Twists of representations of complex reflection groups and rational Cherednik algebras
Abstract
Drinfeld twists, and the twists of Giaquinto and Zhang, allow for algebras and their modules to be deformed by a cocycle. We prove general results about cocycle twists of algebra factorisations and induced representations and apply them to reflection groups and rational Cherednik algebras. In particular, we describe how a twist acts on characters of Coxeter groups of type and and relate them to characters of mystic reflection groups. This is used to characterise twists of standard modules of rational Cherednik algebras as standard modules for certain braided Cherednik algebras. We introduce the coinvariant algebra of a mystic reflection group and use a twist to show that an analogue of Chevalley’s theorem holds for these noncommutative algebras. We also discuss several cases where the negative braided Cherednik algebras are, and are not, isomorphic to rational Cherednik algebras.
1 Introduction
The main theme of the present paper is the use of methods from quantum algebra to achieve results in representation theory. Placing an algebra and its representations in a category of modules over a quasitriangular Hopf algebra, or more generally a Hopf algebra equipped with a -cocycle, leads to a deformation — known as cocycle twist — of the associative product on , as well as of the representations of . This can be used to produce new representation-theoretic constructions and to uncover properties of known algebras and representations by realising them as twists.
A family of examples of cocycle twists arises from our previous paper [2], joint with Berenstein and McGaw. In [2] we demonstrate an isomorphism between braided Cherednik algebras, constructed earlier, and cocycle twists of rational Cherednik algebras of imprimitive complex reflection groups with even . These algebras have an action of the elementary abelian -group whose group algebra carries a non-standard, cohomologically non-trivial quasitriangular structure which is the cocycle we use. A strong property of a rational Cherednik algebra is its PBW-type factorisation, into three subalgebras: polynomial algebras , and the group algebra of . A consequence of this factorisation is that there is a class of representations of called standard modules, which are induced from an irreducible representation of . This motivates us to study how a cocycle twist affects algebra factorisation and induced representations. We do this in Section 2 below.
A result which follows from the theory developed in Section 2 which is not readily obvious is that if is even, then the algebra is isomorphic to its twist by . (Recall that, in the nomenclature of imprimitive complex reflection groups, must be an integer.) This follows from a construction due to Kulish and Mudrov [15] which gives an isomorphism of -algebras whenever the twist is via an adjoint action of a -Hopf algebra on . The action of on for even is indeed adjoint because it factors via the embedding as a subgroup in . Combining this new map with the isomorphism from [2], we conclude that the braided Cherednik algebra of the mystic reflection group is isomorphic to . This is an example where quantum algebra methods are used to establish a result in representation theory.
The isomorphism thus obtained between the rational and braided Cherednik algebras (part of Theorem 4.2 below) is not fully compatible with the PBW factorisation, although it restricts to an isomorphism . Hence, combined with the map constructed in [2], which restricts to an isomorphism , we obtain an isomorphism
between the group algebras of the mystic and the complex reflection group.
Despite having isomorphic group algebras over , the groups and may not be isomorphic. However, if is even, these two groups are the same subgroup of the group of monomial matrices, hence gives us an automorphism of the group algebra of . In particular, induces a permutation of the set of irreducible characters of for even . We study this permutation in Section 3 in the case , , that is, for the Coxeter group of type . The main result is Proposition 3.2 which says that the quantum automorphism of induces the permutation
where the irreducible characters are labelled by bipartitions of , and stands for the partition dual to .
After formally presenting the result of the isomorphism between and the corresponding negative braided Cherednik algebra as part of Theorem 4.2, we introduce a noncommutative analogue of the finite-dimensional coinvariant algebra of a complex reflection group . We show that the noncommutative coinvariant algebra of the mystic reflection group carries a regular representation of , in an analogy to the classical Chevalley theorem for . This is another example of a purely representation-theoretic result which is proved using quantum techniques: an important ingredient is a proof that the above isomorphism intertwines the -action on with the Giaquinto-Zhang twist of the -action on .
