebra is
an interesting object in its own right, and in the G(!, 1, n) case it was introduced in the paper [2]
by Ariki and Koike. A lot is known about the representation theory of this algebra, and we are
able to use this to describe the block structure and composition multiplicities of category O in the
case where O is close to being semisimple (Theorem 3.7). In the other direction, we are also able

vi
to describe the Ariki-Koike algebra in the almost-semisimple case, using the Cherednik algebra
techniques.
The results of Chapter 4 are motivated by a more geometric viewpoint. Just as the Cherednik
algebra may be regarded as a deformation of C[h⊕h∗ ]∗W , it has a subalgebra which may be regarded
as a deformation of the fixed ring C[h ⊕ h∗ ]W , or equivalently, as a noncommutative version of the
singularity (h ⊕ h∗ )/W . This subalgebra (the so-called spherical subalgebra) is just eHk e where
1 !
e = |W | w∈W w is the symmetrising idempotent. The work of Gordon and Stafford [36], [37]

shows that, in the case where W = Sn , the spherical subalgebra is related to the geometry of the
resolution of singularities of (h ⊕ h∗ )/W . However, it is not enough just to study eHk e for one
value of k, rather it is necessary to consider all integer shifts of k at once, to construct a so-called
Z-algebra out of the eHk e. For this purpose, Gordon and Stafford used Heckman-Opdam shift
functors, which are functors
eHk e − Mod → eHk! e − Mod,

where the parameters k$ are obtained from k by an integer shift. In Section 4.1, we define analogues
of these functors for the group G(!, 1, n). The situation here is more complicated, since the Chered-
nik algebra of G(!, 1, n) depends on ! parameters rather than just one. We give some conditions
under which these functors can be shown to be equivalences. Category O is an important tool in
proving these results.
In the second part of Chapter 4, we turn to another kind of shift functor which has been defined
by Gordon [35], following Boyarchenko [10]. This functor relies on a description of eHk e as a factor
of the ring of differential operators on the representation space of a quiver (this is known as a
Hamiltonian reduction, see for example [30]). The quiver here is a cyclic quiver, with one vertex
and one edge attached, which we denote by Q∞ . Let G be the base-change group of the quiver Q∞
and let ε be the dimension vector of Q∞ with 1 at the extra vertex and n at each vertex of the
cycle. Then Gordon’s description of eHk e is
D(Rep(Q∞ , ε))G
eHk e ∼
=
Ik
for some ideal Ik . Unfortunately, in our case this description is not quite complete, since it depends
on a certain homomorphism called the radial part map having the correct image. We are at least
able to show that this holds for W = Z/!Z. Assuming it holds in general (which we conjecture
to be true), we construct our own version of the Boyarchenko-Gordon shift functors. We then in

vii
Section 4.3 study the question of when these coincide with the Heckman-Opdam shift functors. We
are able to show that in some cases, they do coincide (Theorem 4.62). This result is useful because
the Z–algebra associated to the Boyarchenko-Gordon shift functors is related, via an associated
graded construction, to one associated with a Nakajima quiver variety [58]. In some cases, these
quiver varieties are resolutions of the singularity (h⊕h∗ )/W . On the other hand, if the Boyarchenko-
Gordon shift functors are known to be equivalences, one can use the Z–algebra theory of [36] to show
that there is an explicit relationship between the category of coherent sheaves on the quiver variety
and the category of finitely-generated eHk e–modules. Thus, it is important to know when the
Boyarchenko-Gordon and Heckman-Opdam shift functors coincide, since the Boyarchenko-Gordon
shift functors have the correct associated graded properties, whereas it is in the Heckman-Opdam
case that we have a better chance of showing that these functors are equivalences.
It should be possible to go further and study coherent sheaves on the quiver variety using the
Cherednik algebra in the manner of [37]. However, we do not pursue this line of research in this
thesis.
In Chapter 5, we use the results of the earlier chapters to prove a theorem (Theorem 5.2) about
the diagonal coinvariants of G(!, 1, n). This theorem follows a proof of Gordon [34]. It does not use
the Boyarchenko-Gordon shift functors and therefore does not rely on Hypothesis 4.48. Theorem
5.2 states that the diagonal coinvariant ring

C[h ⊕ h∗ ]
,C[h ⊕ h∗ ]W
+-

has a quotient with good combinatorial properties. It is a nice application of Cherednik algebra
techniques to prove a result in commutative algebra.
The structure of this thesis is as follows. In Chapter 1 we give a list of basic definitions, then
define symplectic reflection algebras and rational Cherednik algebras, giving their basic properties.
We discuss the case of the group G(!, 1, n) in detail. In Chapter 2, we study the generic case, in
which the Cherednik algebra is simple and category O is semisimple. In Chapter 3, we present our
results on the almost-semisimple case. In Chapter 4, we define and study the Heckman-Opdam
and Boyarchenko-Gordon shift functors. In order to define the Heckman-Opdam shift functors, we
use Poisson geometry to prove a shift relation, Theorem 4.1, following [6] and [7]. Finally, Chapter
5 contains a proof of Theorem 5.2 on diagonal coinvariants.

viii
Contents

Statement ii

Summary iii

Introduction v

1 The rational Cherednik algebra 1


1.1 Notation and conventions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Symplectic reflection algebras and the PBW theorem . . . . . . . . . . . . . . . . . . 8
1.3 The rational Cherednik algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.1 The spherical subalgebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3.2 Representation theory of Hk . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3.3 Ideals of Hk and category O . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.4 The Dunkl representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.5 The Knizhnik-Zamolodchikov functor . . . . . . . . . . . . . . . . . . . . . . 17
1.3.6 Double centraliser property . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.3.7 Twisting by a linear character . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.3.8 Absence of self-extensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.4 The case of W = G(!, 1, n) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.4.1 The rational Cherednik algebra of G(!, 1, n) . . . . . . . . . . . . . . . . . . . 22
1.4.2 Representation theory of the Ariki-Koike algebra . . . . . . . . . . . . . . . . 24
1.4.3 Fourier map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

ix
2 The semisimple case 29
2.1 The semisimple case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

3 The almost-semisimple case 34


3.1 Finite-dimensional modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2 The main theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.3 Proof of Theorem 3.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3.1 Proof of parts (1) and (2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3.2 Blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.3.3 Proof of parts (3) and (4) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.3.4 Proof of part (5) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.3.5 Characterisations of separating simples . . . . . . . . . . . . . . . . . . . . . 48
3.4 The Ariki-Koike algebra in the almost-semisimple case . . . . . . . . . . . . . . . . . 50

4 Shift functors 55
4.1 The Heckman-Opdam shift functors . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.1.1 A shift relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.1.2 The shift functors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.1.3 The semisimple and almost-semisimple cases . . . . . . . . . . . . . . . . . . 66
4.1.4 The asymptotic parameter case . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.1.5 Shift functors on category O . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.1.6 A commutativity property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.2 The Boyarchenko-Gordon shift functors . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.2.1 Gordon’s construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.2.2 The radial part map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.2.3 A shift functor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.3 Comparison of the shift functors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.3.1 A remark on Z–algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.3.2 Nakajima quiver varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

x
5 Diagonal coinvariants 98
5.1 A quotient ring of the diagonal coinvariants . . . . . . . . . . . . . . . . . . . . . . . 100

Bibliography 104

xi
Chapter 1

The rational Cherednik algebra

1.1 Notation and conventions

The purpose of this section is to set out some notational conventions which will be used throughout
the text, and also to supply some definitions which will be used in the proofs of certain results in
this thesis.

Noncommutative rings

A good reference for standard notions in noncommutative algebra is [55]. In this thesis, we will
work exclusively over the field of complex numbers C. An algebra for us will mean a C-algebra,
associative and unital, and we do not require that a subalgebra of an algebra shares the same
identity element. For an algebra A with subalgebra B, we will say that A is finite over B to mean
that A is a finitely generated left B-module. If A is a Noetherian domain we denote by Frac(A)
the division ring of fractions of A. For a C-algebra A, a filtration is by definition an increasing
collection of subspaces F = {Fi : i ∈ Z} with Fa = 0 for all a < 0 and ∪∞
i=0 Fi = A, such that

Fi Fj ⊂ Fi+j . If a ∈ Fi \ Fi−1 then we say that a has degree deg(a) = i with respect to the
filtration F. We will denote the associated graded algebra with respect to F by grF A, or simply by
grA if there is no risk of confusion. For an algebra A we write GK dim A for the Gelfand-Kirillov
dimension of A; see [47] for the definition and basic properties of this dimension.
By convention, we use A−Mod to denote the category of all left modules over an algebra A, and
A − mod to denote the full subcategory of finitely generated left A–modules. We use the notation

1
Mod − A and mod − A for the analogous categories of right A–modules. A good reference for
standard notions of homological algebra is the book [63]. We will require the notions of Auslander-
regularity and Cohen-Macaulayness for a noncommutative ring. We give the definitions of these
notions following [47, Section 12.9]. We do not discuss these properties in detail since we require
them only to apply some technical theorems of noncommutative algebra (see Lemma 4.38).

Definition 1.1. An algebra R is Auslander-Gorenstein if R R and RR have finite injective dimen-


sion, and given integers 0 ≤ i < j and a finitely generated left or right R–module M , together with
a submodule N of ExtjR (M, R), we have ExtiR (N, R) = 0.

Definition 1.2. An algebra R is Auslander-regular if R is Auslander-Gorenstein and has finite


global dimension.

Definition 1.3. An algebra R is Cohen-Macaulay if for all finitely generated R–modules M , we


have
GK dim(R) = GK dim(M ) + min{j ≥ 0 : ExtjR (M, R) 1= 0}.

Geometry

For basic definitions of algebraic geometry, a good reference is [42]. All varieties in this thesis are
defined over C. For a Zariski-closed subset V of Cn , we write C[V ] for the coordinate ring of V ,
that is, the ring C[x1 , . . . , xn ]/I where I is the ideal of functions which vanish on V . Given a finitely
generated commutative domain A, we write maxspec(A) for the set of maximal ideals of A, which is
an affine variety. Given an affine variety V , we denote by D(V ) the ring of differential operators of
C[V ], as defined in [55, Chapter 15]. We use the term “D–module on V ” as a synonym for “D(V )–
module”. We will often use the fact that if V is a smooth affine variety and f ∈ C[V ] \ {0}, then
D(Vf ) is isomorphic to the localization D(V )[f −1 ], where Vf denotes the open subset of points
x ∈ V where f (x) 1= 0. This follows from [55, Theorem 15.1.25] combined with [55, Corollary
15.5.6].
Given an algebraic variety V , which is not assumed to be affine, a vector bundle on V may be
regarded as a locally free sheaf of OV –modules, where OV denotes the coordinate sheaf of V . For
the fundamental properties of vector bundles on algebraic varieties, see [48, Chapter 1]. Given a
vector bundle B on V , a connection on B is a OV –module map

∇ : B → Ω 1 ⊗ OV B

2
satisfying for any affine open U ⊂ V , ∇(f b) = df ⊗ b + f ∇(b) for all f ∈ OV (U ) and all b ∈ B(U ).
Here Ω1 denotes the cotangent sheaf of V (the sheaf of sections of the cotangent bundle T ∗ V → V ).
Given a vector field v on V , a vector bundle B, and a connection ∇ on B, we may define a
map of sheaves ∇v : B → B by setting ∇v (b) := ∇(b)(v). The connection ∇ is said to be flat if
[∇v , ∇w ] = ∇[v,w] for all vector fields v and w on V . Information about the basic properties of
connections may be found in [19, I.2].
There is an analogous definition of connection on a vector bundle over a complex manifold. In
order to construct the KZ functor in Section 1.29, we will require the notion of a connection having
regular singularities. As this has a lengthy definition, and it will not be used elsewhere in this
thesis, we refer the reader to [19, II, Chapter 4] for the definition.

Quivers

One example of an affine variety that will be used in this thesis is the space of all representations
of a quiver Q with a given dimension vector.

Definition 1.4. A quiver Q is by definition a 4–tuple (Q0 , Q1 , h, t) where:

• Q0 is a finite set, called the set of vertices of Q.

• Q1 is a finite set, called the set of arrows of Q.

• h, t : Q1 → Q0 are functions.

A quiver should be viewed as a finite directed graph, where h, t : Q1 → Q0 are the functions
that assign to an arrow its head and tail respectively.
Given a quiver Q, a dimension vector for Q is a function Q0 → Z≥0 . Given a dimension vector
α for Q, a representation of Q of dimension vector α is a map which assigns to each x ∈ Q0 a
vector space Vx of dimension α(x), and to each a ∈ Q1 a linear map Xa : Vt(a) → Vh(a) . The space
Rep(Q, α) of representations of a quiver Q with dimension vector α may be identified with the
affine variety
"
Rep(Q, α) = Mat(α(t(a)) × α(h(a)), C).
a∈Q1
# −1
The algebraic group GL(α) = x∈Q0 GL(Vx ) acts on Rep(Q, α) via (g · X)a = gh(a) Xa gt(a) for
a ∈ Q1 . Every point of Rep(Q, α) is fixed by the normal subgroup C∗ ∼
= {(λid, . . . , λid) : λ ∈ C∗ }

3
under this action, so this gives an action of the group G(α) = GL(α)/C∗ on Rep(Q, α). Two
representations A, B ∈ Rep(Q, α) are said to be isomorphic if and only if they lie in the same orbit
of G(α).

Category theory

Nearly all the categories considered in this thesis will be subcategories of A − Mod for some algebra
A. Furthermore, most of the categories in this thesis will be abelian. A reference for the basic
properties of abelian categories is [28]. We will require the construction of the quotient of an
abelian category by a Serre subcategory, which we now describe, following the exposition in [62].

Definition 1.5. Let C be an abelian category. An abelian subcategory A of C is said to be Serre if


for every short exact sequence
0 → X $ → X → X $$ → 0

in C, we have X ∈ A if and only if X $ , X $$ ∈ A.

If C is an abelian category and A ⊂ C is a Serre subcategory, let Σ denote the class of morphisms
f in C with ker(f ) ∈ A and cok(f ) ∈ A. Then we define the quotient C/A to be an additive category
C/A together with a functor Q : C → C/A such that Q(f ) is invertible for all f ∈ Σ, and satisfying
the following universal property:
for any additive category D and any additive functor F : C → D such that F (f ) is invertible
for all f ∈ Σ, there exists a unique functor C/A → D making the following diagram commute.
Q
C " C/A
!
!!
F
!!!
#!
! !
D
By [62, I, Theorem 14.1], we have the following proposition.

Proposition 1.6. If C is an abelian category such that the class of subobjects of each object of C
is a set, and A is a Serre subcategory of C, then the quotient category C/A exists. It has objects

ob(C/A) = ob(C)

and morphisms
HomC/A (X, Y ) = limZ →X
s Hom(Z, Y ).
s∈Σ

4
The quotient functor Q is defined to be the identity on objects, and for a morphism f : X → Y one
defines Q(f ) to be (f ◦ s)s∈Σ, Z →X
s

A proof that the above definition really does give a well-defined pair (C/A, Q) satisfying the
universal property may be found in [62, Theorem 14.1, Chapter 1]. In particular, the quotient
functor Q : C → C/A is essentially surjective on objects, ie. every object of C/A is isomorphic to
the image under Q of an object of C. Furthermore, it follows from [62, Exercise 7.3] that C/A is
an abelian category and Q is an exact functor.
If A is an abelian category then we may partition the set S of simple objects of A into equivalence
classes called blocks. We define a relation ∼ on the set S by L1 ∼ L2 if Ext1A (L1 , L2 ) 1= 0. Taking
the reflexive, symmetric and transitive closure of this relation gives an equivalence relation which
we also denote by ∼. The equivalence classes of S under ∼ are called the blocks of A. Furthermore,
if an object M of A has a finite composition series all of whose factors belong to the block B, then
we say that M belongs to the block B. Of course, in general not every object of A need belong to
a block.
The following proposition is standard, but we were not able to find a convenient reference for it.
Recall that a composition series of an object M in an abelian category A is by definition a chain
0 = M0 ⊂ M1 ⊂ · · · ⊂ Mn = M of subobjects of M such that Mi /Mi−1 is simple for each i. We
say that the Jordan-Hölder theorem holds in A if whenever an object M in A has two composition
series 0 = M0 ⊂ M1 ⊂ · · · ⊂ Mn = M and 0 = M0$ ⊂ M1$ ⊂ · · · ⊂ Mm
$ = M , then m = n and the

same composition factors appear in the two composition series with the same multiplicities.

Proposition 1.7. Let A be an abelian category in which the Jordan-Hölder theorem holds and let M
be an indecomposable object of A. Suppose M has a composition series 0 ⊂ M1 ⊂ · · · ⊂ Mn = M .
Then all the composition factors Mi /Mi−1 belong to the same block of A.

Proof. Let {1, 2, . . . , n} = I ∪ J be a partition into disjoint subsets such that

Ext1A (Mi /Mi−1 , Mj /Mj−1 ) = Ext1A (Mj /Mj−1 , Mi /Mi−1 ) = 0

if i ∈ I, j ∈ J. We first show that we may take I = {1, 2, . . . a} for some a. Indeed, suppose there
is j ∈ J, i ∈ I with j < i. Then after relabelling, we may assume that j ∈ J, j + 1 ∈ I. The exact
sequence
Mj Mj+1 Mj+1
0→ → → →0
Mj−1 Mj−1 Mj

5
must split, so we may define a new composition series 0 ⊂ M1$ ⊂ M2$ ⊂ · · · ⊂ Mn$ = M of M
with j ∈ I, j + 1 ∈ J. Therefore, we may assume I = {1, . . . , a}, J = {a + 1, a + 2, . . . , n} for
some a. Now fix r with a + 1 ≤ r ≤ n. Using the long exact sequence for Ext1 , we may show
by induction on t that Ext1A (Mt , Mr /Mr−1 ) = 0, 1 ≤ t ≤ a. Indeed, for t ≤ a, if we have shown
Ext1A (Mt−1 , Mr /Mr−1 ) = 0, then since 0 → Mt−1 → Mt → Mt /Mt−1 → 0 is exact, we get an exact
sequence

Ext1A (Mt−1 , Mr /Mr−1 ) ← Ext1A (Mt , Mr /Mr−1 ) ← Ext1A (Mt /Mt−1 , Mr /Mr−1 ),

which shows Ext1A (Mt , Mr /Mr−1 ) = 0 since t ∈ I.


Then, using the same method, we may show by induction on r that Ext1A (Ma , Mr /Ma ) = 0 for
a + 1 ≤ r ≤ n. Therefore, M = Ma ⊕ M/Ma , which implies that either I or J is empty, since M is
indecomposable.

If A is a finite-dimensional algebra, the blocks of A may also be defined as follows. Write


1 ∈ A as a sum of pairwise orthogonal central idempotents 1 = e1 + · · · + en with n maximal.
This corresponds to a decomposition A = ⊕ni=1 Aei of A into indecomposable subalgebras which are
known as the blocks of A (see [13, I.16.I]). The blocks of the algebra A are in bijection with the
blocks of the category A = A−mod because every simple left A–module S satisfies ei S = S for some
unique ei . Therefore, if S, T belong to distinct blocks of A then Ext1A (S, T ) = Ext1A (T, S) = 0, while
if S, T belong to the same block of A then S and T are composition factors of the indecomposable
module Aei and hence belong to the same block of A by Proposition 1.7.

Invariants and skew group algebras

For a finite group W , we write Irrep(W ) for the set of irreducible representations of W over C. We
refer to the character of a one-dimensional representation as a linear character of W .
Let V be a finite-dimensional complex vector space and let W be a finite subgroup of the general
linear group GL(V ). Then W acts naturally on the coordinate ring C[V ] of V , which is just the
polynomial ring C[x1 , x2 , . . . , xn ] where n = dim(V ), and we can consider the ring of polynomial
invariants C[V ]W . Then C[V ]W may be viewed as the coordinate ring of the orbit space V /W .
See [4, Remark, page 8].

6
Definition 1.8. A linear map s : h → h is called a complex reflection if s fixes a hyperplane H
pointwise, and s has finite multiplicative order. A finite subgroup of GL(h) generated by complex
reflections is called a complex reflection group.

The complex reflection groups were classified by Shepherd and Todd in [66]. They fall into one
infinite family and 34 exceptional cases. Complex reflection groups are important because of the
second part of the following theorem.

Theorem 1.9. Let V be a finite-dimensional complex vector space and W a finite subgroup of
GL(V ).

1. The algebra C[V ]W is a finitely generated domain and C[V ] is a finite C[V ]W –module.

2. C[V ]W is smooth (ie. has finite homological dimension) if and only if W is a complex reflection
group.

Proof. It is obvious that C[V ]W is a domain since C[V ]W ⊂ C[V ]. The Hilbert-Noether theorem
[4, Theorem 1.3.1] states that C[V ]W is finitely generated and C[V ] is a finite C[V ]W –module.
Furthermore, C[V ]W has finite global dimension if and only if it is a polynomial ring, by Serre’s
converse to Hilbert’s syzygy theorem [4, Corollary 4.2.3] and [4, Theorem 6.2.2(b)]. But C[V ]W is
a polynomial ring if and only if W is generated by complex reflections, by [4, Theorem 7.2.1].

The algebra C[V ]W is contained in a larger, noncommutative algebra called the skew group
algebra C[V ] ∗ W . Such an algebra may be defined whenever a finite group W acts on an algebra
A, and we describe it in this more general context.

Definition 1.10. Let A be an algebra and W a finite group which acts faithfully on A by algebra
automorphisms. The skew group algebra A∗W may be defined as the free A–module with basis given
by the elements of W , equipped with the multiplication a1 w1 · a2 w2 = a1 w1 (a2 )w1 w2 for a1 , a2 ∈ A
and w1 , w2 ∈ W .

Some properties of A are inherited by A ∗ W .

Proposition 1.11. Let A be an algebra and W a finite group which acts faithfully on A by algebra
automorphisms. If A is Noetherian then so is A ∗ W . If A is simple and the group of units in A is
central in A, then A ∗ W is simple.

7
Proof. We have that A ∗ W is Noetherian by [61, Proposition 1.6], and the second part follows
from [55, Proposition 7.8.12].

In the case A = C[V ], the ring of invariants and the skew group algebra are related as follows.

Proposition 1.12. Let V be a finite-dimensional complex vector space and W a finite subgroup of
1 !
GL(V ). Let e = |W | w∈W w ∈ CW denote the symmetrising idempotent of W . Then there is an

algebra isomorphism
C[V ]W → e(C[V ] ∗ W )e

defined by z 5→ ze.
$ ! %
Proof. If f ∈ C[V ] then ef e = 1
|W | w∈W w(f ) e.

1.2 Symplectic reflection algebras and the PBW theorem

The aim of this chapter is to introduce the rational Cherednik algebra. Rational Cherednik algebras
are a special case of symplectic reflection algebras, which were introduced in [26]. We will present
the basic facts about symplectic reflection algebras, following the exposition of [26] and [12]. Let
V be an n–dimensional complex vector space equipped with a symplectic form ω : V ⊗ V → C,
that is, a bilinear form on V which is non-degenerate and skew-symmetric. The group of all linear
transformations γ : V → V such that ω(γx, γy) = ω(x, y) for all x, y is called the symplectic group
Sp(V ) of V .

Definition 1.13. A linear transformation γ : V → V is called a symplectic reflection if rank(1 −


γ) = 2.

Let Γ ! Sp(V ) be a finite subgroup of Sp(V ) generated by symplectic reflections. Then the
triple (V, ω, Γ) is called a symplectic triple. We may associate an algebra to a symplectic triple as

i=0 ⊗ V be the tensor algebra of V . Then form the skew group algebra
follows. First, let T (V ) = ⊕∞ i

T (V ) ∗ Γ. Now choose a bilinear map θ : V ⊗ V → CΓ taking values in the group algebra. Define Iθ
to be the 2–sided ideal of T (V ) ∗ Γ generated by all the elements of the form x ⊗ y − y ⊗ x − θ(x, y)
for x, y ∈ V . Finally, define an algebra

Hθ := T (V ) ∗ Γ/Iθ .

8
Let us note that if θ = 0 then Hθ = Sym(V ) ∗ Γ, but that otherwise, Hθ need not be isomorphic
to Sym(V ) ∗ Γ. However, there is in general a relationship between Hθ and Sym(V ) ∗ Γ, since we
may define a filtration on Hθ as follows. Let F−1 (Hθ ) = 0, F0 (Hθ ) = CΓ, F1 (Hθ ) = V + CΓ and
Fi (Hθ ) = (F1 (Hθ ))i for all i ≥ 1. Then consider the associated graded algebra

grF (Hθ ) := ⊕∞
i=0 (Fi (Hθ )/Fi−1 (Hθ )).

Because the image of x ⊗ y − y ⊗ x in grF (Hθ ) is zero for all x, y ∈ V , there is a natural map

ρ : Sym(V ) ∗ Γ → grF (Hθ ).

Definition 1.14. If ρ is an isomorphism, then Hθ is said to have the PBW property.

Note that if Hθ has the PBW property then Hθ has a vector space basis given by the expressions
a · γ where a is an ordered monomial in a basis of V , and γ ∈ Γ. That is, Hθ is isomorphic to
Sym(V ) ⊗C CΓ as vector spaces.
The following result is of fundamental importance.

Theorem 1.15. [26, Theorem 1.3] Let (V, ω, Γ) be a symplectic triple. Let S be the set of symplectic
reflections in Γ. For s ∈ S, let ωs be the symplectic form on V with radical rad ωs = ker(1 − s) and
with ωs |im(1−s) = ω. Suppose there is a t ∈ C and a conjugation-invariant function c : S → C such
that
&
θ(x, y) = tω(x, y)1 + cs ωs (x, y)s
s∈S

for all x, y ∈ V . Then the PBW property holds for Hθ .

If the PBW property holds, then Hθ is said to be a symplectic reflection algebra.

1.3 The rational Cherednik algebra

The rational Cherednik algebras arise as a special case of symplectic reflection algebras. Let h be
a finite-dimensional complex vector space and let W be a finite group acting faithfully on h, with
the action being generated by complex reflections. In this thesis we will be mostly interested in the
group G(!, 1, n) to be defined below. However, the following theory holds for any complex reflection
group.

9
Since W acts on h, we have the contragredient action on h∗ and hence an action of W on h ⊕ h∗ .
Now define a symplectic form ω on h ⊕ h∗ by ω(a + f, b + g) = f (b) − g(a). Then it is easy to
see that ω is nondegenerate. Denote the action map W → GL(h ⊕ h∗ ) by w 5→ ŵ. Then ŵ is
a symplectic reflection if and only if w is a complex reflection. So the group Ŵ = {ŵ|w ∈ W }
is generated by symplectic reflections and therefore (h ⊕ h∗ , ω, Ŵ ) is a symplectic triple. We now
construct a symplectic reflection algebra from this triple.
Each complex reflection w in W has by definition a reflecting hyperplane H. Let A be the set
of reflecting hyperplanes of W . For H ∈ A, let WH be the set of elements of w which fix every
element of H. Then each WH is a cyclic group and the set of complex reflections of W is ∪H∈A WH .
Let nH = |WH | and let vH ∈ h be chosen so that CvH is a WH –stable complement to H. Also, let
αH ∈ h∗ be a linear functional with kernel H. Let S denote the set of complex reflections in W .
For s ∈ S, x ∈ h∗ and y ∈ h, we have in the notation of Theorem 1.15

αH (y)x(vH )
ω(x, y) = y(x), ωŝ (x, y) = .
αH (vH )

Therefore, we may define a symplectic reflection algebra T (h ⊕ h∗ ) ∗ Ŵ /Iθ , where we choose t ∈ C


and a function S → C, w 5→ cw with cγ −1 wγ = cw for all w ∈ S and γ ∈ W , and set
& αH (y)x(vH ) &
θ(x, y) = t · y(x)1 + cw ŵ.
αH (vH )
H∈A w∈WH \{1}

It will be more convenient later on to take a basis of WH given by the idempotents eH,i =
!
w∈WH (det w) w. We fix the parameter t = 1, and define complex parameters kH,i ∈ C such
1 i
nH

that
& H −1
n&
cw w = nH (kH,i+1 − kH,i )eH,i
w∈WH i=0

where kH,0 = kH,nH := 0. We therefore have

H −1
n&
cw = (det w)i (kH,i+1 − kH,i )
i=0

for all w ∈ WH . Since Ŵ ∼


= W , the symplectic reflection algebra we have constructed may be
defined as follows.

Notation 1.16. Write C[h] = Sym(h∗ ) for the algebra of polynomial functions on h. Similarly,
C[h∗ ] = Sym(h).

10
Definition 1.17. Let W be a complex reflection group acting faithfully on a vector space h. Let
k = {kH,i |H ∈ A, 0 ≤ i ≤ nH } be a set of complex parameters with kH,0 = kH,nH −1 := 0 and such
that kw(H),i = kH,i for all H ∈ A, all 1 ≤ i ≤ nH − 1, and all w ∈ W . Then the rational Cherednik
algebra Hk (W ) associated to (h, W, k = {kH,i }) is the algebra generated by h, h∗ and W subject to
the relations:

[x, x$ ] = 0, [y, y $ ] = 0

wxw−1 = w(x), wyw−1 = w(y)


H −1
& αH (y)x(vH ) & n&
[y, x] = y(x) + (kH,i+1 − kH,i )(det w)i w
αH (vH )
H∈A w∈WH i=0

for all x, x$ ∈ h∗ , all y, y $ ∈ h and all w ∈ W .

Let us fix a basis {y1 , . . . yn } of h and the dual basis {x1 , . . . xn } of h∗ . The rational Cherednik
algebra is, by construction, a special case of a symplectic reflection algebra, and we therefore have
the following proposition.

Proposition 1.18. The rational Cherednik algebra Hk has the following properties:

1. Hk has a filtration with h, h∗ in degree 1 and W in degree 0, and grHk ∼


= C[h ⊕ h∗ ] ∗ W .

2. A basis of Hk is given by the set

{p · q · γ : p is an ordered monomial in the xi , q is an ordered monomial in the yi , γ ∈ W }.

In other words, as vector spaces Hk ∼


= C[h∗ ] ⊗ C[h] ⊗ CW via the multiplication map.

3. The ring Hk is Noetherian and has finite homological dimension.

Proof. The first two parts follow from the PBW theorem 1.15. The third part is a general fact
about symplectic reflection algebras, contained in [12, Theorem 4.4].

1.3.1 The spherical subalgebra

There is a certain important subalgebra of Hk called the spherical subalgebra. Let e ∈ CW be the
1 !
idempotent e = |W | w∈W w. Then eHk e is a filtered algebra with a filtration inherited from Hk

and gr(eHk e) = e C[h ⊕ h∗ ] ∗ W e = C[h ⊕ h∗ ]W . This is a domain, so eHk e is a domain. We have


the following proposition.

11
Proposition 1.19. [26, Theorem 1.5] The spherical subalgebra eHk e is a Noetherian domain
which is Auslander-Gorenstein and Cohen-Macaulay.

Proof. We have shown that eHk e is a domain. The same argument shows that it is Noetherian, since
C[h ⊕ h∗ ]W is Noetherian. In order to show that eHk e is Auslander-Gorenstein, we use [8, Theorem
3.9], which states that in this situation, eHk e is Auslander-Gorenstein if gr(eHk e) is Auslander-
Gorenstein. But gr(eHk e) = C[h ⊕ h∗ ]W , which is an Auslander-Gorenstein ring by Watanabe’s
Theorem (see [4, Theorem 4.6.2]), since Ŵ is a subgroup of SL(h ⊕ h∗ ). Furthermore, a theorem
of Hochster and Eagon given in [4, Theorem 4.3.6] implies that C[h ⊕ h∗ ]W is Cohen-Macaulay. It
then follows from the proof of [8, Theorem 3.9], combined with [47, Proposition 6.6], that eHk e is
Cohen-Macaulay.

1.3.2 Representation theory of Hk

We now introduce the topic of this thesis: the representation theory of the rational Cherednik
algebra. The first step is to define some modules which are the analogues of Verma modules in Lie
theory. Given a representation τ of W , we may make τ into a C[h∗ ] ∗ W –module by declaring that
h ⊂ C[h∗ ] acts by zero. Since C[h∗ ] ∗ W ⊂ Hk , we may form the induced module.

Definition 1.20. The standard module associated to τ is the Hk –module

M (τ ) := IndH
C[h∗ ]∗W (τ ) = Hk ⊗C[h ]∗W (τ ).
k

Notice that Proposition 1.18 implies that M (τ ) is isomorphic as a W –module to C[h] ⊗ τ . In


particular, M (τ ) is a free C[h]–module. From now on, we will assume that τ is irreducible.
Now we explain how to give M (τ ) a grading. Recall that we have fixed a basis {yi } of h and
!
the dual basis {xi } of h∗ . Let eu = i xi yi . Define an element
H −1
& n&
z= nH kH,i eH,i
H∈A i=1

where eH,i are the idempotents defined above. Then z is a central element of CW . Let h = eu − z.
Then by [33, (4), Section 3.1], we have [h, y] = −y for all y ∈ h and [h, x] = x for x ∈ h∗ and
[h, w] = 0 for all w ∈ W . Since z ∈ Z(CW ), z acts on an irreducible W –representation τ by a
scalar λ. For i ∈ Z, let
M (τ )i = {x ∈ M (τ ) : hx = (i − λ)x}.

