April 1, 2003
1991 Mathematics Subject Classification. Primary 16D25, 16S38, 18E30.The second author is a director of research at the FWO
.
Ideal classes of three dimensional Sklyanin algebras
Koen De Naeghel and Michel Van den Bergh
Departement WNI, Limburgs Universitair Centrum, Universitaire Campus, Building D, 3590 Diepenbeek, Belgium E-mail address, K. De Naeghel: koen.denaeghel@luc.ac.be E-mail address, M. Van den Bergh: michel.vandenbergh@luc.ac.be
-
Abstract.
In this paper we classify graded reflexive ideals, up to isomorphism and shift, in certain three dimensional Artin-Schelter regular algebras. This classification is similar to the classification of right ideals in the first Weyl algebra, a problem that was completely settled recently. The situation we consider is substantially more complicated however.
Contents
1 Introduction
This paper is motivated by the recent developments on the classification of right ideals in the first Weyl algebra. We start by recalling the (for now) definite result in this subject, as it was formulated by Berest and Wilson.
Theorem 1.1.
[
8]
Let
be the first Weyl algebra.
Put
and let
be the set of isomorphism classes of right
-ideals. Then the orbits of the natural
-action on
are indexed by
, and the orbit corresponding to
is in natural bijection with the
'th Calogero-Moser space
|
(1.1)
|
where
acts by simultaneous conjugation on
.
The fact
has also been proved by Kouakou in his (unpublished) PhD-thesis [18] .
The first proof of Theorem 1.1 used the fact that there is a description of
in terms of the adelic Grassmanian (due to Cannings and Holland [12] ). Using methods from integrable systems, Wilson established a relation between the adelic Grassmanian and the Calogero-Moser spaces [36] .
In [7] Berest and Wilson gave a new proof of Theorem 1.1 using noncommutative algebraic geometry [1, 34] (some of the original proofs in [7] were slightly simplified by the second author). See also [6, 16] . That an approach based on noncommutative geometry should be possible was in fact anticipated very early by Le Bruyn who in [20] already came very close to proving Theorem 1.1 .
Let us briefly indicate how the methods of noncommutative algebraic geometry may be used to prove Theorem 1.1 . We introduce the homogenized Weyl algebra
and then we consider
as the coordinate ring of an open affine part of a noncommutative space
, with “homogeneous coordinate ring”
(see below for more precise definitions). The problem of describing
then becomes equivalent to describing certain objects on the “noncommutative projective plane”
. Objects on
have finite dimensional cohomology groups and these may be used to define moduli spaces, just as in the ordinary commutative case.
The current paper starts from the observation that there are many more noncommutative projective planes than just the one associated to the Weyl algebra (this is in fact a fairly degenerate one) [2, 3, 11] . So below we let
be a so-called “elliptic” quantum projective plane. By definition
is noncommutative projective scheme which has as homogeneous coordinate ring a graded ring
with generators
(in degree one) satisfying the relations
|
(1.2)
|
where
are generic scalars (see below).
Such a
is called an elliptic quantum plane because there is an inclusion (in a noncommutative geometry sense)
where
is a smooth (commutative) elliptic curve.
We let
be the set of reflexive graded
-ideals, considered up to isomorphism and shift of grading. We may think of the elements of
as line bundles on
. In this paper we prove the following result (see Theorem 5.5.5 below).
Theorem 1.2.
There exist smooth affine varieties
of dimension
such that
is naturally in bijection with
.
We would like to think of the
as elliptic Calogero-Moser spaces. We show below that
is a point and
is the complement of
under a natural embedding in
.
Remark 1.3.
In fact
is connected, which we will prove in a subsequent paper [
13]
.
Remark 1.4.
A theorem similar to Theorem 1.2 has been announced by [
24]
. They work in a more general setting where the associated automorphism
of
may have finite order.
The reader will notice that Theorem 1.2 is weaker than Theorem 1.1 but this is probably unavoidable. Although we have a fairly succinct description of the varieties
(see 5.9 below) it is not as explicit as 1.1 . And very likely
can also not be viewed in a natural way as the orbit of a group.
Our proof of Theorem 1.2 is similar in spirit to the proof of Theorem 1.1 . However it is substantially more involved. The reason for this is that the proofs for the Weyl algebra rely heavily on the fact that
contains a central element in degree one (namely
) and the lowest central element in
has degree three.
We also have a result which explicitly describes the elements of
. Recall that a line module over
is a graded
-module of the form
with
.
The following theorem can be deduced easily from Theorem 5.6.6 below.
Theorem 1.5.
Let
. Then there exists an
together with a monomorphism
such that there exists a filtration of reflexive graded
-ideals
with the property that the
are shifted line modules, up to finite length modules.
It seems plausible that this result may be used to obtain an analogue of the Cannings-Holland classification of ideals in the Weyl algebra (see [12] ) but we have not sorted out the details. We hope to come back on this in a subsequent paper.
2 Preliminaries
Throughout we work over an algebraically closed field
of characteristic zero. In this section we recall some basic notions of noncommutative projective geometry.
These are collected from [1, 22, 26, 27, 28, 33] . We use the following convention:
Convention 2.1.
If
denotes an abelian category then
denotes the full subcategory of
consisting of noetherian objects.
To simplify the notations we often use implicitly the following result
Lemma 2.2.
Assume that
is a locally noetherian category and
is the full subcategory of
consisting of noetherian objects. Then the natural map
is an equivalence of categories.
-
Proof.
This follows for example from the dual of [17,1.7.11] . □
2.1 Graded algebras and modules
Let
be a
-graded algebra. If
for all
we say that
is positively graded, and if in addition
we say that
is connected. Any graded connected Noetherian
-algebra
is locally finite, i.e.
, for all
.
We write
for the category of graded right
-modules with morphisms the
-module homomorphisms of degree zero. Let
be a graded right
-module.
We use the notation (for all
)
and
.
We say that
is left (resp. right) bounded if
(resp.
) for some
. For any integer
, define
as the graded
-module that is equal to
with its original
action, but which is graded by
. We refer to the functor
as the
-th shift functor.
Since
is an abelian category with enough injective objects we may define the functors
on
as the right derived functors of
. It is convenient to write (for
)
whence
are the right derived functors of
, for
.
Finally, recall that a module
is reflexive if
where
is the graded dual of
.
2.2 Tails
Let
be a Noetherian connected graded
-algebra. We denote by
the functor that sends a graded right
-module to the the sum of all its finite dimensional submodules.
Denote by
the full subcategory of
consisting of all modules such that
and write
for the quotient category
. We write
for the (exact) quotient functor. By localization theory [29]
has a right adjoint which we denote by
. It is well-known that
. The object
in
will be denoted by
and it is again easy to see that
. The objects in
will be denoted by script letters, like
.
The shift functor induces an automorphism
on
which we also call the shift functor (in analogy with algebraic geometry it should perhaps be called the “twist” functor).
When there is no possible confusion we write
instead of
and
. The context will make clear in which category we work.