There is further evidence to support the view that these new algebras should be treated as a “quantum” analogue of . Namely, the noncommutative multiplication on is the cocycle twist of the product on , see Corollary 5.7. Recall that the coinvariant algebra appears as a factor in the triangular decomposition of the finite-dimensional restricted rational Cherednik algebra . It turns out that the twist is fully compatible with the quotient map from onto , which leads to the twisted version of , constructed in Section 6.
It is the construction of which provides us with an example where a Cherednik-type algebra is not isomorphic to its cocycle twist. The example given in Proposition 6.3 is for algebras of rank over the field ; in fact, the two -dimensional algebras are forms of each other, i.e., they become isomorphic if the scalars are extended to . Since the twist is defined over , this raises a question about a possible interplay between twisting and arithmetic phenomena. This will be explored in further work.
2 Twists of algebras and representations
2.1. -module algebras and smash products.
In this section we recall the notion of cocycle twist. We work in the general setting of Hopf algebras, using the notation of Majid [17]. Let be a Hopf algebra over a field with coproduct , counit and antipode . In this section assume all structures are linear over , with tensor products taken over too. An important example of a Hopf algebra is the group algebra for a finite group , where , and for all , with these maps extending linearly to .
Actions of algebras and Hopf algebras will generally be denoted by , adorned if necessary. If is both an algebra, with product , and an -module where acts by , then is additionally an -module algebra if is an -module homomorphism and . Equivalently, -module algebras can be seen as algebra objects in the category of -modules.
In what follows, the smash product of an -module algebra with will be denoted by . Recall this is the algebra with underlying vector space containing and as subalgebras, and with cross-commutation relation:
(1) |
We use Sweedler notation for the coproduct, where summation is understood here.
2.2. The cocycle twist of an algebra.
The cocycle twist involves deforming the algebra structure of an -module algebra via a -cocycle on :
Definition 2.1 ([17], Example 2.3.1 and Theorem 2.3.4).
A -cocycle on a Hopf algebra is an invertible such that and . The Hopf algebra is defined as having the same algebra structure, and counit, as , but with coproduct and antipode:
where and (summation suppressed).
We can now recall the notion of cocycle twist (also called Drinfeld twist) via the following result. Below let denote the category of left -modules.
Proposition 2.2 ([2], Proposition 4.3).
A -cocycle for a Hopf algebra gives rise to the functor
which takes an object to , and an arrow to , where as -module morphisms.
2.3. Cocycle twists of representations.
Giaquinto and Zhang [9] show that, under the restriction of working within the category of -modules, it is possible to twist a representation of an algebra into a representation of the cocycle twist of that algebra. We outline this procedure next.
Definition 2.3 ([9], Definition 1.6).
If is an -module algebra, define to be the category whose objects are both -modules and -modules where the action is a morphism of -modules. That is, the compatibility condition
(2) |
is satisfied. Morphisms in are simultaneously -module and -module maps.
The following notion of twist, from Giaquinto and Zhang [9], takes objects from to , where is the cocycle twist of , i.e. the image of under the functor given in Proposition 2.2. Suppose , and define as a vector space. Note that is naturally an -module via the actions of on and respectively, and therefore defines an endomorphism of . The map
(3) |
defines an -module structure on . This is easily checked to be a well-defined action on applying the conditions of a -cocycle in Definition 2.1, and using (2).
We emphasise that (3) only applies to -modules and not to arbitrary -modules. The next Proposition implies that in general not all -modules can be given an -module structure. A simple description of the -module category is as follows:
Proposition 2.4 ([7], Exercise 7.8.32).
The category of -modules is equivalent to .
Sketch of proof.
By a result of Kulish and Mudrov, the smash product is stable under the twist:
Proposition 2.5 ([15], Proposition 2.5).
Write as (summation understood). The map
(4) |
is an isomorphism of algebras.
This result provides an alternative means of showing the map defined by Giaquinto and Zhang in (3) is a well-defined action of on : embeds as a subalgebra in via the morphism (4), and the map can be seen to coincide with the restriction of the -action on , arising from Proposition 2.4, onto the image of .