12
Then we have the following lemma due to Dunkl and Opdam.

Lemma 1.21. [23] The grading induced by h has the following properties.

1. For each i ≥ 0, let C[h]i denote the homogeneous polynomials of degree i. Then M (τ )i =
C[h]i ⊗ τ .

2. Every submodule of M (τ ) inherits this grading. That is, if N ⊂ M (τ ) is any Hk –submodule,


then {x ∈ N : hx = (i − λ)x} = N ∩ M (τ )i for all i ∈ Z.

3. M (τ ) has a unique maximal submodule R(τ ), and hence a unique simple quotient L(τ ).

Proof. Since [h, x] = x for all x ∈ h∗ , we see that [h, p] = ip if p ∈ C[h]i . But h · (p ⊗ τ ) =
[h, p] ⊗ τ − pz(1 ⊗ τ ) since each yi acts on τ by zero. This proves the first part. The second part
is [23, 2.27]. The third part is [23, Corollary 2.28].

The simple quotients L(τ ) of the M (τ ) for τ ∈ Irrep(W ) are the building blocks of a very
important category of Hk –modules which we now describe.
If M is a Hk –module and α ∈ C, define the α h–weight subspace of M to be the generalised h–
eigenspace of M with eigenvalue α, that is, the set {x ∈ M : (h−α)N x = 0 for sufficiently large N }.

Definition 1.22. Define a category O(Hk ), “category O”, to be the full subcategory of the category
of Hk –modules consisting of those Hk –modules M such that

1. M is finitely generated;

2. h ⊂ Hk acts locally nilpotently on M . That is, if y ∈ h and m ∈ M then there exists N ∈ N


such that y N m = 0;

3. M is the sum of its h–weight subspaces.

Remark 1.23. It will be shown below in the proof of Theorem 1.25 that part (3) of Definition 1.22
actually follows from parts (1) and (2). We choose to include (3) as part of the definition because
of its extreme importance in the study of category O.

We collect together the basic properties of category O and the modules M (τ ) in the following
proposition.

13
Proposition 1.24. Write O = O(Hk ).

1. O is an abelian category, closed under extensions, subobjects and quotients.

2. For all τ ∈ Irrep(W ), the standard module M (τ ) belongs to O.

3. The set of simple objects of O is {L(τ ) : τ ∈ Irrep(W )}.

4. Every object M of O has a composition series of the form

M = Mk ⊃ Mk−1 ⊃ · · · ⊃ M1 ⊃ M0 = 0

with Mi /Mi−1 ∼
= L(τi ) for some τi ∈ Irrep(W ), 1 ≤ i ≤ k. We write [M : L(τ )] for the
composition multiplicity of L(τ ) in M .

5. Let cτ ∈ C be the scalar by which z acts on τ . Then if σ 1= τ and [M (τ ) : L(σ)] 1= 0, then


cτ − cσ ∈ N.

6. For all τ , [M (τ ) : L(τ )] = 1.

7. M (τ ) is an indecomposable Hk –module

8. M (τ ) ∼
= M (σ) if and only if L(τ ) ∼
= L(σ) if and only if τ ∼
= σ as W –modules.

9. For each τ ∈ Irrep(W ), let ∇(τ ) denote the submodule of HomC[h∗ ]∗W (Hk , τ ) consisting of
those elements m for which dim(C[h] · m) < ∞. Then ∇(τ ) ∈ O and L(τ ) is the unique
simple submodule of ∇(τ ) and [∇(τ ) : L(τ )] = 1.

10. Every simple object L(τ ) of O has a projective cover P (τ ). Furthermore, P (τ ) has a series
of submodules
0 = P0 ⊂ P1 ⊂ · · · ⊂ Pn = P (τ )

such that Pi /Pi−1 is a standard module M (τi ) for each i, Pn /Pn−1 = M (τ ), and the BGG
reciprocity formula [P (τ ) : M (σ)] = [M (σ) : L(τ )] holds for all σ, τ ∈ Irrep(W ).

11. Dually, every L(τ ) has an injective envelope I(τ ) and I(τ ) has a series of submodules

0 = I0 ⊂ I1 ⊂ · · · ⊂ In = I(τ )

such that Ii /Ii−1 is a costandard module ∇(τi ) for all i, I1 /I0 = ∇(τ ), and the BGG reciprocity
formula [I(τ ) : ∇(σ)] = [∇(σ) : L(τ )] holds for all τ and σ.

14
Proof.

1. Follows from [33, Theorem 2.19].

2. This is [33, Lemma 2.3].

3. See [33, Proposition 2.11].

4. See [33, Corollary 2.16].

5. See [23, (32)].

6. See [23, (31)].

7. This is [23, Corollary 2.28].

8. If M (τ ) ∼
= M (σ) then L(τ ) ∼
= L(σ) since L(τ ) is the unique simple quotient of M (τ ). If
L(τ ) ∼
= L(σ) then let λ be the eigenvalue of z on τ . Since L(τ ) is a quotient of M (τ ), the set
{x ∈ L(τ ) : hx = −λx} equals 1 ⊗ τ . This is W –stable since h commutes with W , so there
is an isomorphism of W –modules τ ∼
= σ.

9. The module ∇(τ ) is defined in [33, Section 2.3.1]. The statement that L(τ ) is the unique
simple submodule of ∇(τ ) is given in [33, Section 2.5.1].

10. The statement about projective covers is [33, Corollary 2.10]. By [33, 2.6.2], there is a
reciprocity formula [P (τ ) : M (σ)] = [∇(σ) : L(τ )]. But by [33, Proposition 3.3], we have
[∇(σ)] = [M (σ)] in the Grothendieck group K0 (O), which yields the desired reciprocity
formula.

11. It is shown in [33, Theorem 2.19] that O is a highest-weight category in the sense of Cline-
Parshall-Scott [17]. The existence of a filtration of the given form follows from [17, Definition
3.1 (c)], while the BGG reciprocity is [17, Theorem 3.11] combined with the above-noted fact
that [M (σ)] = [∇(σ)] for all σ ∈ Irrep(W ).

The modules ∇(σ) appearing in the proof of Proposition 1.24 are called costandard modules
We work mostly with standard modules in this thesis, and will use costandard modules only in the
proof of Theorem 3.26.

15
1.3.3 Ideals of Hk and category O

One reason why category O is useful is that it can be related to the ideal structure of Hk by the
following theorem, which will be used in the next chapter to determine when Hk is simple.

Theorem 1.25 (Ginzburg). Let I be a proper ideal of Hk . Then I annihilates a nonzero object
of category O.

Proof. The proof follows from [31, Theorem 2.3]. In the notation of [31], take A = Hk , A+ = C[h]+
(meaning the element of positive degree in C[h]), A− = C[h∗ ]+ . Then (A± , h) is a commutative
triangular structure on A in the sense of [31]. In [31, Definition 2.2], a category ↑ O is defined as
the category of finitely generated A–modules such that the action of A− is locally nilpotent. But
in [31, Theorem 2.5], it is proved that if M ∈ ↑ O and m ∈ M then dim(C[h] · m) < ∞, and hence
m belongs to a sum of generalised h–eigenspaces. Thus, the category O of Definition 1.22 is the
same as ↑ O. Now by [31, Theorem 2.3], any primitive ideal of A annihilates a nonzero object of
↑ O. But any nonzero ideal is contained in a primitive ideal.

1.3.4 The Dunkl representation

Of particular interest is the module M (triv) where triv denotes the trivial representation of W .
Since C[h] ⊗ triv ∼
= C[h] as vector spaces, we get an action of Hk on C[h]. This action may be given
by the following homomorphism Hk → EndC (C[h]).

x 5→ x x ∈ h∗

w 5→ w w∈W
H −1
& αH (y) n&
y 5→ ∇y = ∂y + nH kH,i eH,i y∈h
αH
H∈A i=1

where w denotes the endomorphism p 5→ w(p) = p ◦ w−1 , and x denotes the endomorphism p 5→ xp.
! !
Also, if y = i αi yi then ∂y := i αi ∂x

i
.
The operators ∇y are called Dunkl operators, and this map is called the Dunkl representation.
#
Write δ = H∈A αH . Then the Dunkl operators may be viewed as elements of D(h)[δ−1 ] ∗ W .
Write hreg = h \ ∪H∈A H = {x ∈ h : δ(x) 1= 0}. Then D(hreg ) = D(h)[δ−1 ] by [55, Theorem 15.1.25,
Corollary 15.5.6]. So the Dunkl representation defines a homomorphism

p : Hk → D(hreg ) ∗ W.

16
Proposition 1.26. The set {δn }∞
n=0 of powers of δ forms a left and right Ore set in Hk .

Proof. We must show that, given z ∈ Hk and N ∈ N, there exists M ∈ N and t ∈ Hk with
δM z = tδN . By the PBW Theorem 1.15, it suffices to show this when z is a monomial in {xi },
{yi } and w. But δ commutes with each xi , and every w ∈ W acts by a scalar on δ, so it suffices
to check that, for each 1 ≤ i ≤ n and r ∈ N, there is an M with δM yir ∈ Hk δN . But the
commutation relations of Definition 1.17 imply that [yi , δM ] ∈ (C[h] ∗ W )δM −1 for all M , and
therefore δN +r yir ∈ Hk δN as required. A similar proof shows that the powers of δ form a right Ore
set as well.

Because of Proposition 1.26, we may form the localisation Hk [δ−1 ], and for any Hk –module M ,
we may consider the localisation M [δ−1 ]. We often write these localisations as Hk |hreg and M |hreg
respectively.

Proposition 1.27. [33, Theorem 5.6] The map p is injective and the localization p|hreg : Hk |hreg →
D(hreg ) ∗ W is an algebra isomorphism.

Note that Proposition 1.27 says that we may regard Hk as the subalgebra of D(hreg ) ∗ W
generated by C[h], W and the Dunkl operators.

1.3.5 The Knizhnik-Zamolodchikov functor

The Dunkl representation provides a link between the Cherdnik algebra and differential operators.
This may be used to construct a functor KZ from category O to a category of modules over a certain
finite-dimensional algebra, which is one of the most powerful tools for studying category O, and
will be put to extensive use in the next chapter. We now give a description of how to define this
functor.
Let M be an object of O and consider the localization M [δ−1 ] = M |hreg . This is a module
for Hk |hreg , which is isomorphic to D(hreg ) ∗ W by Proposition 1.27. So M |hreg may be regarded
as a W –equivariant D(hreg )–module. The fixed point set M |W
hreg is then a D(h
reg )W –module. The

algebras D(hreg )W and D(hreg /W ) are isomorphic1 , so M |W


hreg may be regarded as a D–module

on hreg /W , which is a smooth complex variety. By [9, VI, 1.6, 1.7], there is an equivalence of
categories between the category of D–modules on hreg /W and the category of vector bundles on
1
This follows from [49, Théorème 4.2].

17
hreg /W endowed with a flat connection. We show that the connection associated with any object
M of O has regular singularities. Indeed, by [33, Proposition 5.7], the connection associated to
M (τ ) has regular singularities. Now, [19, Proposition 4.6] states that if V $ → V → V $$ is an exact
sequence of vector bundles with flat connection on a smooth complex manifold and V $ , V $$ have
regular singularities, then so does V . From the sequence M (τ ) → L(τ ) → 0 and the fact that the
functors of localization to hreg and taking W –invariants are exact (because W is a finite group),
we get that the connection associated to L(τ ) has regular singularities. Hence, the connection
associated to any object M of O has regular singularities as well. Now by [9, IV, 7.1.1], there is an
equivalence of categories between the category of algebraic vector bundles on hreg /W endowed with
a flat connection with regular singularities, and the category of holomorphic vector bundles on the
complex manifold hreg /W endowed with a flat connection. Finally, by [9, IV, 1.1] and [19, Corollaire
1.4], there is an equivalence of categories between the category of holomorphic vector bundles on
hreg /W endowed with a flat connection, and the category of finite-dimensional representations of
the fundamental group π1 (hreg /W, ∗), where ∗ is any choice of basepoint. In this way, we obtain an
exact functor O → π1 (hreg /W, ∗) − mod. The group π1 (hreg /W, ∗) is called the braid group BW of
W.
Now, following [33, 5.2.5], we define an algebra H to be the quotient of CBW by the relations
H −1
n'
(T − 1) (T − det(s)−j e−2πikH,j ) = 0 (1.1)
j=1

where H ∈ A, s ∈ W is the reflection about H with eigenvalue e2πi/nH , and T is an s–generator


of the monodromy around H in the parlance of [11, Section 4.C]. Note that the sign of kH,j differs
from that given in [33, Section 5.2.5], because there is a sign error in [33], as remarked in [64, Section
5.2.1].
By [33, Theorem 5.13], the functor O → CBW − mod constructed above factors through H, and
in this way we obtain an exact functor

KZ : O → H − modf

where H − modf denotes the category of finite-dimensional H–modules.

Definition 1.28. We call the algebra H = Hk (W ) the Hecke algebra of W with parameters k.

We list the most important properties of the KZ functor in the following proposition.

18
Proposition 1.29. There exists an exact functor (the Knizhnik-Zamolodchikov functor) KZ : O →
H − modf . Denote by Otor the full subcategory of O consisting of those objects M such that
M |hreg = 0. Then KZ factorises as

Q KZ
KZ : O −→ O/Otor −−→H
0
− modf

where Q is the quotient functor. Furthermore, the following properties hold.

1. KZ0 is an equivalence.

2. If kH,i −kH,j + i−j


nH ∈/ Z for all H ∈ A and all i 1= j, then for any M ∈ O and any τ ∈ Irrep(W ),
we have a bijection

HomO (M, M (τ )) → HomH (KZ(M ), KZ(M (τ ))).

Proof. The first part is [33, Theorem 5.14]. The second part follows from [33, Proposition 5.9],
which states that, under the given hypotheses,


HomO (M, M (τ )) → HomO/Otor (QM, QM (τ ))

where Q : O → O/Otor is the quotient functor. Since KZ0 is an equivalence, we get part (2) of the
proposition.

Note that since Hk is Noetherian, for any two finitely generated Hk –modules M, N , there is a

natural isomorphism HomHk (M, N )|hreg → HomHk |h reg (M |hreg , N |hreg ) (see for example [63, 3.84]).
It is easy to check using this fact that the full subcategory of Hk |hreg –modules whose objects are
those modules of the form M |hreg for some M ∈ O satisfies the universal property of O/Otor .
Therefore, we get the following corollary of [33, Proposition 5.9].

Corollary 1.30. Suppose kH,i − kH,j + i−j


nH ∈/ Z for all H ∈ A and all i 1= j. Then for any M ∈ O
and any τ ∈ Irrep(W ), we have a bijection


HomO (M, M (τ )) → HomHk |h reg (M |hreg , M (τ )|hreg ).

19
1.3.6 Double centraliser property

One important property of category O which follows from the existence of the KZ functor is the
so-called double centraliser property. This states that category O is the category of modules over
a certain finite-dimensional algebra, and also that H is the endomorphism ring of an object of
category O. The double centraliser property is not known to hold for all complex reflection groups
since it requires that the dimension of the Hecke algebra H is |W |. This is known to be true in all
cases, except for some of the exceptional complex reflection groups, for which it is conjectured to
be true (see [33, Remark 5.12]).
The double centraliser property is the following pair of theorems.

Theorem 1.31. [33, Theorem 5.15] Let W be a complex reflection group, k a vector of complex
parameters for the Cherednik algebra of W , and Hk the associated Hecke algebra. Suppose that
dim(Hk ) = |W |. For each τ ∈ Irrep(W ), let P (τ ) be the projective cover of L(τ ) in the category
O = O(Hk (W )). Define an object PKZ ∈ O by

PKZ = ⊕τ ∈Irrep(W ) dim KZ(L(τ )) · P (τ ).

Then there is an isomorphism of algebras

Hk ∼
= EndO (PKZ )opp .

Dually, we have the following theorem.

Theorem 1.32. [33, Theorem 5.16] Let W be a complex reflection group, k a vector of complex
parameters for the Cherednik algebra of W , and Hk the associated Hecke algebra. Suppose that
dim(Hk ) = |W |. Suppose X ∈ O is a projective generator for the category O (for example, X can
be taken to be ⊕τ ∈Irrep(W ) P (τ )). Then there is an equivalence of categories

O∼
= EndHk (KZ(X))opp − mod.

20
1.3.7 Twisting by a linear character

Following [33, Section 5.4.1], let ζ be a linear character of W (that is, a group homomorphism
W → C∗ ), and define an automorphism of T (h ⊕ h∗ ) ∗ W via

x 5→ x x ∈ h∗

y 5→ y y∈h

w 5→ ζ(w) · w w ∈ W.

From the generators and relations 1.17, it is easy to check that this defines a map Hk → Hζ(k)
where ζ(k) is defined as follows. There is some u(H) such that ζ|WH = det−u(H) |WH . Then

ζ(k)H,i = kH,u(H)+i − kH,u(H)

for 1 ≤ i ≤ nH − 1, where the subscripts are taken modulo nH . By [33, 5.4.1], this isomorphism
Hk → Hζ(k) gives an equivalence of categories Fζ : O(Hk ) → O(Hζ(k) ) which satisfies Fζ V (τ ) =
V (τ ⊗ ζ −1 ) where V stands for any of the symbols M , L or P . Furthermore, if Hk (W ) denotes the
Hecke algebra associated to O(Hk ) via the KZ functor, then there is a commutative diagram


O(Hk ) " O(Hζ(k) )

KZ KZ

! Gζ !
Hk (W ) − mod " Hζ(k) (W ) − mod

where Gζ is an equivalence induced by an isomorphism of algebras Hk (W ) ∼


= Hζ(k) (W ).

1.3.8 Absence of self-extensions

Before specialising to the case W = G(!, 1, n), we give one more useful property of category O for
a general complex reflection group W , namely the absence of self-extensions of simple objects. We
give the proof, which is identical to the proof given in [7] in the symmetric group case.

Theorem 1.33. [7, Proposition 1.12] For any τ ∈ Irrep(W ), Ext1O (L(τ ), L(τ )) = 0.

21
i π
Proof. Suppose there is a short exact sequence 0 → L(τ ) → N → L(τ ) → 0. Then the set of
generalised h–eigenvalues of N is equal to the set of generalised h–eigenvalues of L(τ ). Consider
the lowest weight space N0 ⊂ N , that is, if z acts on τ by λ ∈ C, then

N0 := {x ∈ N : (h + λ)k x = 0 for sufficiently large k}.

If x ∈ N0 and y ∈ h, then yx is a generalised h–eigenvector with a lower eigenvalue than x, since


[h, y] = −y. Therefore, yx = 0. So hN0 = 0. Now choose a 0 1= v ∈ i(1 ⊗ τ ) ⊂ N0 and a v $ ∈ N0
with 0 1= π(v $ ) ∈ 1 ⊗ τ . Then Hk v $ is a quotient of M (τ ), since hv $ = 0. If Hk v $ has length > 1 then
Hk v $ = N and so M (τ ) surjects onto N , which contradicts [M (τ ) : L(τ )] = 1. Therefore, Hk v $ is
simple. Also, Hk v = im(i) is simple, since it is a nonzero quotient of L(τ ). Then if Hk v ∩ Hk v $ 1= 0
then Hk v = Hk v $ and so 0 1= π(v $ ) ∈ π(Hk v) = 0, a contradiction. Therefore, Hk v ∩ Hk v $ = 0
which forces N = Hk v ⊕ Hk v $ = im(i) ⊕ Hk v $ , and the sequence splits.

1.4 The case of W = G(!, 1, n)

Let us introduce the infinite family of complex reflection groups G(!, p, n). Let !, n ≥ 1 and p|! be
natural numbers. Then G(!, 1, n) may be defined to be the set of all n × n complex matrices with
exactly one nonzero entry in each row and column, such that the nonzero entries are powers of
2πi
ε := e # . In this way, G(!, 1, n) has a natural defining representation h = Cn . Note that G(!, 1, n)
is isomorphic to the group (Z/!Z) ! Sn = (Z/!Z)n ! Sn . This representation is irreducible if ! > 1.
The group G(1, 1, n) is the symmetric group of degree n, and h = Cn is its natural representation.
The group G(2, 1, n) is the Weyl group of type Bn . If ! > 2 then G(!, 1, n) is not a Coxeter group.
Further, define G(!, p, n) to be the normal subgroup of G(!, 1, n) consisting of those matrices in
G(!, 1, n) such that the pth power of the product of the nonzero entries is 1.

1.4.1 The rational Cherednik algebra of G(!, 1, n)

From now on, we take W = G(!, 1, n). Let {y1 , y2 , . . . , yn } denote the standard basis of h and
{x1 , . . . , xn } the dual basis of h∗ . Then the complex reflections in W may be described as follows.
First, for 1 ≤ i ≤ n and 1 ≤ t ≤ ! − 1, we have an element sti ∈ W defined by

sti (yi ) = εt yi

sti (yj ) = yj j 1= i

22
(t)
and for 1 ≤ i < j ≤ n and 0 ≤ t ≤ ! − 1 we have an element σij defined by
(t)
σij (yi ) = ε−t yj
(t)
σij (yj ) = εt yi
(t)
σij (yk ) = yk k 1= i, j
(t)
The set of complex reflections of W is {sti : 1 ≤ i ≤ n, 1 ≤ t ≤ ! − 1} ∪ {σij : i < j, 0 ≤ t ≤ ! − 1}.
(t)
The reflecting hyperplane of sti is Hi = {v ∈ h : xi (v) = 0} while the reflecting hyperplane of σij is
Ii,j,t = {v ∈ h : xi (v) = ε−t xj (v)}. For each reflecting hyperplane H, we choose a linear functional
αH and a vector vH as in Section 1.3. We can choose αHi = xi , αIi,j,t = xi − εt xj , vHi = yi and
vIi,j,t = yi − ε−t yj . The hyperplanes Hi belong to a single W –orbit Cs while Cσ = {Ii,j,t} form
another W –orbit. We have nCs = ! and nCσ = 2. We write κ00 = kIi,j,t ,1 for all i, j, t and κj = kHi ,j
for all i, where κ0 = κ( = 0. Then the rational Cherednik algebra Hκ = Hκ (W ) of W is the
quotient of the C–algebra T (h ⊕ h∗ ) ∗ W by the relations [x, x$ ] = 0 for x, x$ ∈ h∗ , [y, y $ ] = 0 for
y, y $ ∈ h, together with the commutation relations
n
& (−1
& (−1
&
[y, x] = y(x) + y(xi )x(yi ) (κj+1 − κj ) εrj sri
i=1 j=0 r=0

& (−1
& (t)
+ κ00 y(xi − εt xj )x(yi − ε−t yj )σij (1.2)
1≤i<j≤n t=0

for all x ∈ h∗ and all y ∈ h. Usually we write the parameters as a vector of complex numbers
κ = (κ00 , κ1 , . . . , κ(−1 ) ∈ C( and we write Hκ = Hκ (W ) for the Cherednik algebra. We also
write Oκ , Hκ and KZκ when we want to emphasise the dependency on the parameters, but we will
sometimes drop κ from the notation when we are working with a fixed choice of κ and there is no
ambiguity.
With respect to the standard basis {y1 , . . . , yn } of h and the dual basis {x1 , . . . , xn } of h∗ , we
may rewrite the commutation relations (1.2) as follows:
(−1
& (−1
& (−1
&& (t)
[yi , xi ] = 1 + (κp+1 − κp ) ε si + κ00
pr r
σir 1≤i≤n
p=0 r=0 r+=i t=0
(−1
& (t)
[yi , xj ] = −κ00 εt σji i 1= j. (1.3)
t=0

For the algebra Hκ = Hκ (G(!, 1, n)) we have all the theory of the preceding sections: category
O = Oκ , the Dunkl representation and the KZ functor O → H − mod. We now identify the algebra

23
H through which the KZ functor factors in the W = G(!, 1, n) case. Recall that this algebra is the
quotient of CBW by the relations (1.1) of Section 1.3.5.

Definition 1.34. [2] Let !, n be positive integers and let q, u1 , u2 , . . . , u( ∈ C be complex numbers.
The Ariki-Koike algebra with parameters (q, u1 , . . . , u( ) is the algebra generated by Ts , Tt2 , . . . , Ttn
subject to the relations:

Ts Tt2 Ts Tt2 − Tt2 Ts Tt2 Ts = 0

[Ts , Tti ] = 0 i≥3

Tti Tti+1 Tti − Tti+1 Tti Tti+1 = 0 2≤i≤n−1

[Tti , Ttj ] = 0 |i − j| > 1

(Tti − 1)(Tti + q) = 0 2≤i≤n


(
'
(Ts − 1) (Ts − uj ) = 0
j=2

From the braid diagram in [11, Table 1], we see that the Hecke algebra H through which
KZκ factors is the Ariki-Koike algebra with parameters (q, u1 , . . . , u(−1 ) where q := e2πiκ00 and
uj := ε−j+1 e−2πiκj−1 for 1 ≤ j ≤ !, where as before, ε = e2πi/( . Note in particular that q 1= 0 and
uj 1= 0 for all j. We denote this algebra by H(q,u1 ,...,u#−1) (G(!, 1, n)). The Ariki-Koike algebra will
be our main tool for obtaining information about category O, so we now describe its representation
theory.

1.4.2 Representation theory of the Ariki-Koike algebra

A good reference for the Ariki-Koike algebra is the survey article [54]. We remark that H is a
finite-dimensional algebra of dimension n!!n = |W |, by [54, Theorem 2.2]. If we put q = 1 and
ui = εi−1 for all i in Definition 1.34, then the Ariki-Koike algebra reduces to the group algebra
CW . Thus, H may be regarded as a deformation of CW .
Let n be a a positive integer. A partition of n is by definition a sequence of nonnegative
!
integers λ1 " λ2 " · · · " λt with i λi = n. We ignore trailing zeroes, so that the partition

λ1 " λ2 " · · · " λt is identified with λ1 " λ2 " · · · " λt " 0 " 0 " · · · . Each partition has a
Young diagram, which is a left-justified array of n boxes in the plane, with λi boxes in the ith row.

24
For example, the Young diagram

!
corresponds to the partition 5 " 3 " 2 of 10. For a partition λ of n, we write |λ| = n = i λi .
We identify λ with its Young diagram and label the rows of the Young diagram from the top down
and the columns from left to right. The box in the (i, j) position is called the (i, j)–node of λ. For
example, in the Young diagram of 5 " 3 " 2,

a
λ=
b

the box labelled a is the (1, 2)–node of λ, and b is the (3, 1)–node. In general, if x is the (i, j)–node,
we write row(x) = i and col(x) = j.
A multipartition of n with ! parts is an !–tuple of partitions λ = (λ(1) , λ(2) , . . . , λ(() ) with
!( (i)
i=1 |λ | = n. We may identify a multipartition λ with an !–tuple of Young diagrams, and hence
regard λ as a subset of N × N × N. A node (i, j, k) of λ is defined to be a node (i, j) of λ(k) for some
k. More generally, a node is any element of N × N × N.
By [54], for each multipartition λ of n with ! parts, the Ariki-Koike algebra H has a Specht
module S λ . These modules are the cell modules with respect to a certain cellular basis of H.
Each Specht module has a quotient Dλ which is either 0 or absolutely irreducible, and the set
{D λ : Dλ 1= 0} is a complete set of pairwise nonisomorphic simple H–modules (see [54, Theorem
3.12]). We will need a parametrisation of this set. There are two different parametrisations,
depending on whether q = 1 or q 1= 1.

Lemma 1.35. [52, Theorem 3.7] If q = 1 and λ is a multipartition of n with ! parts, then Dλ 1= 0
if and only if λ(s) = ∅ whenever s < t and us = ut .

If q 1= 1 then the description, due to Ariki and stated in [54, Theorem 3.24] is more complicated.
The nonzero Dλ are in bijection with the set of Kleshchev multipartitions, which we now describe.
Given a multipartition λ, the residue of a node x in row i and column j of λ(k) is defined to be
uk q j−i . A node x in λ with residue a is called a removable a–node if λ \ {x} is a multipartition. A
node x not in λ with residue a is called an addable a–node if λ ∪ {x} is a multipartition.
Say a node y ∈ λ(r) is below a node x ∈ λ(k) if either r > k, or r = k and row(y) > row(x).

25
A removable a–node x is called normal if whenever x$ is an addable a–node below x then there
are more removable a–nodes between x and x$ than there are addable a–nodes. The highest normal
a node in λ is called the good a–node.
The set of Kleshchev multiparitions with ! parts is defined inductively as follows: the empty
partition (∅, . . . , ∅) is Kleshchev, and otherwise λ is Kleshchev if and only if there is some a ∈ C
and a good a–node x ∈ λ such that λ \ {x} is Kleshchev.

Lemma 1.36. Suppose q 1= 1. Then the set {D λ : Dλ 1= 0} of nonisomorphic irreducible


H(q,u1 ,...,u# ) (G(!, 1, n))–modules is in bijection with the set of Kleshchev multipartitions of n with !
parts.

Example 1.37. We now give an example to illustrate the above definitions. Consider the multi-
partition λ of n = 2 defined by
$ %
λ= Ø .

The addable nodes of λ are the nodes labelled x, y, z on the following diagram. The node labelled
r is the only removable node. ( )
r x
z
y
Let q be a non-root of unity and suppose u1 = u2 = q. Then the residues of the nodes of λ and the
addable nodes of λ are given by the following diagram.
( )
q q2 q3
q
1

The node r is a removable q 2 –node. There is no addable q 2 –node below r, and so r is a good q 2 –
$ %
node. Thus, λ is Kleshchev if and only if µ := Ø is Kleshchev. But µ has one removable
node, and this node has residue q, and there is an addable node below it with residue q, and no
removable nodes lie between them. Thus, the unique removable node of µ cannot be a good node.
Therefore, µ is not Kleshchev and so λ is not Kleshchev.

By [54, Theorem 3.13], the Ariki-Koike algebra is semisimple if and only if S λ = Dλ for all
λ. Furthermore, semisimplicity can be expressed entirely in terms of the parameters thanks to the
following semisimplicity criterion.

26
Theorem 1.38. [1, Main Theorem] The algebra H = H(q,u1 ,...,u# ) (G(!, 1, n)) is semisimple if and
only if
' d=(−1
'
[n]q ! (ui − q d uj ) 1= 0,
i+=j d=−(+1
#n
where [n]q ! = j=1 (1 + q + · · · + q j−1 ).

Finally we need a description of the blocks of H. This is given in [51, Corollary 2.16]. Define
the content cont(λ) of a multipartition λ to be the multiset of residues of λ, ie. the set of residues
counted according to multiplicity. Then by [51, Lemma 2.2], all the composition factors of a Specht
module S λ for H belong to the same block, so we can say a Specht module belongs to a block B if
all its composition factors belong to B.

Theorem 1.39. [51, Corollary 2.16] Let λ and µ be multipartitions of n with ! parts. If q 1= 1,
two Specht modules S λ and S µ of H belong to the same block if and only if cont(λ) = cont(µ).

1.4.3 Fourier map

We close this chapter with an analogue for W = G(!, 1, n) of the so-called Fourier automorphism of
the Cherednik algebra of Sn , which will be needed in the proof of Lemma 4.39. It is easy to see that
(0)
the group W is generated by the elements s1 and σi := σi−1,i , 2 ≤ i ≤ n. In fact, by [2, Proposition
2.1], W has a presentation as the abstract group generated by symbols s1 and σi , 2 ≤ i ≤ n, subject
to the relations

s(1 = 1

s 1 σ2 s 1 σ2 = σ2 s 1 σ2 s 1

σi2 = 1 1 ≤i ≤n−1

σi σi+1 σi = σi+1 σi σi+1 2 ≤i ≤n−1

σi σj = σj σi |i − j| > 1

s 1 σi = σi s 1 i ≥ 3. (1.4)

Calculating with the defining representation of W shows that these relations imply sp1 σ2 sp1 σ2 =

σ2 sp1 σ2 sp1 for all integers p. It follows that there is an automorphism (sic) W → W defined on the
(k) (0)
generators by s1 5→ s−1 −k
1 and σi 5→ σi for all i. Note that since σij = si σij si , this automorphism
k

27
(k) (−k)
sends σij to σij for all i, j, k. This can be extended to a map

ψ : T (h ⊕ h∗ ) ∗ W → T (h ⊕ h∗ ) ∗ W

by setting ψ(yi ) = xi and ψ(xi ) = −yi . This map descends to the Cherednik algebra, and by
checking the relations (1.3) of Section 1.4, we get the following theorem (which has not appeared
before in the literature).

Theorem 1.40. Let κ = (κ00 , κ1 , . . . , κ(−1 ) ∈ C( and define κ = (κ00 , κ1 , κ1 −κ(−1 , κ1 −κ(−2 , . . . , κ1 −
κ2 ). Then there is an algebra isomorphism

ψ : Hκ → Hκ

given on the generators as above. In particular, ψ(C[h]) = C[h∗ ] and ψ(C[h∗ ]) = C[h].

If ! = 1 or ! = 2 then κ = κ, and ψ coincides with the Fourier automorphism of Hκ (W )


defined in [33, Remark 4.6]. Otherwise we may write ψ = ψκ : Hκ → Hκ and since κ = κ, we get
(ψκ ψκ )2 = idHκ .