If
then
is left exact, so we may define its right derived functors
. We also use the notation
and we set
.
Our Convention 2.1 fixes the meaning of
,
and
. It is easy to see that
consists of the finite dimensional graded
-modules.
Furthermore
.
If
is finitely generated and
is arbitrary we have
|
(2.1)
|
If
and
are both finitely generated, then ( 2.1 ) implies
explaining the word “tails”.
For
there is an exact sequence (see [1] , Proposition 7.2)
|
(2.2)
|
An object
is said to be reflexive if
for some reflexive
.
We say that
satisfies condition
if
for all
and all
.
In case
satisfies condition
then for every
the cokernel of the map
in the exact sequence ( 2.2 ) is right bounded. In particular, for
we have
for some
.
Every graded quotient of a polynomial ring satisfies condition
and so do most noncommutative algebras of importance. The condition is essential to get a theory for noncommutative schemes which resembles the commutative theory.
Proposition 2.2.1.
[
1]
Let
be a right Noetherian connected
-algebra satisfying condition
. Then
is finite dimensional for all
and all
.
2.3 Serre duality
It was shown in [38] that under reasonable hypotheses the category
satisfies a classical form of Serre duality. However we will need a stronger form of Serre duality introduced by Bondal and Kapranov in [10] . Let
be a
-linear
-finite triangulated category. By this we mean that for all
we have
. The category
is said to satisfy Bondal-Kapranov-Serre (BKS) duality if there is an auto-equivalence
together with for all
natural isomorphisms
(where
denotes the
-dual).
Let
be an abelian category. We say that
has finite global dimension if there exists an
such that
for all
and for all
. The minimal such
is called the global dimension of
.
In this section we assume that
is a connected graded noetherian ring over a
.
By
we denote the functor on graded vector spaces which sends
to
.
If we use notations which refer to the left structure of
then we adorn them with a superscript “
”.
We make the following additional assumptions on
-
(1)
satisfies
and the functor
has finite cohomological dimension.
-
(2)
satisfies
and the functor
has finite cohomological dimension.
-
(3)
has finite global dimension.
Note that if
has finite global dimension then so does
by 2.1 .
Put
. According to [32]
is a complex of bimodules with finitely generated cohomology on the left and on the right, which in addition has finite injective dimension, also on the left and on the right. We now have the following result
Theorem 2.3.1.
(Serre duality) The category
satisfies BKS-duality with Serre functor defined by
This result is certainly not unexpected but as far as we know a written proof does not exist in the literature. We prove a more general version of Theorem 2.3.1 in Appendix A .
2.4 Projective schemes
We use the definition of
(for a noncommutative algebra
) suggested by Artin and Zhang (see [1] ). Let
be a Noetherian graded
-algebra. We define the (polarized) projective scheme
of
as the triple
. In what follows we shall refer to the objects of
(resp.
) as the coherent (resp. quasicoherent) sheaves on
, even when
is not commutative, and we shall use the notation
,
. By analogy we sometimes write
.
The following definitions agree with the classical ones for projective schemes.
If
is be a quasicoherent sheaf on
, we define the cohomology groups of
by
We refer to the graded right
-modules
as the full cohomology modules of
.
Finally, we mention the cohomological dimension of
It is easy to prove that
2.5 The Grothendieck group, the Euler form and Hilbert series
In this subsection
will be a Noetherian connected graded
-algebra with finite global (homological) dimension. We recall some basic tools.
2.5.1 The Euler form
Let
be an
-finite
-linear abelian category of finite global dimension. We define
for
. It is clear that
defines a bilinear form
which we call the Euler form for
.
2.5.2 Hilbert series
The Hilbert series of
is the Laurent power series
This definition makes sense since
is right Noetherian. Let
. Given a resolution
we have
Since
is connected, left bounded graded right
-modules are projective if and only if they are free hence isomorphic to a sum of shifts of
. So if we write
we obtain the formula
|
(2.3)
|
where
is the characteristic polynomial of
, it is defined by
2.5.3 The Grothendieck group
and the rank function
Set
. If
is a coherent sheaf on
, we denote by
its image in
.
The shift functor on
induces an automorphism of
. Following [22] , we view
as a
-module with
acting as the shift functor
. Now
may be described in terms of the Hilbert series of
.
Theorem 2.5.1 ([22] , Theorem 2.3).
Let
be a Noetherian connected graded
-algebra of finite global dimension. Set
and let
. Then
and for each
, the isomorphism sends
to the characteristic polynomial
of
:
In particular,
is sent to
.
Let us assume that
is a domain generated in degree one, so it has a graded division ring of fractions (graded version of Goldie's Theorem; see [23,Ch.C,Cor.I.1.7] )
The degree zero component
of
is a division algebra which is called the function field of
.
is isomorphic to a skew Laurent extention
where
has degree one (see [23] , Chapter A, Corollary I.4.3). The rank of a finitely graded right
-module
is
This also defines an additive rank fuction on
and hence a homomorphism
also denoted by “
”. Obviously
and
.
2.6 Artin-Schelter regular algebras
Definition 2.6.1.
[
1]
A connected graded
-algebra
is called an Artin-Schelter regular algebra of dimension
if it has the following properties:
-
(i)
A has finite global dimension
;
-
(ii)
A has polynomial growth, that is, there exists positive real numbers
such that
for all positive integers
;
-
(iii)
A is Gorenstein, meaning there is an integer
such that
where
is called the Gorenstein parameter of
.
If
is commutative, then the condition (i) already implies that
is isomorphic to a polynomial ring
with some positive grading. If in this case the grading is standard then
.
The Gorenstein property determines the full cohomology groups of
.
Theorem 2.6.2.
[
1]
Let
be a Noetherian Artin-Schelter regular algebra of dimension
, and let
. Then
, and the full cohomology modules of
are given by
Let
be an Artin-Schelter regular algebra as in the previous Theorem and put
. It is easy to see that
satisfies the hypotheses for Theoreom 2.3.1 . In this case the Serre functor has a particularly simple form: indeed in [1] it is shown that
as left
-modules and in [32] it is proved that
as complexes of bimodules. Thus we also have that
as right
-modules. In other words
where
is some graded automorphism of
. The automorphism
of
passes to an automorphism
for which we also use the notation
.
We find the the following formula for the Serre functor on
.
From this we easily obtain:
Proposition 2.6.3.
One has
.
-
Proof.
As above put
. The inequality
follows directly from BKS-duality and the above discussion. The other inequality follows from Theorem 2.6.2 . □
2.7 Dimension and multiplicity
Let
be a noetherian Artin-Schelter regular algebra. If
then the Gelfand-Kirilov dimension
of
[19] can be computed as the order of the pole of
in
[3] (in particular it is an integer). If
then we define
as
. Clearly
is additive on short exact sequences of objects with
. We have
and furthermore
if and only if
. If
then we put
and we call this the multiplicity of
. If
and if
is a domain generated in degree one then it is easy to see that
.
If
then we put
and
,
.