Proposition 2.5 also directly explains the following equivalence of categories proved by Giaquinto and Zhang; they are equivalent to module categories over isomorphic algebras and :
Corollary 2.6 ([9], Theorem 1.7).
The categories and are equivalent.
2.4. Twists via adjoint actions.
One can say much more about the twist and the semidirect product when has the following special -module structure:
Definition 2.7 ([19], Example 7.3.3).
An action of on is called adjoint (or strongly inner) if, for some algebra homomorphism , is given by:
Proposition 2.8.
Assume that is an -module algebra with an adjoint action of via the homomorphism , and retain the above notation. Then,
-
1.
The smash product and the tensor product (where the two factors commute) are isomorphic as algebras, via the map .
-
2.
The map
(5) is an isomorphism of algebras.
-
3.
Let be an -module, and denote the action of as . Define an action of on as . Then is an object in , and the twisted action , is equal to the pullback of along , i.e. .
Proof.
Part 1 follows from [19, Example 7.3.3], or see [3, Theorem 3.18] for a proof by explicit calculation. Part 2 is [15, Theorem 2.1]. Alternatively, the isomorphism in (5) is obtained as the composite map , given by (4) followed by the map which, by part 1, is a homomorphism of algebras. For Part 3, we first must check equation (2) holds. Firstly note , whilst for ,
where in the second, third and fourth equalities we use coassociativity, and the antipode and counit axioms for Hopf algebras respectively. Now note , as required. ∎
Remark 2.9.
An immediate corollary to Proposition 2.8(3) is that, if is the representation corresponding to , and is the representation corresponding to , arising as a result of the Giaquinto and Zhang twist, then we have that is given by the pullback of along , i.e. . Additionally, if and are the characters of and respectively, then .
2.5. Twists of algebra factorisations.
Later we wish to discuss rational Cherednik algebras, which, owing to their PBW property, are examples of algebra factorisations. In this section, we prove a general result that twisting preserves the algebra factorisation structure.
Definition 2.10.
If is an algebra, then an algebra factorisation is a pair of subalgebras , of an associative unital -algebra such that the restriction of the multiplication map to is an isomorphism
(6) |
of -vector spaces. To an algebra factorisation there is associated a linear map given by
(7) |
The map obeys the “generalised braiding” equations given by Majid in [18, Proposition 21.4].
If is a Hopf algebra, we consider an algebra factorisation in to be an algebra factorisation in which is an -module algebra, and , are -module subalgebras of . In this case, the isomorphism (6) and the map given by (7) are -module homomorphisms. Furthermore, given a -cocycle , the Drinfeld twists of are clearly -module subalgebras of the Drinfeld twist of . Additionally, the following holds,
Proposition 2.11.
If is an algebra factorisation in , then is an algebra factorisation in , and .
2.6. Induced representations.
Induced representations of associative algebras were defined by Higman [11], extending the notion from group theory. Given a subalgebra of an associative unital algebra (so that is a – bimodule via the regular action), the induction functor
is a left adjoint of the restriction functor from to , see [22]. If is an algebra factorisation and is a -module, one has
as vector spaces, and moreover, as left -modules. The action of on is given by
(8) |
where is the action of on . As pointed out above, when is an algebra factorisation in , then (6) and (7) are -module homomorphisms. Therefore, if , then, in view of (8), we see that .
We now show that for algebra factorisations, the induction functor commutes with the twist functor. More precisely,
Proposition 2.12.
If is an algebra factorisation in and is a -module, then
(9) |
as -modules, where and are Drinfeld twists, and and are Giaquinto and Zhang twists, as in (3). The two -module structures on the underlying vector space are intertwined by the map .
Proof.
Above we explained that , so firstly we find that the Giaquinto and Zhang twist is a well-defined object.
Now by Proposition 2.11, the algebra is generated by and , therefore it is enough to show that the two actions of on the space are intertwined by , and to do the same for .