28
Chapter 2

The semisimple case

In this short chapter, we consider the Cherednik algebra Hκ of the group W = G(!, 1, n). We
determine when Hκ is simple and show that the simplicity of the ring Hκ is equivalent to the
semisimplicity of the category Oκ (recall that an abelian category A is said to be semisimple if and
only if every short exact sequence in A splits).
In fact, we will prove that simplicity of the Cherednik algebra Hk (W ) is equivalent to semisim-
plicity of the Hecke algebra Hk (W ) for any complex reflection group W and parameters k such that
dim Hk = |W | (recall that by [33, Remark 5.12], this is known to hold for all complex reflection
groups apart from some of the exceptional types, and is conjectured to hold for these also).

2.1 The semisimple case

Let W be a complex reflection group and let k be a vector of parameters for the Cherednik algebra
of W . By Proposition 1.27, Hk |hreg is isomorphic to D(hreg ) ∗ W , a simple ring. So Hk is in some
sense close to being simple. Before we begin the study of the more interesting cases where Hk is
not simple, it is helpful to know precisely when Hk is a simple ring. We will answer this in the
G(!, 1, n) case with the following theorem.

Theorem 2.1. Let W be a complex reflection group and k a vector of complex parameters for
the Cherednik algebra of W . Write Hk = Hk (W ), O = O(Hk (W )) and Hk = Hk (W ), the Hecke
algebra associated to k. Suppose dim Hk = |W |.
Then the following are equivalent.

29
1. Hk is a simple ring.

2. Category Ok is a semisimple category.

3. The Hecke algebra Hk is semisimple.

In particular, if W = G(!, 1, n) and κ ∈ C( , then Hκ is a simple ring if and only if Oκ is semisimple


if and only if the Ariki-Koike algebra Hκ is semisimple.

We begin with a lemma first proved in [6, Remark, page 9].

Lemma 2.2. [6] Let W be an arbitrary complex reflection group and k a vector of complex
parameters for the Cherednik algebra. Category O = O(Hk ) is semisimple if and only if M (τ ) =
L(τ ) for all τ ∈ Irrep(W ).

Proof. Suppose M (τ ) = L(τ ) for all τ . By [33, Corollary 2.10], the projective cover P (τ ) of L(τ )
has a filtration 0 = Q0 ⊂ Q1 ⊂ · · · ⊂ Qd = P (τ ) with Qi /Qi−1 a standard module for all i,
and Qd /Qd−1 = M (τ ). By BGG reciprocity, [P (τ ) : M (σ)] = [M (σ) : L(τ )] for all τ, σ. Thus,
[P (τ ) : M (σ)] = 0 if τ 1= σ, while [P (τ ) : M (τ )] = 1. So P (τ ) ∼
= M (τ ) ∼
= L(τ ). Therefore, all the
L(τ ) are projective and, since every object of O has a finite filtration by L(τ ), every object of O is
projective. So O is semisimple.
Conversely, suppose O is semisimple. Then M (τ ) = L(τ ) ⊕ R(τ ) for all τ . But M (τ ) is
indecomposable, so M (τ ) = L(τ ) for all τ .

Our next lemma relates semisimplicity of O to localisation. Recall that Otor is the full subcat-
egory of objects M in O with M |hreg = 0.

Lemma 2.3. Let W be an arbitrary complex reflection group and k a vector of complex parameters.
Category O = O(Hk ) is semisimple if and only if Otor = 0.

Proof. Suppose O is semisimple. Then by Lemma 2.2, L(τ ) = M (τ ) for all τ . So L(τ ) is a free
C[h]–module for all τ , and so L(τ )|hreg = C[hreg ] ⊗C[h] L(τ ) 1= 0 for all τ . Therefore, M |hreg 1= 0 for
all nonzero M in O.
Conversely, suppose Otor = 0. Then L(τ )|hreg 1= 0 for all τ . Now apply [33, Proposition 5.21],
which states that if L(τ )|hreg 1= 0 then L(τ ) is a submodule of a standard module. By Proposition
1.24, we may choose an ordering τ1 < τ2 < · · · < τm on Irrep(W ) such that [M (τi ) : L(τj )] 1= 0 only

30
if i ≤ j. We show by induction on this ordering that M (τ ) = L(τ ) for all τ . First, consider L(τ1 ).
Since L(τ1 )|hreg 1= 0, L(τ1 ) ⊂ M (τi ) for some i. So i ≤ 1, therefore i = 1 and L(τ1 ) ⊂ M (τ1 ). But
[M (τ1 ) : L(τ1 )] = 1 and L(τ1 ) is a quotient of M (τ1 ). So we conclude that M (τ1 ) = L(τ1 ). Now
consider L(τ2 ). Then L(τ2 ) ⊂ M (τi ) for some i, so i ≤ 2. But i 1= 1 since we have already shown
that the only composition factor of M (τ1 ) is L(τ1 ). Therefore i = 2 and M (τ2 ) = L(τ2 ). Continuing
inductively, we get M (τ ) = L(τ ) for all τ . Therefore, O is semisimple by Lemma 2.2.

We can now prove that parts (2) and (3) of Theorem 2.1 are equivalent. This was first proved
in [35] by a different method.

Lemma 2.4. Let W be a complex reflection group and let k be a vector of parameters for the
Cherednik algebra of W . Let Hk (W ) be the Cherednik algebra of W and let O = O(Hk (W )). Let
Hk be the Hecke algebra associated to k. Suppose dim Hk = |W |. Then O is a semisimple category
if and only if Hk is a semisimple algebra.

Proof. Suppose O is semisimple. Then Otor = 0 by Lemma 2.3, and it follows that the functor KZ0
is an equivalence O → Hk − mod. So Hk − mod is a semisimple category and therefore Hk is a
semisimple algebra.
Conversely, suppose Hk is semisimple. By the hypothesis, dim(Hk ) = |W |, and so we may
apply Theorem 1.32 to get that

O∼
= EndHk (KZ(X))opp − mod

for some object KZ(X) of Hk −mod. Since Hk is semisimple, KZ(X) is a direct sum of simple modules,
and hence EndHk (KZ(X)) is a direct product of matrix algebras Matri (C), so is a semisimple algebra
by Wedderburn’s theorem. Therefore, EndHk (KZ(X))opp − mod is a semisimple category, so O is
also semisimple.

Lemma 2.5. Let W be a complex reflection group and k a vector of complex parameters for the
rational Cherednik algebra of W . The rational Cherednik algebra Hk = Hk (W ) is a simple ring if
and only if O = O(Hk (W )) is a semisimple category.

Proof. Suppose Hk is a simple ring. Let τ ∈ Irrep(W ). Then L(τ ) is a nonzero Hk –module, so
the left annihilator of L(τ ) is a proper two-sided ideal of Hk , and is therefore zero. We show
that L(τ )|hreg 1= 0. If not, then L(τ )|hreg = 0. But L(τ ) = M (τ )/R(τ ), so as a vector space

31
L(τ ) = (C[h] ⊗ τ )/R(τ ). Let v ∈ τ be nonzero, and consider 1 ⊗ v + R(τ ) ∈ L(τ ). Then there is
some t with δt (1 ⊗ v) + R(τ ) = R(τ ), and therefore if f ∈ C[h] then δt (f ⊗ v) = 0 in L(τ ). Also, we
#
may as well take t to be a multiple of H∈A |WH |, so that δt is W –invariant. Then δt annihilates
L(τ ), a contradiction. Therefore, L(τ )|hreg 1= 0. Therefore, O is semisimple by Lemma 2.3.
Finally, if O is semisimple, we show that Hk is simple. Let I ⊂ Hk be a proper two-sided ideal.
Then I annihilates a simple object L(τ ) of O by Theorem 1.25. But L(τ ) = M (τ ) by Lemma 2.2.
So IM (τ ) = 0. Therefore, I|hreg annihilates M (τ )|hreg 1= 0. So I|hreg 1= Hk |hreg and since Hk |hreg
is a simple ring, we get I|hreg = 0. But I ⊂ Hk and, by the PBW Theorem, Hk is a torsionfree
C[h]–module, so I|hreg = 0 implies I = 0 as required.

Proof of Theorem 2.1

Theorem 2.1 follows immediately from Lemma 2.4 combined with Lemma 2.5. In the case of
W = G(!, 1, n), it is known that dim H = |W | by [54, Theorem 2.2], so Theorem 2.1 applies to W .

Example 2.6 (the cyclic case). Some insight into the Cherednik algebra of G(!, 1, n) can be
obtained by studying the simplest case, namely when n = 1. In this case, all calculations can be
done explicitly.
We let W = G(!, 1, 1) ∼
= Z/!Z = ,s-, acting on h = C via s(y) = εy where {y} is a basis of
h. Let κ1 , . . . , κ(−1 ∈ C. From the relations (1.2), we see that the Cherednik algebra Hκ (W ) is
generated by three elements x, y, s with the relations

sxs−1 = ε−1 x

sys−1 = εy
(−1
&
[y, x] = 1 + (κp+1 − κp )!ep (2.1)
p=0
!(−1
where ep = 1
( j=0 ε s .
pj j

The group W has ! irreducible representations which we denote τi , 0 ≤ i ≤ ! − 1. We take


τi = Cei . We wish to consider the standard modules M (τi ). The module M (τi ) is isomorphic to
C[x] as a vector space and has a basis {xa ⊗ τi : a ≥ 0}. From the commutation relation (2.1) and
induction, we obtain  
(−1
& a−1
&
[y, xa ] = a + (κp+1 − κp )! ep+r  xa−1 .
p=0 r=0

32
It follows that the action of y on xa ⊗ τi ∈ M (τi ) is given by

y · (xa ⊗ τi ) = (a + !(κi+a − κi ))xa−1 ⊗ τi , (2.2)

where the subscript i of κi is taken modulo !. Since any Hκ –submodule of M (τi ) is a C[x]–
submodule, we get that any proper quotient of M (τi ) must be finite-dimensional, and furthermore
from equation (2.2), we see that M (τi ) has a proper quotient if and only if a + !(κi+a − κi ) = 0 for
some a ≥ 1.
Now we describe the Hecke algebra of W . The braid group of W is by definition the fundamental
group π1 (C∗ /W ) ∼
= π1 (C∗ ) = Z, so the Hecke algebra has one generator T , and using the relations
(1.1), may be described as
C[T ]
H = #(
j=1 (T − uj )
where uj = ε−(j−1) ε−2πiκj for all j. This algebra is semisimple unless ut+r = ut for some t and
some ! > r > 0. This happens if and only if κt+r − κt + r( ∈ Z. So !(κt+r − κt ) + r − α! = 0 for some
α ∈ Z. If r − α! > 0 then by equation (2.2), there is a finite-dimensional simple module in category
O. If r − α! < 0 then !(κt+r+((−r) − κt+r ) + (α − 1)! + (! − r) = 0, and so again by equation (2.2),
category O contains a finite-dimensional module. So we see that M (τ ) = L(τ ) for all τ ∈ Irrep(W )
if and only if the algebra H is semisimple. This confirms Lemma 2.4.

2.2 Remarks

1. Let W = G(!, 1, n). From the semisimplicity criterion 1.38, we get that Hκ is simple if and
only if
' $ %
[n]e2πiκ00 ! ε−i e−2πiκi − ε−j e−2πi(κj +cκ00 ) 1= 0.
i+=j,−(≤c≤(

In particular, the set of κ ∈ C( such that Hκ is simple, is the complement of a countable union
of hyperplanes, and hence is a dense subset of C( in the Euclidean topology (such parameters
are sometimes said to be Weil generic).

2. Our proof of Lemma 2.5 is original. The implication Hk semisimple =⇒ Hk simple was
proved in [6, Theorem 3.1]

3. The calculations of Example 2.6 are standard. See for example [16, Section 2.2].

33
Chapter 3

The almost-semisimple case

3.1 Finite-dimensional modules

In this chapter, we take W = G(!, 1, n). For κ ∈ C( , we write as usual Hκ for the Cherednik
algebra, Hκ for the Ariki-Koike algebra, Oκ for the category O(Hκ ), and KZκ for the KZ functor.
We have seen in Chapter 2 that Hκ is almost always simple. Furthermore, Hκ |hreg ∼
= D(hreg ) ∗
W = (D(h) ∗ W )[δ−1 ], which is a simple ring by Proposition 1.11. So Hκ is close to being simple.
However, it is an interesting fact that not only is Hκ not always simple, in fact, it can have finite-
dimensional modules. These modules are of great interest and we will give one application of them
in Chapter 5. Much of the research on Cherednik algebras to date has involved constructing and
classifying their finite-dimensional modules, see for example [7], [15], [20] and [34]. The study
of finite-dimensional modules can be reduced to category O because of the following well-known
theorem.

Theorem 3.1. Every finite-dimensional Hκ –module belongs to category O.

Proof. A finite-dimensional Hκ –module M is clearly a direct sum of generalised h–eigenspaces.


Since [h, y] = −y for all y ∈ h, if Wα denotes the generalised eigenspace of M with eigenvalue α,
then hWα ⊂ Wα−1 . But the set of the real parts of the generalised eigenvalues of h on a finite-
dimensional module is bounded below, so each y ∈ h must act locally nilpotently, as required.

Another useful fact about finite-dimensional modules is that they are not “seen” by the KZ
functor.

34
Proposition 3.2. If L is a finite-dimensional Hκ –module then KZ(L) = 0.

Proof. Suppose L is a finite-dimensional Hκ –module. Then the left annihilator of L is nonzero,


since Hκ is an infinite-dimensional algebra. So there is some x ∈ Hκ with xL = 0. Therefore,
annD(hreg )∗W (L|hreg ) 1= 0. But D(hreg ) ∗ W is a simple ring by Proposition 1.11. So L|hreg = 0 and
hence KZ(L) = 0.

In the W = G(!, 1, n) case, a family of finite-dimensional modules has been constructed by


Chmutova and Etingof, which will be very useful to us. We summarise the results we need in the
following theorem.

Theorem 3.3. [16] Suppose !(n − 1)κ00 + !κs = −s + t! < 0 for some t ∈ Z and some 1 ≤ s ≤ !,
and suppose [n]q ! 1= 0 where q = e2πiκ00 . Then there exists a finite-dimensional quotient Ỹc of
M (triv) and, as graded W –modules, Ỹc ∼
= Ur⊗n where r := s − t! > 0 and Ur is the representation
C[u]/(ur ) of ,s1 - ∼
= Z/!Z with s1 u = ε−1 u, while Sn acts by permuting the factors of the tensor
product Ur⊗n . In particular, dim(Ỹc ) = (s − t!)n .

Proof. By the definition of the Cherednik algebra in [16, Section 2.1, Section 4.1], the parameters
k and cj of [16] are related to κ00 , κi by

κ00 = −k
r−1 (−1
2 & & −at
κr = − ε ct .
! a=0 t=1

Translating our parameters into the language of [16], we therefore get


(−1
& 1 − ε−ts
!(n − 1)k + 2 ct =r
1 − ε−t
t=1

where r = s − !t is a positive integer of the form (p − 1)! + s for some nonnegative integer p and
some 1 ≤ s ≤ ! − 1. Then we have the module Y˜c defined in [16, Theorem 4.2] which is a quotient
of M (triv). Furthermore, since [n]q ! 1= 0, we may apply [16, Theorem 4.3] to conclude that Ỹc is
finite-dimensional, and by [16, Theorem 4.3 (ii)] and [16, Theorem 4.2] we get that Y˜c ∼
= Ur⊗n .

In the case of W = Sn , the rational Cherednik algebra depends on only one parameter c = κ00 ,
and the finite-dimensional modules have been completely classified in the paper [7]. The result is
as follows.

35
Theorem 3.4. [7] Suppose the rational Cherednik algebra Hc (Sn ) has a finite-dimensional module.
Then the following hold.

1. There are r, n ∈ N with (r, n) = 1 and c = ±r/n.

2. For some linear character χ of Sn , L(χ) is finite-dimensional, and if τ 1= χ then dim L(τ ) =
∞.

3. Category O splits as O = O∧ ⊕ Oss , where O∧ is generated by L(∧k h ⊗ χ), 0 ≤ k ≤ n, and


Oss is a semisimple category generated by the other simple objects.

4. The composition multiplicities in the nontrivial block O∧ are




1 if j = i, i + 1
[M (∧ h ⊗ χ) : L(∧ h ⊗ χ)] =
i j

0 otherwise

Proof. Parts 1 and 2 follow from [7, Theorem 1.2]. Part 3 is [7, Theorem 1.3 (iii)]. Part 4 is [7,
Corollary 3.10].

We cannot hope that Theorem 3.4 will hold in the case of G(!, 1, n) for any !. For example,
consider the case of the cyclic group W = G(!, 1, 1). We use the notation of Example 2.6. We
have the standard module M (τi ), 0 ≤ i ≤ ! − 1. Equation (2.2) shows that if we take κi = −i/!
for 1 ≤ i ≤ ! − 2, then in fact Hκ (G(!, 1, 1)) has ! − 1 finite-dimensional simple modules and just
one infinite dimensional simple. If, however, we insist that there is exactly one finite-dimensional
simple module in category O, then we have the following analogue of Theorem 3.4. The proof of
the following proposition can be viewed as a toy version of the proof of Theorem 3.7 to follow.

Proposition 3.5. Suppose W = G(!, 1, 1) = Z/!Z and there is exactly one finite-dimensional
simple module in the category O = O(Hκ (W )). Then the unique finite-dimensional simple module
is L(τi ) for some i, and there exists a 1= 0 such that there is an exact sequence

0 → M (τa+i ) = L(τa+i ) → M (τi ) → L(τi ) → 0,

implying that L(τi ) and L(τa+i ) belong to the same block. Furthermore, for each j 1= i, i+a, {L(τj )}
is a semisimple block of O.

36
Proof. In the notation of Example 2.6, if L(τi ) is finite-dimensional then we must have y·(xa ⊗τi ) = 0
for some a ≥ 1. The generator s of W acts on xa ⊗ τi via the scalar ε−(i+a) . Let R(τi ) denote
the unique maximal submodule of M (τi ). Then we have a map M (τi+a ) # R(τi ) defined by
1 ⊗ τi+a 5→ xa ⊗ τi . Since we assumed that L(τi+a ) was infinite-dimensional, M (τi+a ) must be
simple, and so this map is an isomorphism.
In order to prove the statement about the blocks, note from Example 2.6 that if u1 , . . . , u(
denote the parameters in the Hecke algebra H, then two of the ui must coincide, since there is
a finite-dimensional simple module in category O. Furthermore, no more than two of the ui can
coincide, or else there would be more than one finite-dimensional simple module, by Equation (2.2).
Relabelling the parameters, let us say u1 = u2 := u, and {u, u3 , u4 , . . . , u( } are all distinct. Then
the block decomposition of H is
( )
∼ C[T ] ( C[T ]
H= × ×i=3 .
(T − u)2 (T − ui )

Therefore, H has ! − 1 blocks. By [33, Corollary 5.18], there is a bijection between blocks of O and
blocks of H. Since we have already shown that one block of O contains at least two simples, all the
other blocks must be singletons. Furthermore, these blocks are semisimple by Theorem 1.33.

3.2 The main theorem

We return to the case of W = G(!, 1, n). We have seen in Section 2.2 that Oκ is semisimple for
almost all values of κ. Now let us consider the case where Oκ is not semisimple. It turns out
that if Oκ = O is close to being semisimple then we can still completely describe the structure of
category O, and yet category O can contain finite-dimensional modules. Note that by Lemma 2.3,
O is semisimple if any only if KZ is an equivalence. Motivated by this, we look for conditions under
which KZ is close to being an equivalence. We make the following definition.

Definition 3.6. We say that the KZ functor separates simples if whenever S # T are simple objects
of O, then KZ(S) # KZ(T ).

Most of this chapter will be devoted to the proof of the following theorem.

Theorem 3.7. Suppose ! > 1 and KZ separates simples. Then either O is semisimple, or the
following hold:

37
1. There exists a linear character χ of W such that L(χ) is finite-dimensional and all the other
simple objects in O are infinite-dimensional.

2. There exists a positive integer r not divisible by !, such that dimL(χ) = r n .

3. Let s ∈ N be the residue of r modulo !, 1 ≤ s ≤ ! − 1. Then there is a representation hs of


W with dim hs = dim h such that if τ ∈
/ {∧i hs ⊗ χ| 0 ≤ i ≤ n} , then M (τ ) = L(τ ).

4. O = O∧ ⊕ Oss where O∧ is generated by the L(∧i hs ⊗ χ) and Oss is a semisimple category


generated by the other simple objects.

5. The composition multiplicities in O∧ are




1 if j = i, i + 1
[M (∧i hs ⊗ χ) : L(∧j hs ⊗ χ)] =

0 otherwise

Before proving Theorem 3.7, we make some remarks. Theorem 3.7 is an analogue for G(!, 1, n)
of Theorem 3.4. Although the methods we use for proving Theorem 3.7 are similar to those of [7],
we have to use different arguments to get round the problem that in the G(!, 1, n) case, it is not
known whether the functor KZ takes standard modules M (λ) in O to the corresponding Specht
modules S λ for H, ie. we do not know an analogue of [7, Lemma 3.2]. We also have to do some
work to calculate the blocks of the Hecke algebra at the parameters that we are interested in.
One reason why Theorem 3.7 is of interest is that it gives a source of examples of choices of
κ such that there is a finite-dimensional object in category O, and yet category O is completely
understood. We will later, in Theorem 3.23, give examples of choices of parameters such that KZ
separates simples.

Remark 3.8. It would be interesting to know whether, as in the ! = 1 and n = 1 cases, KZ


is guaranteed to separate simples whenever there is just one finite-dimensional simple object in
category O. We cannot prove this, but note that Theorem 3.7 is true in the n = 1 case, because in
this case either L(τ ) = M (τ ) or L(τ ) is finite-dimensional. So in the n = 1 case, if KZ(L(τ )) 1= 0
then L(τ ) = M (τ ). Hence if KZ separates simples then there can be at most one finite-dimensional
simple object in category O, and Theorem 3.7 reduces to Proposition 3.5. Therefore, in the proof
of Theorem 3.7, we may assume that n > 1.

38
3.3 Proof of Theorem 3.7

We begin with a lemma. As usual, write H for the Ariki-Koike algebra defined in Definition 1.34,
with parameters q = e2πiκ00 and ui = ε−(i−1) e−2πiκi−1 , 1 ≤ i ≤ !.

Lemma 3.9. Suppose that KZ separates simples. Then H has at least |Irrep(W )|−1 simple modules.

Proof. Suppose KZ separates simples. For any simple object S of O, KZ(S) is either 0 or simple
because KZ induces an equivalence O/Otor → H − mod, and if S is simple then either S|hreg = 0 or
S|hreg is simple. Furthermore, the simple objects KZ(S) are pairwise nonisomorphic. If KZ(S) = 0 for
some S then since KZ separates simples, we must have KZ(T ) 1= 0 for all simple T # S. Therefore,
H has at least |Irrep(W )| − 1 simple modules.

Note that when H has |Irrep(W )| simple modules, we must have L(τ )|hreg 1= 0 for all τ , and so
Otor = 0, and so O is semisimple by Lemma 2.3. From now on, we will assume that we are not in
the semisimple case. We prove a series of lemmas which will give (at least in principle) a description
of all the possible values of the parameters such that H has |Irrep(W )| − 1 simple modules. First,
we show that q 1= 1.

Lemma 3.10. Suppose that H has |Irrep(W )| − 1 simple modules. Then q 1= 1.

Proof. Suppose q = 1. Then by Lemma 1.35, since H is not semisimple, there must be some s < t
with us = ut . Under the assumption that n > 1 and ! > 1, there are at least three multipartitions
λ with λ(s) 1= ∅. Hence, by Lemma 1.35, there are at least three Dλ which are zero and so H
cannot have |Irrep(W )| − 1 simple modules. So q 1= 1.

Therefore, the simple H–modules are in bijection with Kleshchev multipartitions by Lemma
# #
1.36. Now, Ariki’s semisimplicity criterion (Theorem 1.38) tells us that [n]q ! i<j −n<c<n (ui −
q c uj ) = 0. Therefore, either there are i, j, c with ui = q c uj , or else [n]q ! = 0. We show [n]q ! 1= 0.

Lemma 3.11. Suppose H has |Irrep(W )| − 1 simple modules. Then [n]q ! 1= 0.

Proof. Suppose that [n]q ! = 0. Then there is a k, 1 ≤ k ≤ n with q k = 1 and q t 1= 1, 0 < t < k.
Since q 1= 1, the simple H–modules are in bijection with Kleshchev multipartitions. Let ρk be the
partition of k whose Young diagram is a row of k boxes. Then ρk is not Kleshchev, because the
only removable node of ρk , call it µ, cannot be good, because it is not normal. Indeed, the node

39
labelled λ in the diagram below is an addable node below µ with the same residue as µ, and there
are no removable nodes between them.
k boxes in row
2 34 5
µ
λ .

Hence, ρk is not Kleshchev and therefore ρn , a row of n boxes, is not Kleshchev. So we may define
multipartitions λ1 = (ρn , ∅, . . . , ∅) and λ2 = (∅, ρn , ∅, . . . , ∅), neither of which is Kleshchev (here
we use the hypothesis that ! > 1). This contradicts the fact that there is only one non-Kleshchev
multipartition, and so [n]q ! 1= 0.

Therefore, there exist integers 1 ≤ i, j ≤ ! and −n < c < n such that ui = q c uj . Writing what
this means in terms of the κi , we get

!(κj − κi ) − !cκ00 − (i − j) ∈ !Z. (3.1)

The next step is to show that |c| = n − 1.

Lemma 3.12. Suppose H has |Irrep(W )| − 1 simple modules. Then there exist 1 ≤ i 1= j ≤ ! such
that ui = q n−1 uj .

Proof. Redefining c if necessary, we have that there are i < j with q c ui = uj . Either c ≥ 0 or c ≤ 0.
Consider the case c ≥ 0. In this case, let ρc+1 be a row of c + 1 boxes, and take a multipartition
τ with ρc+1 as its ith part and ∅ everywhere else. If c < n − 1 then consider two multipartitions
defined as follows: λ is the multipartition of n whose ith part is ρn and µ is the multipartition of
n whose ith part is
n − 1 boxes in row
2 34 5

Then τ is not Kleshchev, and so λ is clearly not Kleshchev. Also, µ is not Kleshchev. (To see this,
consider the (1, c + 1)–node of µ(i) . This node cannot be a good node in any multiparition obtained
by deleting some nodes from µ. This is because there is an addable node below it with the same
residue, namely the unique node that can be added to the empty diagram µ(j) . The only removable
node between them is the (2, 1)–node of µ(i) . But this node cannot have residue q c because we
have already shown that q c+1 1= 1). Hence there are two non-Kleshchev multipartitions, which
contradicts our hypothesis. So c = n − 1.

40
In the c ≤ 0 case, we take γc+1 to be a column of −c + 1 boxes, and do a similar argument to
show that c = −(n − 1).

The above argument shows that the multiplicative order of q must be at least 2n − 1.

Lemma 3.13. Suppose H has |Irrep(W )| − 1 simple modules. Then q k 1= 1, 1 ≤ k ≤ 2n − 2.

Proof. Indeed, suppose q n+a = 1 where a is a nonnegative integer. Then if q n−1 ui = uj for some
i, j, we get q −a−1 ui = uj . But the above argument in the c ≤ 0 case shows that −a − 1 ≤ −n or
else we would have more than one non-Kleshchev multipartition.

3.3.1 Proof of parts (1) and (2)

Now we may rewrite our condition (3.1) on the parameters as

!(κj − κi ) + (−1)a !(n − 1)κ00 = (i − j) + !t

for some a ∈ {0, 1} and some t ∈ Z. Note that (i − j) + !t cannot be zero because 1 ≤ i, j ≤ ! and
i 1= j. If it is positive, multiply through by −1 (possibly interchanging the roles of i and j, and
changing a), so assume that (i − j) + !t < 0. Now we apply one of the twists from Section 1.3.7.
(t) (t)
Consider the linear character of W which sends σrs to (−1)a σrs for all r, s, t, and which sends sm
to ε−i sm . Explicitly checking with the set of generators and relations of W given in Equation 1.4
of Section 1.4.3 shows that this is a well-defined character of W . Now by Section 1.3.7, we have an
isomorphism of Cherednik algebras ψ : Hκ → Hκ! where κ$00 = (−1)a κ00 and κ$u = κu+i − κi . The
twist ψ induces an auotequivalence of category O which preserves the dimension of the objects.
Our new parameters satisfy

!κ$j−i + !(n − 1)κ$00 = (i − j) + !t < 0. (3.2)

Now we are in a position where we can use Theorem 3.3. There is a finite-dimensional quotient Y˜c
of Mκ! (triv). Therefore, Lκ! (triv) is finite-dimensional. By Section 1.3.7, twisting by ψ sends Lκ (χ)
to Lκ! (triv) for some linear character χ of W . Furthermore, dim Lκ (χ) = dim Lκ! (triv) = r n by
Theorem 3.3. Since Lκ (χ) is finite-dimensional, KZκ (Lκ (χ)) = 0 by Proposition 3.2, and therefore
KZκ (Lκ (τ )) 1= 0 for τ 1= χ, by our assumption that KZ separates simples. Therefore Lκ (τ ) is
infinite-dimensional if τ 1= χ. We have proved parts (1) and (2) of Theorem 3.7.

41
3.3.2 Blocks

To proceed further, it is necessary to calculate the blocks of the Hecke algebra.

Standing assumption 3.14. We have parameters q and u1 , . . . , u( for the Hecke algebra. We
are assuming that there is exactly 1 non-Kleshchev multipartition, and we have already shown that
q n−1 ui = uj for some i 1= j.

First, we prove the following lemma.

Lemma 3.15. Under assumption 3.14, if k 1= i, j then for each t 1= k, we have uk /ut 1= q c for any
−n < c < n.

Proof. Suppose uk = q c ut for some −n < c < n. If t 1= i, j then it would follow from the earlier
calculations that there is another non-Kleshchev multipartiton, so we need only consider the case
where t = i or t = j. Suppose i < j. If t = i then suppose there is −n < c < n with uk = q c ui , and
uj = q n−1 ui . Suppose k > j. If c < 0 then there are at least two non-Kleshchev multipartitions:
one is the multipartition whose only nontrivial part is a row ρn of n boxes in the ith position,
and the other is the multipartition whose only nontrivial part is a column γn of n boxes in the ith
position. Similarly, if k < j then there are at least two non-Kleshchev multipartitions. On the other
hand, if c ≥ 0 then uk = q c ui = q c−(n−1) uj and hence there exists a non-Kleshchev multipartition
which is ∅ except in the j th position, and one which is ∅ except in the ith position. Similarly, if
t = j, we reach the same conclusion, and so such a c cannot exist. Similar arguments deal with the
i > j case.

Recall from Theorem 1.39 that if α and β are multipartitions then the Specht modules S α and
S β belong to the same block if and only if cont(α) = cont(β). The next lemma is needed to study
the content of a multipartition.

Lemma 3.16. Under assumption 3.14, let α = (α(1) , α(2) , . . . , α(() ) be a multipartition of n. Then
cont(α(r) ) ∩ cont(α(s) ) = ∅ for all r 1= s.

Proof. By Lemma 3.15 and our assumption that q n−1 ui = uj , we get that for all r, s, ur /us 1= q c
for any −(n − 1) < c < n − 1. Now, if the residue of some node x in α(r) is equal to the residue of
some other node y in α(s) , then

ur q col(x)−row(x) = us q col(y)−row(y) .

42
But if t := col(x) + row(y) − row(x) − col(y) then us /ur = q t . However, t ! n − 2 and t ≥ −(n − 2),
a contradiction.

The next lemma is useful in determining a multipartition from its content.

Lemma 3.17. Under assumption 3.14, if α and β are multipartitions of n and 1 ≤ k ≤ !, then
cont(α(k) ) = cont(β (k) ) implies α(k) = β (k) .

Proof. We show that if two nodes of α(k) have the same residue, then they lie on the same diagonal.
It will follow that the multiplicity of a residue in cont(α) is equal to the length of the corresponding
diagonal of α. The same is true of β. Thus under the hypothesis, the Young diagrams α and β
have diagonals of the same lengths, so they are equal.
! !
Suppose then that nodes (i, j) and (i$ , j $ ) in α(k) have the same residue. Then uk q j−i = uk q j −i .
! !
Thus q j−i−j +i = 1 and therefore if j − i 1= j $ − i$ then either z := j − i − j $ + i$ ≥ n or z ≤ −n. But
2 ≤ j + i$ , j $ + i ≤ n + 1 and so z cannot be either ≥ n or ≤ −n. Therefore, z = 0 and j − i = j $ − i$ .
In other words, (i, j) and (i$ , j $ ) lie on the same diagonal.

We are finally in a position to calculate the blocks of the Hecke algebra. In order to determine
the blocks of H, we first note that if ρa denotes a row of length a and γb a column of length b,
then we may define a multipartition λa to have ρa in the ith place and γn−a in the j th place. For
example, if ! = 3, n = 3, i = 3, j = 2 then
$ % 6 7
λ0 = ∅ ∅ , λ1 = ∅ , λ2 = ( ∅ ) , λ3 = ( ∅ ∅ ).