An object in
or
is said to be pure if it contains no subobjects of strictly smaller dimension. It is critical if all non-trivial subobjects have the same multiplicity. It is easy to see that if
is pure or critical then so is
, and conversely if
is pure or critical then there exists a module
which has the corresponding property such that
.
2.8 Three dimensional Artin-Schelter regular algebras
There exists a complete classification for Artin-Schelter regular algebras of dimension three [2, 3, 30, 31] :
Theorem 2.8.1.
The Artin-Schelter regular algbras
of dimension three can be classified. They are all Noetherian domains with Hilbert series of a weighted polynomial ring
.
It is known that three dimensional Artin-Schelter regular algebras have all expected nice homological properties. For example they are both left and right noetherian domains.
In this paper we restrict ourselves to Koszul three dimensional Artin-Schelter regular algebra. These have three generators and three defining relations in degree two. The minimal resolution of
has the form
hence
and the Hilbert series of
is the same as that of the commutative polynomial algebra
with standard grading.
Such algebras are also referred to as quantum polynomial ring in three variables.
The corresponding
will be called a quantum projective plane and will be denoted by
.
So let
be a quantum polynomial ring in three variables. A linear module of dimension
over
is a cyclic
-module generated in degree zero with Hilbert series
. Linear modules of dimension one and two are respectively called point and line modules. The images of these objects in
will be called point and line objects respectively. Line and point modules are classified in [2, 3] .
Line modules are of the form
with
. Hence line modules correspond naturally to lines in
. To classify point modules we write the relations of
as
. Set
. We introduce auxiliary (commuting) variables
(for
) and for a monomial
we define the multilinearization of
as
as
. We extend this operation linearly to homogeneous polynomials in the variables
.
Let
denote the locus of common zeros of the
. It turns out that
is the graph of an automorphism
of
, the locus of zeros of the multihomogenized polynomial
. If
is not identically zero then
is a divisor of degree
in
. We then say that
is elliptic. Otherwise,
is all of
and we call
linear in this case.
The connection between
and point modules is as follows: let
be a point module where
. Put
with
. From the fact that
we deduce that
and hence
. This construction is reversible and defines a bijection between the closed points of
and the point modules over
. If
is the pointmodule corresponding to
then we have
.
Let
be the inclusion and put
. Associated to the geometric data
is a so-called “twisted” homogeneous coordinate ring
. This is a special case of a general construction in [1] . See also [4] . Denote the auto-equivalence
by
. For
put
and
. It is easy to see that
has a natural ring structure and
is a right
-module. A straightforward verification shows
In [3] it is shown that there is a surjective morphism
of graded
-algebras.
Its kernel is trivial in the linear case and it is generated by a regular normalizing element
of degree three in the elliptic case. All point modules are
-modules. In other words:
annihilates all point modules.
By analogy with the commutative case we may say that
contains
as a “closed” subscheme. Though the structure of
is somewhat obscure, that of
is well understood.
Indeed it follows from [1, 4] that the functor
defines an equivalence
. The inverse of this equivalence and its composition with
are both denoted by
For further properties of point modules and line modules over three dimensional quantum polynomial algebras we refer to [2, 3] .
We will frequently use the following result
Lemma 2.8.2.
Assume that we are in the elliptic case. Let
be such that
. Then
. If
has infinite order then
.
-
Proof.
Multiplication induces an isomorphism
for large
. Hence
. Furthermore
is a finite dimensional representation of
.
It is is shown in [3] that if
has infinite order then
is a simple ring. In particular it has no finite dimensional representations. Thus
. This implies
. □
In the sequel it will be useful to cast the relationship between the noncommutative graded ring
and the commutative scheme
into the language of noncommutative algebraic geometry exhibited in [26, 33] although we will use this language only in an intuitive way. Let
,
.
We define a map of noncommutative schemes
by
| |
We will call
the restriction of
to
.
is clearly an exact functor.
For the left derived functor of
we have:
Lemma 2.8.3.
If
then
-
Proof.
One shows first that the objects
where
is a finitely generated graded free
-module are acyclic for
in the sense of [14] . Then the lemma follows by replacing
by a resolution of finitely generated free
-modules. □
We easily obtain the following consequence:
Lemma 2.8.4.
Assume that we are in the elliptic case and let
. Then there are short exact sequences:
-
Proof.
Take
such that
. We may assume that
is given by a right bounded complex of graded projective
-modules. The lemma now follows by applying
to the long exact homology sequence associate to the short exact sequence of complexes
2.9 Three dimensional Sklyanin algebras
Below we are interested in Sklyanin algebras of dimension three which are elliptic Artin-Schelter regular algebras such that the corresponding elliptic curve
is smooth and the automorphism is a translation. More specificly, we are interested in the algebras
where
are the quadratic equations
|
(2.4)
|
and
where
The algebras
are elliptic quantum polynomial rings. They correspond to Artin-Schelter algebras of dimension three where, in the associated geometric data,
is a smooth elliptic curve and
is given by translation under the group law. We refer to [2] for the description of
and
. The regular normalizing element
of degree three turns out to be central in this case.
Put
. Combining the results in [32] with Theorem 2.3.1 we see that Serre duality for
takes a particularly simple form:
Theorem 2.9.1.
Let
. Then there are natural isomorphisms
Corollary 2.9.2.
Let
and let
be a point object corresponding to
. Then
where
is the point object corresponding to
.
3 Cohomology of rank one sheaves on a quantum projective plane
In this section,
will be a quantum polynomial ring in three variables, and
the associated quantum projective plane. As usual
.
We say that a graded right
-module
is torsion if
.
is called torsion-free if
contains no torsion submodule. This is the same as saying that
is pure three dimensional. We use the same terminology for objects in
.
The graded right ideals of
are, up to isomorphism, precisely the shifts of torsion-free rank one right
-modules.
A torsion-free rank one graded
-module
gives rise to a torsion-free coherent sheaf
on
of rank one. Conversely, every torsion-free
determines a torsion-free rank one graded
-module
.
Any shift
of a torsion-free rank one graded
-module
gives rise to a torsion-free rank one coherent sheaf
on
. Our first aim is to normalize this shift.
We will use the following natural basis for
.
Proposition 3.1.
Let
be a point module and
a line module over
. Denote the corresponding objects in
by
and
.
Then
is a
-module basis of
, which does not depend on the particular choice of
and
, and the action of the shift functor on that basis is
|
(3.1)
|
-
Proof.
It follows from Theorem 2.5.1 that the class in
of an object
depends only on the Hilbert series of
. Thus
and
are indeed independent of the particular choice of
and
.
Using a computation with Hilbert series we see that the images of
and
under the isomorphism
of Theorem 2.5.1
are respectively
,
,
. Furthermore the shift functor corresponds to multiplication by
. This easily yields what we want. □
From now on, we fix such a
-module basis
of
. For any coherent sheaf
on
we may write
where
is the rank of
.
It follows from ( 3.1 ) that we have
|
(3.2)
|
for all integers
.