In the following diagram, the top row represents the action of on the induced module : the composite arrow is , which is indeed the action of on the free left module . The bottom row is the action of on , twisted by .
The leftmost square of the diagram commutes by the cocycle equation, and the rightmost square commutes because is an -module morphism. Hence the diagram commutes, proving that intertwines the two -actions on .
The second diagram deals with the two actions of . The top row consists in applying , which by Proposition 2.11 is to the first two legs, followed by the action on . By (8), this gives the action of on . The bottom row is the -twisted action of on . To make the diagram more compact, we write to denote acting on the first two legs of the tensor product, i.e., the operator ; similarly for other operators:
The first (leftmost) square commutes by the cocycle equation, the second square commutes because is an -module morphism, the third square is the cocycle equation again, and the rightmost square commutes because the action of is a morphism in . Hence the diagram commutes, and the two actions of are indeed intertwined by , as claimed. ∎
3 Twisting irreducible characters of Coxeter groups
In [2, Theorem 6.3] we showed Drinfeld twists induce a non-trivial permutation of the four linear characters of the Coxeter group of type . Here we will extend this result, showing how twisting permutes all of the irreducible characters of . Additionally, we show how twisting induces a bijective correspondence between the irreducible characters of the Coxeter group of type and those of its mystic partner . In particular we describe the permutation for , and the bijective correspondence between and , using a partition conjugation action.
Before discussing the Coxeter groups and let us first recall the family of complex reflection groups , and their mystic counterparts . For , let and denote the groups of -monomial, and permutation, matrices over respectively.
Definition 3.1.
For with , define the following groups:
-
•
is the subgroup of formed by matrices where all the non-zero entries are -th roots of unity, and the product of all non-zero entries is an -th root of unity.
-
•
is the subgroup of with non-zero entries being -th roots of unity, and such that the determinant of the matrix is an -th root of unity.
We write to denote the subgroup of all diagonal matrices in and put . The diagonal matrix whose -th entry is and the rest of diagonal entries are will be denoted by . The elements play a special role in the paper, and we abbreviate to .
Note that if is even, then . Bazlov and Berenstein [1] defined a family of algebra isomorphisms , which for certain choices of and , restrict to an isomorphism from to . For , the map is given by
(10) |
When , the set is empty, and so . Therefore the interesting behaviour of comes from how it evaluates on the standard generators , of . We will be particularly interested in the cases where and . These evaluate on as follows:
(11) | ||||
(12) |
Note that when is even, restricts to an isomorphism . Later in Section 4.1 we will see that, when is even, in fact coincides precisely with the embedding map of Theorem 4.2, after restricting to the subalgebra of the negative braided Cherednik algebra . Here is an isomorphism, constructed in [2], between the and the twist of the rational Cherednik algebra, and is a new map which arises from the results given in Section 2.
Now, since is an isomorphism, the pullback of a representation along gives rise to a representation of the group . Based on Remark 2.9, pulling back along corresponds to twisting the representation of (in the sense of Giaquinto and Zhang) into a representation . Pulling back along the map (from [2, Theorem 5.2]) allows us to interpret the twisted representation as another group representation, in particular of . Therefore we interpret the action of pulling back a representation of along as the action on representations induced by twisting.
In Section 3.1 we describe this action explicity for the groups , which are the Coxeter groups of type . We then turn our attention to the map in (12). Although this map doesn’t directly arise from any twisting constructions, it turns out that pulling back the irreducible characters of via , instead of , induces the same permutation of characters. The advantage of however is that it restricts to an isomorphism when is odd and even. We can therefore use to map the irreducible characters of Coxeter group of type , given by , to those of . We describe this mapping explicity in Section 3.2.
3.1. Twisting irreducible characters of .
Recall the group is isomorphic to the Coxeter group of type . This group is well-known [8, Section 1.6.3] to have irreducible characters labelled by bipartitions of , i.e. ordered pairs where and are partitions of respectively, where . Since is even for this group, the map of Theorem 4.2 restricts to an isomorphism . It is clear from Definition 3.1 that , and therefore we see gives rise to an automorphism of . So the pullback of an irreducible character of along will be another irreducible character of . The following result describes this permutation of the characters explicity.