Then if q n−1 ui = uj , then cont(λa ) = {ui q x |0 ≤ x ≤ n − 1} and hence all the λa belong to the
same block. It remains to show that if α, β are multipartitions and one of them is not of the form
λa , then they belong to distinct blocks.

Claim 1. Under assumption 3.14, if α and β are multipartitions of n and cont(α) = cont(β) and
α, β are not both of the form λa , then α = β.

Our aim is now to prove Claim 1, so we suppose that we have two multipartitions α =
(α(1) , . . . , α(() ) and β = (β (1) , . . . , β (() ) with cont(α) = cont(β). We will show that if k 1= i, j
then α(k) = β (k) .

Lemma 3.18. Let k 1= i, j. If x ∈ cont(α(k) ) then x ∈


/ ∪t+=k cont(β (t) ).

43
Proof. There is a unique integer b with −n + 1 ≤ b ≤ n − 1 such that x = uk q b . We consider the
cases b ≥ 0 and b ≤ 0 separately. In the case b ≥ 0, we now prove by induction that x ∈
/ cont(β (t) )
for any t 1= k. The proof for b ≤ 0 is very similar, so we omit it.
For the base step, suppose b = 0. Then uk ∈ cont(α(k) ). Hence uk is a residue of β. If
uk ∈ cont(β (t) ) where t 1= k then uk = ut q c−r for some column c and row r of β (t) . But clearly
−n < c − r < n which contradicts Lemma 3.15. Therefore x = uk ∈
/ ∪t+=k cont(β (t) ) and so
uk ∈ cont(β (k) ).
Now we do the inductive step. Suppose b > 0. Suppose uk q b is a residue of β (t) with t 1= k. Then
uk q b = ut q c−r for some c, r. So uk /ut = q c−r−b . Since c − r < n and b > 0, we have c − r − b < n.
So by Lemma 3.15, c − r − b ≤ −n. Therefore, r ≥ n + c − b ≥ n + 1 − b. But β (t) contains at least
r boxes, by definition of r. So |β (t) | ≥ n + 1 − b.
Next, we note that since uk q b is the residue of a node in α(k) , this node must lie on the diagonal
containing (1, b + 1). So there are at least b + 1 boxes in the first row of α(k) and hence there is a
node in the first row of α(k) with residue uk q b−1 . By induction on b, this is also a residue of β (k) . So
there is a box in column b and row 1 of β (k) . Therefore, |β (k) | ≥ b. So |β| ≥ |β (k) | + |β (t) | ≥ n + 1,
a contradiction.

It follows from Lemma 3.18 that if cont(α) = cont(β) then cont(α(k) ) = cont(β (k) ) for all
k 1= i, j. Then applying Lemma 3.17, we get α(k) = β (k) . It remains to deal with α(i) and α(j) . The
proof of this case will be very similar to Lemma 3.18, but slightly more complicated.
Given multipartitions α = (α(1) , . . . , α(() ) and β = (β (1) , . . . , β (() ), with cont(α) = cont(β), let
a1 be the length of the first row of α(i) and a2 be the length of the first column of α(j) and define
b1 , b2 similarly for β. First we prove a technical lemma.

Lemma 3.19. Under assumption 3.14, suppose a1 + a2 < n. Then ui q a1 ∈


/ cont(α).

/ cont(α(k) ) when k 1= i, j. So let k 1= i, j and suppose there


Proof. First, we show that ui q a1 ∈
is a node of α(k) with residue ui q a1 . Say this node lies in column c and row r of α(k) . Then
ui q a1 = uk q c−r . So ui /uk = q a1 −(c−r) . We show that a1 − (c − r) lies between −n and n. If
a1 − (c − r) ≥ n then c + n ≤ r + a1 ≤ n, a contradiction. While if a1 − (c − r) ≤ −n then
c ≥ n + a1 + r ≥ n + 1, a contradiction. So −n < a1 − (c − r) < n, which violates Lemma 3.15.
Hence, ui q a1 is not a residue of α(k) .

44
Next, we show that ui q a1 is not a residue of α(i) . If it is, then there is a node in column c and
row r of α(i) whose residue is ui q a1 = ui q c−r . So q a1 −(c−r) = 1. So by Lemma 3.13, if a1 −(c−r) 1= 0
then either a1 − (c − r) ≥ 2n − 1 or a1 − (c − r) ≤ −(2n − 1). If a1 − (c − r) ≤ −(2n − 1) then
2n ≤ a1 +r −1+2n ≤ c, which is impossible. If a1 −(c−r) ≥ 2n−1 then c+2n ≤ a1 +r +1 ≤ n+2,
which is impossible if n > 1. Therefore, a1 = c − r. But c ≤ a1 and r ≥ 1, so this is also impossible.
Therefore, ui q a1 cannot be a residue of α(i) .
The argument that ui q a1 is not a residue of α(j) is very similar. We use the fact that a1 <
n − a2 .

The Claim 1 follows from the next lemma. We use the same notation as Section 3.19.

Lemma 3.20. Under assumption 3.14, if a1 + a2 < n then if x ∈ cont(α(i) ) then x ∈


/ cont(β (j) ).

Proof. By Lemma 3.18, cont(α(k) ) = cont(β (k) ) for k 1= i, j. Therefore, by Lemma 3.16, we get
cont(α(i) ) ∪ cont(α(j) ) = cont(β (i) ) ∪ cont(β (j) ). This is a disjoint union.
If x ∈ cont(α(i) ) then x = ui q b for some b with −n + 1 ≤ b ≤ n − 1. As in the proof of Lemma
3.18, we consider the cases b ≥ 0 and b ≤ 0 separately. We give the proof only for the b ≥ 0 case.
The proof is by induction on b.
For the base step, if b = 0 then ui is a residue of α(i) . If this is a residue of β (j) , then it has the
form ui = ui q n−1 q c−r for some c, r. So q n−1+c−r = 1. Now, n−1+c−r ≥ 0. If n−1+c−r ≥ 2n−1
then c − r ≥ n which is impossible. So n − 1 + c − r = 0. Hence, c = 1, r = n, and β (j) must be a
column of n boxes. But then cont(β (j) ) = {ui q n−1 , ui q n−2 , . . . , ui q, ui }. Since 0 ≤ a1 < n, we have
ui q a1 ∈ cont(β (j) ) = cont(β) = cont(α), which contradicts Lemma 3.19. Therefore ui must be a
residue of β (i) , which proves the base step.
For the inductive step, suppose b > 0 and ui q b is a residue of α(i) . If ui q b is a residue of a node
in column c and row r of β (j) , then ui q b = ui q n−1 q c−r . So q c−r+n−1−b = 1. Since c − r < n and
b > 0, we have c − r − b < n. So c − r − b + n − 1 < 2n − 1. Therefore, either c − r − b + n − 1 = 0
or c − r − b + n − 1 ≤ −(2n − 1). If the latter holds then c + 3n ≤ r + b + 2 ≤ 2n + 1 since we may
take b ≤ n − 1. Hence 1 + n ≤ c + n ≤ 1, a contradiction. We therefore get c − r − b + n − 1 = 0. So
r ≥ n − b. But β (j) has at least r nodes. Therefore, |β (j) | ≥ n − b and has at least n − b rows. But
since ui q b ∈ cont(α(i) ), we get ui q b−1 ∈ cont(α(i) ), as in the proof of Lemma 3.18. By induction on
b, ui q b−1 ∈ cont(β (i) ). So, as in the proof of Lemma 3.18, there is a box in row 1 and column b of
β (i) . Therefore, |β (i) | ≥ b and β (i) has at least b columns. So β = λb in the notation of Section 1.

45
Therefore cont(β) = {ui , qui , . . . , q n−1 ui }. So ui q a1 ∈ cont(β) = cont(α). This contradicts Lemma
3.19. Therefore, ui q b must be a residue of β (i) and this proves the inductive step.

Now we prove Claim 1. Suppose we have a multipartition α not of the form λa . Suppose β 1= α.
We show that cont(α) 1= cont(β). Indeed, if β 1= λb for any b, then by Lemmas 3.18 and 3.20,
cont(α(k) ) = cont(β (k) ) for all k. Therefore, by Lemma 3.17, α(k) = β (k) for all k, so α = β, a
contradiction. On the other hand, if β = λb for some b, then ui q a1 ∈ cont(β) \ cont(α) by Lemma
3.19. So cont(α) 1= cont(β).
Therefore, S α is the unique Specht module in its block. Furthermore, {S λa |0 ≤ a ≤ n} form a
block, by the same reasoning.

3.3.3 Proof of parts (3) and (4)

By Theorem 1.39, we get that there is one block of the Hecke algebra containing n + 1 of the
Specht modules, and all the other blocks are singletons. Hence, there are |Irrep(W )| − n blocks.
By [33, Corollary 5.18], the blocks of O are in bijection with blocks of H and hence O also has
|Irrep(W )| − n blocks. We work in the category O(Hκ! ). Now by [16, Theorem 2.3], there is a
BGG-resolution of Y˜c , ie. an exact sequence

0 ← Y˜c ← M (triv) ← M (hs ) ← · · · ← M (∧n hs ) ← 0 (3.3)

where hs is a certain n–dimensional irreducible representation of W . By Proposition 1.24 parts (5)


and (6), there is an ordering on Irrep(W ) such that the matrix whose entries are the composition
multiplicities [M (τ ) : L(σ)] is upper triangular with ones on the diagonal. Hence, it has an
inverse with integer entries, and it follows from the fact that the classes [L(µ)] form a basis of
the Grothendieck group K0 (O), that the classes [M (τ )] form a basis of the Grothendieck group
K0 (O) as well. Therefore, none of the maps in this sequence (3.3) can be zero, or else there would
be a nontrivial linear relation amongst the [M (∧i hs )], contradicting the fact that {[M (µ)] : µ ∈
Irrep(W )} is a basis of K0 (O). By Proposition 1.7, for any τ , all the composition factors of M (τ )
belong to the same block, and hence all the L(∧i hs ) belong to the same block. There are n + 1
simples in this block and hence by counting we see that all the other blocks must be singletons.
Using the fact that simple objects in O have no self-extensions (Theorem 1.33), we get that these
blocks are semisimple. Translating back to category O(Hκ ), we get parts (3) and (4) of Theorem
3.7.

46
In order to prove part (5) of Theorem 3.7, we require the following easy lemma.

Lemma 3.21. The module Ỹc is isomorphic to L(triv).

Proof. Since Ỹc is finite-dimensional, its only composition factor can be L(triv), by part (1) of
Theorem 3.7. But M (triv) # Ỹc and [M (triv) : L(triv)] = 1.

3.3.4 Proof of part (5)

It remains to compute the composition multiplicities in the one nontrivial block O∧ . Again we work
in the category O(Hκ! ). [33, Proposition 5.21(ii)] tells us that each L(∧i hs ), i > 0 is a submodule
of a standard module. Write Li = L(∧i hs ) and Mi = M (∧i hs ). Let Ri be the radical of Mi . We
cannot have a nonzero map Li → Mj if j > i for the following reason.

Lemma 3.22. If j > i then [Mj : Li ] = 0.

Proof. The argument is based on [34, Lemma 4.2].


Recall Proposition 1.24, part (5), which states that [Mj : Li ] 1= 0 only if c∧j hs − c∧i hs ∈ N,
! !nH −1
where cτ denotes the scalar by which the element z := H∈A i=1 nH kH,i eH,i ∈ CW acts on
the irreducible representation τ of W . We calculate c∧i hs , 0 ≤ i ≤ n. In our situation,
(−1
&& & (−1 &
n & (−1
(k)
z= κ$00 (1 − σab ) + κ$r εrt stu .
a<b k=0 u=1 r=1 t=0

By [16, Section 4.1], hs = h as a vector space, but each su acts on hs by multiplying the uth
(0)
standard basis vector by ε−s , while σab acts in the same way as on h. Since z acts by a scalar,
6 7
it acts by tr(z)/ dim(∧i hs ) = tr(z)/ ni . Calculating with respect to the natural choice of basis of
6 7 6n−17 (k) (k)
∧i hs , the trace of stu is ε−st n−1
i−1 + i . Also, since σab acts by a real reflection, the trace of σab
6 7 6n−17
is n−1 i − i−1 . Therefore, c∧i hs = i(!(n − 1)κ$00 + !κ$s ) = i(−s + !t) by Equation 3.2 of Section
3.3.1. Since −s + !t < 0, we get that if j > i then c∧j hs − c∧i hs ∈
/ N.

So L1 is a submodule either of M0 or M1 . It can’t be a submodule of M1 because [M1 : L1 ] = 1,


so L1 3→ M0 . So L1 3→ R0 . But by Lemma 3.21, we have Ỹc ∼
= L(triv). Hence Ỹc is simple and
so R0 = ker(M0 → Ỹc ) = im(M1 → M0 ) is a quotient of M1 . Hence [R0 : L1 ] = 1. If we had
[R0 : Li ] 1= 0 for some i > 1 then R0 would have Li as a quotient for some i > 1. Therefore, so would
M1 . But M1 has a unique simple quotient L1 . Therefore, it is impossible to have [R0 : Li ] 1= 0 for
i > 1 and we conclude that R0 = L1 .

47
We have shown that the composition factors of M0 are L0 and L1 . To conclude the argument,
we show by induction that the composition factors of Mi are Li and Li+1 . Consider first Li+1 .
Then, by [33, Proposition 5.21(ii)], Li+1 is a submodule of some Mj . We cannot have j ≥ i + 1, and
by induction, we cannot have j < i. Hence, Li+1 is a submodule of Mi and so Li+1 3→ Ri . Now
Ri = ker(Mi → Mi−1 ) by induction and so Ri is a quotient of Mi+1 . Therefore, [Ri : Li+1 ] = 1. If
there was a j > i + 1 with [Ri : Lj ] 1= 0 then we would have that for some j > i + 1, Lj would be a
quotient of Ri and hence a quotient of Mi+1 , contradicting the fact that Mi+1 has a unique simple
quotient. Therefore, Ri = Li+1 and we are done. This proves part (5) of Theorem 3.7.

3.3.5 Characterisations of separating simples

Now that we have completed the proof of Theorem 3.7, let us turn our attention to the question of
when KZ separates simples, and whether it is possible to choose κ such that KZ separates simples.

Theorem 3.23. The following are equivalent

1. KZ separates simples.
#
2. If q, u1 , . . . , u( are the parameters of the Ariki-Koike algebra H, then (q+1) i<j (ui −uj ) 1= 0,
and furthermore,

#{τ ∈ Irrep(W ) : L(τ )|hreg 1= 0} ≥ |Irrep(W )| − 1.

3. The algebra H has at least |Irrep(W )| − 1 nonisomorphic simple modules.

Proof. First, we show that (2) implies (1). We must show that if L(σ)|hreg ∼
= L(τ )|hreg 1= 0 then
σ = τ . Suppose then that L(σ)|hreg ∼
= L(τ )|hreg 1= 0. By [33, Proposition 5.21(ii)], there exists a
standard module M (λ) such that L(σ) 3→ M (λ). Let t = dimHom(L(σ), M (λ)). Then M (λ) must
have t submodules isomorphic to L(σ), because the only automorphisms of L(σ) are the scalars.
Therefore, L(σ)⊕t ⊂ M (λ) and M (λ) has no submodule isomorphic to L(σ)⊕(t+1) . Now since
L(σ)|hreg ∼
= L(τ )|hreg , we have Hom(L(τ )|hreg , M (λ)|hreg ) = Hom(L(σ)|hreg , M (λ)|hreg ) 1= 0 and hence
by Corollary 1.30, Hom(L(τ ), M (λ)) 1= 0 (using the condition on the parameters). Therefore, M (λ)
has a submodule isomorphic to L(τ ) and hence a submodule isomorphic to L(τ ) + L(σ)⊕t . This
sum must be direct if L(σ) # L(τ ), hence M (λ) has a submodule L(τ ) ⊕ L(σ)⊕t and M (λ)|hreg has

48
⊕(t+1)
a submodule L(τ )|hreg ⊕ L(σ)|⊕t
hreg = L(σ)|hreg . Therefore, dim(Hom(L(σ)|hreg , M (λ)|hreg )) ≥ t + 1
and so dim(Hom(L(σ), M (λ))) ≥ t + 1, a contradiction. So L(σ) ∼
= L(τ ) and hence σ = τ .
Next, (1) implies (3) by Lemma 3.9.
Finally, to show (3) implies (2), note that under the hypothesis that H has |Irrep(W )| − 1
simple modules, it has already been shown in Lemma 3.11 that [n]q ! 1= 0, hence q 1= −1 since we
assume n ≥ 2, and in Lemma 3.15 that ui 1= uj for all i 1= j, so the condition on the parameters
holds. Furthermore, by the essential surjectivity of KZ, if H has |Irrep(W )| − 1 simple modules then,
because KZ is essentially surjective on objects and exact, there are at least |Irrep(W )| − 1 of the
L(τ ) with KZ(L(τ )) 1= 0 and hence with L(τ )|hreg 1= 0.

Having characterised when it is possible for KZ to separate simples, let us turn our attention to
constructing examples of choices of the parameters κ so that KZ separates simples.

/ Q, and suppose there is some 1 ≤ b ≤ ! − 1 with


Theorem 3.24. Let κ00 ∈

!κb + !(n − 1)κ00 = −b + !t < 0

/ q Z uj . Then KZ
for some t ∈ Z. Suppose further that if i 1= j and {i, j} 1= {1, b + 1} then ui ∈
separates simples.

Proof. Recall that u1 = 1 and uj = ε−(j−1) e−2πiκj−1 for j ≥ 2. Then the condition on the
parameters says that ub+1 = q n−1 . It follows that the multipartition

λ=(4 52· · · 3 , ∅, ∅, . . . , ∅ )
n boxes

is not Kleshchev, because the box at the right hand end of the row λ(1) has residue q n−1 and is
not a normal node. Next, we show that all the other Dλ are nonzero. By [3, 1.2], Dλ will be
(1) ,λ(b+1) )
nonzero if and only if D(λ is a nonzero module for the Ariki-Koike algebra of Z2 ! Sn , with
parameters q and u1 = 1, u2 = q n−1 . Since q 1= 1, this is the case if and only if the multipartition
(λ(1) , λ(b+1) ) is a Kleshchev multipartition
8 of the 9
integer |λ(1) | + |λ(b+1) | ≤ n. So it suffices to show

that every bipartition (λ, µ) 1= 4 52· · · 3 , ∅ with |λ| + |µ| = t ≤ n is Kleshchev. This is a
n boxes
straightforward induction argument: if µ 1= ∅ then the rightmost node of the bottom row of µ is a
good node and so may be removed. This reduces us to the case where µ = ∅. But then the same
procedure may be applied to λ, proving that (λ, µ) is Kleshchev.

49
We have shown that H has |Irrep(W )|−1 nonisomorphic irreducible modules, and so KZ separates
simples.

Note that in the situation of Theorem 3.24, it follows from Theorem 3.3 that the module L(triv)
is finite-dimensional, so L(triv) must be the unique finite-dimensional simple module in O. If χ is an
arbitrary linear character of W , then we may twist the parameters by χ to obtain new parameters
κ$ . By Section 1.3.7, twisting by χ induces an isomorphism of Ariki-Koike algebras and so by
Theorem 3.23, KZκ! separates simples. This shows the following.

Corollary 3.25. For every linear character χ of W and every r > 0 with ! $ r, there is a choice of
parameters κ such that KZκ separates simples and the unique finite-dimensional simple module in
category O is L(χ) of dimension r n .

3.4 The Ariki-Koike algebra in the almost-semisimple case

We close this chapter by using the facts proved about category O in Theorem 3.7 to prove a
theorem about the Hecke algebra which does not mention the Cherednik algebra in its hypothesis
or conclusion. This theorem is an example of a general philosophy suggested by Rouquier in [64] of
using the Cherednik algebra and the KZ functor as a tool to prove theorems about Hecke algebras.
We know that Hκ is semisimple if and only if the number of irreducible modules |Irrep(Hκ )|
of Hκ equals the number of irreducible modules of CW , and that in this case Hκ ∼
= CW . So
the property of having |Irrep(W )| simple modules determines the algebra Hκ up to isomorphism.
We show that the property of having |Irrep(W )| − 1 simple modules also determines Hκ up to
isomorphism.

Theorem 3.26. Suppose Hκ and Hµ are Ariki-Koike algebras corresponding to some parameters
κ, µ ∈ C( and that |Irrep(Hκ )| = |Irrep(Hµ )| = |Irrep(W )| − 1. Then there is an isomorphism of
algebras Hκ ∼
= Hµ .

Proof. By Theorem 1.31, there is an algebra isomorphism Hκ ∼


= EndO (PKZ )opp where
"
PKZ = dim KZ(L(τ ))P (τ ).
τ ∈Irrep(W )

Here, P (τ ) is the projective cover of L(τ ). The strategy of the proof is to calculate PKZ in the case
where KZκ separates simples, and show that its endomorphism ring can be written in a way that

50
does not depend on κ. We work in the category O = Oκ and write KZ = KZκ , M (τ ) = Mκ (τ ),
and so forth. By Theorem 3.7, there is a linear representation χ of W with O = O∧ ⊕ Oss , where
O∧ is the subcategory of O generated by {L(∧i hs ⊗ χ) : 0 ≤ i ≤ n}. Let λi = ∧i hs ⊗ χ and let
S = {λi : 0 ≤ i ≤ n}. Write Mi = M (λi ), Li = L(λi ) and Pi = P (λi ) (the projective cover of Li ).
For σ, τ ∈ Irrep(W ), we have in general

dim Hom(P (σ), P (τ )) = [P (τ ) : L(σ)]


&
= [P (τ ) : M (γ)][M (γ) : L(σ)]
γ
&
= [M (γ) : L(τ )][M (γ) : L(σ)]
γ
& &
= [M (γ) : L(τ )][M (γ) : L(σ)] + [M (γ) : L(τ )][M (γ) : L(σ)].
γ∈S γ ∈S
/

If γ ∈
/ S then M (γ) = L(γ), so we get
n
& &
dim Hom(P (σ), P (τ )) = [Mi : L(τ )][Mi : L(σ)] + δγτ δγσ .
i=0 γ ∈S
/

Now, if σ ∈ / S, this sum must be δστ . Otherwise, σ, τ ∈ S and so σ = λa , τ = λb for some


/ S or τ ∈
a, b. We get
n
&
dim Hom(P (λa ), P (λb )) = [Mi : La ][Mi : Lb ]
i=0

which equals 2 if a = b and 1 if |a − b| = 1 and 0 otherwise. So we get





2 if σ = τ ∈ S






1 if σ = τ ∈ /S
dim Hom(P (σ), P (τ )) =



 1 if {σ, τ } = {λa , λa+1 }, 0 ≤ a ≤ n − 1





0 otherwise

The ring EndO (PKZ ) is a matrix algebra with entries in the various Hom-spaces Hom(P (σ), P (τ )).
We calculate the multiplication relations between basis elements of the Hom(P (σ), P (τ )) and show
that these relations do not depend on κ. It will follow that the structure constants of EndO (PKZ ) do
not depend on κ, which will prove the theorem provided that the multiplicity of each P (τ ) in PKZ
$ 6n−17 %
is also independent of κ. But in our situation PKZ = ⊕τ ∈S
/ (dim τ ) · P (τ ) ⊕ ⊕ 1≤i≤n i−1 Pi since

51
6n−17
dimC KZ(Li ) = i−1 as a vector space1 . By BGG reciprocity, we have [Pi : Mi ] = [Mi : Li ] = 1 =
[Mi−1 : Li ] = [Pi : Mi−1 ], and [Pi : M (σ)] = [M (σ) : Li ] = 0 if σ 1= λi , λi−1 . Therefore, the factors
in any filtration of Pi by standard modules are Mi and Mi−1 . But by [33, Corollary 2.10], Pi has
Mi
a filtration by standard modules with Mi as the top factor, so Pi may be described as Pi = Mi−1 ,

meaning that there is a series 0 = Pi0 ⊂ Pi1 ⊂ Pi2 = Pi with Pi1 ∼


= Mi−1 and Pi2 /Pi1 ∼
= Mi . We may
write the resulting composition series of Pi as

Li
Li+1
Pi =
Li−1
Li

This description of Pi makes it easy to write down the nontrivial maps Pi → Pi .


First, there are two obvious maps Pi → Pi , namely the identity map idi and the map ξi which
is projection onto the bottom composition factor Li followed by inclusion. Note that ξi2 = 0 and
therefore EndO (Pi ) = C[ξi ]/(ξi2 ), since we have already shown that dim Hom(Pi , Pi ) = 2.
Mi Mi+1
Next, we describe the map Pi → Pi+1 . This is a map Mi−1 → Mi . So we may construct a
map fi,i+1 : Pi → Pi+1 by factoring out the copy of Mi−1 and then embedding Mi in Pi+1 . This
map is nonzero, so Hom(Pi , Pi+1 ) = Cfi,i+1 , 1 ≤ i ≤ n − 1.
Now we describe the map Pi → Pi−1 , n ≥ i ≥ 2. By [33, Proposition 5.2.1 (ii)], Pi ⊃ Li is
injective and therefore Pi contains the injective envelope Ii = I(λi ) of Li . Therefore, since Pi is
indecomposable, Pi = Ii . Now, recall that category O contains a costandard module ∇(τ ) ⊃ L(τ )
for every τ ∈ Irrep(W ), with [∇(τ )] = [M (τ )] in K0 (O). Write ∇i = ∇(λi ). Then Li ⊂ ∇i
Li+1
by Proposition 1.24, so ∇i has a composition series of the form ∇i = Li . Furthermore, by
∇i−1
Proposition 1.24, ∇i ⊂ Ii and Ii has a filtration by costandard modules of the form Ii = ∇i
∇i−1 ∇i−2
Since Ii = Pi , to get a map ∇i = Pi → Pi−1 = ∇i−1 , we may factor out the copy of ∇i and then
embed ∇i−1 in Pi−1 . This gives a nonzero map fi,i−1, and therefore Hom(Pi , Pi−1 ) = Cfi,i−1. In
particular, this shows that the image of fi,i−1 has length 2.
1
this can be readily shown using the following argument: since for any τ ∈ Irrep(W ), M (τ ) ∼
= C[h] ⊗ τ as C[h]–
modules, we get that the vector bundle associated to M (τ ) on hreg /W has rank dim(τ ). Therefore, dim KZ(M (τ )) =
dim(τ ) for all τ . Now recall from Theorem 3.7 that the composition factors of Mi are Li and Li+1 . Write di =
! ! ! 6 7
dim KZ(Li ). Then di = n j=i (−1)
j−i
(di + di+1 ) where dn+1 := 0. So di = nj=i (−1)
j−i
dim Mi = nj=i (−1)
j−i n
i
=
6n−17
i−1
.

52
Now we calculate multiplication relations between the various fi,i+1 , fi,i−1 and ξi . First, it is
immediate from the definitions that ξi+1 fi,i+1 = fi,i+1 ξi = 0. We need to do a little more work
∇i−1
to show that the same holds for fi,i−1. Take the description of Ii as Ii = ∇i . Then Ii has a
composition series
Li
Li−1
Ii =
Li+1
Li
So there is a map ζi : Ii → Ii defined by projection onto the bottom composition factor Li followed
by the embedding Li 3→ Ii . Clearly, ζi fi−1,i = fi−1,i ζi−1 = 0. But since Pi = Ii , we may regard ζi
as a map Pi → Pi . Therefore, there are a, b ∈ C with ζi = aidi + bξi . Since ζi2 = 0, we get a2 = 0
and hence ζi is a nonzero multiple of ξi . This shows that ξi fi−1,i = fi−1,i ξi−1 = 0.
Finally, we need to calculate fi+1,i fi,i+1 and fi−1,i fi,i−1. Consider first fi−1,i fi,i−1. By the
definition of fi,i−1 above, we have [im(fi,i−1 ) : Li ] 1= 0. Hence, im(fi,i−1 ) cannot be contained
in the submodule of Pi−1 isomorphic to Mi−2 , and therefore fi−1,i fi,i−1 must be nonzero. Since
fi−1,i fi,i−1 ξi = 0, fi−1,i fi,i−1 must be a nonzero multiple of ξi . Let us replace ξi by fi−1,i fi,i−1. So
we may assume that fi−1,i fi,i−1 = ξi , and this does not change any of the relations which have
already been calculated. Now consider fi+1,i fi,i+1 . We show that this composition is nonzero.
Indeed, the image im(fi,i+1 ) has composition factors Li and Li+1 . If fi+1,i fi,i+1 were zero, then we
would get that im(fi+1,i ) could only have composition factors Li+1 and Li+2 . But we have shown
that im(fi+1,i ) has length 2, and [Pi : Li+2 ] = 0, a contradiction. Therefore, fi+1,i fi,i+1 1= 0 and so
there is a nonzero bi,i+1 ∈ C, n − 1 ≥ i ≥ 1, such that

fi+1,i fi,i+1 = bi,i+1 ξi = bi,i+1 fi−1,i fi,i−1 .

It remains to do some rescaling. Let

1
ξi$ = ξi , 1≤i≤n
b12 b23 · · · bi−1,i
$
fi,i−1 = fi,i−1 2≤i≤n
$ 1
fi,i+1 = fi,i+1 1 ≤ i ≤ n − 1.
b12 b23 · · · bi,i+1

53
Then we have the following relations:

ξi$ fi−1,i
$ $
= fi−1,i $
ξi−1 =0
$ $ $
ξi+1 fi,i+1 = fi,i+1 ξi$ = 0
$ $ $ $
fi−1,i fi,i−1 = fi+1,i fi,i+1 = ξi$ . (3.4)

These are the only nontrivial relations between the various Hom(P (σ), P (τ )). This shows that we
may choose a basis of Hom(P (σ), P (τ )) for each σ, τ such that the composition relations between
the basis elements are independent of κ. Hence, we may choose a basis of the algebra EndO (PKZ )
such that the structure constants are independent of κ. This proves the theorem.

Remark 3.27. By variations on the arguments given in the above proof, it is possible to show that


1 j = i + 1, i − 1
dimC ExtO (Li , Lj ) =
1

0 otherwise

Li
and so the composition series of Pi may be written more symmetrically as Pi = Ii = Li−1 ⊕Li+1 .
Li

Note that since Theorem 3.7 implies that the Ariki-Koike algebra has |Irrep(W )| − n blocks, by
6 7
counting we get that the algebra Bn := EndO (⊕ni=1 n−1
i−1 Pi ) is a block of the Ariki-Koike algebra.

From the relations (3.4), it is clear that Bn is independent both of κ and !. So we have the following
corollary.

Corollary 3.28. Let !1 , !2 > 1 and for i = 1, 2 let κi ∈ C(i and suppose Hκi (G(!i , 1, n)) has
|Irrep(G(!i , 1, n))| − 1 simple modules. Then the unique nonsemisimple blocks of Hκ1 (G(!1 , 1, n))
and Hκ2 (G(!2 , 1, n)) are isomorphic algebras.

Remark 3.29. The representation theory of the algebra Bn is described in [7, 5.3] and [25, 3.2].

54
Chapter 4

Shift functors

In this chapter, we take W = G(!, 1, n). For κ ∈ C( , we write as usual Hκ for the Cherednik
algebra, Hκ for the Ariki-Koike algebra, Oκ for the category O(Hκ ), and KZκ for the KZ functor.
The aim of this chapter is to study certain relationships between category Oκ and Oκ! for
κ 1= κ$ . Recall that the parameters for the associated Ariki-Koike algebra are the exponentials
of the κi . Therefore, if κ$i = κi + ai for all i = 00, 1, 2, . . . , ! − 1 for some ai ∈ Z, then the
Ariki-Koike algebras Hκ and Hκ! are equal. It turns out that in some cases this isomorphism of
Ariki-Koike algebras extends to an equivalence Oκ ∼
= Oκ! . In this chapter, we consider two different
notions of shift functor. Both of these are really defined as functors eHκ e − Mod → eHκ! e − Mod,
but we will show in Section 4.1.5 that in some circumstances these can be extended to functors
Hκ − Mod → Hκ! − Mod and then give functors Oκ → Oκ! . First, in Section 4.1, we consider the
Heckman-Opdam shift functors. These functors appear to have been first defined in [7, Lemma
4.7]. Our aim is to give some conditions on the parameters which guarantee that these functors
are equivalences. Second, in Section 4.2, we consider a different notion of shift functor, which we
call the Boyarchenko-Gordon shift functor. We are not able to construct these functors in all cases,
but we will show that they exist provided a hypothesis (Hypothesis 4.48) holds. Having defined
the Boyarchenko-Gordon shift functors, we address in Section 4.3 the question of whether the two
notions of shift functor coincide.

55
4.1 The Heckman-Opdam shift functors

In this section we will define and study the Heckman-Opdam shift functors for G(!, 1, n), largely
following [36]. For each value of κ = (κ00 , κ1 , . . . , κ(−1 ) ∈ C( , we will define functors

Fκa : eHκ[a] e − Mod → eHκ e − Mod

where κ[a] is obtained from κ by incrementing the values of some of the parameters by integers. We
are interested in when these functors are equivalences, since in such cases, they give a powerful tool
for studying category O. We will show that there are two cases of interest in which the functors
Fκa are equivalences. The first case is when the KZ functor separates simples. The second case is
when the parameters are “asymptotic”, in a sense to be defined below.