Proposition 3.2.
-
(1)
Let
be a coherent sheaf on
of rank one, and write
. Then there is an unique shift
(namely
) and an integer
such that
Moreover,
.
-
(2)
Let
be a coherent sheaf on
of rank zero, and write
. Then
. If
, then
.
-
Proof.
For the first part, use ( 3.2 ). The uniqueness is easy to see.
For the second statement, take the image of
under the isomorphism
of Theorem 2.5.1 . Take
such that
. We obtain
for a suitable
. Multiplying both sides with
yields (see ( 2.3 ))
We find
and if
then
. □
We call the integer
appearing in Proposition 3.2 the “invariant” of
(or of the corresponding torsion-free rank one graded
-module
such that
). Note that two torsion-free rank one graded
-modules
have the same invariant if and only if
for
and for a fixed integer
.
We will call a torsion-free rank one coherent sheaf
on
normalized if
for an integer
. We will prove later that this
is actually positive.
We will call a torsion-free reflexive rank one sheaf on
a line bundle. Our aim is to classify line bundles on
up to shift. By the above discussion this is equivalent to classifying normalized line bundles up to isomorphism.
It is also easy to see that through the functors
and
classifying line bundles up to shift is equivalent to classifying reflexive torsion-free rank one graded
-modules, also up to shift.
We recall two elementary lemmas.
Lemma 3.3.
Let
be torsion-free coherent sheaves on
of rank one. Then every nonzero morphism in
is injective.
-
Proof.
and
are critical of the same dimension. It is well-known that this implies that any map between them must be injective [3] . □
Lemma 3.4.
Let
. Then
is reflexive if and only if
is torsion-free and
for all
of dimension zero.
-
Proof.
Assume that
is a reflexive coherent sheaf on
. By 2.1 we need to prove the corresponding statement for
. Thus assume that
is a reflexive
-module and
. Assume that there is a non-split exact sequence
|
(3.3)
|
By [3,Theorem4.1] one has
. Hence we obtain
and thus
. Thus the composition of
is an isomorphism, implying that the first map splits. This contradicts the non-triviality of the extention 3.3 .
For the other implication, let
be torsion-free and
for all
of dimension zero.
is pure and
since
is pure two dimensional. By [3,Corollary4.2] there is a canonical map
and
is the maximal submodule of
which has
. Hence
is injective, and we have an exact sequence
|
(3.4)
|
where
. Applying
on ( 3.4 ) yields
|
(3.5)
|
where
. Now
must be zero, otherwise
and since
the sequence ( 3.5 ) would split, which is impossible because
is pure three dimensional. Hence
and thus
is reflexive. □
Now we can partially compute the cohomology of line bundles on
. This computation is similar to the one for the homogenized Weyl algebra in [20] . However the computations for the homogenized Weyl algebra rely on the existense of a central element in degree one. So they do not apply in a straightforward way to the case we consider.
Theorem 3.5.
Let
be a rank one torsion-free coherent sheaf on
where
. Assume that
. Then
-
(1)
-
(2)
-
(3)
-
(4)
for
.
As a consequence,
is positive and nonzero.
If
is a line bundle then we have in addition:
and
.
-
Proof.
That
for
is part of Theorem 2.6.2 .
To prove the rest of the current theorem we first let
. Suppose
is a nonzero morphism in
. By lemma 3.3
is injective and from the exact sequence
|
(3.6)
|
we get
. Using Proposition 3.2 gives
, thus
and
. Hence by the discussion in § 2.7 together with Proposition 3.2 we obtain
. By lemma 3.4
.
This means that the exact sequence ( 3.6 ) splits hence
is not torsion-free. A contradiction. We conclude that
for
.
Second, let
. Serre duality (Theorem 2.9.1 ) yields
If
is a nonzero morphism in
then
is injective, and from the exact sequence
|
(3.7)
|
we get
where
and
.
By Proposition 3.2
but
implies
. This yields a contradiction.
Assume now
and
reflexive. By the same reasoning as above we obtain
and thus the dimension of
is zero. By lemma 3.4 it follows that 3.7 splits. But this contradicts the fact that
is torsion-free.
For the second part we use Theorem 2.6.2 to obtain
and from Proposition 3.1 we deduce
| |
| |
Combining these results yields
and
. Now we use ( 3.2 ) to obtain
| |
Finally, we combine the first two results of the theorem. If
(or
if
is reflexive) the first statement gives
| |
and comparing with the expression
completes the proof. □
Using Theorem 3.5 the torsion-free rank one graded
-modules having invariant zero are easy to determine.
Corollary 3.6.
Let
be a torsion-free coherent sheaf of rank one on
with invariant
. Then
for some integer
.
-
Proof.
If
then clearly
. Assume conversely
. We may assume that
is normalized. If
then by Theorem 3.5
. Since
we obtain
by contraposition. □
4 Restriction of coherent sheaves
In this section,
will be a Sklyanin algebra
as defined in § 2.9 . We recycle the notations of sections § 2.6 -§ 2.9 . In particular the symbols
have their usual meaning.
Note that
is a smooth elliptic curve. We fix a grouplaw on
. Then
is a translation by some element
.
The dimension of objects in
or
will be computed in
or
. The dimension of objects in
is the dimension of their support.
There is a group homomorphism
which as usual is also denoted by
.
Lemma 4.1.
We have
| |
| |
-
Proof.
This follows easily from lemma 2.8.3 □
According to [15,ExII.6.11] we have
. The projection
is given by the rank and the projection
is given the first Chern class. If
is a vector bundle on
then
. We also have for
:
.
There is a homomorphism
which assigns to a line bundle its degree. For simplicity we will denote the composition
also by
. If
is a line bundle then
. If
has finite length then
[15,Ex.6.12] . From lemma 4.1 we deduce that if
then
|
(4.2)
|
Lemma 4.2.
-
(1)
If
is pure two dimensional then
is pure one dimensional.
-
(2)
If
is pure one dimensional then
is pure two dimensional.
-
Proof.
The indecomposable objects in
are vector bundles and finite length objects. Using Riemann-Roch it is easy to see that if
then
. From this we deduce that if
is not in
then
. The lemma now easily follows. □
We deduce
Proposition 4.3.
-
(1)
If
is reflexive then
is a vector bundle on
and
for
.
-
(2)
If
is a line bundle then so is
.
-
(3)
If
is a line bundle then
is normalized if and only if
.
-
(4)
If
is a normalized line bundle with invariant
then
where “
” is the origin for the group law.
-
Proof.
-
(1)
We have
where
is reflexive. In particular
is torsion-free.
By lemma 2.8.3 it follows that
for
and
.
If
contains a nonzero submodule
of GK-dimension
then
represents an element of
which must be zero by lemma 3.4 (or rather its proof ). Thus
. This is impossible since
is torsion-free.
Hence
is pure of GK-dimension 2. By the previous lemma it follows that
is a vector bundle.
-
(2)
This follows from 4.1 .