Proposition 3.2.
where and with .
Proof.
By [8, Section 5.5.4], the irreducible characters of are given as
where is the pullback of the irreducible character (see [8, Definition 5.4.4]) along the projection map , denotes outer tensor product, and is the restriction to of the linear character on which sends and . Precomposing with we find:
The first equality just uses the definition of , whilst the second rewrites as , i.e. via the unique bipartition of which characterises it. We do similarly for . In the 3rd equality we apply [8, Theorem 5.5.6(c)] which says .
Now note that the functor which sends a character of to the character is a composition of two functors, the induction functor followed by the autofunctor of given by precomposing a representation of with . Note that is a permutation of irreducible characters of which is involutive, because . Hence is a self-adjoint functor. It follows, by Frobenius reciprocity, that is adjoint to the functor . But since restricted to is an automorphism of the group algebra , same for , it is obvious that this functor is the same as . Taking adjoints again, we conclude that
At this point we claim and prove the following,
(13) |
The left hand identity follows since for the standard projection map, and on noting that , we deduce .
It remains to check the right hand identity. Firstly, . Also, . Since and are algebra homomorphisms which agree on generators, one has . So , where the final equality again applies [8, Theorem 5.5.6(c)] which says .
With these two identities we find,
as required. ∎
The above result shows how gives rise to permutation of the irreducible characters of the Coxeter group of type . In the next section we consider the Coxeter groups of type , which are isomorphic to the groups . Since is odd in this case, we cannot pull back the irreducible characters of along . However, it was shown in [1, proof of Theorem 2.8] that the map in (10) with restricts to an isomorphism when is both even and odd. Since , we find restricts to an isomorphism . We can therefore ask how precomposition with acts on the irreducible characters of both and . It follows from the next result that permutes the characters of the in exactly the same way as (or ).
Proposition 3.3.
.
Proof.
This follows by almost exactly the same argument as used in the proof of Proposition 3.2, with replaced with . The only part of the proof which does not immediately follow also for is the version of equations (13) with in place of . But we see these hold too since, for the first equation, and , so as required. For the second equation: and also, . ∎
What is the reason why acts the same way on irreducible characters of as ? Using the fact , we can reframe Proposition 3.3 as saying
(14) |
So the characters of are invariant under . This means that is an inner automorphism of :
Lemma 3.4.
Suppose that is a finite group and is an automorphism of its group algebra such that for all irreducible characters of . Then there exists invertible such that for all .
Proof.
Note that the condition means that for all finite-dimensional -modules and for all , the trace is equal to the trace .
The centre of is spanned by primitive idempotents labelled by irreducible characters of . As an automorphism, must permute the set . Suppose that , and let be a -module which affords . Then . However, if then acts on by zero. Hence , meaning that is the identity map.
But then fixes each component in the decomposition of into the direct product of central simple algebras. By the Skolem-Noether theorem, is given by for all . Put . ∎
We can conjecture from this that and twist the characters of all the groups , for even, in the same way.
Conjecture 3.5.
If is even and is a character of , then .
3.2. Twisting irreducible characters of .
Having addressed the Coxeter group of type , we turn to two of its index normal subgroups, the Coxeter group of type and its mystic counterpart , given by and respectively. We will use the map to twist the characters of into those of . Let us recall how the irreducible characters of these groups relate to those of .
The irreducibles of are indexed by bipartitions , and, by [13, Corollary 6.19], the restriction to of an irreducible character of is either irreducible, or the sum of two distinct irreducibles. More specifically, for an irreducible of , if , then the restriction to remains irreducible. Whereas when , the restriction is a direct sum of two non-isomorphic irreducibles of .
Now is also an index subgroup of , so again by [13, Corollary 6.19] the restriction to of an irreducible character of will either be an irreducible, or the sum of two distinct irreducibles. In particular, is the kernel of the linear character , and . We therefore find that, for , the irreducible of restricts to an irreducible of . Whilst for , the restriction decomposes as a direct sum of two non-isomorphic irreducibles of .