4.1.1 A shift relation

The theorem which allows the Heckman-Opdam shift functors to be constructed is the following
so-called shift relation.

Proposition 4.1. Let κ = (κ00 , κ1 , . . . , κ(−1 ) ∈ C( and define κ[a] ∈ C( by κ[a]00 = κ00 + 1,
κ[a]i = κi + 1 for 1 ≤ i ≤ a, and κ[a]i = κi for a + 1 ≤ i ≤ ! − 1. Let θκ be the Dunkl representation
of Hκ and θκ[a] be the Dunkl representation of Hκ[a] . Then there is an equality of subsets of
D(hreg ) ∗ W
eθκ[a] (Hκ[a] )e = eµ−1
a θκ (Hκ )µa e,
# #
where µa = ( ni=1 xi )a i<j (x(i − x(j ).

We now give a proof of Proposition 4.1. Since Proposition 4.1 is known to be true in the case
! = 1 (see [7, Proposition 4.6]), we will assume in this proof that ! > 1. We remark that a much more
general shift relation, which holds for all complex reflection groups, appears in unpublished work
of Berest and Chalykh [5]. Our proof of Proposition 4.1 will follow the argument of [7, Proposition
4.6] in the ! = 1 case. The strategy is to prove Proposition 4.1 first in the case where Hκ is simple,
and then to extend to all values of κ using a specialisation argument.

Notation 4.2. Say κ ∈ C( is regular if Hκ is simple.

In the proof of Proposition 4.1, we fix a and write e− := µa eµ−1


a (we have e ∈ CW because

µa is a W –semiinvariant). So we wish to show that θκ[a] (eHκ[a] e) = µ−1


a θκ (e Hκ e )µa .
− −

56
In the regular case, we will first show that eHκ e is generated as an algebra by the subset
C[h]W e ∪ C[h∗ ]W e. The proof will require the notion of a Poisson bracket, which we now explain.

Definition 4.3. Let R be a commutative algebra. A Poisson bracket on R is a bilinear map


{−, −} : R × R → R which is a Lie bracket and which satisfies the identity

{xy, z} = x{y, z} + y{x, z}

for all x, y, z ∈ R. We refer to (R, {−, −}) as a Poisson algebra.

i=0 F R be a
A standard example of a Poisson algebra can be obtained as follows. Let R = ∪∞ i

filtered algebra and suppose that grF R is commutative. For x ∈ F i R/F i−1 R and y ∈ F j R/F j−1 R,
let x̂ ∈ F i R, ŷ ∈ F j R be lifts of x and y respectively. Then define

{x, y} := [x̂, ŷ] + F i+j−2 R.

Then it is routine to check that {−, −} defines a Poisson bracket on R.


A concrete example of a Poisson algebra is the following. Let (V, ω) be a symplectic vector
space and let W be a finite subgroup of Sp(V ). For x, y ∈ V , set {x, y} = ω(x, y). Then {−, −}
may be extended to C[V ] in a natural way, and this gives a Poisson bracket. The only hard part
to check is the Jacobi identity, but it follows from the axioms for a Poisson algebra that one only
needs to check the Jacobi identity on a set of algebra generators of C[V ]. This bracket also induces
a bracket on C[V ]W , which we denote by {−, −}ω .
We require the notion of the degree of a Poisson bracket.

i=0 Ri be a graded commutative algebra with a Poisson bracket {−, −}.


Definition 4.4. Let R = ⊕∞
We say that {−, −} has degree k if for all i and j and for all x ∈ Ri and all y ∈ Rj , we have
{x, y} ∈ Ri+j+k , and furthermore there exist i, j with x ∈ Ri , y ∈ Rj and {x, y} ∈
/ Ri+j+k−1.

For more information on Poisson brackets, we refer the reader to the book [71].

Notation 4.5. Let X = {(a, b) : a ∈ h, b ∈ h∗ }. By convention, we will generally write X = h ⊕ h∗


when we consider X as a representation of W , as in Chapter 1, and X = h × h∗ when we are
thinking of X as an algebraic variety.

In the proof of Proposition 4.1, we will consider two filtrations on Hκ . One is the filtration F
defined in Section 1.2. The other is a filtration F with F 0 = C[h] ∗ W and F i = (C[h] + h)i ∗ W for

57
i ≥ 1. In other words, we place the elements of h∗ and W in filtration degree 0, and the elements
of h in filtration degree 1. We require a lemma concerning the filtration F.

Lemma 4.6.
grF (Hκ ) ∼
= C[h × h∗ ] ∗ W.

Proof. If x ∈ h∗ and y ∈ h, then [y, x] ∈ F 0 , so there is a natural map

φ : C[h × h∗ ] ∗ W # grF (Hκ )

which is clearly surjective. We wish to show that φ is injective. Let α, β ∈ Zn!0 be multiindices
!
and write xα for xα1 1 · · · xαnn and y β for y1β1 · · · ynβn . Suppose an element λαβw xα y β w ∈ ker(φ)
where the sum runs over α, β ∈ Zn!0 and w ∈ W .Then for a fixed natural number a, writing
! ! ! ! !
|β| := i βi , we have |β|=a α,w λαβw xα y β w ∈ ker(φ) and so |β|=a α,w λαβw xα y β w ∈ F a−1 .
But this contradicts the PBW theorem. So ker(φ) = 0 and φ is an isomorphism.

The filtration F induces a filtration on eHκ e, and grF (eHκ e) ∼


= eC[h × h∗ ] ∗ W e ∼
= C[h × h∗ ]W .
In this way, we get a Poisson bracket on C[h × h∗ ]W . Similarly, e− Hκ e− also defines a Poisson
bracket on C[h × h∗ ]W .
Also, recall that h × h∗ is a symplectic vector space with symplectic form ω((a, α), (b, β)) =
β(a) − α(b). This defines a third Poisson bracket {−, −}ω on C[h × h∗ ]W .

Lemma 4.7. The Poisson bracket on C[h × h∗ ]W induced from the isomorphism grF (eHκ e) ∼
=
C[h × h∗ ]W coincides with the bracket defined by grF (e− Hκ e− ) ∼
= C[h × h∗ ]W , and furthermore, both
of these brackets coincide with the natural bracket {−, −}ω .

Proof. The algebra C[h × h∗ ]W is graded by polynomial degree. Note that this grading is not
the grading inherited from the isomorphism C[h × h∗ ]W ∼
= grF (eHκ e). The idea of the proof is to
calculate the degree of the three Poisson brackets with respect to the grading by polynomial degree.
Let p, q be homogeneous polynomials in C[h × h∗ ]W . Then, using the notation of the proof of
Lemma 4.6, there are scalars λαβ , α, β ∈ Zn≥0 such that
&
p= λαβ xα y β .
|α|+|β|=deg(p)

Let p̂ ∈ eHκ e be the sum of noncommutative monomials


&
p̂ = λαβ xα y β e.
|α|+|β|=deg(p)

58
Then p is the image of p̂ in grF (eHκ e). Similarly, we may define q̂. The defining relations of the
Cherednik algebra show that [p̂, q̂] ∈ F deg(p)+deg(q)−2 (Hκ ) and hence {p, q} has polynomial degree
≤ deg(p) + deg(q) − 2. Therefore, the degree of the bracket {−, −} induced from eHκ e is ≤ −2.
To show that this bracket has degree exactly −2, we calculate
&n n
&
(
{ xi , xj yj }.
i=1 j=1

In the Cherednik algebra, we can compute


&n n
& n &
& n
[ (
xi , xj yj ] = [x(i , xj yj ]
i=1 j=1 i=1 j=1

& n &
n & (−1
= xbi xj [xi , yj ]x(−1−b
i
i=1 j=1 b=0

A tedious calculation using the relations 1.3 then yields


n n
8 n
9
& & &
[ (
xi , xj yj ]e = −! (
xi e.
i=1 j=1 i=1

Thus, multiplying on the left and right by e, we get that


&n n
& n
&
{ x(i , xj yj } = −! x(i .
i=1 j=1 i=1

So {−, −} has degree exactly −2. Since h = Cn is an irreducible representation of G(!, 1, n),
it now follows from [26, Theorem 2.23] that {−, −} is a scalar multiple of {−, −}ω . But since
! ! !
{ ni=1 x(i , nj=1 xj yj }ω = −! ni=1 x(i , we must have {−, −} = {−, −}ω as required.
The proof that the bracket induced from e− Hκ e− also coincides with {−, −}ω is identical.

We need one more definition before proceeding with the proof of Proposition 4.1.

Definition 4.8. If S ⊂ R where R is a Poisson algebra, then we say that the Poisson subalgebra
of R Poisson-generated by S is the smallest subalgebra of R which contains S and which is closed
under the Poisson bracket.

Our next lemma is based on [6, Lemma 4.7]. We follow the proof of [6] very closely, except we
have to make some minor modifications since W is not a Coxeter group.
Note that if R is a Poisson algebra which is a domain, then the Poisson structure on R can be
extended to any localisation of R in a natural way. For s, x ∈ R, define {s−1 , x} = −s−2 {s, x}. In

59
this way, the Poisson structure on C[h × h∗ ]W considered in Lemma 4.7 induces a Poisson structure
on C[hreg × h∗ ]W , which we also denote by {−, −}.

Lemma 4.9. The algebra C[hreg × h∗ ]W is Poisson-generated by C[hreg ]W ∪ C[h∗ ]W .

Proof. Let R := C[hreg × h∗ ]W and let A be the Poisson subalgebra of R Poisson-generated by


C[hreg ]W ∪ C[h∗ ]W . Since C[h× h∗ ] = C[h]⊗ C[h∗ ], we have that C[h× h∗ ] is a finite C[h]W ⊗ C[h∗ ]W –
module, by Theorem 1.9. So C[hreg × h∗ ] is a finite C[hreg ]W ⊗ C[h∗ ]W –module and it follows that R
is also a finite C[hreg ]W ⊗ C[h∗ ]W –module. Therefore, R is a finite A–module and R is a Noetherian
C[hreg ]W ⊗ C[h∗ ]W –module. Since A is a C[hreg ]W ⊗ C[h∗ ]W –submodule, it follows that A is a
Noetherian C[hreg ]W ⊗ C[h∗ ]W –module and in particular, A is a Noetherian ring. Furthermore, A
is a domain since A ⊂ R.
In geometric language, we have an affine variety Y = (hreg × h∗ )/W and another affine variety
Y $ := maxspec(A). The inclusion A 3→ R induces a map f : Y → Y $ . This is a finite map since
R is a finite A–module, hence f is surjective by [24, Corollary 9.3]. We aim to show that f is an
isomorphism. First, we show that f is injective and then that f is an isomorphism.
Before beginning the proof, we note that the Poisson bracket on C[hreg × h∗ ] satisfies {f, yi } =
∂f
− ∂x i
for all f ∈ C[hreg ]. Furthermore, let L := (! − 1)(! − 2) and define D : A → A by D =
! ! ! !n
{ ni=1 x(i , −}. Then D(−2 ( nj=1 yj( ) is a nonzero scalar multiple of ni=1 xL
i yi . Hence,
2 L 2
i=1 xi yi ∈

A.
Now we show that f is injective. Let (qi , pi ), i = 1, 2, be points of hreg × h∗ and denote by [qi , pi ]
their images in (hreg × h∗ )/W . We have f ([q1 , p1 ]) = f ([q2 , p2 ]) if and only if a(q1 , p1 ) = a(q2 , p2 )
for all a ∈ A. If this holds, then since A ⊃ C[hreg ]W ∪ C[h∗ ]W , we have q2 = wq1 , p2 = up1 for
some u, w ∈ W . So [q2 , p2 ] = [q1 , w−1 up1 ]. Thus it suffices to show that if (q, p) ∈ hreg × h∗ and Wp
denotes the stabiliser of p in W , then there exists a ∈ A such that the values a(q, gp) are distinct
as g runs over a set of coset representatives W/Wp of Wp in W .
Let ,−, −- be a W –invariant Hermitian inner product on h such that the basis {yi } is or-
thonormal, and use the same notation to denote a W –invariant Hermitian inner product on h∗
such that the basis {xi } is orthonormal. We may choose z ∈ h∗ such that the inner products
! !
,z, gp- are distinct for distinct gp. Write z = zi xi and gp = (gp)i xi with zi , (gp)i ∈ C, so that
!
,z, gp- = zi (gp)i where the bar stands for complex conjugation. Now, since hreg ⊂ {(a1 , . . . , an ) ∈
Cn |ai 1= 0 for all i}, we have qi 1= 0 for all i (where qi are the coordinates of q with respect to the

60
dual basis of {xi }). Therefore, there is a well-defined linear functional
&
ẑ = (qi−L zi )xi ∈ h∗ = Tq∗ (hreg ) ∼
= Tq∗ (hreg /W ).
i

(Here, we used the fact that Tq∗ (hreg ) ∼


= Tq∗ (hreg /W ) since W acts freely on hreg ). Choose b ∈
C[hreg ]W such that dbq = ẑ, ie. ∂b
∂xi |q = qi−L zi for 1 ≤ i ≤ n. Let

1& L 2 & ∂b
a = {b, xi y i } = − xL
i yi ∈ A.
2 ∂xi
i i
! !
Then a(q, gp) = − i qi (gp)i ∂xi |q
L ∂b
= − i (gp)i zi , which are distinct for distinct gp. So f is
injective.
Now we show that df is injective on tangent spaces. Let a1 , . . . , an ∈ C[hreg ]W be chosen so
! ∂ak
that {(dai )q |1 ≤ i ≤ n} is a basis for h∗ = Tq∗ (hreg /W ). Note that (dak )q = i xi ( ∂xi |q ), so
1! ! L ∂ai
det( ∂a
∂xi |q ) 1= 0. For 1 ≤ i ≤ n, define bi := {ai , 2
k
j xj y j } = −
L 2
j xj yj ∂xj ∈ A.

Then for (q, p) ∈ hreg × h∗ ,


& ∂ak
(dak )(q,p) = xi ( |q )
∂xi
i
and, for 1 ≤ k, i ≤ n, there exist scalars αki such that
& & ∂ak ∗
(dbk )(q,p) = αki xi + yi (qiL |q ) ∈ T(q,p) (hreg × h∗ /W ) = h∗ ⊕ h.
∂xi
i i

Therefore, the vectors {(dak )(q,p) } ∪ {(dbk )(q,p) }, 1 ≤ k ≤ n are linearly independent in h∗ ⊕ h, since
 
∂ak
| 0 '
det  ∂xi q  = ( qiL )(det( ∂ak |q ))2 1= 0.
∗ qiL ∂a ∂xi
∂xi |q
k
i

Hence, if (q, p) ∈ Y then T(q,p)


∗ Y is spanned by {(da)(q,p) |a ∈ A} so f induces a surjection on cotan-
gent spaces, hence an injection on tangent spaces. So by [41, Theorem 14.9], f is an isomorphism.
So the inclusion A 3→ R is an isomorphism, hence is surjective. So A = R.

Now we require the following lemma due to Levasseur-Stafford.

Lemma 4.10. [50, Lemma 9] Let R ⊂ S be two Noetherian domains such that S is simple and is
a finite left and right R–module. Suppose Frac(R) = Frac(S). Then R = S.

Our next lemma is exactly [6, Theorem 4.6]. The proof is identical to the proof in [6], but for
completeness we give the argument anyway.

61
Lemma 4.11. [6, Theorem 4.6] If κ is regular then eHκ e is generated as an algebra by C[h]W e ∪
C[h∗ ]W e and e− Hκ e− is generated as an algebra by C[h]W e− ∪ C[h∗ ]W e− .

Proof. We give the proof for eHκ e, the proof for e− Hκ e− being the same but with e replaced by
e− . Let S = eHκ e and R the subalgebra of S generated by C[h]W e ∪ C[h∗ ]W e. Then with respect to
the filtration F on Hκ which gives h, h∗ degree 1, we get gr(S) = C[h × h∗ ]W . Now, C[h × h∗ ]W is a
finite C[h]W ⊗ C[h∗ ]W –module, and so gr(S) is a finite gr(R)–module. So S is a finite left and right
R–module. Furthermore, S is a Noetherian domain and S is simple by [6, Lemma 4.1]. Clearly,
gr(R) is a domain. Also, since gr(S) is a Noetherian C[h]W ⊗ C[h∗ ]W –module, so is gr(R). Hence,
gr(R) is a Noetherian ring and therefore so is R. In order to apply Lemma 4.10, it remains to show
that Frac(R) = Frac(S).
#
Now we use the second filtration F on Hκ . Recall the element δ = H∈A αH from Chapter
1. The element δ2( ∈ C[h] is W –invariant, so we may localise R and S at δ2( . Write R|hreg /W and
S|hreg /W for these localisations. It suffices to show that R|hreg /W = S|hreg /W , and since the filtration
F induces a nonnegative filtration1 on S|hreg /W , it suffices to show that gr(R|hreg /W ) = gr(S|hreg /W ).
We have gr(S|hreg /W ) = gr(eHκ [δ−2( ]e). By [36, Lemma 6.8], this equals egr(Hκ [δ−2( ])e ∼
= eC[hreg ×
h∗ ] ∗ W e ∼
= C[hreg × h∗ ]W . The algebra C[hreg × h∗ ]W inherits a Poisson bracket from this construc-
tion. By Lemma 4.7, this bracket coincides with the standard bracket. Furthermore, gr(R|hreg /W )
is a Poisson subalgebra of C[hreg × h∗ ]W containing C[hreg ]W and C[h∗ ]W . So by Lemma 4.9,
gr(R|hreg /W ) = gr(S|hreg /W ). So R = S as required.

We are now in a position to prove Proposition 4.1 in the case when κ is regular.
Recall that θκ : Hκ → D(hreg ) ∗ W and θκ[a] : Hκ[a] → D(hreg ) ∗ W denote the Dunkl rep-
κ[a]
resentations. Write Tiκ for the Dunkl operator θκ (yi ) and similarly for Ti . If f ∈ C[h]W then
θκ (f ) = θκ[a] (f ) = f so eθκ[a] (f )e = eµ−1
a θκ (f )µa e trivially. Now suppose f ∈ C[h ] . Let
∗ W

g ∈ C[h]W . Then g is a symmetric polynomial in x(1 , . . . , x(n . We have


' ' ' '
µ−1
a θκ (f )µa (g) = (x(i − x(j )−1 ( xi )−a θκ (f )( xi )a ( (x(i − x(j )g)
i<j i<j
' '
= (x(i − x(j )−1 θ(κ00 ,κ1 +1,...,κa +1,κa+1,...,κ#−1 ) (f )( (x(i − x(j )g) by [23, Prop 3.23]
i<j i<j

= θκ[a] (f )(g) by [23, Prop 3.26].


1
This is why we have to use the filtration F and not F .

62
Therefore, if we denote by res(D) the restriction of a W –invariant map D : C[h] → C[h] to
C[h]W , then we have shown that res(µ−1
a θκ (f )µa ) = res(θκ[a] (f )). But if D : C[h] → C[h] then

eDe(g) = e(D ◦ av)(g) where av : C[h] → C[h]W denotes the averaging map. So res(µ−1
a θκ (f )µa ) =

res(θκ[a] (f )) implies that µ−1


a e θκ (f )e µa = eµa θκ (f )µa e = eθκ[a] (f )e. By Lemma 4.11, we obtain
− − −1

µ−1 − −
a θκ (e f e )µa = θκ[a] (ef e)

for all f ∈ Hκ . By the PBW theorem, eHκ e is spanned by the W –invariant noncommutative
polynomials p(x, y) in the variables x1 , . . . , xn and y1 , . . . , yn . We have shown the following.

Lemma 4.12. Let p(x, y) be any W –invariant noncommutative polynomial in the variables x1 , . . . , xn
and y1 , . . . , yn . Then for regular κ, we have

ep(x, T κ )e = µ−1 −
a e p(x, T
κ[a] −
)e µa ,

where p(x, T κ ) ∈ D(hreg ) ∗ W denotes p with the Dunkl operator Tiκ substituted for yi , 1 ≤ i ≤ n.

It remains to extend to the case of non-regular κ. This part of the argument is based on an
argument of Berest and Chalykh [5]. Let κ 00 , κ 1 , . . . , κ (−1 be commuting indeterminates and let
κ00 , . . . , κ (−1 ]⊗C (D(hreg )∗W ). Define Dunkl operators Tiκ ∈ D by the usual formula for the
D = C[κ
Dunkl operators from Section 1.3.4, but with the parameters κ replaced by the indeterminates κ .
Let {bλ } be a C–basis of D(hreg )W ∼
= e(D(hreg )∗W )e. Then there are aλ (κ
κ), a$λ (κ
κ) ∈ C[κ
κ00 , . . . , κ(−1 ]
with
&
ep(x, Tiκ )e = aλ (κ
κ)bλ
λ
κ [a]
&
eµ−1
a p(x, Ti )µa e = a$λ (κ
κ)bλ
λ

where κ [a]00 = κ 00 + 1 and κ [a]i = κ i + 1, 1 ≤ i ≤ a and κ [a]i = κ i otherwise. Now by Lemma


4.12, for all λ we have aλ (κ) = a$λ (κ) for all regular κ. Therefore, aλ and a$λ are equal on a dense
subset of C( , so they are equal everywhere. So aλ (κ
κ) = a$λ (κ
κ). This proves the following theorem.

Theorem 4.13. Let p(x, y) be any W –invariant noncommutative polynomial in the variables
x1 , . . . , xn and y1 , . . . , yn . Then for all κ ∈ C( , we have

ep(x, T κ )e = µ−1 −
a e p(x, T
κ[a] −
)e µa ,

where p(x, T κ ) ∈ D(hreg ) ∗ W denotes p with the Dunkl operator Tiκ substituted for yi , 1 ≤ i ≤ n.

63
Since eHκ e is generated by W –invariant noncommutative polynomials in the xi ’s and yi ’s,
Proposition 4.1 follows immediately from Theorem 4.13.

4.1.2 The shift functors

We use the following notation.

Notation 4.14. In general, we have a parameter shift pa : κ 5→ κ[a] for each 0 ≤ a ≤ ! − 1. Clearly
pa has an inverse p−a which we denote by κ 5→ κ[−a]. If a1 , a2 , . . . ak are integers with the same
sign, then we will write κ[a1 a2 · · · ak ] as shorthand for ((κ[a1 ])[a2 ] · · · )[ak ].

For the rest of this chapter, we will identify θκ (Hκ ) with Hκ , so that the Cherednik algebra Hκ
will be regarded as the subset of D(hreg ) ∗ W generated by C[h], W and the Dunkl operators. We
also write Uκ = eθκ (Hκ )e = θκ (eHκ e). With this notation, Proposition 4.1 reads

Uκ[a] = eµ−1
a Hκ µa e.

a := µa eµa . Then recall that ea is an idempotent in CW , since µa is a W –semiinvariant,


We write e− −1 −

a ∈ Hκ , and it follows that


so e−
eHκ µa e

is a Uκ − Uκ[a] –bimodule.

Definition 4.15. The Heckman-Opdam shift functor Fκa associated to κ ∈ C( and 0 ≤ a ≤ ! − 1


is the functor Uκ[a] − Mod → Uκ − Mod defined by

M 5→ eHκ µa e ⊗Uκ[a] M.

We may also define a related functor Fκ−a : Uκ −Mod → Uκ[a] −Mod by Fκ−a (M ) = eµ−1
a Hκ e⊗Uκ M .

Remark 4.16. Note that the functors Fκa and Fκ−a also give functors Uκ[a] − mod ↔ Uκ − mod.
This is because eµa Hκ e and eµ−1
a Hκ e are finitely-generated right eHκ e–modules. Finite generation

follows by considering the associated graded modules with respect to the order filtration F. For
example, grF (eµa Hκ e) = (µa C[h ⊕ h∗ ])W . Since µa is a W –semiinvariant, Cµa is a one-dimensional
−1
representation of W , say with character χ. Then (µa C[h ⊕ h∗ ])W = µa C[h ⊕ h∗ ]W,χ , which is a
finitely-generated C[h ⊕ h∗ ]W –module by Theorem 1.9.

64
We are interested in the question of when the functor Fκa is an equivalence of categories (note
that we cannot hope that Fκa will always be an equivalence, as shown, for example, in [36, Section
3.14]). The first step in addressing this question is to write the functor in another way. Using

a Hκ ea defined by ψ(x) = µa xµa . Write


Proposition 4.1, we have an isomorphism ψ : Uκ[a] → e− − −1

Uκ− := e−
a Hκ ea . There is a twist functor Gψ : M 5→ M from Uκ[a] − Mod to Uκ − Mod defined
− ψ −

by M ψ = µa M . We may rewrite Fκa (M ) as

Fκa (M ) = eHκ µa e ⊗Uκ[a] M

= eHκ µa eµ−1
a ⊗µa eHκ[a] eµ−1
a
µa M

= eHκ e−
a ⊗Uκ− M
ψ

= Pe ◦ Qe−
a
◦ Gψ (M )

where Qe−
a
: Uκ− − Mod → Hκ − Mod is defined by Qe−
a
= Hκ e−
a ⊗Uκ− (−) and Pe : Hκ − Mod →

Uκ − Mod is defined by Pe = e(−).


Therefore, Fκa = Pe ◦ Qe−
a
◦ Gψ . The functor Gψ is an equivalence because ψ is an isomorphism.
We turn to the problem of showing that Pe and Qe−
a
are equivalences. The following standard
lemma is very useful to us.

Lemma 4.17. Let A be an algebra and x ∈ A an idempotent such that AxA = A. Then the
functors Px : M 5→ xM and Qx : N 5→ Ax ⊗xAx N are inverse equivalences between A − Mod and
xAx − Mod.

Proof. We only check on objects, since checking that certain maps are natural etc. is straightfor-
ward. Clearly Px Qx N = N for all N . It remains to show that if M ∈ A − mod then Ax ⊗xAx xM ∼
=
M . The multiplication map Ax ⊗xAx xM → M is surjective because AxM = AxAM = AM = M ,
!
so we need only show that it is injective. Suppose ai ∈ A, bi ∈ M and ai xbi = 0. We need to
! !
show that ai x ⊗ xbi = 0. Since AxA = A, we have 1 = wi xzi for some wi , zi ∈ A. Then
! ! ! ! ! ! !
ai x ⊗ xbi = i j wj xzj ai x ⊗ xbi = j wj x ⊗ i xzj ai xbi = j wj x ⊗ xzj i ai xbi = 0.

We have Fκa = Pe ◦ Qe−


a
◦ Gψ . Similarly, we may write Fκ−a = Gψ−1 ◦ e−
a (−) ◦ Hκ e ⊗eHκ e (−).

Therefore, Lemma 4.17 gives us the following result.

Lemma 4.18. The functors Fκa and Fκ−a are equivalences whenever Hκ eHκ = Hκ = Hκ e−
a Hκ .

Furthermore, if these conditions hold then Fκa and Fκ−a are quasi-inverses.

65
Now, following [36], we relate the conditions of Lemma 4.18 to category O. From Theorem 1.25
we get the following.

Lemma 4.19. Let eχ ∈ CW be an idempotent corresponding to a linear character χ of W . Then


for any choice of parameter κ ∈ C( , Hκ eχ Hκ = Hκ if and only if eχ Lκ (τ ) 1= 0 for all τ ∈ Irrep(W ).

Proof. By Theorem 1.25, if Hκ eχ Hκ 1= Hκ , then eχ would annihilate some simple object of category
O.

4.1.3 The semisimple and almost-semisimple cases

The easiest case in which the Heckman-Opdam shift functors are equivalences is the case where
category O is semisimple. We know by Theorem 2.1 that this happens if and only if Hκ is simple.
But if Hκ is simple then Hκ xHκ = Hκ for any nonzero x ∈ Hκ . This proves the following theorem.

Theorem 4.20. Suppose Oκ is semisimple. Then the Heckman-Opdam shift functors Fκa and Fκ−a
are equivalences.

The next situation in which we wish to check that Fκa is an equivalence is the case when the
KZ functor separates simples. By [36, Remark 3.14], it is not true that Fκa is always an equivalence
whenever KZκ separates simples, so it will be necessary to impose a further condition. We first
need a lemma, which shows that the condition of Lemma 4.19 will be satisfied for all τ such that
L(τ )|hreg 1= 0.

Lemma 4.21. Suppose M is an object of O and KZ(M ) 1= 0. Then for any linear character χ of
W , if eχ denotes the corresponding idempotent, then eχ M 1= 0.

Proof. Consider the two-sided ideal I = Hκ eχ Hκ . Localising to hreg , I|hreg is a nonzero ideal of
Hκ |hreg , which is a simple ring. Therefore, I|hreg = Hκ |hreg and therefore I contains a power of
#
the element δ = H∈A αH . Say δk ∈ I. Then if eχ M = 0 then δk M = 0 and M |hreg = 0 so
KZ(M ) = 0.

Therefore, if KZκ separates simples, then Fκa will be an equivalence provided that if L is the
unique finite-dimensional simple object in Oκ , then eL 1= 0 and e−
a L 1= 0.

Theorem 4.22. Suppose KZκ separates simples, and suppose further that the unique finite-dimensional
simple object L in the category Oκ has dimension dim(L) > (n!)n . Then eL 1= 0 and e−
a L 1= 0.

66
Proof. We will show in fact that the unique finite-dimensional simple object L in category O
contains a copy of every linear representation of W if dim(L) > (n!)n . By Theorem 3.7, the
dimension of L is of the form r n = ((p − 1)! + s)n for some p ≥ 1 and some 1 ≤ s ≤ ! − 1.
So our hypothesis on the dimension of L says that p ≥ n + 1. Now recall that L is constructed
as the module Y˜c of [16, Section 4.1]. By Theorem 3.3, there exists a linear character χ of W
such that L is isomorphic as a W –module to χ ⊗ (C[u]/(ur ))⊗n , where the generators of W act
as follows: the generator s1 with s(1 = 1 acts by st1 (uj ) = ε−jt uj , while Sn acts by permuting the
factors of the tensor product. Now consider an arbitrary linear character ζ of W . Such a character
(0) (0) (0)
acts on s1 by ε−t for some t, and takes all the transpositions σij either to σij or −σij (ie. its
restriction to the symmetric group is either the trivial or the sign representation). Consider the
elements ut , ut+( , . . . , ut+(n−1)( ∈ C[u]. Provided r ≥ n!, these are guaranteed to remain linearly
independent in C[u]/(ur ). Write ai = ut+(i−1)( , 1 ≤ i ≤ n. Then consider the element
&
a := sgn(σ)b aσ(1) ⊗ aσ(2) ⊗ · · · ⊗ aσ(n) ,
σ∈Sn

where b = 0 if ζ|Sn is the trivial representation of Sn and b = 1 if it is the sign representation. Then
Ca is a copy of the representation corresponding to χ ⊗ ζ, so eχ⊗ζ L 1= 0. Since ζ was arbitrary,
this shows that L contains a copy of every linear representation of W .

Corollary 4.23. If KZκ separates simples and the unique finite-dimensional simple object L in Oκ
has dimension dim(L) > (n!)n , then Fκa and Fκ−a are quasi-inverse equivalences.

4.1.4 The asymptotic parameter case

Now let us turn to the other situation in which we can prove that Fκa is an equivalence. This is the
situation where the parameters are asymptotic, in the sense of Definition 4.24 below.

Definition 4.24. Say the parameter tuple κ = (κ00 , κ1 , . . . , κ(−1 ) is asymptotic if κ00 ∈ R and
κi ∈ R for all i, and the following two conditions hold.

1
κ00 < − n(n + 1)
2
κi − κi−1 ≥ −nκ00

for 0 ≤ i ≤ ! − 2, where as usual we put κ0 = 0 and κ−1 = κ(−1 .

67
The point of this definition is that Theorem 4.26 below shows that when the parameters are
asymptotic, category O is equivalent to the category of modules over an algebra called the cy-
clotomic q–Schur algebra. We will exploit the combinatorics of the category of modules over the
q–Schur algebra in order to prove that the Heckman-Opdam shift functors are equivalences.
Note that, in this section, we will freely identify a multipartition of n with ! parts with the
corresponding representation of W .
We begin by giving the definition of the cyclotomic q–Schur algebra. Let q, u1 , . . . , u( be nonzero
complex numbers. Then we have the Ariki-Koike algebra H of Definition 1.34. Recall that this
algebra is generated by elements Ts and Tti , 2 ≤ i ≤ n. By [54, Section 2.2], the subalgebra of
H generated by Tt2 , . . . , Ttn is isomorphic to the Hecke algebra of the group Sn , as defined in [53,
Chapter 1], with parameter q. Let s1 , . . . , sn−1 be the usual generators of the symmetric group, so
that si = (i, i + 1) is the transposition. For a word w in these generators, let w = si1 si2 · · · sik be a
reduced expression in the terminology of [53]. Then define Tw := Tti1 −1 · · · Ttik −1 ∈ H. The element
Tw is well-defined by [54, Section 2.2]. Furthermore, for k = 1, . . . , n, define an element Lk ∈ H by
Lk := q 1−k Ttk · · · Tt2 Ts Tt2 · · · Ttk . Now for a multipartition λ = (λ(1) , . . . , λ(() ) of n with ! parts,
define  ! i−1 
(r) |
& d
' r=1 |λ
'
mλ = ( Tw )  (Lj − ui )
w∈Sλ i=2 j=1

where Sλ = Sλ(1) × Sλ(2) × · · · × Sλ(#) , and Sλ(i) denotes the group of permutations of the Young
diagram λ(i) which stabilise each row of λ(i) .