-
(3)
This follows from 4.2 .
-
(4)
We have
. By lemma 4.1 we obtain
. Hence
. Now
and
are both divisors of degree zero which have the same sum for the group law. Hence they are linearly equivalent by [15,IVThm4.13B] . This finishes the proof. □
Now we prove a converse of Proposition 4.3 .
Proposition 4.4.
Assume that
has infinite order and that
is such that
is a vector bundle on
. Then
is a reflexive object in
.
-
Proof.
It follows from lemma 2.8.4 that
for
. Then it follows from lemma 2.8.2 that
and
, using lemma 2.8.4 again.
Pick an object
in
such that
. We may assume that
contains no subobject in
. By lemma 2.8.3 we have
. Thus
. Since
contains no subobject in
it follows that
is
-torsion free. Furthermore by lemma 4.2
is pure two dimensional. If
is the maximal submodule of
which is in
then since
we obtain that
is pure two dimensional.
We now claim that
is pure three dimensional. Let
be the maximal submodule of
of dimension
. Then
is pure three dimensional and in particular
-torsion-free. Hence we have a short exact sequence
By the purity of
it follows that
and hence
.
It follows from lemma 2.8.2 than
and hence
. This shows that
is pure.
Put
. Thus we obtain an exact sequence
By [3] we have
. Thus we have
.
So by the proof of lemma 4.2 ,
. Since
is a vector bundle by hypotheses it contains no finite dimensional subobjects and we obtain
. Thus
. Thus, in high degree, multiplication by
is an isomorphism on
. But then by lemma 2.8.2
. Hence
and thus
is reflexive. □
5 Elliptic quantumspaces
5.1 Generalities
Let
be a Sklyanin algebra
. We use again our standard notations as in the previous section.
We set
and
the algebra of endomorphisms of
. We consider the left exact functor
which takes coherent sheaves on
to right
-modules.
extends to a functor
on bounded derived categories
|
(5.1)
|
This is done as follows:
has enough injectives and this yields a functor
. Now
and
are noetherian abelian categories and this yields equivalences
and
(lemma 2.2 ). The functor 5.1 is obtained by composing with these equivalences.
In a similar way as in [9,Theorem6.2] one shows that
is an equivalence of derived categories. The inverse functor is given by
. For a non-negative integer
the equivalence restricts to an equivalence between
and
where
is the full subcategory with objects
and
the full subcategory with objects
The inverse equivalences between these categories are given by
and
.
Let
be the quiver
|
(5.2)
|
with relations
|
(5.3)
|
We write
for the category of representations of the quiver
(representations are always assumed to satisfy the relations 5.3 ). If
then we denote by
respectively the projective representation and the simple representation corresponding to
.
It is easy to see that
. Since the category
of representations of
is equivalent to the category of right
-modules we deduce
.
Let
and
. By functoriality, multiplication by
induces linear maps
and
(
). Hence
is determined by the following representation of
| |
For further use we note that the Euler form
is given by the following matrix
|
(5.4)
|
where
refers to the columns and
refers to the rows.
Let
be the full subquiver of
consisting of the vertices
and let
be the obvious restriction functor.
has a left adjoint which we denote by
. If
is the sum of the vertices of
then
.
Note that
.
The following was already observed by Le Bruyn in the case of the homogenized Weyl algebra.
Lemma 5.1.1.
If
is a normalized line bundle on
and
then
.
-
Proof.
This follows from an argument by Baer [5,Corollary7.2] . For the convenience of the reader we repeat this argument.
We say that two objects
,
in an abelian category are orthogonal (
) if
.
We have
,
. Thus
. In particular
and
where we have used Serre duality and Theorem 3.5 . We conclude by lemma 5.1.2 below. □
Lemma 5.1.2.
Let
. Then
if and only if
.
-
Proof.
First assume
. Put
and take a projective resolution
Applying
we get a projective resolution of
of the form
for some
where
. The fact that
(by adjointness) implies
and
.
To prove the converse let
. By adjointness we have a map
whose kernel
and cokernel
are direct sums of
. We have
and hence
. Thus
and
is surjective.
Applying
to the short exact sequence
and using
(by adjointness) yields
and hence
. Thus
is an isomorphism and we are done. □
Lemma 5.1.3.
Let
and put
.
corresponds to a point module
of
. Put
and
.
-
(1)
for all
and
. In particular
.
-
(2)
for all
and
is a shifted point module for all
. In particular
.
-
(3)
.
-
(4)
The representation of
corresponding to
is
| |
-
(5)
Denote the representation in the previous diagram also by
.
We have
.
-
Proof.
-
(1)
Since the
are all obtained from point modules, it suffices to treat the case
. We use lemma 2.8.3 and the discussion before that. We have
and hence
for
.
-
(2)
This is easy to check.
-
(3)
Use
.
-
(4)
This follows from the previous step.
-
(5)
According to lemma 5.1.2 we need
for
. This follows from the fact that we have
for
by lemma 3.4 . □
To simplify the discussion below we define
(for
) as the category in which the objects are the normalized line bundles on
with invariant
and the morphisms are the isomorphisms in
. Thus
is a groupoid. Note that we do not know yet if
. This question will be addressed below.
5.2
is non-empty
To simplify things we assume that
has infinite order. The following result is necessary for the dimension computations in § 5.5 .
Lemma 5.2.1.
The set
is not empty.
-
Proof.
Let
be a line object on
. Writing
as the cokernel of a map
we find by Theorem 2.6.2 that if
then
has dimension
.
By [3] there exist at most three line objects
such that
is a subobject of
and furthermore these three line objects contain in turn any other object contained in
.
Hence if
then we may pick an epimorphism
(a generic
will do). Put
. Using Proposition 3.1 we find
and hence
. It is easy to see that
is reflexive. Thus
. □
Below we will show that
is parametrized by an algebraic variety of dimension
. The amount of freedom in the construction exhibited in the proof of lemma 5.2.1 is less than or equal to
parameters, hence for
this construction can not possibly yield all elements of
. In § 5.6 we will exhibit a related construction which works for all
.
5.3 First description of
Let
be the image of
under the equivalence
.
Theorem 5.3.1.
Let
. There is an equivalence of categories
|
(5.5)
|
where
-
Proof.
First, let
be an object of
. By Proposition 4.3 we have
for a line bundle
of degree zero on
. Hence (for all
) we have
. Since
we obtain
where
is the corresponding point object on
. Writing
this means that
, proving that
and
. By BKS-duality (Theorem 2.9.1 ) we obtain
for some other point
determined by
(and determining
). Hence
.
Conversely, let
be an object of
. Thus (using Serre duality on
again)
for all
. Now
so we may compute
using the Euler form 5.4 on
. We obtain
. In other words
.
Put
. By the category equivalence between
and
we obtain
, giving (by adjointness)
. Since
is a smooth elliptic curve it is easy to see that this implies that
is a line bundle. Hence by Propositions 4.3 and 4.4 the same is true for
. □
5.4 Application
Using the material in the previous sections it is now easy to parametrize the line bundles on
with invariant one.