In the following, let denote the character of the irreducible representation of . When , let denote the irreducible character arising from the restriction of to , i.e. . Likewise, if , let .
Proposition 3.6.
induces a bijection between the irreducible characters of and the irreducible characters of . In particular, for , .
Proof.
Precomposition with indeed gives a bijection of characters since restricts to an isomorphism . For brevity let us identify with this restriction. We then have,
where the third equality applies Proposition 3.2 and Proposition 3.3. Since , we have , and therefore restricting onto gives an irreducible character, in particular . ∎
4 Rational Cherednik algebras
Let us recall the result [2, Theorem 5.2], which showed that certain rational and negative braided Cherednik algebras are related by a Drinfeld twist. For , let be an -dimensional -vector space with basis , and also let be the dual basis for . For and define the following maps in :
Fix parameters and and .
Definition 4.1.
-
•
The rational Cherednik algebra is the -algebra generated by , , and , subject to the relations:
for with .
-
•
The negative braided Cherednik algebra is the -algebra generated by , , and , subject to the relations:
for with .
Let , isomorphic to the -fold direct product of the cyclic group of order . By [2, Proposition 5.1], for even, is a -module algebra under the following action:
for and . Also recall from [2, Lemma 4.5] that the following is a counital -cocycle of ,
(15) |
and by [2, Theorem 5.2], we have where and . Below in Theorem 4.2 we will denote the Drinfeld twist by , where is the twisted product of .
Note that when is even, we have a natural embedding , and the action of on is seen to be adjoint with respect to this embedding. Therefore, by Proposition 2.8(2), , and the module categories of these algebras must therefore be equivalent. In Theorem 4.2 we compute the isomorphism explicitly, and compose it with the isomorphism of [2, Theorem 5.2] in order to establish an explicit isomorphism between the negative braided Cherednik algebra and the rational Cherednik algebra . Note this isomorphism becomes an embedding when is odd.
4.1. Application to Cherednik algebras.
Theorem 4.2.
Consider the negative braided Cherednik algebra where is even and . Define , and as follows,
Then embeds inside the rational Cherednik algebra via the following mapping of generators:
Proof.
Using the isomorphism from [2, Theorem 5.2] we have an isomorphism where and . Note that the rational Cherednik algebra is a subalgebra of , with and
Additionally for taken to be the natural embedding, acts adjointly (in the sense of Definition 2.7) on . On restricting to the subalgebra this action coincides with the action given in [2, Proposition 5.1] which is used for twisting . Therefore the twisted algebra is a subalgebra of twisted algebra .
Using the fact that the action of on is adjoint, we can apply Proposition 2.8(1) to find . We deduce that is isomorphic to a subalgebra of as required.
Next we inspect how the generators of are mapped into . Recall that under , . It remains to show how the map defined in Proposition 2.8 maps and . Let us use the cocycle given in (15). Let . Then, using the commutativity of , we see , i.e. the composition (in any order) of for . Now , and it is easy to verify that
Therefore,
One similarly shows that . Also, since is commutative, the adjoint action of the subalgebra on is trivial, so and . Finally we check . Using we find
Additionally,
So (using the fact and are orthogonal idempotents). Therefore . ∎
Since the embedding map of Theorem 4.2 is an isomorphism when is even, we deduce:
Corollary 4.3.
For even, the category of modules over the negative braided Cherednik algebra is equivalent to the category of modules over the rational Cherednik algebra .
4.2. Twisting standard modules of rational Cherednik algebras.
Standard modules for rational Cherednik algebras are modelled after induced modules introduced by Verma in [24] and further studied by Bernstein, Gelfand and Gelfand in [4]. Here we define a natural analogue of standard modules for negative braided Cherednik algebras, and show that the twist (in the sense of (3)) of a standard module of a rational Cherednik algebra is precisely one of these standard modules for the corresponding negative braided Cherednik algebra.