Definition 4.25. The cyclotomic q–Schur algebra S = S(q, u1 , . . . , u(−1 ) is the algebra
 
"
EndH  mλ H 
λ∈Π#n

where Π(n denotes the set of mulipartitions of n with ! parts.

By [54, Theorem 4.13, Theorem 4.14], for each λ ∈ Π(n there is a module W λ for S called a Weyl
module. Each Weyl module has a quotient F λ which is absolutely irreducible, and furthermore,
{F λ : λ ∈ Π(n } is a complete set of nonisomorphic irreducible S–modules.
We have the following theorem due to Rouquier.
#
Theorem 4.26. [64, Theorem 6.8] Suppose that the parameters are asymptotic and (q+1) i<j (ui −

68
uj ) 1= 0. Then there is an equivalence O → S − mod that sends the standard module M (λ) to the
Weyl module W λ for all λ ∈ Irrep(W ).

Recall the dominance ordering on multipartitions. Given two multipartitions λ = (λ(1) , . . . , λ(() )
and µ = (µ(1) , . . . , µ(() ) we say λ ≤dom µ if for all 0 ≤ k ≤ ! and all t, we have
k
& t
& k
& t
&
|λ(j) | + λr(k+1) ≤ |µ(j) | + µr(k+1) .
j=1 r=1 j=1 r=1

Note that the dominance ordering is not a total ordering. Theorem 4.26 allows us to prove the
following very useful result.
#
Lemma 4.27. Suppose the parameters are asymptotic and (q + 1) i<j (ui − uj ) 1= 0. Suppose λ,
µ are multipartitions of n with ! parts and HomO (M (λ), M (µ)) 1= 0. Then λ ≤dom µ.

Proof. By Theorem 4.26, there is an equivalence O → S − mod which sends M (λ) to W λ for all λ.
So HomS (W λ , W µ ) 1= 0. Now, F λ is the unique irreducible quotient of W λ , so if φ : W λ → W µ is
a nonzero map, then F λ is a composition factor of im(φ). Therefore, F λ is a composition factor of
W µ . Now [21, Corollary 6.17] yields λ ≤dom µ.

The point of Lemma 4.27 is that it enables us to apply the method of [36] to show that the
Faκ are equivalences. The rest of our proof follows the proof in [36] almost word-for-word, but we
include it here for completeness.
#
Lemma 4.28. [36, Corollary 4.6] Suppose the parameters are asymptotic and (q + 1) i<j (ui −
uj ) 1= 0.
Suppose [M (µ) : L(λ)] 1= 0. Then λ ≤dom µ.

Proof. The proof is by induction on µ. We usually write λ ! µ for λ !dom µ. Suppose that
[M (µ) : L(λ)] 1= 0 with λ 1= µ, and the conclusion of the lemma holds for all ν < µ. The
module L(λ) has a projective cover P (λ), by Theorem 1.24. Since L(λ) is a composition factor
of M (µ), there is a submodule A of M (µ) such that L(λ) 3→ M (µ)/A. Therefore, we get a map
P (λ) → M (µ)/A, and the projectivity of P (λ) yields a map P (λ) → M (µ), which is certainly
nonzero. Now by [33, Corollary 2.10], there is a filtration of P (λ) of the form

P (λ) = A0 ⊃ A1 ⊃ · · · ⊃ At = 0

69
where At /At+1 = M (λt ) for some λt . Let K be the kernel of the map P (λ) → M (µ) and choose i
so that (K + Ai )/K 1= 0 and (K + Ai+1 )/K = 0. This gives a map

ψ : M (λi ) = Ai /Ai+1 → (Ai + K)/K 3→ P (λ)/K → M (µ).

This map is nonzero, and it follows that λi ≤ µ by Lemma 4.27. Now, if λi = µ then we have
a nonzero map ψ : M (µ) → M (µ). Such a ψ must be surjective. Indeed, if we write µ for
the irreducible representation of W corresponding to µ, then for some nonzero v ∈ µ we have
1 ⊗ v ∈ M (µ). This vector generates M (µ) as a Hκ –module, so ψ(1 ⊗ v) must be nonzero. But
ψ(1 ⊗ v) must belong to the lowest weight space of M (µ) for the h–action, which is 1 ⊗ µ. Any
nonzero vector in 1 ⊗ µ generates the whole of M (µ), and therefore ψ must be surjective. It follows
that the map P (λ)/K → M (µ) is surjective, and therefore so is the map L(λ) → M (µ)/A. So L(λ)
is a simple quotient of M (µ). Therefore, L(λ) ∼
= L(µ) so λ = µ, a contradiction. This shows that
λi < µ.
Now by the BGG reciprocity, [P (λ) : M (λi )] = [M (λi ) : L(λ)] 1= 0. So by induction, λ ≤ λi
and therefore λ < µ.

Given a Hκ –module M and a scalar α ∈ C, we have the weight space Wα (M ), which is defined
to be the generalised eigenspace for the action of h with eigenvalue α. Because h commutes with
W , the space Wα (M ) is W –stable for all α. Following [36, Section 3.10], we define the graded
Poincaré series of a Hκ –module M to be the expression
& &
p(M, v, W ) := vα [Wα (M ) : λ][λ]
α∈C λ∈Irrep(W )

where v is an indeterminate, and the [λ] denote the isomorphism classes of irreducible representa-
tions of W .
We also make the following definition.

Definition 4.29. If W ! GL(h) is a complex reflection group acting faithfully on a finite-dimensional

+ - the ideal of C[h] generated by the invariant polynomials


complex vector space h, denote by ,C[h]W
of positive degree. Then the ring of coinvariants of W is the ring

C[h]
.
,C[h]W
+-

This ring is graded by polynomial degree and we denote the ith graded component by (C[h]/,C[h]W
+ -)i .

70
Proposition 4.30. [36, Proposition 3.10] The graded Poinaré series of M (λ) is
!
µ Gµ (v)[λ ⊗ µ]
−cλ #
p(M (λ), v, W ) = v n ,
i=1 (1 − v )
i(

where Gµ (v) is the polynomial in v defined by


> ?
& ( C[h] )
Gµ (v) = W-
: µ vi .
i≥0
,C[h]+ i

Proof. First of all, the lowest weight space of M (λ) is 1 ⊗ λ. The eigenvalue of h on this space is
!
just the eigenvalue of −z, because h = i xi yi − z. In our notation, this is −cλ .
As a graded W –module, M (λ) = (C[h] ⊗ λ)[−cλ ] where [−cλ ] denotes a grading shift. So it
!
remains to compute the graded Poincaré series i,τ v i [C[h]i : τ ][τ ] of C[h]. By the proof of [68,
Theorem 2.2], there is a decomposition as graded W –modules,

C[h]
C[h] = C[h]W ⊗ ,
,C[h]W
+-

and the graded Poincaré series of C[h] is the product of the graded Poincaré series of the two
factors. It is known (see [11, Table 1]) that the degrees of a set of algebraically independent
homogeneous generators of C[h]W are !, 2!, . . . , n!, which says that the graded Poincaré series of
# C[h] !
C[h]W is ni=1 (1 − v i( )−1 . On the other hand, the graded Poincare series of 0C[h] W 1 is µ Gµ (v) by
+

definition.

Remark 4.31. The polynomials Gµ (v) are referred to as fake degrees and appear in work of
Opdam [60].

Now we explain how Lemma 4.27 and Proposition 4.30 enable us to prove that the shift functors
are equivalences. Recall from Lemma 4.19 that, to prove that Fκa is an equivalence, it suffices to
show that eLκ (τ ) 1= 0 and e−
a Lκ (τ ) 1= 0 for all τ . That is, no simple object of the category Oκ is

annihilated by e or e−
a.

Suppose we have a linear character χ of W and an M (τ ). Then the composition factors of M (τ )


are L(τ ) and L(µ) for some set of µ < τ . The generalised eigenvalues of h on L(µ) are a subset of
those on M (µ). Let w(τ ) denote the set of weights (generalised h–eigenvalues) of M (τ ) and let dχτ
denote inf{d ∈ w(τ ) | [Wd (M (τ )) : χ] 1= 0}. If we can show that dχτ < dχµ for all µ < τ , then we
will be done, because there will be a copy of χ in M (τ ) which cannot lie in the radical R(τ ), and
hence from the exact sequence 0 → R(τ ) → M (τ ) → L(τ ) → 0, there will be a copy of χ in L(τ ).

71
Proposition 4.32. Let χ be a one-dimensional representation of W and let µ < τ be two repre-
sentations of W . Then
dχτ < dχµ .

Proof. By [68, Theorem 5.3], the polynomials Gµ (v) may be explicitly written as follows. Fix a
!(−1
multipartition µ = (µ(1) , µ(2) , . . . , µ(() ) with ! parts. Define r(µ) = i=1 (i − 1)|µ(i) |. Furthermore,
define (v ( )n = (1 − v ( )(1 − v 2( ) · · · (1 − v n( ), and if η is a partition with parts η1 , η2 , . . . (ie. ηi is the
!
length of the ith row of the Young diagram of η), define n(η) = i (i − 1)ηi , the partition statistic.
Also, define a polynomial
'
Hη (v) = (1 − v hooklength(i,j) ),
(i,j) a box in η

where hooklength(i, j) is by definition the cardinality of the set of boxes which lie to the right of
or below the box (i, j). Then [68, Theorem 5.3] states that
(−1
' (i)
v (n(µ )
Gµ (v) = v r(µ) (v ( )n .
Hµ(i) (v ( )
i=0

Therefore,

& (−1
' (i)
v (n(λ ) 1
p(M (µ), v, W ) = v −cµ v r(λ) (v ( )n # [λ ⊗ µ].
Hλ(i) (v ( ) ni=1 (1 − v i( )
λ i=0

Now, the linear character χ appears in λ ⊗ µ if and only if λ = µ∗ ⊗ χ. The term (v ( )n cancels with
#n
i=1 (1 − v ) = (v )n , and by taking the leading term, we see that the smallest weight of a copy of
i( (

χ in M (µ) is
(−1
&

dχµ = −cµ + r(χ ⊗ µ ) + ! n((χ ⊗ µ∗ )(i) ).
i=0

Therefore, we wish to show that if µ < τ then


& &
−cτ + r(χ ⊗ τ ∗ ) + ! n((χ ⊗ τ ∗ )(i) ) < −cµ + r(χ ⊗ µ∗ ) + ! n((χ ⊗ µ∗ )(i) ).
i i

Since n(−) and r(−) are nonnegative, this will certainly hold if we can show that
&
cτ − cµ > r(θ) + ! n(θ (i) ) (4.1)
i

for every multipartition θ.

72
To verify Equation 4.1, we use the calculations in [64, Sections 6.3, 6.4]. In the notation of [64],
h = κ00 , hi = −κ−i and his cχτ is the negative of our cτ . Let us write cRou
τ = −cτ = cχτ for the
parameter appearing in [64].
By [64, Proposition 6.4], if h ≤ 0 and hi − hi−1 ≥ −nh for 1 ≤ i ≤ ! − 1 and µ < τ , then
cRou
µ ≥ cRou
τ . We want to show that the difference can be made large enough so that Equation 4.1
holds. Given µ < τ , there are µ1 , µ2 , . . . , µt such that µ = µ1 < µ2 < · · · < µt = τ and for each j,
!
there is no multipartition ξ with µi < ξ < µi+1 . Then cτ − cµ = j (cµj − cµj−1 ) ≥ cµt − cµt−1 . So
we may assume that there is no µ$ with µ < µ$ < τ . Now, [64, Lemma 6.3] gives three possible case
in which τ > µ and there is no other multipartition strictly between them. The cases are labelled
(a), (b) and (c). Then using the proof of [64, Proposition 6.4], we have:
in case (c), there are positive integers i < i$ with

cτ − cµ = cRou
µ − cRou
τ = −!h(i$ − i + 1) ≥ −!h

because h < 0.
(s) (s)
in case (b), there is an s with τi > τi+1 and

(s) (s)
cτ − cµ = cRou
µ − cRou
τ = −!h(τi − τi+1 ) ≥ −!h

again because h < 0.


in case (a), let ps be the number of nonzero parts of µ(s) . Then there is an s such that

(s+1)
cτ − cµ = cRou
µ − cRou
τ = !(hs − hs−1 ) + !h(ps + µ1 − 1) ≥ −!nh + !h(n − 1) ≥ −!h

(s+1) !
where we used the fact that the parameters are asymptotic, and that ps + µ1 ≤ i |µi | = n.
Thus in all cases cτ − cµ ≥ −!h = −!κ00 > ! n(n+1)
2 since the parameters are assumed to
be asymptotic. To verify Equation 4.1, it remains to show that, for all multipartitions θ, r(θ) +
! !
! i n(θ (i) ) ≤ ! n(n+1)
2 . The maximum possible value of r(θ) = (i − 1)|θ (i) | is (! − 1)n, while
!
the maximum possible value of i n(θ (i) ) is attained when θ (i) is a column of length n for exactly
!
one i, and the other θ (j) are ∅. Therefore, the maximum value of r(θ) + ! i n(θ (i) ) is attained
when θ = (∅, ∅, · · · , ∅, γn ), where γn denotes a column of n boxes, and this maximum value equals
(! − 1)n + ! 12 n(n − 1) = ! n(n+1)
2 . This verifies Equation 4.1 and finishes the proof.

From the remarks proceeding Proposition 4.32, we get the following theorem.

73
#
Theorem 4.33. Suppose the parameter κ is asymptotic and (q + 1) i<j (ui − uj ) 1= 0. Then the
shift functors Fκa and Fκ−a are equivalences.

4.1.5 Shift functors on category O

We may construct functors Sκa : Hκ[a] − Mod → Hκ − Mod which are very similar to the functors
Faκ . They may be defined by
Sκa (M ) = Hκ µa e ⊗eHκ[a]e (eM ).

Note that eSκa (M ) = Fκa (eM ). The functor Sκa will be an equivalence provided M 5→ eM and
Hκ µa e ⊗eHκ[a] e (−) are both equivalences. By Lemma 4.18, Sκa will therefore be an equivalence if

a Hκ = Hκ and Hκ[a] eHκ[a] = Hκ[a] . Parts (1), (2) and (3) of the following theorem follow from
Hκ e−
the proofs of Theorem 4.20, 4.22 and 4.33 respectively.

Theorem 4.34. The functor Sκa is an equivalence in any of the following cases:

1. Oκ is semisimple.

2. The functor KZκ separates simples, and eL 1= 0, where L is the unique finite-dimensional
simple object in Oκ[a] , and furthermore, e−
a L 1= 0 where L is the unique finite-dimensional
$ $

simple object in category Oκ .


#
3. The parameters κ and κ[a] are asymptotic and (q + 1) i<j (ui − uj ) 1= 0.

One reason why the functors Sκa are of interest is that they can easily be shown to preserve
category O.

Lemma 4.35. [36] The functor Sκa restricts to a functor Oκ[a] → Oκ .

Proof. Let M ∈ Oκ[a] . Then Sκa (M ) = Hκ µa e ⊗eHκ[a]e eM . First, Sκa (M ) is finitely-generated,


because Hκ µa e is a finite eHκ[a] e–module (this follows from considering the associated graded
modules with respect to the filtration F on Hκ , and using Theorem 1.9). Now by the proof of
Theorem 1.25, we need only show that every y ∈ h acts locally-nilpotently on Sκa (M ). Let y ∈ h.
It suffices to show that, given m ∈ M and p ∈ C[h], there exists N such that y N pµa e ⊗ (em) = 0.
Since C[h∗ ] is a finite C[h∗ ]W –module by Theorem 1.9, it suffices to show that every r ∈ C[h∗ ]W
+

acts nilpotently on pµa e ⊗ (em). The proof of this is identical to the proof of the analogous fact
for W = Sn given in [36, Proposition 3.16].

74
It is also possible to prove various other properties of the functors Sκa that hold when they are
equivalences. However, we will not use these properties in the sequel, and their proofs are identical
to the proofs of [36] for the ! = 1 case, so we omit them.

4.1.6 A commutativity property

Finally, we turn to an obvious question about the functors Fκa . If we have two integers a, b with
0 ≤ a, b ≤ ! − 1 then κ[ab] = κ[ba]. This suggests to ask the question whether any route from κ[ab]
to κ gives the same shift functor, ie. if we shift first to κ[a] and then to κ, is this the same as
shifting first to κ[b] and then to κ? We will show that this is the case when the functors involved
are all known to be equivalences. Although the theorem is original, the argument we use is due to
Ginzburg-Gordon-Stafford [32].
We start with a lemma.

Lemma 4.36. Suppose eHκ µa e induces a Morita equivalence Uκ[a] − Mod → Uκ − Mod and that
eHκ[a] µb e induces an equivalence Uκ[ab] −Mod → Uκ[a] −Mod. Let P1 = eHκ[a] µb e and P2 = eHκ µa e.
Let T = P2 ⊗Uκ[a] P1 . Then the multiplication map

m : T = P2 ⊗ P1 → P2 P1

is an isomorphism of Uκ − Uκ[ab] –bimodules.


!
Proof. Clearly m is surjective, so we need only show that it is injective. Suppose m ( ai ⊗ bi ) = 0.
Then since by the Morita theorem, each of the modules P1 , P2 is a progenerator, so is their product
P2 ⊗ P1 . The map m becomes, on localising by δ, the natural map D(hreg )W ⊗D(hreg )W D(hreg )W →
!
D(hreg )W , which is an isomorphism. So ( ai ⊗ bi )|hreg = 0. But since P2 ⊗ P1 is projective on
!
both sides, it is a submodule of a free C[h]–module. Therefore, ai ⊗ bi = 0 as required.

The next argument is due to [32], and uses Lemma 4.38 below.

Definition 4.37. [55, Section 5.1.7] Let R be a ring. A left R–module M is said to be torsionless
if for every m ∈ M \ {0} there exists α : M → R with α(m) 1= 0. A torsionless left R–module M
is reflexive if the natural map

M → M ∗∗ = HomR (HomR (M, R), R)

is an isomorphism of left R–modules.

75
Note that the class of reflexive modules is closed under direct sums and direct summands.
Therefore, every projective R–module is reflexive.
If R is a Noetherian domain and Frac(R) denotes the quotient division ring of R, then for a
left R–module M ⊂ Frac(R), the right R–module M ∗ := HomR (M, R) may be identified with
{q ∈ Frac(R) : M q ⊂ R} (see [55, Section 5.1.8]). Therefore, M ∗∗ may be naturally identified with
a subset of Frac(R). We will need to use the following Lemma due to Gabber.

Lemma 4.38. [67, Theorem 2.2] Suppose k is a field and R is a prime Noetherian ring which is
finitely-generated as a k–algebra and is Auslander-regular and Cohen-Macaulay.
If M is a finitely-generated left R–submodule of Frac(R) then M ∗∗ is the unique largest left
R–submodule M $ of Frac(R) such that M ⊂ M $ and GK dim(M $ /M ) ≤ GK dim(R) − 2.

Now we give Stafford’s argument. First, we give a bound on the GK dimension of a proper
quotient of the spherical subalgebra.

Lemma 4.39 (Stafford). Let κ ∈ C( . Let I be a nonzero two-sided ideal of Uκ . Then GK dim(Uκ /I) ≤
2n − 2.

Proof. Since Uκ |hreg ∼


= D(hreg )W is a simple ring, I|hreg = D(hreg )W . Hence, I contains a power
of δ. Now recall the Fourier map ψ from Theorem 1.40. This map ψ takes δ to an element
δ ∈ C[h∗ ] ⊂ Hκ . Considering ψ(I), we get that I contains some power of δ as well. Now with
respect to the filtration F on Uκ from Section 1.15, grUκ ∼
= C[h × h∗ ]W which is a finite module
N
over C[h]W ⊗ C[h∗ ]W . Hence gr(Uκ /I) is a finite module over C[h]W /(δM ) ⊗ C[h∗ ]W /(δ ) for some
M, N . We show that this algebra has Gelfand-Kirillov dimension ≤ 2n − 2. First
8 9 ( ) 8 9
C[h]W C[h∗ ]W C[h]W C[h∗ ]W
GK dim ⊗ ≤ GK dim + GK dim
(δM ) N
(δ ) (δM ) N
(δ )
by [47, Lemma 3.10]. Next, GK dim(C[h]W ) = GK dim(C[h∗ ]W ) = dim(h∗ /W ) by [47, Theo-
rem 4.5]. Since W is a finite group, this dimension is n. By [47, Proposition 3.15], we have
N
GK dim(C[h]W /(δM )) ≤ n−1 and similarly GK dim(C[h∗ ]W /(δ )) ≤ n−1. So GK dim(C[h]W /(δM )⊗
N
C[h∗ ]W /(δ )) ≤ 2n − 2. Applying [47, Lemma 4.3] now gives GK dim(gr(Uκ /I)) ≤ 2n − 2. But
by [47, Proposition 6.6], we have GK dim(Uκ /I) = GK dim(gr(Uκ /I)) ≤ 2n − 2 as required.

Lemma 4.40. [32] Let κ , κ$ ∈ C( be any two parameter tuples, and let R = {x ∈ Uκ! | xUκ ⊂ Uκ! }.
Suppose Uκ! has finite homological dimension. Then if Q is a nonzero reflexive left ideal of Uκ! with
QUκ ⊂ Q, then Q = R.

76
Proof. By definition, Q ⊂ R, so we may consider the Uκ! − Uκ –bimodule R/Q. Since Q 1= 0 and
D(hreg )W = Uκ |hreg = Uκ! |hreg is a simple ring, we have (R/Q)|hreg = 0.
We show that R/Q is a Noetherian right Uκ –module. It suffices to show that R is a Noetherian
right Uκ –module. Take the filtration F on D(hreg )W by order of differential operators. This
induces the filtration F on Uκ and Uκ! which was considered in Lemma 4.6 above. Then grF (R)
is a submodule of grF (Uκ! ), which is a Noetherian grF (Uκ )–module since grF (Uκ! ) = grF (Uκ ). So
!
R is a Noetherian Uκ –module and so is R/Q. This shows that R/Q = N i=1 vi Uκ for some vi .

Therefore, the fact that (R/Q)|hreg = 0 implies that R/Q is annihilated on the left by a sufficiently
high power of δ2( ∈ Uκ! . So I := annUκ! (R/Q) 1= 0.
Therefore, by Lemma 4.39, the GK dimension of Uκ! /I is less than or equal to 2n − 2. But R/Q
is a finitely generated Uκ! /I–module and so GKdim(R/Q) ≤ 2n − 2 by [47, Proposition 5.1].
Next Uκ! is a Cohen-Macaulay and Auslander-Gorenstein algebra by Proposition 1.19. By
assumption, Uκ! has finite homological dimension. Hence, by Definition 1.2, Uκ! is Auslander-
regular and Cohen-Macaulay. Furthermore, Uκ! is a prime ring since it is a domain. Now we may
apply Lemma 4.38 and conclude that R ⊂ Q∗∗ = Q. So R = Q.

Remark 4.41. Note that the hypothesis that Uκ! has finite homological dimension holds whenever
Hκ! eHκ! = Hκ! , since in this case, Hκ! is Morita equivalent to Uκ! , and Hκ! always has finite
homological dimension.

Now we are in a position to prove the commutativity result.

Theorem 4.42. Suppose Fκa , Fκ[a]


b , F b and F a are equivalences, and that U has finite homological
κ κ[b] κ

dimension. Then the following diagram commutes


b
Fκ[a]
Uκ[ab] − Mod " Uκ[a] − Mod

a
Fκ[b] Fκa

! Fκb !
Uκ[b] − Mod " Uκ − Mod

Or, more succintly, the Heckman-Opdam shift functors commute.

Proof. By Lemma 4.36, we have eHκ µa e ⊗ eHκ[a] µb e = eHκ µa eHκ[a] µb e and eHκ µb e ⊗ eHκ[b] µa e =

77
eHκ µb eHκ[b] µa e. So we show that we have an equality of bimodules

eHκ µa eHκ[a] µb e = eHκ µb eHκ[b] µa e.

By Lemma 4.36, both of these bimodules are projective on both sides, so they are both reflexive.
We show that they are both contained in Uκ . We have

eHκ µa eHκ[a] µb e ⊂ eHκ µa eHκ[a] e

= eHκ µa eµ−1
a H κ µa e by Proposition 4.1

= eHκ e−
a H κ µa e

⊂ eHκ e

Similarly, eHκ µb eHκ[b] µa e ⊂ eHκ e. So eHκ µa eHκ[a] µb e and eHκ µb eHκ[b] µa e are left ideals of eHκ e
which are both reflexive and both stable under right multiplication by eHκ[ab] e. Hence by Lemma
4.40, they are both equal to {x ∈ Uκ | xUκ[ab] ⊂ Uκ }.

Remark 4.43. Theorem 4.34 may be viewed as an analogue for the group G(!, 1, n) of [7, Theorem
4.8] for W = Sn , where the condition that KZ separates simples is analogous to the condition
c∈ 1
h + Z≥0 , while the condition that |c| is sufficiently large is analogous to the parameters κ being
asymptotic. It is known that there are many other situations in which Oκ and Oκ! are equivalent,
for κ 1= κ$ . Indeed, define a partial order on Irrep(W ) by λ <κ µ if and only if cλ − cµ ∈ Z>0 .
#
Suppose (q + 1) i<j (ui − uj ) 1= 0. Then [64, Theorem 5.5] states that, for τ ∈ Z( , Oκ and Oκ+τ are
equivalent if the orders <κ and <κ+τ are equal. It is not known whether these equivalences come
from Heckman-Opdam shift functors, although Rouquier conjectures that they do [64, Remark 5.6].
If this is the case, then we would expect a much more general result than Theorem 4.34 to hold.

4.2 The Boyarchenko-Gordon shift functors

Now we consider a second notion of shift functor. This comes from a homomorphism constructed
by Gordon [35], following Boyarchenko [10]. Write D(V ) for the ring of differential operators on
an affine variety V . The idea is to define a quiver Q∞ with a dimension vector ε, such that there
is an isomorphism ( )G
D(Rep(Q∞ , ε))
→ Uκ (4.2)

78
where G = G(ε) is the base-change group of the quiver Q, and Iκ is an ideal of D(Rep(Q∞ , ε))
which depends on the parameters κ. With this description of Uκ , it is easy to construct a functor
Uκ −Mod → Uκ! −Mod for some shifted values of the parameters κ$ . We refer to the functors defined
in this way as the Boyarchenko-Gordon shift functors. In [35, Section 4.5], it is asked whether the
Boyarchenko-Gordon and Heckman-Opdam shift functors coincide. Our aim is to investigate this
question.
The first step is to explain the isomorphism (4.2). This ismorphism comes from a map

D(Rep(Q∞ , ε))G → Uκ

called the radial part map. Such a map was constructed in [35], following Oblomkov [59]. In our
situation it is convenient to alter slightly the definition of the radial part map that was given in [35]
and [59]. We will therefore explain in full how to define this map.
Unfortunately, the radial part map is by definition a map D(Rep(Q∞ , ε))G → D(hreg )W , and it
is not clear that the image of our version of the radial part map will always be equal to Uκ (although
we conjecture that this is so). Our results will therefore depend on the hypothesis that the image
of the radial part map does coincide with Uκ (Hypothesis 4.48). We will prove the hypothesis holds
when n = 1, so that our results at least apply to the case of W a cyclic group.

4.2.1 Gordon’s construction

We now quote some of the main results of [35]. Choose once for all n, ! ≥ 1 and define some quivers
as follows.
Let Q be the quiver with ! vertices labelled 0, 1, . . . , ! − 1 with cyclic orientation 0 ← 1 ← · · · ←
(! − 1) ← 0. Let Q∞ be the quiver whose set of vertices is {0, 1, . . . ! − 1, ∞} and whose arrows
consist of the arrows of Q together with one extra arrow ∞ → 0. Let δ be the dimension vector
of Q with δi = 1 for all i, and let ε be the dimension vector of Q∞ with ε∞ = 1 and εi = nδi for
i 1= ∞. For example, if ! = 4 then

$1 ## $1 ##
""" ## """ ##
"" #% "" #%
Q = 0 '# 2, Q∞ = ∞ "0
'## 2
## " ## "
## """ # """
& " & "
3 3
and δi = 1, i = 0, 1, 2, 3, while ε∞ = 1, εi = n, i = 0, 1, 2, 3.

79
We consider the space RQ∞ := Rep(Q∞ , ε) of representations of Q∞ with dimension vector ε.
The points of this space have the form

(X0 , X1 , . . . , X(−1 , v)

where Xi are n×n matrices and v ∈ Cn . Here, Xj is considered to be attached to the arrow j → j+1
#(−1 #(−1
and v to the arrow ∞ → 0. We write G for the group GL(ε) = ( i=0 GLn × C∗ )/C∗ ∼ = i=0 GLn ,
which acts naturally on Rep(Q∞ , ε) by change of basis. Explicitly, for (g0 , . . . g(−1 ) ∈ G, we have

(g0 , . . . g(−1 ) · (X0 , X1 , . . . X(−1 , v) = (g0 X0 g1−1 , g1 X1 g2−1 , . . . , g(−1 X(−1 g0−1 , g0 v).

Contained within RQ∞ is the space Rep(Q, nδ) which also has a natural action of G which is the re-
striction of the action on RQ∞ . We identify Rep(Q, nδ) with the subspace {(X0 , X1 , . . . , X(−1 , 0)} ⊂
Rep(Q∞ , ε).
We may relate these representation spaces to the space h = Cn with its natural action of
W = G(!, 1, n) by defining a subset Ŝ := {(X, X, . . . , X) : X = diag(x1 , . . . xn ) for some xi ∈ C} ⊂
Rep(Q, nδ). Then Ŝ is isomorphic to h, and if we embed W in G by sending a permutation matrix
σ to (σ −1 , σ −1 , . . . , σ −1 ) and a diagonal matrix s to (1, s−1 , s−2 , . . . , s−((−1) ), then the embedding
Ŝ 3→ Rep(Q, nδ) is W –equivariant. We denote by S the image of hreg under this embedding. We
loosely write S = hreg .
Let T∆ denote the subgroup of G consisting of the elements (T, T, . . . , T ) where T is a diagonal
matrix. The action map π : G × hreg → G · hreg ⊂ Rep(Q, nδ) induces a map (G/T∆ ) × hreg →
Rep(Q, nδ). Here, G/T∆ denotes the set of right cosets of T∆ in G, which can be given the structure
of an affine variety by [27, Theorem 11.4.4].
The group W normalises T∆ . Hence, W acts on (G/T∆ ) × hreg via w · [gT∆ , x] = [gT∆ w−1 , wx].
Therefore, π induces a map

π : (G/T∆ ) ×W hreg → Rep(Q, nδ)

where (G/T∆ ) ×W hreg denotes the quotient space of (G/T∆ ) × hreg by the W –action.

Proposition 4.44. [35, Lemma 2.2] The image of π is an open set, which we denote Rep(Q, nδ)reg ,
and π induces an isomorphism


ω : (G/T∆ ) ×W hreg → Rep(Q, nδ)reg .

80
In [35, Section 2.5], this setup is extended to Rep(Q∞ , ε) = Rep(Q, nδ) × Cn as follows. Let
S∞ := S × (1, 1, . . . 1) ∼
= hreg , a subset of RQ∞ , and define

$
U∞ = {([gT∆ , x], v) ∈ ((G/T∆ ) ×W hreg ) × Cn : g0−1 v is a cyclic vector for diag(x)}.

Let U∞ = (ω × idCn )(U∞


$ ). The open subset U
∞ ⊂ RQ∞ is useful because of the following

proposition.

Proposition 4.45. [35, Lemma 2.5] The G–action on U∞ is free, and the projection U∞ → hreg /W
is a principal G–bundle.

4.2.2 The radial part map

We now explain how to define the radial part map, following an idea shown to the author by
Ginzburg. Recall that we wish to construct a map D(RQ∞ )G → D(hreg )W = D(hreg /W ). One
way of doing this is to start with a function f ∈ C[hreg ]W and extend f to Rep(Q, nδ)reg × Cn via

f ([gT∆ , x], v) := f (x)

for all g ∈ G and all v ∈ Cn . The function f makes sense1 because of Proposition 4.44. By
restriction to the open set U∞ , we may apply any differential operator D ∈ D(RQ∞ ) to f . Since
S∞ ∼
= hreg , we may therefore define a differential operator D on hreg by D(f ) = D(f )|S∞ . The map
D 5→ D induces a map D(RQ∞ )G → D(hreg /W ) which is clearly a homomorphism of algebras.
The image of this map does not yet depend on κ, so we twist by something that depends on κ. We
now describe this twist.
!(−1
Define Γi = κ(−i − κ(−i+1 − 1( , 1 ≤ i ≤ ! − 1 and let Γ0 = − i=1 Γi = −κ1 + ( .
(−1
Define a
function s on RQ∞ by

s(X0 , X1 , . . . , X(−1 , v) = v ∧ X0 X1 · · · X(−1 v ∧ · · · ∧ (X0 X1 · · · X(−1 )n−1 v,

so that s(X0 , X1 , · · · X(−1 , v) 1= 0 if and only if v is a cyclic vector for X0 X1 · · · X(−1 . For 1 ≤ i ≤ !−
1, we also define functions si by si (X0 , X1 , . . . X(−1 , v) = det(Xi Xi+1 · · · X(−1 )s(X0 , X1 , . . . , X(−1 , v).
Define
(−1
'
ζκ := sΓi i sκ00 +Γ0 −1 .
i=1
1
The symbol f should be read “f antlers”.