Theorem 5.4.1.
The representations in
are the representations
|
(5.6)
|
for some
-
Proof.
First let
.
is given by a representation as in 5.6 . Then the condition
for
implies
.
Conversely let
be as in 5.6 with
. Then we immediately have
for
. □
5.5 Second description of
Although the category
has a fairly elementary description, it is not so easy to handle. In particular the analogy with the Weyl algebra case is not obvious. We will now give another description of
which is more similar to the one used for the Weyl algebra. In particular it will follow that the isomorphism classes of objects in
are parametrized by smooth affine varieties of dimension
.
In general, let
be a quiver without oriented cycles and write
,
for respectively the set of vertices and edges of
. Let “
” be the standard scalar product on
:
.
Let
. A representation
of
is called
-semistable (resp. stable) if
and
(resp.
) for every non-trivial subrepresentation
of
.
We have
, canonically. It is a fundamental fact [25] that
is semistable for some
if and only there exists
such that
. The relation between
and
is such that the forms
and
are proportional.
Fix a dimension vector
and let
be the corresponding representation space, i.e.
where the maps
associate to an arrow its begin and end vertex. The isomorphism class of representations of dimension vector
are in one-one correspondence with the orbits of the group
acting on
by conjugation.
Associated to
there is a semi-invariant function
on
such that the set
|
(5.7)
|
coincides with
. In particular 5.7 is affine.
Lemma 5.5.1.
There exists
with
such that
-
(1)
for all
we have
where
and
-
(2)
if
then
.
-
Proof.
-
(1)
Pick a degree zero line bundle
on
which is not of the form
for
(where
,
are as in Proposition 4.3 ).
Let
. Then we have by adjointness
. By Proposition 4.3 we have
.
We conclude by Serre duality for
that
. Now put
and
. We obtain
.
What is
? By adjointness we have
.
An easy verification shows that
. Thus by Riemann-Roch and Serre duality
where
.
Put
. Thus
.
Replacing
with a projective resolution it is easy to see that
. It follows that
and
.
-
(2)
Put
for
. Then
.
Now
is a nonzero complex whose homology has finite length. It is easy to deduce from this
. Hence we are done □
We obtain the following consequence.
Lemma 5.5.2.
If
and
then
is
-semistable for
.
-
Proof.
This is a straightforward verification. □
Lemma 5.5.3.
Assume that
has infinite order. Let
be a representation of
of dimension vector
,
. If
for all
then
.
-
Proof.
Assume the lemma is false. Thus
. Then we may construct a surjective map
where
. We will consider
and
. Note that since
is smooth,
is the sum of its homology.
We have for
:
|
(5.8)
|
Now a simple computation shows that
. Furthermore we have
. Finally by Serre duality on
(see Theorem 2.9.1 ) we have
. We conclude that also
. It follows from 5.8 that
. Hence by lemmas 2.8.4 and 2.8.2 we deduce
and hence
which is a contradiction. □
We can now prove our main result.
Theorem 5.5.4.
Assume that
has infinite order. Let
be as in lemma 5.5.1 .
-
(1)
The functors
and
define inverse equivalences between
and the following category
-
(2)
The representations in
are
-stable for
.
-
Proof.
Below we use often implicitly the already proved equivalence
(Theorem 5.3.1 ).
Step 1.
. This follows from lemmas 5.1.1 and 5.5.1 .
Step 2.
Let
. If
is such that
then
. This follows easily by using the Euler form.
Step 3.
Let
and
. Then
. Both claims are similar so we only consider the first one. We have
hence
is
-semistable and
is obviously stable. So if
then there is an epimorphism
. By Step 2 we obtain
. But this contradicts the choice of
, finishing the argument.
Step 4.
. Let
. By Step 3 and lemma 5.5.3 we obtain
.
It remains to show that for
we have
. By lemma 5.1.3 we have
. Thus
and similarly
where we have used Step 3 again.
Step 5.
and
are inverses to each other. To prove this we only need to show
for
. This follows from lemma 5.1.1 .
Step 6.
Let
. Then
is
-stable. Put a filtration
on
such that
is
-stable.
With the same proof as Step 3 it follows that
for
. Assume
. Then by lemma 5.5.3 we have
. From the right exactness of
we deduce
. Hence
and thus
is stable. □
Below we will define some varieties. We take the classical viewpoint. So they are always reduced.
Let
as in lemma 5.5.1 . Let
and put
|
(5.9)
|
It is clear that
is a closed subset of
so in particular
is affine.
Put
.
Theorem 5.5.5.
The affine variety
is smooth of dimension
.
The isomorphism classes in
(and hence in
and
) are in natural bijection with the points in
.
-
Proof.
Since all representions in
are stable by Theorem 5.5.4 , all
-orbits on
are closed and so
is really the orbit space for the
action on
.
This proves that the isomorphism classes in
are in natural bijection with the points in
.
To prove that
is smooth it suffices to prove that
is smooth (this follows for example using the Luna slice theorem [21] ).
We first estimate the dimension of
. We write the equations of
in the usual form
. Given
-matrices
,
,
let
be obtained from
by replacing
by
(thus
is a
-matrix). Then
has the following alternative description:
By § 5.2
is non-empty. The triples satifying
are a dense open subset of
and hence they represent a variety of dimension
. Imposing that
should have corank
represents
independent conditions. So the irreducible components of
have dimension
.
Define
by
where
(as usual we assume the points of
to satisfy the relation imposed on
).
To extend
to a point in
we need to choose a basis in
.
Thus
is a principal
fiber bundle over
. In particular
is smooth if and only
is smooth and the irreducible components of
have dimension
. Note that by the description of
in Theorem 5.3.1 it follows that
is an open subset of
.
Let
. The stabilizer of
consists of scalars thus if we put
then we have inclusions
. Voigt in [35,Ch.2,§3.4] has shown that there is a natural inclusion
(Voigt actually obtains an isomorphism since he is not assuming his representation spaces to be reduced). Now
corresponds to some line bundle
on
and we have
. An easy computation shows
.
We have
and by Serre duality
.
Thus
.
Hence we obtain
.
Thus
is constant and hence
is smooth. We also obtain
.
The dimension of
is equal to
. This finishes the proof. □
5.6 Explicit construction of the elements in
For simplicity we assume throughout that
has infinite order.
In the discussion below we have to compute the cohomology of a line object.
Lemma 5.6.1.
Let
be a line object on
. Let
. Then
and
for
. Furthermore if
then the induced linear map
corresponds to
on
.
-
Proof.
That
follows by writing
as the cokernel of a map
and invoking Theorem 2.6.2 . That
follows by Serre duality (Theorem 2.9.1 ).
Using Theorem 2.6.2 we find
. Using Serre duality (Theorem 2.9.1 ) this translates into
. Dualizing yields that indeed
. That
acts in the indicated way follows by inspecting the appropriate commutative diagram. □
Corollary 5.6.2.