For an irreducible complex reflection group with reflection representation , let be the associated rational Cherednik algebra (with ).
Definition 4.4.
Let be a simple -module, and extend to a -module in which acts by zero, i.e. for , . Then the standard module associated to is the -module .
In the following we suppose is even, and let and . Let be the negative braided Cherednik algebra isomorphic to the twist of under the map from [2, Theorem 5.2]. Note that contains the subalgebra , and restricts to an isomorphism . We have also have the following natural analogue of Definition 4.4: for a simple -module, extend to a -module by letting act on by . Then define the standard -module as .
In the following we will prove several cases where the standard -module is (up to an isomorphism of categories) isomorphic to the Giaquinto and Zhang twist of the standard -module, . To prove this we first require the following elementary result.
Suppose is an algebra isomorphism, is a subalgebra of , and . defines an isomorphism of categories , whereby a -module is sent to the -module . Note that the map also defines an isomorphism of categories . The image of an -module under this functor will be similarly denoted .
Lemma 4.5.
If is an -module, then as -modules.
Proof.
We already proved a special case of this in the proof of Proposition 3.2. Note the functor has an adjoint given by . Also and are adjoint to and respectively. Therefore has adjoint , which is easily seen to be equal to . The adjoint of this is then , as required. ∎
Recall the map of (12) restricts to an isomorphism (regardless of the parity of ). This map therefore induces an isomorphism of categories , where as a -vector space, and if is the action of on , then the action of on is given by
In the above notation, . Similarly, the map defines an isomorphism of categories , and below we take to denote the image of under this functor.
Proposition 4.6.
Suppose that is either a simple -module, or a simple -module corresponding to a bipartition of where (see Section 3.2). Then as -modules.
Proof.
In following we will denote by , by , and by . Also let denote an arbitrary group , where is even. Note and .
We wish to apply Proposition 2.12 to the Giaquinto-Zhang twist . To apply the proposition we first must note that is indeed an algebra factorisation in , where recall is the group isomorphic to . We also require , which we check next. Note this category is well-defined since is a -module algebra under the adjoint action of (in particular it is a -submodule algebra of ). When is even, then is a subalgebra of , and therefore is naturally a -module by restricting the action to . This action can equivalently be seen as pulling back the action of along the embedding map , and therefore we can apply Proposition 2.8(3) to deduce .
However does not embed in . By assumption though, is equal to the simple -module , where . On recalling Section 3.2, this module is the restriction to of either of the simple -modules or . Since is a subalgebra of , each of these -modules defines an action of on . Note that on extending to one these -modules, we can apply the even case above to deduce that . But then the compatibility condition (2) will still hold if we restrict the action of on to , and therefore as required. Later we show that this proposition holds independently of which action is chosen.
We can now apply Proposition 2.12 to deduce:
Here is the Giaquinto and Zhang twist of (see (3)), so the action of on is given by , where denotes the action of on , and denotes the action on . We now apply Lemma 4.5 to find,
where as a -vector space, and has action of given by .
To complete the proof it remains to check as -modules. We consider the cases of even, and , separately.
Let be even. The action of on is just a restriction of the action of on , and the action of on is adjoint, so we can apply Proposition 2.8(3). We deduce that the action of is given by the pulling back along the map , i.e. , where was defined in the proof of Theorem 4.2. Then , where is regarded here as a map .
Note that the map is degree-preserving, and, by definition, is such that elements of degree in act on by . Therefore is determined by how it behaves on the subalgebra of . Recall that the restriction of the map to the subalgebra coincides precisely with the map , given in (11). Also, in Proposition 3.3 it was proven that , for an irreducible character of . Therefore, as actions of on , we have . Hence, as -modules, . This implies they are also equal as -modules, since acts by for both and . This proves the result for the group .
Let us now check , as -modules, when for . As mentioned before, can be extended to either of the simple -modules or . Suppose, for now, that we extend to . Let us consider . From the even case above we know that the action of on is given by . Therefore by Proposition 3.2, . Since , restricting the -action on to gives the irreducible representation . Let us now inspect , which is defined to be . By Proposition 3.6, . Therefore as -modules, and so also as -modules, since each of these -modules is extended by taking to act by .