81
Then ζκ is not a well-defined function on RQ∞ , but for any point p ∈ S∞ , there exists a small
Euclidean neighbourhood of p such that there is a well-defined branch of ζκ in this neighbourhood.
Also, for any D ∈ D(RQ∞ ), the conjugation ζκ−1 ◦ D ◦ ζκ makes sense as a differential operator
#
on the open set of points where s (−1
i=1 si 1= 0. Note that since the Γi have been chosen to sum to

zero, ζκ may also be expressed as


n
' !i
Γj κ00 −1
ζκ = det(Xi ) j=1 s .
i=1

Definition 4.46. The radial part map associated to κ is the map

Rκ : D(RQ∞ )G → D(hreg )W

given for D ∈ D(RQ∞ )G and f ∈ C[hreg ]W by

Rκ (D)(f ) = ζκ−1 Dζκ (f )|S∞ .

It is useful to note that Rκ preserves the order filtration on differential operators. Here, the
order filtration on D(RQ∞ )G is by definition the filtration induced from the order filtration on
D(RQ∞ ), while the order filtration on D(hreg )W = D(hreg /W ) is the natural order filtration.

Lemma 4.47. If D ∈ D(RQ∞ )G is a differential operator of order n, then the order of Rκ (D) is
at most n.

Proof. We may ignore the conjugation by ζκ , since conjugation does not change the order of a
differential operator. Write R for the map D(RQ∞ )G → D(hreg )W defined by R(D)(f ) = D(f )|S∞ .
Let g ∈ C[hreg ]W and let D ∈ D(RQ∞ )G be a differential operator of order n. Then we must show
that [R(D), g] has order ≤ n − 1. By definition, for f ∈ C[hreg ]W , we have

[R(D), g](f ) = R(D)(gf ) − g · R(D)(f )

= D(g f )|S∞ − g |S∞ · D(f )|S∞

= [D, g ](f )|S∞

So [R(D), g] = R([D, g ]). By induction on the order of D, we get the result.

Now we introduce our hypothesis.

82
Hypothesis 4.48. The image of the map Rκ : D(RQ∞ )G → D(hreg )W is the spherical subalgebra
Uκ = eHκ e.

In [59, Theorem 2.5], Oblmokov has proved a similar statement to Hypothesis 4.48 for the quiver
Q rather than Q∞ . However, Oblomkov’s methods do not seem to work in our case.
We now present some evidence for believing Hypothesis 4.48 to be true.

Theorem 4.49. If ! = 1 then Hypothesis 4.48 holds.

Proof. In the ! = 1 case, the group G is GLn and RQ∞ = gln × Cn . Also, ζκ reduces to sκ00 −1 . In
the notation of [32], we have c = −κ00 so [32] shows that ImRκ = eHκ e.

Hypothesis 4.48 also holds in the n = 1 case. We give the argument for this in full, following
very closely [59, Proposition 3.3].

Theorem 4.50 (Oblomkov). If n = 1 then Hypothesis 4.48 holds.

Proof. In the n = 1 case, we have W = Z/!. The defining representation h is C, and hreg = C∗ .
The space RQ∞ is isomorphic to C(+1 , and G = (C∗ )( . We continue to write points of RQ∞ as
(X0 , X1 , . . . , X(−1 , v) ∈ C(+1 . We write xi for the coordinate function corresponding to Xi and ν for
the coordinate function corresponding to v. Recall from Example 2.6 that the rational Cherednik
algebra eHκ e is generated by three elements x, y, s, where s(x) = ε−1 x and s(y) = εy. We may
ignore the parameter κ00 and regard κ = (κ1 , . . . , κ(−1 ) as an element of C(−1 . With respect to the
standard filtration on eHκ e, we have gr(eHκ e) = C[h⊕ h∗ ]W = C[x, y]Z/( = C[x( , y ( , xy]. Therefore,
in order to show that ImRκ contains eHκ e, it suffices to show that x( e, y ( e and xye all lie in the
image. We do this by an explicit calculation.
→ D(C∗ )Z/( may be written explicitly as
∗ )#
The radial part map Rκ : D(C(+1 )(C

! #−1 ! i (−1
' !i
− Γj Γj Γ −1
Rκ (D)(f (x)) = x i=1 j=1 D( xi j=1
ν 0
· f )|S∞ .
i=1
#(−1
We have Rκ (F ) = F |S∞ for all F ∈ C[RQ∞ ]G , so Rκ ( i=0 xi ) = x( .
Now we show that y ( e lies in the image of Rκ . Via the Dunkl embedding, y ( is regarded as the
!(−1 !(−1 ij j
Dunkl operator ∇( , where ∇ = dx d
+ x1 i=1 !κi ei , where ei := 1( j=0 ε s . A calculation gives
ei ∇ = ( dx
d
+ x1 !κi+1 )ei+1 for all i, and so
d 1 d 1 d 1 d
e∇( e = ( + !κ1 )( + !κ2 ) · · · ( + !κ(−1 ) .
dx x dx x dx x dx

83
#(−1
This operator maps a monomial xr to r i=1 (!κ(−i +r −i)x
r−( . Now consider Rκ ( ∂x∂ 0 ∂x∂ 1 · · · ∂x∂#−1 ).
r/( r/( r/(
If f = xr then f = x0 x1 · · · x(−1 , so
8(−1 9@
∂ ∂ ∂ ∂ ∂ ∂ ' !i @
r/(+ j=1 Γj r/( Γ0 −1 @
Rκ ( ··· )(f ) = ζκ−1 ··· xi x0 ν @
∂x0 ∂x1 ∂x(−1 ∂x0 ∂x1 ∂x(−1 @
i=1 S∞
(−1
' i
&
= r/! (r/! + Γj )xr−(
i=1 j=1
(−1 i
1 ' &
= r (r + ! Γj )xr−( .
!(
i=1 j=1
!i
But by definition of Γj , ! j=1 Γj = !κ(−i − i, and so Rκ ( ∂x∂ 0 ∂x∂ 1 · · · ∂x∂#−1 ) = 1 (
(#
y e.
Finally, we show that xy lies in the image of Rκ . The element exye ∈ eHκ e corresponds under
the Dunkl embedding to the differential operator x dx
d
, which maps xr to rxr . In a similar way to
the above calculation, we have
&(−1 (−1
& i
& !i
∂ Γj +r/(−1 Γ0 −1
Rκ ( xi )(xr ) = ζκ−1 (r/! + Γj )xi xi j=1 ν |S∞
∂xi
i=0 i=0 j=1
(−1
& i
&
= (r/! + Γj )xr
i=0 j=1
(−1 &
& i
= rxr + Γj xr .
i=1 j=1
! !(−1 !i
Hence, Rκ ( (−1 ∂
i=0 xi ∂xi − i=1 j=1 Γj · 1) = xye as required. This shows that ImRκ ⊃ eHκ e.

To obtain the reverse inclusion, we follow Oblomkov’s argument. Take the filtration by order of
differential operators on D(RQ∞ ) and D(hreg )W . The map Rκ preserves this filtration by Lemma
4.47. The associated graded map
grRκ
−→ C[x±1 , y]Z/( = grD(hreg )W
∗ )#
grD(RQ∞ )G = C[x0 , . . . , x(−1 , y0 , . . . y(−1 , ν, ω](C

is induced by the map which maps xi to x, yi to 1


(y and ν, ω to 0. Note also that the Dunkl
embedding identifies gr(eHκ e) with C[x( , y ( , xy] ⊂ C[x±1 , y]Z/( . We now define an ideal Iκ of
D(RQ∞ )G as follows: Iκ is the 2–sided ideal generated by the elements xi ∂x

i
− xi−1 ∂x∂i−1 − Γi ,
1 ≤ i ≤ ! − 1, together with ν ∂ν

− (Γ0 − 1). Then by similar calculations to the above, we get that
Iκ ⊂ ker Rκ , and grIκ = ker grRκ , because
∗ #
' ' (−1 (−1
C[x0 , . . . , x(−1 , y0 , . . . , y(−1 , ν, ω](C ) ∼
= C[ xi , yi , x0 y0 ] −→ C[x( , y ( , xy].
({xi yi = xj yj }1≤i,j≤(−1 , νω = 0)
i=0 i=0

84
We obtain grImRκ = Im(grRκ ) = gr(eHκ e), which together with ImRκ ⊃ eHκ e, shows that
ImRκ = eHκ e as required.

As stated above, we have not been able to show that Hypothesis 4.48 holds for general !, n and
κ. For the rest of this chapter, we make the standing assumption that we have chosen !, n and κ
so that Hypothesis 4.48 holds.

Remark 4.51. In order to prove Hypothesis 4.48 in general, it would suffice to show that imRκ ⊂
Uκ . An associated graded argument like that at the end of Theorem 4.50 would then complete
the proof, using the fact that Gan has shown [29, Theorem 3] that the associated graded map of
Rκ (with respect to the order filtration) would be an isomorphism. The method of [59, Lemma
3.1] may be used to show that imRκ is contained in the localisation (Uκ )# n x#i , but this does not
i=1

appear to be enough.

4.2.3 A shift functor

Before we can construct the shift functor, we must calculate the kernel of the map Rκ . Since G
acts on C[RQ∞ ], we get an action of the Lie algebra g of G by differential operators on C[RQ∞ ]
d@
@
which we denote by g < X 5→ τ̂ (X). This action is given by f 5→ dt t=0
eXt · f . We get X(ζκ ) =
$ ! %
− (−1i=1 Γi Tr(Xi ) − (Γ0 + κ00 − 1)Tr(X0 ) ζκ . We define a character χκ of g by

(−1
&
χκ (X) = (Γ0 + κ00 − 1)Tr(X0 ) + Γi Tr(Xi ).
i=1

Then we get that (D(RQ∞ )(τ̂ + χκ )(g))G ⊂ ker Rκ . Because G is a reductive group, the functor
of taking G–invariants is exact. We write Rκ to denote the induced map
( )G
D(RQ∞ )
# eHκ e.
D(RQ∞ )(τ̂ + χκ )(g)

We wish to show that this map is an isomorphism. In order to show this, we require the concept of
the moment map µ : T ∗ RQ∞ → g∗ . This map, which appears in work of Crawley-Boevey [18] and
Holland [43], arises from the fact that the cotangent bundle T ∗ RQ∞ is a symplectic vector space
equipped with a Hamiltonian G–action, but we define µ via the following explicit formula: a point
of T ∗ RQ∞ may be regarded as a tuple of the form

(X0 , . . . , X(−1 , Y0 , . . . , Y(−1 , v, w)

85
where Xi , Yi are n × n matrices and v ∈ Cn , w ∈ (Cn )∗ . Then from [35, Section 2.6], we have

µ(X0 , . . . , X(−1 , Y0 , . . . , Y(−1 , v, w) =

(Y0 X0 − X(−1 Y(−1 + vw, Y1 X1 − X0 Y0 , . . . , Y(−1 X(−1 − X(−2 Y(−2 ).

Theorem 4.52. [35] Under Hypothesis 4.48, the map Rκ is an isomorphism of algebras.

Proof. By Hypothesis 4.48, Rκ is a surjection, so it suffices to show that ker(Rκ ) = (D(RQ∞ )(τ̂ +
χκ )(g))G . To show this, we use Gelfand-Kirillov dimension.
Recall that the Bernstein filtration on the Weyl algebra Ar is defined as the filtration with the
coordinate functions xi and derivations ∂i in degree 1. See [43, Section 2.1].
$ %G
The space D(RQD(RQ ∞)
∞ )(τ̂ +χκ )(g)
inherits a filtration from the Bernstein filtration on D(RQ∞ ),
$ %G
and we denote the associated graded algebra with respect to this filtration by grB D(RQD(RQ ∞)
∞ )(τ̂ +χκ )(g)
.
By [43, Proposition 2.4], we have
( )G
D(RQ∞ )
grB = C[µ−1 (0)]G ,
D(RQ∞ )(τ̂ + χκ )(g)

where µ is the moment map. (In fact, [43, Proposition 2.4] is stated with respect to the order
filtration. But, as remarked in [43, Section 2.1] and [10, Section 3.1], the proposition is still true
with the order filtration replaced by the Bernstein filtration). By [35, Theorem 2.6], C[µ−1 (0)]G is
$ %G
a domain of Gelfand-Kirillov dimension 2n. So D(RQD(RQ ∞)
∞ )(τ̂ +χκ )(g)
is also a domain of Gelfand-
Kirillov dimension 2n by [47, Proposition 6.6]. Furthermore, with respect to the filtration F of
Section 1.15, grUκ = C[h⊕h∗ ]W and so GK dim(Uκ ) = 2n as well. Therefore, Rκ is a surjection from
a domain of GK-dimension 2n to an algebra of GK-dimension 2n. It follows from [47, Proposition
3.15] that Rκ must be an isomorphism.

Given a vector space V with a G–action and a linear character θ of G, we denote by V G,θ the
space of (G, θ)–semiinvariants V G,θ = {x ∈ V : g(x) = θ(g)x for all g ∈ G}.
Following [35], given a character θ of G, we now consider the space
( )G,θ
D(RQ∞ )
Bκ :=
θ
D(RQ∞ )(τ̂ + χκ )(g)
of (G, θ)–semiinvariants.
We may consider θ to be an element of Z( , since the character group of G is isomorphic to
#
Z( via (θ0 , . . . , θ(−1 ) 5→ (g 5→ (−1
i=0 det(gi ) ). Assuming Hypothesis 4.48, Theorem ?? holds, and
θi

86
so the space Bκθ is clearly a right Uκ –module. Also , for X ∈ D(RQ∞ ) and X ∈ g, we have
$ %G
τ̂ (X)D = [τ̂ (X), D] + Dτ̂ (X), so Bκθ is a left D(RQ∞D(RQ ∞)
)(τ̂ +χκ −dθ)(g) –module. We calculate the
κ$ ∈ C( such that χκ − dθ = χκ! . From the definition of χκ we get

κ$00 + Γ$0 − 1 = κ00 + Γ0 − 1 − θ0

Γ$i = Γi − θi , i " 1.

!(−i
Using j=1 !Γj = !κi − (! − i) we obtain

(−1
&
κ$00 = κ00 − θi
i=0
(−i
&
κ$i = κi − θj , i"1 (4.3)
j=0

So Bκθ is a Uκ! − Uκ –bimodule.

Definition 4.53. We call the functor Uκ − Mod → Uκ! − Mod defined by Bκθ ⊗ − the Boyarchenko-
Gordon shift functor.

Remark 4.54. Note that of course the Boyarchenko-Gordon shift functor is only well-defined when
Hypothesis 4.48 holds.

4.3 Comparison of the shift functors

We now come to the main aim of this chapter. We wish to prove that under certain conditions the
Boyarchenko-Gordon shift functor coincides with one of the Heckman-Opdam shift functors defined
above. The idea of the proof is to modify the embedding Rκ to get a map from Bκθ to D(hreg )W
which is a map of Uκ! − Uκ –bimodules, and then to show that in certain circumstances the image
of this map is precisely the bimodule defining a Heckman-Opdam shift functor. We can only show
this for some values of κ and θ as explained below.
Given θ ∈ Z( and κ ∈ C( , the first step is to define a map Pκθ : D(RQ∞ )G,θ → D(hreg )W .
Since s(g · p) = det(g0 )s(p) and si (g · p) = det(gi )s(p) for all p ∈ RQ∞ and all g ∈ G, the function
#
ξ := (−1 θi θ0
i=1 si s is (G, −θ)–semiinvariant. Therefore, if D ∈ D(RQ∞ )G,θ then Dξ ∈ D(U )G ,
#(−1
where U is the open set of points where s · i=1 si 1= 0 (we must restrict to U because the

87
θi are allowed to be negative), so we may set Pκθ (D) := Rκ! (Dξ) ∈ D(hreg )W . If X ∈ g then
(τ̂ (X) + χκ (X))ξ ∈ ker Rκ! , so Pκθ induces a map

Pθκ : Bκθ → D(hreg )W .

Proposition 4.55. The map Pθκ is a map of Uκ! − Uκ –bimodules. That is, Rκ! (E)Pκθ (D)Rκ (F ) =
Pκθ (EDF ) for all E, F ∈ D(RQ∞ )G and all D ∈ D(RQ∞ )G,θ .

Proof. It is clear that Pθκ is a left Uκ! –module map, so it suffices to show that Rκ! (Dξ)Rκ (F ) =
Rκ! (DF ξ) for all D ∈ D(RQ∞ )G,θ and all F ∈ D(RQ∞ )G . But if f ∈ C[hreg ]W then

Rκ! (Dξ)(f ) = ζκ−1


! D(ζκ! ξf )|S∞

= ξ|S∞ (ζκ! ξ)−1 D(ζκ! ξf )|S∞

= ξ|S∞ ζκ−1 D(ζκ f )|S∞

= Rκ (ξD)(f )

Therefore, Rκ! (Dξ)Rκ (F ) = Rκ (ξD)Rκ (F ) = Rκ (ξDF ) = Rκ! (DF ξ) as required.

In order to compare the Boyarchenko-Gordon and Heckman-Opdam shift functors, we first


specialise to the case θ = −εa , where εa is the character of G with (εa )i = δai . In this case, the
shift (4.3) is

κ$00 = κ00 + 1

κ$i = κi + 1 1≤i≤!−a

κ$i = κi ! − a < i ≤ ! − 1.

This is the same as the shift κ 5→ κ[! − a] in the notation of Notation 4.14. We wish to show that
the map P−ε
κ
a
gives an isomorphism

P−ε
κ
a
: Bκ−εa → eµ−1
(−a Hκ e

# #
where µ(−a := ( ni=1 xi )(−a · i<j (x(i − x(j ) as in Proposition 4.1. Recall that we say κ is regular if
Hκ is simple. We first prove that P−ε
κ
a
is surjective in the regular case.
For a Uκ! − Uκ –bimdoule M , we write M ∗ for the dual with respect to the right Uκ –module
structure. The assigment M 5→ M ∗ = HomUκ (M, Uκ ) is clearly functorial. We have the following
lemma.

88
Lemma 4.56. For all κ ∈ C( ,

(imP−ε
κ )
a ∗∗
= (eµ−1 ∗∗
(−a Hκ e) .

Proof. If θ = −εi , i ≥ 1, then ξ = s−1


i while if θ = −ε0 then ξ = s−1 . So if D ∈ D(RQ∞ ) then
ξ −1 D ∈ D(RQ∞ ), and Pκ−εa (ξ −1 D) = Rκ (D) for all D ∈ D(RQ∞ )G , since Pκ (D) = Rκ (ξD) by
the proof of Proposition 4.55. Therefore, imP−ε
κ
a
⊃ Uκ . Taking duals with respect to the right
Uκ –module structure, we get (imP−ε
κ ) ⊂ {x ∈ Frac(Uκ ) : Uκ x ⊂ Uκ } = Uκ . So (imPκ ) is a
a ∗ −εa ∗

Uκ − Uκ[(−a] –bisubmodule of Uκ . But for any module X, X ∗ is a reflexive module by [55, Section
−1
5.1.7]. So by Lemma 4.40, (imP−ε
κ ) = {x ∈ Uκ : xUκ[(−a] ⊂ Uκ }. Similarly, (eµ(−a Hκ e) is a
a ∗ ∗

reflexive Uκ − Uκ[(−a] –bisubmdoule of Uκ . So (imP−ε −1 ∗ ∼


κ ) = (eµ(−a Hκ e) = (Bκ ) . This gives the
a ∗ −εa ∗

result.

Lemma 4.57. Suppose eµ−1


(−a Hκ e is a reflexive right Uκ –module. Then the image of the map

P−ε
κ
a
: Bκ−εa → D(hreg )W is contained in eµ−1
(−a Hκ e.

Proof. If eµ−1 −εa ∗∗ → eµ−1 H e. The map


(−a Hκ e is reflexive, then Lemma 4.56 yields a map (Bκ ) (−a κ

P−ε
κ
a
is the composition of this map with the natural map Bκ−εa → (Bκ−εa )∗∗ .

Now we can prove that the image of P−ε


κ
a
is eµ−1
(−a Hκ e for regular κ.

Theorem 4.58. Suppose κ ∈ C( is regular. Then

imPθκ = eµ−1
(−a Hκ e.

Proof. If κ is regular then Hκ is simple, hence Uκ is simple by [6, Lemma 4.1]. Since imP−ε
κ
a

contains Uκ , we get imP−ε reg )W . But imP−εa ⊂ eµ−1 H e by Lemma 4.57 (since
κ |hreg = D(h
a
κ (−a κ

eµ−1
(−a Hκ e is reflexive by Theorem 4.20). So
@
eµ−1 H e @
(−a κ @
@ = 0.
imP−ε
κ
a @
hreg

Now, we note that eµ−1


(−a Hκ e is a Noetherian left Uκ[(−a] –module. This follows by considering the

associated graded modules with respect to the filtration F on Hκ , as in Remark 4.16. Hence,
eµ−1
#−a Hκ e
−εa is a Noetherian left Uκ[(−a] –module, and so is finitely generated. If {vi } denotes a finite
imPκ
set of generators, then we may choose N ∈ N large enough such that vi δN = 0 for all i. Therefore,

89
( )
eµ−1
#−a Hκ e eµ−1
#−a Hκ e
some power of δ annihilates −εa on the right. So annMod−Uκ −εa 1= 0 and therefore,
imPκ imPκ
since Uκ is a simple ring, we get

eµ−1
(−a Hκ e
= 0,
imP−ε
κ
a

as required.

We may extend Theorem 4.58 from regular κ to a larger class of κ by considering the associated

κ . Consider the filtration by order of differential operators on D(RQ∞ ) and


graded map of P−ε a

D(hreg ). The filtration on D(RQ∞ ) induces a filtration on Bκθ for any θ, which we also refer to
as the filtration by order. It is clear from Lemma 4.47 that Rκ respects the filtration by order.
Therefore, so does the map Pθκ : Bκθ → D(hreg )W .
Now suppose that eµ−1
(−a Hκ e is a rexflexive right Uκ –module (this happens for example if

eµ−1
(−a Hκ e induces a Morita equivalence). Then by Lemma 4.57,there is a map Pκ
−εa : B −εa →
κ

eµ−1
(−a Hκ e.

Lemma 4.59. Suppose κ ∈ C( is chosen so that eµ−1


(−a Hκ e is a rexflexive right Uκ –module. Then

with respect to the order filtrations on Bκ−εa and eµ−1


(−a Hκ e, the map

grP−ε
κ
a
: grBκ−εa → gr(eµ−1
(−a Hκ e)

is independent of κ.

Proof. By [36, Lemma 6.8], gr(eµ−1 −1


(−a Hκ e) = eµ(−a C[h × h ]e, while grBκ
∗ −εa = C[µ−1 (0)]G,−εa

by [35, Lemma 4.1]. Furthermore, suppose D ∈ D(RQ∞ )G is a differential operator of order n.


Then for any f ∈ C[hreg ]W , we have

P−ε −1
κ (D)(f ) = ζκ[(−a] D(ξζκ[(−a] f )|S∞
a

−1
= ζκ[(−a] [D, ξζκ[(−a] ](f )|S∞ + ξ|S∞ D(f )|S∞ .

Modulo differential operators of order ≤ n − 1, this is just D 5→ (f 5→ ξ|S∞ D(f )|S∞ ), which is
independent of κ.

Theorem 4.60. Suppose κ ∈ C( is chosen so that eµ−1


(−a Hκ e is a reflexive right Uκ –module.Then

P−ε
κ
a
: Bκ−εa → eµ−1
(−a Hκ e

is a surjection of Uκ[(−a] − Uκ –bimodules.

90
Proof. Let µ ∈ C( be regular. Then grP−ε
κ
a
= grP−ε
µ
a
by Lemma 4.59. But grP−ε
µ
a
is surjective,
since P−ε
µ
a
is graded-surjective by Theorem 4.58 combined with [65, Corollary 4.5]. Therefore,
grP−ε
κ
a
is also surjective, so P−ε
κ
a
is surjective, as required.

We close this section by explaining how to extend Theorem 4.60 from the character −εa to an
arbitrary θ with θi ≤ 0 for all i. First of all, we consider the case θ = −εa −εb where 0 ≤ a, b ≤ !−1.

Lemma 4.61. Suppose eµ−1 −1


(−a Hκ[(−b] e and eµ(−b Hκ e both induce Morita equivalences. Then

imP−ε
κ
a −εb
= eµ−1 −1
(−a Hκ[(−b] eµ(−b Hκ e.

Proof. By the proof of Lemma 4.36, the multiplication map

m : eµ−1 −1 −1 −1
(−a Hκ[(−b] e ⊗ eµ(−b Hκ e → eµ(−a Hκ[(−b] eµ(−b Hκ e

is an isomorphism. This implies that eµ−1 −1


(−a Hκ[(−b] eµ(−b Hκ e is a reflexive module on both sides, so

we have

P−ε
κ
a −εb
: Bκ−εa −εb → (Bκ−εa −εb )∗∗ → (eµ−1 −1
(−a Hκ[(−b] eµ(−b Hκ e)
∗∗
= eµ−1 −1
(−a Hκ[(−b] eµ(−b Hκ e.

Therefore, imP−ε
κ
a −εb
⊂ eµ−1 −1
(−a Hκ[(−b] eµ(−b Hκ e. There is a commutative diagram

−εa
Bκ[(−b] ⊗ Bκ−εb " B −εa −εb
κ

−εa −εb −εa −εb


Pκ[#−b] ⊗Pκ Pκ

! !
m
eµ−1
(−a Hκ[(−b] e ⊗ eµ−1
(−b Hκ e
" eµ−1 Hκ[(−b] eµ−1 Hκ e
(−a (−b

−εa
where the map Bκ[(−b] ⊗ Bκ−εb → Bκ−εa −εb is the multiplication map induced by the multiplication
of D(RQ∞ ). The map P−ε −εb is surjective by Theorem 4.60. Also, m has been shown to
κ[(−b] ⊗ Pκ
a

be an isomorphism. Therefore, P−ε


κ
a −εb
is surjective, which shows that it is an isomorphism.
!N
Now suppose we are given an arbitrary θ ∈ Z( with θi ≤ 0 for all i. Write θ = k=1 −ε(−ik .

Define κ[θ] = κ[i1 i2 · · · iN ]. Let Fκθ denote the composition

−iN −i −i2 −i1


Fκθ := Fκ[i1 i2 ···iN−1 ]
◦ Fκ[i1N−1
i2 ···iN−2 ] ◦ · · · Fκ[i1 ] ◦ Fκ

−ik
which is a functor Uκ −Mod → Uκ[θ] −Mod. The functor Fκ[i1 ···ik−1 ]
is by definition eµ−1
ik Hκ[i1 ···ik−1 ] e⊗

−. Let Ak = eµ−1
ik Hκ[i1 ···ik−1 ] e, 1 ≤ k ≤ N and let Aκ = AN −1 AN −2 · · · A1 .
θ

91
Theorem 4.62. Suppose Ak induces an equivalence for each k. Then the image of

Pθκ : Bκθ → D(hreg )W

is Aθκ .

Proof. The argument is the same as the proof of Lemma 4.61.

4.3.1 A remark on Z–algebras

In this section, we give an interpretation of Theorem 4.62 in terms of Z–algebras, as promised in


the introduction. In the definition of Z–algebras, it is necessary to drop our assumption that all
algebras have to have an identity element.

Definition 4.63. [36, Section 5.1] A Z–algebra is an algebra (which is not assumed to have a
unit) of the form
"
B= Bij
i≥j≥0

where Bij Bjk ⊂ Bik for all i ≥ j ≥ k and Bij Bak = 0 if j 1= a. Furthermore, each of the subalgebras
Bii is required to have a unit element 1i such that 1i bij = bij 1j = bij for all bij ∈ Bij for any j.

A Z–algebra may be thought of as an algebra of infinite lower triangular matrices, with the
diagonal entries being the algebras Bii .

Example 4.64. If

"
S= Si
i=0

is a graded commutative ring, then there is a Z–algebra Ŝ associated to S, given by Ŝ = ⊕i≥j≥0 Ŝij
where Ŝij = Si−j for all i, j.

We consider some categories of modules over a Z–algebra B, following [36, Section 5.1]. First,

i=0 Mi such that


let B − grmod denote the category of Noetherian graded left B–modules M = ⊕∞
Bij Mj ⊂ Mi for all i, j, and Bik Mj = 0 if j 1= k. The homomorphisms from M to N are B–module
maps f such that f (Mk ) ⊂ Nk for all k.
Say M ∈ B − grmod is bounded if Mn = 0 for all but finitely many n. Denote by B − tors the
subcategory of B − grmod consisting of those M which are bounded. (Note that in [36], B − tors is
taken to be the subcategory of B − grmod consisting of those M which are the direct limit of their

92
bounded submodules. But in the Noetherian case, this is the same as being bounded, as remarked
in the proof of [10, Theorem 4.4]). Then B − tors is a Serre subcategory of B − grmod, and we
may form the quotient category
B − grmod
B − qgr := .
B − tors
#(−1
Let θ ∈ Z( with θi ≤ 0 for all i. Regard θ as a character of G = i=0 GLn . Recall the moment
map µ : T ∗ RQ∞ → g from Section 4.2. Consider the graded ring

" i
S := C[µ−1 (0)]G,θ .
i=0

Let Ŝ be the Z–algebra associated to this ring, in the notation of Example 4.64. We have the
following theorem. Both the theorem and its proof are modelled very closely on the main theorem
of [36].

i−j
Theorem 4.65. Suppose Hypothesis 4.48 holds and that Pθκ[θj ] is injective for all i ≥ j. Let κ ∈ C(
i−j
be regular and define a Z–algebra Bκ by (Bκ )ij = Bκ[θ
θ
j ] for i ≥ j ≥ 0. Define another Z–algebra

grBκ by (grBκ )ij = gr(Bκ )ij for all i ≥ j ≥ 0, where the associated graded space is taken with
respect to the order filtration. Then there are equivalences of categories

1. Uκ − mod ∼
= Bκ − qgr.

2. grBκ − qgr ∼
= Ŝ − qgr.

Proof. The proof is an application of results of [36]. By Theorem 4.62, Bκ is a Morita Z–algebra
in the sense of [36, Section 5.4]. Therefore, there is an equivalence of categories (Bκ )00 − mod →
Bκ − qgr by [36, Lemma 5.5]. The second part follows from the fact that grBκθ = C[µ−1 (0)]G,θ for
i−j
any θ, by [35, Section 4.1]. The maps induced by this isomorphism on the grBκθ are the natural
multiplications, by [36, Lemma 6.7]. So grBκ ∼
= Ŝ as Z–algebras.

Remark 4.66. The algebra S has a projective variety Proj(S) associated to it. By [58, 3.i],
Proj(S) is the Nakajima quiver variety Mθ with stability condition θ. By [58, 3.iii], there is a
natural map π : Mθ → µ−1 (0)//G, where µ−1 (0)//G denotes the affine space whose coordinate
ring is C[µ−1 (0)]G . By [18, Theorem 1.1], C[µ−1 (0)]G is isomorphic to C[h ⊕ h∗ ]W . Thus, there
is a map π : Mθ → (h ⊕ h∗ )/W . For some θ, this is known to be a resolution of singularities
(see [39, Proposition 7.2.7]). Furthermore, the category Ŝ − qgr may be identified with the category

93
coh − Proj(S) of coherent sheaves on Mθ , by [42, Exercise 5.9] 1 . Thus Theorem 4.65 gives a
relationship between the geometry of resolutions of (h ⊕ h∗ )/W and the Cherednik algebra.

4.3.2 Nakajima quiver varieties

In view of Theorem 4.65, we now wish to address the question of when the map Pθκ is injective.
We have already shown that, under Hypothesis 4.48, Pθκ has image Aθκ provided κ is regular and
θi ≤ 0 for all i. Therefore, if we can show that Pθκ is injective, then Pθκ will be an isomorphism of
Bκθ with Aθκ , and the two notions of shift functor will coincide.
In the previous section we have introduced the Nakajima quiver variety
$ %
G,θ i
Mθ := Proj ⊕∞
i=0 C[µ −1
(0)] .

We aim to prove the following theorem.

Theorem 4.67. Suppose θi < 0 for all i. Then

Pθκ : Bκθ → D(hreg )W

is injective.