Let
be a line object on
. Then
corresponds to
where
is the representation of
given by
|
(5.10)
|
Since line objects on
are of the form
they are naturally parametrized by points in
.
Proposition 5.6.3.
Let
be a normalized line bundle on
with invariant
. Then the set of line objects
such that
is a curve of degree
in
. In particular this set is non-empty.
-
Proof.
Let
with
. Put
,
. Then
where
,
are the restrictions of
and
to
.
Assume that
is given by matrices
. Then an easy verification shows that
if and only if
. This is a homogeneous equation in
and we have to show that it is not identically zero, i.e. we have to show that there is at least one
such that
.
This follows from lemma 5.6.4 below. □
Lemma 5.6.4.
Let
be a normalized line bundle on
with invariant
and let
be a point object on
. Then, modulo zero dimensional objects, there exist at most
different line objects
such that
and such that
.
-
Proof.
We use induction on
. Writing
as the cokernel of a map
we deduce by Theorem 2.6.2 that
. So the case
is clear by Corollary 3.6 . Assume
. Let
be the different line objects (modulo zero dimensional objects) satisfying
and
. If
then we are done. So assume
. Let
be the kernel of a non-trivial map
. It is proved in [3] that there is some different point object
such that for all
:
. Let
be the kernel of a non-trivial map
. The subobjects of line objects are shifted line objects and hence the image of
in
is a shifted line object. We find by 3.1 :
with
. From this we deduce that the invariant of
is
.
Since
for some point
it follows by adjointness and by Proposition 4.3 that
. Hence the composition
maps
to
. We claim that for
this map must be nonzero. If not then there is a non-trivial map
and since
is also subobject of
it follows that
and
have a common subobject. But this is impossible since
and
are different modulo zero dimensional objects.
Hence
for
. Since the
are still different modulo zero dimensional objects, we obtain
and hence
. □
The following lemma shows how to reduce the invariant of a line bundle.
Lemma 5.6.5.
Let
be a normalized line bundle on
with invariant
. Then there exists a line object
on
such that
. If
is the middle term of a corresponding non-trivial extension and
then
is a normalized line bundle with invariant
. Furthermore
is a shifted line object.
-
Proof.
Using Serre duality we have
. Also using Serre duality we deduce
. Then a simple computation using the Euler form shows that
. Hence it follows from Proposition 5.6.3 that there exist
such that
.
Now let
be the middle term of a non-trivial extension of
by
.
Then we have
.
We claim that
is torsion-free. Assume this is not the case and let
be a maximal subobject of
of dimension
. So
. Since
is torsion-free we have
. So we may consider
as a subobject of
. Hence we obtain an extention
|
(5.11)
|
According to lemma 3.4 this extension is split. But this means that
is a subobject of
of dimension
, contradicting the maximality of
.
It follows from [3] that
. Thus
for some
by Proposition 3.2 . Hence
.
Let
. Then by lemma 3.4
is pure and furthermore we have
. It then follows easily using the methods of [3] that
is a shifted line object. This finishes the proof. □
We can now prove another main result.
Theorem 5.6.6.
Let
be a normalized line bundle on
. Then there exists an
together with a monomorphism
such that there exists a filtration of line bundles
on
with the property that the
are shifted line objects.
-
Proof.
This follows easily from the lemma 5.6.5 and Corollary 3.6 . □
Remark 5.6.7.
There is some freedom in choosing the line objects occurring in Theorem 5.6.6 . We may assume for example that they all map to the same point object.
A Serre duality for graded rings
In this section we prove that (a generalization of ) BKS-duality holds for graded rings. For the convenience of the reader we restate some definitions so that this appendix can be read independently of the rest of this paper.
Let
be a
-linear
finite triangulated category. By this we mean that for all
we have
. The category
is said to satisfy Bondal-Kapranov-Serre (BKS) duality if there is an autoequivalence
together with for all
natural isomorphisms
(where
denotes the
-dual).
Let
be an abelian category. An object
in
is said to have finite projective (injective) dimension if
(
) for
for some
. The minimal such
we call the projective (injective) dimension of
.
In this appendix we assume that
is a connected graded noetherian ring over a
. By
we denote the functor on graded vectorspaces which sends
to
. If we use notations which refer to the left structure of
then we adorn them with a superscript “
”.
We make the following additional assumptions on
:
-
(1)
satisfies
and the functor
has finite cohomogical dimension.
-
(2)
satisfies
and the functor
has finite cohomogical dimension.
These conditions imply that
has a balanced dualizing complex [37] given by
[32, 37] . Below we freely use the properties of such dualizing complexes.
We let
be the derived category of graded right
-modules.
will be the full subcategory of objects in
with finitely generated homology.
The category
is the full subcategory of
consisting of complexes with homology in
.
We let
(
) be the full category of
consisting of objects of finite projective (injective) dimension. The categories
and
are defined in a similar way. The fact that
has finite cohomological dimension implies
.
We will denote the functors
and
by
. Since they define a duality between
and
it is clear that they define a duality between
and
and between
and
.
It is also clear that these functors induce a duality between
and
. We denote these induced functors also by
. Again they define a duality between
and
and between
and
. Recall the following:
Lemma A.1.
Let
. Then there exists an object
such that
is a direct summand of
.
-
Proof.
This can be deduced from general results about compact objects in triangulated categories. For simplicity we give a direct proof based on a trick which the authors learned from Maxim Kontsevich. Take
arbitrary such that
. Take a quasi-isomorphism
where
is a right bounded complex of finitely generated projective modules. This yields a triangle:
where
. This triangle corresponds to an element of
which must be zero for large
. Hence
. This proves the lemma. □
We recall the following fact.
Proposition A.2.
The functors
and
induce inverse equivalences between
and
.
-
Proof.
If
then it is quasi-isomorphic to a bounded complex of finitely generated projective
-modules. For such such a complex it is clear that
has finite injective dimension. There is a canonical map
which is an isomorphism for
. By induction over triangles one shows that it is an isomorphism for all
.
Conversely assume
. Then by duality
.
By the above discussion
. Hence
.
We also find
. □
The functor
induces a functor
which we denote by
. Similarly the functor
induces a functor
which we denote by
.
Proposition A.3.
The functors
and
induces inverse equivalences between
and
.
-
Proof.
If
then by lemma A.1
is direct summand of some
with
.
Using the proof of the previous proposition this easily implies that
and
(essentially because we may reduce to
for some
).
Conversely assume
. Then
.
We have by definition
, and hence
. Then it follows from lemma A.1 that
is a direct summand of some
with
.
We easily deduce from this that
is a direct summand of
and hence
.
The proof now continous as the proof of the previous proposition. □
Theorem A.4.
(Serre duality) For all
,
there are natural isomorphisms
where
|
(A.1)
|
Furthermore the functor
defines an equivalence between
and
.
-
Proof.