In order to prove Theorem 4.67, we will require a series of lemmas. Let us first note that, by [38,
! !
Lemma 4.3], Mθ is smooth if (i=0 θi 1= 0 and θi + · · · θj 1= −m (i=0 θi for any 1 ≤ i < j ≤ ! − 1
and any −(n − 1) ≤ m ≤ (n − 1). In particular, if θi < 0 for all 0 ≤ i ≤ ! − 1, then Mθ is smooth.
Also, we see that if θi < 0 for all i, then both Theorem 4.67 and Theorem 4.62 apply, and for
such θ, Pθκ is an isomorphism.
By [46, Section 2], Mθ may be identified with the set of closed orbits of G on the set

i
µ−1 (0)ss −1 ∞ −1
θ := {x ∈ µ (0) : ∃f ∈ ∪i=1 C[µ (0)]
G,θ
with f (x) 1= 0}

of θ–semistable points of µ−1 (0). We may also consider the set of θ–stable points µ−1 (0)sθ of
µ−1 (0), which is the set of points x ∈ µ−1 (0)ss
θ such that the stabiliser of x is finite and there exists
i
i=1 C[µ (0)]
f ∈ ∪∞ −1 G,θ such that the orbit of x is a closed subset of {p ∈ µ−1 (0) : f (p) 1= 0}. Our

1
This requires that S be generated by S1 as an S0 –algebra. In our case, this is true because the proof of
Theorem 4.36 combined with [36, Lemma 6.7] implies that the multiplication map C[µ−1 (0)]G,−εa ⊗C[µ−1 (0)]G,−εb →
C[µ−1 (0)]G,−εa −εb is surjective for all a, b.

94
first goal is to show that the sets of stable and semistable points coincide. This result is based on
an argument of Cassens and Slodowy [14].
We may identify T ∗ RQ∞ with Rep(Q∞ , ε) where Q∞ is the double of the quiver Q∞ (that is,
the quiver obtained from Q∞ by adding an edge in the opposite direction for each edge of Q∞ ).
Similarly, we have the double Q of the cyclic quiver Q. For a representation A of Q∞ of dimension
! !(−1
vector α, we write θ(A) = ∞ i=0 αi θi , where θ∞ := −n j=0 θj . Then by [46, Proposition 3.1], we

have that a representation A of Q∞ is semistable if and only if θ(A) = 0 and θ(M ) ≥ 0 for all
M ⊂ A, while A is stable if and only if A is semistable, and θ(M ) > 0 for all 0 1= M ⊂ A with
M 1= A.

Lemma 4.68. If θi < 0 for all 0 ≤ i ≤ ! − 1, then µ−1 (0)ss


θ = µ (0)θ .
−1 s

Proof. Let V be a θ–semistable representation of Q∞ that lies in µ−1 (0). Suppose V is not stable.
Then there exists a submodule U ≤ V such that θ(U ) = 0 and U 1= 0, V . We define a representation
N of Q∞ by N = U if U∞ = 0 and N = V /U otherwise. In either case, θ(N ) = 0 and N∞ = 0, so
N = 0 by the choice of θ, a contradiction.

The variety µ−1 (0) is not irreducible. By [30, Theorem 3.3.3], it has a decomposition

µ−1 (0) = M0 ∪ · · · ∪ Mn

into irreducible components, each of which satisfies G(Mi ) ⊂ Mi (this follows from the explicit
description of Mi given in [30]). Our next aim is to show that the semistable points of µ−1 (0) are
all contained in a single Mi .

Lemma 4.69. Suppose Mθ is smooth. Then there is some i with 0 ≤ i ≤ n with µ−1 (0)ss
θ ⊂ Mi .

Proof. Since µ−1 (0)ss


θ ⊂ µ (0), we have µ (0)θ = ∪i=0 (µ (0)θ ∩ Mi ). Let π : µ (0)θ → M
−1 −1 ss n −1 ss −1 ss θ

be the quotient map. Since µ−1 (0)ss


θ = µ (0)θ , it follows from geometric invariant theory (see for
−1 s

example [57, Definition 0.6, Theorem 1.10]) that π takes closed G–invariant subsets of µ−1 (0)ss
θ to

closed subsets of Mθ . So we have


A
Mθ = π(µ−1 (0)ss
θ ∩ Mi ).
i

Since we have seen that Mθ is smooth, and Mθ is known to be connected by [38, Section 3.9], we
see that Mθ is irreducible. Hence there exists i with

Mθ = π(µ−1 (0)ss
θ ∩ Mi ).

95
Let x ∈ µ−1 (0)ss
θ . Then there exists y ∈ Mi ∩ µ (0)θ with π(x) = π(y). So by the standard
−1 ss

geometric invariant theory description of the quotient Mθ (see [46, Section 2]), we get Gx∩Gy 1= ∅.
But Gx and Gy are closed orbits in µ−1 (0)ss
θ , since x and y are stable by Lemma 4.68. Therefore,

Gx = Gy and so x ∈ G(Mi ) = Mi as required.

We spend the rest of this subsection proving Theorem 4.67. Let D ∈ D(RQ∞ )G,θ and suppose
that Pκθ (D) = 0. Then Rκ (Dξ) = 0. Considering the restriction of Dξ to V := U ∩ U∞ , we have

ζκ−1 Dξζκ ∈ ker(R : D(V )G → D(hreg /W )).

Therefore by [65, Corollary 4.5], we get

ζκ−1 Dξζκ ∈ (D(V )τ̂ (g))G .

Therefore, D ∈ D(RQ∞ )G,θ ∩ (D(V )(τ̂ + χκ )(g))G,θ . So it suffices to show that

(D(RQ∞ )(τ̂ + χκ )(g))G,θ = (D(RQ∞ ) ∩ D(V )(τ̂ + χκ )(g))G,θ .

In order to prove the equality, we consider the filtration by order on D(RQ∞ ), which induces a
filtration on D(V ), the ring of differential operators on an open subset of RQ∞ . Since D(RQ∞ )(τ̂ +
χκ )(g) is clearly contained in D(RQ∞ ) ∩ D(V )(τ̂ + χκ )(g), it is enough to show that

gr(D(RQ∞ )(τ̂ + χκ )(g))G,θ ⊃ gr(D(RQ∞ ) ∩ D(V )(τ̂ + χκ )(g))G,θ .

The right hand side is contained in gr(D(RQ∞ ))G,θ ∩ gr(D(V )(τ̂ + χκ )(g))G,θ , so it suffices to show
that
(C[T ∗ RQ∞ ]µ∗ (g))G,θ ⊃ (C[T ∗ RQ∞ ] ∩ C[T ∗ V ]µ∗ (g))G,θ . (4.4)

The right hand side of (4.4) consists of regular (G, θ)–semiinvariant functions on T ∗ RQ∞ which
vanish on T ∗ V ∩ µ−1 (0). The cotangent bundle T ∗ V of V is just the set of points (X, Y, i, j) ∈
Rep(Q∞ , ε) such that (X, i) ∈ V . In particular, T ∗ V ∩ µ−1 (0) contains the point (X, 0, i, 0) for
each (X, i) ∈ V , and hence is a nonempty open subset of µ−1 (0). If p ∈ T ∗ V , then since θi < 0 for
# −θi −θ0
all i, (−1
i=1 si s is a regular (G, θ)–semiinvariant function on T ∗ RQ∞ with s(p) 1= 0. Therefore,
by definition of µ−1 (0)ss
θ , we get T V ∩ µ (0) ⊂ µ (0)θ . By Lemma 4.69, there is an Mi with
∗ −1 −1 ss

µ−1 (0)ss
θ ⊂ Mi . So T V ∩ µ (0) is an open subset of the irreducible variety Mi and therefore
∗ −1

any (G, θ)–semiinvariant function that vanishes on T ∗ V ∩ µ−1 (0) must vanish on Mi and hence

96
on µ−1 (0)ss
θ as well. But every (G, θ)–semiinvariant function vanishes on µ (0) \ µ (0)θ by
−1 −1 ss

definition. So any such function vanishes on the whole of µ−1 (0). Now by [35, Theorem 2.6], the
ring
C[T ∗ RQ∞ ]
C[µ−1 (0)] =
C[T ∗ RQ∞ ]µ∗ (g)
is known to be reduced, and so it follows from Hilbert’s Nullstellensatz that any function that
vanishes on µ−1 (0) belongs to C[T ∗ RQ∞ ]µ∗ (g), as required. This verifies (4.4) and completes the
proof of Theorem 4.67.

97
Chapter 5

Diagonal coinvariants

The purpose of this chapter is to give an application of the results of the previous chapters. We
will prove a result in classical invariant theory by using the Cherednik algebra. The fact that this
is a purely commutative result indicates the usefulness of the Cherednik algebra.
Let W be a complex reflection group acting faithfully on a finite-dimensional complex vector
space h. Recall from Definition 4.29 that we may define the ring of coinvariants

C[h]
CW := .
,C[h]W
+-

It is known that CW is a finite–dimensional algebra isomorphic to CW as a W –module (see [45,


Theorem 24-1]). There is interest in analogues of this construction with the representation h ⊕ h∗
in place of h, see for example [40]. The ring

C[h ⊕ h∗ ]
DW :=
,C[h ⊕ h∗ ]W
+-

is called the ring of diagonal coinvariants of W . The ring DW has a natural grading with deg(h∗ ) = 1
and deg(h) = −1. The following result was conjectured by Haiman in [40] and proved in Gordon [34]:

Theorem 5.1. [34] Let W be a finite Coxeter group of rank n with Coxeter number h and sign
representation ε. Then there exists a W –stable quotient ring RW of DW with the properties:

1. dim(RW ) = (h + 1)n

2. RW is graded with Hilbert series t−hn/2 (1 + t + · · · + th )n

3. The image of C[h] in RW is C[h]/,C[h]W


+ -

98
4. The character χ of the W –module RW ⊗ ε is χ(w) = (h + 1)dim ker(1−w)

In [34], a method for generalising this result to the groups W = G(!, 1, n) was outlined. It is
the aim of the present chapter to carry this out, using the theory developed in the earlier chapters.
The following result will be proved:

Theorem 5.2. Let W = G(!, 1, n) where ! > 1 and let h be the reflection representation of W .
Then there exists a W –stable quotient ring SW of DW with the properties:

1. dim(SW ) = (!n + 1)n

2. SW is graded with Hilbert series t−n−(( 2 ) (1 + t + · · · + t(n )n


n

3. The image of C[h] in SW is C[h]/,C[h]W


+ -

4. The character χ of SW ⊗ ∧n h∗ as a W –module is χ(w) = (!n + 1)dim ker(1−w)

Theorem 5.1 is proved by obtaining RW as the associated graded module of a finite–dimensional


module over the rational Cherednik algebra of W . The properties of this module are derived by
studying the category O for the rational Cherednik algebra. This is also the method that will be
used to prove Theorem 5.2.
Choose κ00 , κ1 , . . . , κ(−1 ∈ C such that

!κ1 + !(n − 1)κ00 = −1 − !n

and KZκ separates simples. Such a choice is possible by Theorem 3.24. Consider the parameters
κ[1], where recall we put κ[1]00 = κ00 + 1, κ[1]1 = κ1 + 1 and κ[1]i = κi for i > 1. Then

!κ[1]1 + !(n − 1)κ[1]00 = −1.

Applying Theorem 3.3, we get that there is a one-dimensional module Lκ[1] (triv) in Oκ[1] , and as
a W –module, Lκ[1] (triv) is trivial, so eLκ[1] (triv) 1= 0. Recall the shift functor Sκ1 : Oκ[1] → Oκ
constructed in Section 4.1.5. Since !κ1 + !(n − 1)κ00 = −1 − !n, we see from Theorem 3.3 that the
unique finite-dimensional simple object in the category Oκ has dimension (1 + !n)n . So by Lemma
4.22, e−
1 Lκ (triv) 1= 0. Therefore, Theorem 4.34 tells us that Sκ is an equivalence. Therefore,
1

Sκ1 (Lκ[1] (triv)) is a finite-dimensional simple module in Oκ . But KZκ separates simples, so by
Theorem 3.7 we have
Sκ1 (Lκ[1] (triv)) = Lκ (triv).

99
5.1 A quotient ring of the diagonal coinvariants

We follow the proof of [34, Section 5] to obtain the desired ring SW of Theorem 5.2. Choose κ as
above, and write L = Lκ (triv). Since L = Sκ1 (Lκ[1] (triv)), we may write L as

L = Hκ e−
1 ⊗e− Hκ e− (eLκ[1] (triv))
ψ
(5.1)
1 1

where (−)ψ denotes twisting by the isomorphism eHκ[1] e → e− −


1 Hκ e1 coming from Proposition 4.1
#n #
(recall that e− −1
1 = µ1 eµ1 where µ1 = i=1 xi · i<j (xi − xj ) = δ). Write Λ for the one-dimensional
( (

eHκ[1] e–module eLκ[1] (triv). Consider the filtration F on Hκ with deg(h) = deg(h∗ ) = 1 and
deg(W ) = 0, and the associated graded module grL.

Lemma 5.3. There is a surjection of C[h ⊕ h∗ ] ∗ W –modules:

1 ⊗C[h⊕h∗ ]W grΛ # grL.


C[h ⊕ h∗ ] ∗ W e− ψ

Proof. This follows from taking the associated graded of equation (5.1) and applying [36, Lemma
6.7, Lemma 6.8].

The definition of the shift isomorphism from Proposition 4.1 implies that C[h ⊕ h∗ ]W
+ acts on

grΛψ by 0. Also, Ce−


1 is the representation ∧ h of W . Hence SW := grL ⊗ ∧ h is a quotient
n ∗ n

ring of DW . Most of the properties of SW listed in Theorem 5.2 are consequences of the following
theorem. Recall that DW is graded with deg(h) = −1 and deg(h∗ ) = 1 and that this grading is
W –stable. In general, if M is a Z–graded module and χk : w 5→ tr(w|Mk ) is the character of the
!
kth graded piece then the graded character of M is defined to be the formal power series i χi ti .

Theorem 5.4. The graded character of grL = SW ⊗ ∧n h∗ is

det |h∗ (1 − t(n+1 w)


w 5→ t−n−((2 )
n

det |h∗ (1 − tw)

Proof. Recall the element h ∈ Hκ from Chapter 1. By the proof of Lemma 3.3.4, z acts by 0 on
the trivial representation triv = ∧0 h∗ of W and hence h also acts by 0 on the trivial representation
triv. Hence, the eigenvalue of h on the subspace C[h]d ⊗ triv ⊂ M (triv) is d.
By Theorem 3.3, the representation Lκ (triv) of Hκ is isomorphic as a graded W –module to
⊗n
U(n+1 where U(n+1 = C[u]/(u(n+1 ), regarded as a representation of Z/!Z = ,s1 - via s1 (u) = ε−1 u,
and where Sn acts by permuting the factors of the tensor product. Since {ui |0 ≤ i ≤ !n} is a basis

100
of U(n+1 , we may define distinct basis elements ai by ai := u((i−1)+1 , 1 ≤ i ≤ n. Then the element
!
v := σ∈Sn sgn(σ)aσ(1) ⊗ aσ(2) ⊗ · · · ⊗ aσ(n) affords the representation ∧n h∗ of W , and lies in degree
! 6 7 6 7
n + ni=1 !(i − 1) = n + ! n2 . Hence, hv = (n + ! n2 )v.
6 7
We define h$ = h − n − ! n2 ∈ Hκ . We now calculate the graded character of L = Lκ (triv) with
respect to the h$ –eigenspaces1 . By [16, Thm 2.3], Lκ (triv) admits a BGG-resolution:

0 ← Lκ (triv) ← M (triv) ← M (h∗ ) ← M (∧2 h∗ ) ← · · · ← M (∧n h∗ ) ← 0.

Hence, in the Grothendieck group of O,


n
&
[Lκ (triv)] = (−1)i [M (∧i h∗ )]
i=0

Now, again by the proof of Lemma 3.3.4, z acts by −i(!n + 1) on ∧i h∗ , and hence the lowest
6 7
eigenvalue of h$ on M (∧i h∗ ) is i(!n + 1) − n − ! n2 . Therefore, the graded character of M (∧i h∗ ) is
n χ i ∗ (w)ti(#n+1) !
t−n−(( 2 ) ∧deth | ∗ (1−tw) . But det |h∗ (1−t(n+1 w) = i (−1)i χ∧i h∗ (w)ti((n+1) (this follows readily from
h

diagonalising w) which gives the graded character of Lκ (triv) with respect to the h$ –eigenspaces as

det |h∗ (1 − t(n+1 w)


w 5→ t−n−(( 2 )
n
.
det |h∗ (1 − tw)

By the definition of the diagonal coinvariant ring DW , there is a unique copy of the trivial rep-
resentation in DW , which lies in degree 0, and hence a unique copy of ∧n h∗ in SW ⊗ ∧n h∗ , and
hence a unique copy of ∧n h∗ in Lκ (triv) (since SW ⊗ ∧n h∗ = grL, which is isomorphic to L as a
W –module). The unique copy of ∧n h∗ in Lκ (triv) must be spanned by v. But h$ v = 0 and hence h$
must act by 0 on the element e−
1 ⊗ 1 ∈ L which affords the unique copy of ∧ h in L. But because
n ∗

the grading induced by h$ on L gives e−


1 ⊗ 1 degree 0 and x ∈ h degree 1, and y ∈ h degree −1,

we see that grL has the same graded character as Lκ (triv), which proves the theorem.

Proof of Theorem 5.2 (3)

This is similar to a proof in [34, Section 5]. Let us recall what it means for a finite-dimensional
graded algebra to satisfy Poincaré duality.
1
the point of the proof is that, while equation 5.1 says that L and Lκ (triv) are isomorphic as W –modules, we must
compare them as graded W –modules.

101
Definition 5.5. Let A = ⊕N
i=0 Ai be a finite-dimensional graded C–algebra with Ai = 0, i > N .

Then A is said to satisfy Poincaré duality if dim(AN ) = 1 and for all 0 ≤ r ≤ N , the multiplication
pairing
Ar × AN −r → AN

is nonsingular (that is, if x ∈ Ar and xAN −r = 0 then x = 0).

It is well-known (see for example [44]) that the ring of coinvariants A = C[h]/,C[h]W
+ - satisfies

Poincaré duality. Therefore the highest degree graded component of A, which by [45, Theorem 20-3]
!
lies in degree i (di − 1) where the di are the degrees of the fundamental invariants of W , is an ideal
of A which is contained in every nonzero ideal. This ideal is called the socle of A. In the case of
W = G(!, 1, n) the degrees are !, 2!, . . . !n by [11, Table 1], so the socle lies in degree 12 !n(n + 1) − n.
The image of C[h] in SW corresponds to the subspace C[h]e− W −
1 ⊗ Λ of L. If p ∈ C[h]+ e1 then
ψ

p ⊗ Λψ = e− − − − − ψ −
1 pe1 ⊗ Λ = e1 ⊗ e1 pe1 Λ = e1 ⊗ epe · Λ = 0.
ψ


+ annihilates e1 ⊗Λ . On the other hand, the quotient C[h]/,C[h]+ -
Thus the ideal generated by C[h]W ψ W

contains a unique (up to scalar) element of maximal degree 12 !n(n + 1) − n, say q. The space Cq

is the socle of C[h]/,C[h]W
+ -. We claim qe1 ⊗ Λ 1= 0. By the PBW theorem, any element of
ψ

Hκ can be written as a sum of terms of the form p− wp+ where p− ∈ C[h∗ ], p+ ∈ C[h] and
w ∈ W . Since p− and w do not increase degree, it would follow if qe−
1 ⊗ Λ were zero, then L
ψ

could have no subspace in degree 1


2 !n(n + 1) − n. But the Hilbert series of L has highest order
(n2 −n−( 21 n(n−1) 1
term t =t 2
(n(n+1)−n
. Thus qe− −
1 ⊗ Λ is non–zero and C[h]e1 ⊗ Λ is isomorphic to
ψ ψ

+ -)e1 ⊗ Λ . This proves Theorem 5.2 (3). $



(C[h]/,C[h]W ψ

In the case W = Bn = G(2, 1, n), Gordon has previously constructed a ring RW in [34], having
the same properties as SW (see Theorem 5.1). As mentioned above, our proof is modelled on [34],
and we can check that in the Bn case RW and SW coincide.

Corollary 5.6. If W = Bn then RW = SW .

Proof. Haiman, in [39, Conjecture 7.25], constructs a ring R̃W which is the quotient of DW by the
largest ideal J such that J ∩ (DW )ε = 0. He states that R̃W has the same Hilbert series as RW
and is isomorphic to RW as a W –module. He proves that these conditions imply that R̃W = RW .
But the ring SW also has the same Hilbert series as RW and is isomorphic to RW as a W –module.
Hence, R̃W = SW by the same proof. Therefore, SW = RW .

102
Remark 5.7. We close with a remark about the meaning of Theorem 5.2. Note that the diagonal
coinvariant ring DW is in fact a bigraded ring, with h in degree (1, 0) and h∗ in degree (0, 1). It is
an interesting problem to calculate the Frobenius series of DW and its quotient SW , that is, the
series,
&
dim((SW )(i,j) )pi q j
(i,j)∈Z2

where p and q are indeterminates. Theorem 5.2 would follow from a description of the Frobenius
series and bigraded character of SW analogous to that given in the ! = 1 case in [39, Theorem 4.2.5],
since the Hilbert series can be obtained from the Frobenius series by setting q = p−1 . It is not clear
how the Frobenius series of SW could be calculated using Cherednik algebra methods. Nevertheless,
the existence and nice combinatorial properties of the ring SW may be seen as evidence that an
analogue of the combinatorial part of Haiman’s n! theorem may hold for the groups G(!, 1, n).

103
Bibliography

[1] S. Ariki. On the semi-simplicity of the Hecke algebra of (Z/rZ)!Sn . J. Algebra, 169(1):216–225,
1994.

[2] S. Ariki and K. Koike. A Hecke algebra of (Z/rZ) ! Sn and construction of its irreducible
representations. Adv. Math., 106(2):216–243, 1994.

[3] S. Ariki and A. Mathas. The number of simple modules of the Hecke algebras of type G(r, 1, n).
Math. Z., 233(3):601–623, 2000.

[4] D. J. Benson. Polynomial invariants of finite groups, volume 190 of London Mathematical
Society Lecture Note Series. Cambridge University Press, Cambridge, 1993.

[5] Y. Berest and O. Chalykh. Quasi-invariants of complex reflection groups. In preparation.

[6] Y. Berest, P. Etingof, and V. Ginzburg. Cherednik algebras and differential operators on
quasi-invariants. Duke Math. J., 118(2):279–337, 2003.

[7] Y. Berest, P. Etingof, and V. Ginzburg. Finite-dimensional representations of rational Chered-


nik algebras. Int. Math. Res. Not., (19):1053–1088, 2003.

[8] J.-E. Björk. The Auslander condition on Noetherian rings. In Séminaire d’Algèbre Paul Dubreil
et Marie-Paul Malliavin, 39ème Année (Paris, 1987/1988), volume 1404 of Lecture Notes in
Math., pages 137–173. Springer, Berlin, 1989.

[9] A. Borel, P.-P. Grivel, B. Kaup, A. Haefliger, B. Malgrange, and F. Ehlers. Algebraic D-
modules, volume 2 of Perspectives in Mathematics. Academic Press Inc., Boston, MA, 1987.

[10] M. Boyarchenko. Quantization of minimal resolutions of Kleinian singularities. Adv. Math.,


to appear. arXiv:math/0505165.

104
[11] M. Broué, G. Malle, and R. Rouquier. Complex reflection groups, braid groups, Hecke algebras.
J. Reine Angew. Math., 500:127–190, 1998.

[12] K. A. Brown. Symplectic reflection algebras. In Proceedings of the All Ireland Algebra Days,
2001 (Belfast), number 50, pages 27–49, 2002.

[13] K. A. Brown and K. R. Goodearl. Lectures on algebraic quantum groups. Advanced Courses
in Mathematics. CRM Barcelona. Birkhäuser Verlag, Basel, 2002.

[14] Heiko Cassens and Peter Slodowy. On Kleinian singularities and quivers. In Singularities
(Oberwolfach, 1996), volume 162 of Progr. Math., pages 263–288. Birkhäuser, Basel, 1998.

[15] T. Chmutova. Representations of the rational Cherednik algebras of dihedral type. J. Algebra,
to appear. arXiv:math.RT/0405383.

[16] T. Chmutova and P. Etingof. On some representations of the rational Cherednik algebra.
Represent. Theory, 7:641–650 (electronic), 2003.

[17] E. Cline, B. Parshall, and L. Scott. Finite-dimensional algebras and highest weight categories.
J. Reine Angew. Math., 391:85–99, 1988.

[18] W. Crawley-Boevey. Decomposition of Marsden-Weinstein reductions for representations of


quivers. Compositio Math., 130(2):225–239, 2002.

[19] P. Deligne. Équations différentielles à points singuliers réguliers. Springer-Verlag, Berlin, 1970.
Lecture Notes in Mathematics, Vol. 163.

[20] C. Dezélée. Représentations de dimension finie de l’algèbre de Cherednik rationnelle. Bull.


Soc. Math. France, 131(4):465–482, 2003.

[21] R. Dipper, G. James, and A. Mathas. Cyclotomic q-Schur algebras. Math. Z., 229(3):385–416,
1998.

[22] C. F. Dunkl. Differential-difference operators associated to reflection groups. Trans. Amer.


Math. Soc., 311(1):167–183, 1989.

[23] C. F. Dunkl and E. M. Opdam. Dunkl operators for complex reflection groups. Proc. London
Math. Soc. (3), 86(1):70–108, 2003.

105
[24] D. Eisenbud. Commutative algebra with a view toward algebraic geometry, volume 150 of
Graduate Texts in Mathematics. Springer-Verlag, New York, 1995.

[25] K. Erdmann and D. K. Nakano. Representation type of Hecke algebras of type A. Trans.
Amer. Math. Soc., 354(1):275–285 (electronic), 2002.

[26] P. Etingof and V. Ginzburg. Symplectic reflection algebras, Calogero-Moser space, and de-
formed Harish-Chandra homomorphism. Invent. Math., 147(2):243–348, 2002.

[27] Walter Ferrer Santos and Alvaro Rittatore. Actions and invariants of algebraic groups, volume
269 of Pure and Applied Mathematics (Boca Raton). Chapman & Hall/CRC, Boca Raton, FL,
2005.

[28] P. J. Freyd. Abelian categories. Repr. Theory Appl. Categ., (3):1–190 (electronic), 2003.

[29] Wee Liang Gan. Chevalley restriction theorem for the cyclic quiver. Manuscripta Math.,
121(1):131–134, 2006.

[30] Wee Liang Gan and Victor Ginzburg. Almost-commuting variety, D-modules, and Cherednik
algebras. IMRP Int. Math. Res. Pap., pages 26439, 1–54, 2006. With an appendix by Ginzburg.

[31] V. Ginzburg. On primitive ideals. Selecta Math. (N.S.), 9(3):379–407, 2003.

[32] V. Ginzburg, I. Gordon, and J. T. Stafford. In preparation.

[33] V. Ginzburg, N. Guay, E. Opdam, and R. Rouquier. On the category O for rational Cherednik
algebras. Invent. Math., 154(3):617–651, 2003.

[34] I. Gordon. On the quotient ring by diagonal invariants. Invent. Math., 153(3):503–518, 2003.

[35] I. Gordon. A remark on rational Cherednik algebras and differential operators on the cyclic
quiver. Glasg. Math. J., 48(1):145–160, 2006.

[36] I. Gordon and J. T. Stafford. Rational Cherednik algebras and Hilbert schemes. Adv. Math.,
198(1):222–274, 2005.

[37] I. Gordon and J. T. Stafford. Rational Cherednik algebras and Hilbert schemes. II. Represen-
tations and sheaves. Duke Math. J., 132(1):73–135, 2006.

106
[38] Iain Gordon. Quiver varieties, category O for rational Cherednik algebras, and Hecke algebras,
2007. arXiv.org:math/0703150.

[39] M. Haiman. Combinatorics, symmetric functions, and Hilbert schemes. In Current develop-
ments in mathematics, 2002, pages 39–111. Int. Press, Somerville, MA, 2003.

[40] Mark D. Haiman. Conjectures on the quotient ring by diagonal invariants. J. Algebraic
Combin., 3(1):17–76, 1994.

[41] J. Harris. Algebraic geometry. A first course, volume 133 of Graduate Texts in Mathematics.
Springer-Verlag, New York, 1992.

[42] R. Hartshorne. Algebraic geometry. Springer-Verlag, New York, 1977. Graduate Texts in
Mathematics, No. 52.

[43] M. P. Holland. Quantization of the Marsden-Weinstein reduction for extended Dynkin quivers.
Ann. Sci. École Norm. Sup. (4), 32(6):813–834, 1999.

[44] R. Kane. Poincaré duality and the ring of coinvariants. Canad. Math. Bull., 37(1):82–88, 1994.

[45] R. Kane. Reflection groups and invariant theory. CMS Books in Mathematics/Ouvrages de
Mathématiques de la SMC, 5. Springer-Verlag, New York, 2001.

[46] A. D. King. Moduli of representations of finite-dimensional algebras. Quart. J. Math. Oxford


Ser. (2), 45(180):515–530, 1994.

[47] G. R. Krause and T. H. Lenagan. Growth of algebras and Gelfand-Kirillov dimension, vol-
ume 22 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI,
revised edition, 2000.

[48] J. Le Potier. Lectures on vector bundles, volume 54 of Cambridge Studies in Advanced Math-
ematics. Cambridge University Press, Cambridge, 1997. Translated by A. Maciocia.

[49] T. Levasseur. Relèvements d’opérateurs différentiels sur les anneaux d’invariants. In Opera-
tor algebras, unitary representations, enveloping algebras, and invariant theory (Paris, 1989),
volume 92 of Progr. Math., pages 449–470. Birkhäuser Boston, Boston, MA, 1990.

107
[50] T. Levasseur and J. T. Stafford. Invariant differential operators and an homomorphism of
Harish-Chandra. J. Amer. Math. Soc., 8(2):365–372, 1995.

[51] S. Lyle and A. Mathas. Blocks of affine and cyclotomic Hecke algebras.
arXiv:math.RT/0607451.

[52] A. Mathas. Simple modules of Ariki-Koike algebras. In Group representations: cohomology,


group actions and topology (Seattle, WA, 1996), volume 63 of Proc. Sympos. Pure Math., pages
383–396. Amer. Math. Soc., Providence, RI, 1998.

[53] A. Mathas. Iwahori-Hecke algebras and Schur algebras of the symmetric group, volume 15 of
University Lecture Series. American Mathematical Society, Providence, RI, 1999.

[54] A. Mathas. The representation theory of the Ariki-Koike and cyclotomic q-Schur algebras. In
Representation theory of algebraic groups and quantum groups, volume 40 of Adv. Stud. Pure
Math., pages 261–320. Math. Soc. Japan, Tokyo, 2004.

[55] J. C. McConnell and J. C. Robson. Noncommutative Noetherian rings. Pure and Applied
Mathematics (New York). John Wiley & Sons Ltd., Chichester, 1987. With the cooperation
of L. W. Small, A Wiley-Interscience Publication.

[56] R. V. Moody and A. Pianzola. Lie algebras with triangular decompositions. Canadian Math-
ematical Society Series of Monographs and Advanced Texts. John Wiley & Sons Inc., New
York, 1995. , A Wiley-Interscience Publication.

[57] David Mumford. Geometric invariant theory. Ergebnisse der Mathematik und ihrer Grenzge-
biete, Neue Folge, Band 34. Springer-Verlag, Berlin, 1965.

[58] Hiraku Nakajima. Quiver varieties and Kac-Moody algebras. Duke Math. J., 91(3):515–560,
1998.

[59] A. Oblomkov. Deformed Harish-Chandra homomorphism for the cyclic quiver.


arXiv:math/0504395.

[60] E. M. Opdam. Complex reflection groups and fake degrees. arxiv:math/9808026.

[61] D. S. Passman. Infinite crossed products, volume 135 of Pure and Applied Mathematics. Aca-
demic Press Inc., Boston, MA, 1989.

108
[62] N. Popescu and L. Popescu. Theory of categories. Martinus Nijhoff Publishers, The Hague,
1979.

[63] J. J. Rotman. An introduction to homological algebra, volume 85 of Pure and Applied Mathe-
matics. Academic Press Inc. [Harcourt Brace Jovanovich Publishers], New York, 1979.

[64] R. Rouquier. q-Schur algebras and complex reflection groups I. arXiv:math.RT/0509252.

[65] Gerald W. Schwarz. Lifting differential operators from orbit spaces. Ann. Sci. École Norm.
Sup. (4), 28(3):253–305, 1995.

[66] G. C. Shephard and J. A. Todd. Finite unitary reflection groups. Canadian J. Math., 6:274–
304, 1954.

[67] J. T. Stafford. Auslander-regular algebras and maximal orders. J. London Math. Soc. (2),
50(2):276–292, 1994.

[68] J. R. Stembridge. On the eigenvalues of representations of reflection groups and wreath prod-
ucts. Pacific J. Math., 140(2):353–396, 1989.

[69] R. Vale. On category O for the rational Cherednik algebra of G(m, 1, n): the almost semisimple
case. Submitted. arXiv.org:math/0606523.

[70] R. Vale. Rational Cherednik algebras and diagonal coinvariants of G(m, p, n). J. Algebra, to
appear. arXiv:math.RT/0505416.

[71] P. Vanhaecke. Integrable systems in the realm of algebraic geometry, volume 1638 of Lecture
Notes in Mathematics. Springer-Verlag, Berlin, 1996.

109

You might also like