As in [38] our proof of Serre duality is based on the local duality formula [32, 37] . The formulation of local duality in [32] used the functor
but the same proof works for the functor
where
. Furthermore it is possible to throw an extra perfect complex into the bargain. If we do this we obtain canonical isomorphisms
|
(A.2)
|
for
and
. By adjointness
. In addition, if we apply A.2 with
finite dimensional then we find
. Thus using lemma A.1 we obtain for
:
.
Now the standard triangle for local cohomology yields
and thus
. Using the fact that
is torsion we easily obtain from this:
where
is defined as in the statement of the theorem. So now we have shown
|
(A.3)
|
Now we obtain from lemma A.1 that
is a direct summand of a complex
with
. Thus A.3 is true for
and this finishes of the the first part of the theorem. The last part is Proposition A.3 . □
Corollary A.5.
If
has finite global dimension then
satisfies BKS-duality.
References
-
M. Artin and J. J. Zhang, Noncommutative projective schemes, Adv. in Math. 109 (1994), no. 2, 228–287.
-
M. Artin, J. Tate, and M. Van den Bergh, Some algebras associated to automorphisms of elliptic curves, The Grothendieck Festschrift, vol. 1, Birkhäuser, 1990, pp. 33–85.
-
, Modules over regular algebras of dimension 3, Invent. Math. 106 (1991), 335–388.
-
M. Artin and M. Van den Bergh, Twisted homogeneous coordinate rings, J. Algebra 133 (1990), 249–271.
-
D. Baer, Tilting sheaves in representation theory of algebras, Manuscripta Math. 60 (1988), no. 3, 323–347.
-
V. Baranovsky, V. Ginzburg, and A. Kuznetsov, Quiver varieties and a non-commutative
, Compositio Math. 134 (2002), no. 3, 283–318.
-
Y. Berest and G. Wilson, Ideal classes of the Weyl algebra and noncommutative projective geometry (with an appendix by Michel Van den Bergh), Int. Math. Res. Not. (2002), no. 26, 1347–1396.
-
, Automorphisms and ideals of the Weyl algebra, Math. Ann. 318 (2000), no. 1, 127–147.
-
A. I. Bondal, Helices, representions of quivers and Koszul algebras, Helices and Vector bundles, 75–95, London Math. Soc. Lecture Note Ser., 148, Cambridge Univ. Press, Cambridge, (1990).
-
A. I. Bondal and M. M. Kapranov, Representable functors, Serre functors, and reconstructions, Izv. Akad. Nauk SSSR Ser. Mat. 53 (1989), no. 6, 1183–1205, 1337.
-
A. I. Bondal and A. E. Polishchuk, Homological properties of associative algebras: the method of helices, Russian Acad. Sci. Izv. Math 42 (1994), 219–260.
-
R. C. Cannings and M. P. Holland, Right ideals of rings of differential operators, J. Algebra 167 (1994), no. 1, 116–141.
-
K. De Naeghel and M. Van den Bergh, Ideal classes of three dimensional Sklyanin algebras 2, in preparation.
-
R. Hartshorne, Residues and duality, Lecture notes in mathematics, vol. 20, Springer Verlag, Berlin, 1966.
-
, Algebraic geometry, Springer-Verlag, 1977.
-
A. Kapustin, A. Kuznetsov, and D. Orlov, Noncommutative instantons and twistor transform, Comm. Math. Phys. 221 (2001), no. 2, 385–432.
-
M. Kashiwara and P. Schapira, Sheaves on manifolds, Die Grundlehren der Mathematischen Wissenschaften, vol. 292, Springer Verlag, 1994.
-
K. M. Kouakou, Isomorphismes entre algèbres d'opérateurs différentiels sur les courbes algébriques affines, Ph.D. thesis, Université Claude Bernard-Lyon 1, 1994.
-
G. R. Krause and T. H. Lenagan, Growth of algebras and Gelfand-Kirillov dimension, Research Notes in Mathematics, vol. 116, Pitman, Boston, 1985.
-
L. Le Bruyn, Moduli spaces for right ideals of the Weyl algebra, J. Algebra 172 (1995), 32–48.
-
D. Luna, Slices étales, Bull. Soc. Math. France 33 (1973), 81–105.
-
I. Mori and S. P. Smith, Bézout's theorem for non-commutative projective spaces, J. Pure Appl. Algebra 157 (2001), no. 2-3, 279–299.
-
C. Nastasescu and F. Van Oystaeyen, Graded and filtered rings and modules, Springer, Berlin, 1979.
-
T. A. Nevins and J. T. Stafford, Sklyanin algebras and Hilbert schemes of points, to appear.
-
A. Schofield, Semi-invariants of quivers, J. London Math. Soc. (2) 43 (1991), no. 3, 385–395.
-
S. P. Smith, Subspaces of non-commutative spaces, Trans. Amer. Math. Soc. 354 (2002), no. 6, 2131–2171 (electronic).
-
J. T. Stafford and M. Van den Bergh, Noncommutative curves and noncommutative surfaces, Bull. Amer. Math. Soc. (N.S.) 38 (2001), no. 2, 171–216 (electronic).
-
S. T. Stafford, Noncommutative projective geometry (ICM '2002 talk).
-
B. Stenström, Rings of quotients, Die Grundlehren der Mathematischen Wissenschaften in Einzeldarstellungen, vol. 217, Springer Verlag, Berlin, 1975.
-
D. R. Stephenson, Artin-Schelter regular algebras of global dimension three, J. Algebra 183 (1996), 55–73.
-
, Algebras associated to elliptic curves, Trans. Amer. Math. Soc. 349 (1997), 2317–2340.
-
M. Van den Bergh, Existence theorems for dualizing complexes over non-commutative graded and filtered rings, J. Algebra (1997), 662–679.
-
, Blowing up of non-commutative smooth surfaces, Mem. Amer. Math. Soc. 154 (2001), no. 734, x+140.
-
F. Van Oystaeyen and L. Willaert, Grothendieck topology, coherent sheaves and Serre's theorem for schematic algebras, J. Pure Appl. Algebra 104 (1995), no. 1, 109–122.
-
D. Voigt, Induzierte Darstellungen in der Theorie der endlichen, algebraischen Gruppen, Springer-Verlag, Berlin, 1977, Mit einer englischen Einführung, Lecture Notes in Mathematics, Vol. 592.
-
G. Wilson, Collisions of Calogero-Moser particles and an adelic Grassmannian, Invent. Math. 133 (1998), no. 1, 1–41, With an appendix by I. G. Macdonald.
-
A. Yekutieli, Dualizing complexes over noncommutative graded algebras, J. Algebra 153 (1992), 41–84.
-
A. Yekutieli and J. J. Zhang, Serre duality for non-commutative projective schemes, Proc. Amer. Math. Soc. (1997), no. 125, 697–707.
Departement WNI, Limburgs Universitair Centrum, Universitaire Campus, Building D, 3590 Diepenbeek, Belgium E-mail address, K. De Naeghel: koen.denaeghel@luc.ac.be E-mail address, M. Van den Bergh: michel.vandenbergh@luc.ac.be