Finally suppose we had instead extended to . Then on applying Proposition 3.2 we would have found . Upon restricting this to we arrive at the irreducible representation . But we see that since by [8, Theorem 5.5.6(c)], , where is the linear character which satisfies .
∎
5 Twisted coinvariant algebras
5.1. Classical (commutative) coinvariant algebras.
Let be a finite-dimensional vector space over a field . Each finite subgroup of acts on the symmetric -algebra of . Classically, invariant theory of finite groups studies the graded algebra of -invariants. Of particular interest is the case when is a finite reflection group, that is, when generates . In the non-modular case (when the characteristic of does not divide ), the work of Chevalley, Sheppard and Todd, and Serre in the 1950s established that is a reflection group, if and only if is a polynomial algebra. This is known as the Chevalley-Shephard-Todd theorem, see for example [20, Theorem 7.1.4].
The coinvariant algebra is defined as the quotient algebra where is the ideal of generated by homogeneous -invariants of positive degree. The following result was initially proved by Chevalley [6] under more restrictive assumptions, but holds for all non-modular reflection groups, see [20, Theorem 7.2.1]. See also [5] for a discussion of the modular case.
Theorem 5.1.
Supposing is an arbitrary field and is a finite reflection group generated by reflections whose orders are prime to the characteristic of . Then the coinvariant algebra, viewed as a -module, affords the regular representation of .
5.2. The noncommutative coinvariant algebra of a mystic reflection group.
Let be a field of characteristic . An extension of the Chevalley-Shephard-Todd theorem to skew-polynomial -algebras is given by Kirkman, Kuzmanovich and Zhang in [14]. It characterises mystic reflection groups as finite subgroups which act by algebra automorphisms on such that the subalgebra is again skew-polynomial. Set ; it turns out that the “building blocks” of mystic reflection groups are
-
•
the usual complex reflection groups, which act on the commutative polynomial algebra ,
-
•
the family of Definition 3.1 where is even, which act on the skew-polynomial algebra .
Recall that the group is defined via its faithful action on a space with a fixed basis . We regard the graded vector space
spanned by the standard monomials, as the underlying space for both and . The space contains
(16) |
which are homogeneous algebraically independent generators, in , of the subalgebra of -invariants, see [16, Chapter 2, Section 8]. The subalgebra of -invariants in happens to form a commutative algebra isomorphic to , as noted in [14].
Theorem 5.2 ([1], Theorem 2.6).
Let . Then is a commutative polynomial algebra, freely generated by and .
Given that the mystic reflection groups have polynomial invariants in , we define their coinvariant algebras and obtain the noncommutative analogue of Chevalley’s classical result.
In the following, take and .
Definition 5.3.
Let be the two-sided ideal of the algebra generated by homogeneous -invariants of positive degree. The noncommutative coinvariant algebra of is the quotient algebra .
It is clear from the definition that the ideal is a -submodule of , and so acts on by algebra automorphisms — the action descends from .
Theorem 5.4.
The noncommutative coinvariant algebra of the mystic reflection group affords the regular representation of .
We prove Theorem 5.4 over the next few sections, using the techniques we developed above for working with cocycle twists.
5.3. -actions and Giaquinto-Zhang twists of -actions.
We are going to use the rational Cherednik algebra of as well as the negative braided Cherednik algebra . We will make use only of the semidirect product relations in these algebras, so for simplicity we set the parameters to . To state the next Lemma, we need to recall from [2, Remark 5.10] that the group algebra is the middle term in the PBW-type algebra factorisation
(17) |
The result [2, Theorem 5.2], establishes an algebra isomorphism between and the cocycle twist of by . The map is defined on generators of by
for , and . In particular, restricts to an isomorphism . Furthermore,
(18) |
i.e., is the identity on the underlying vector space because it maps every standard monomial in to itself [2, section 5.9]. We show that intertwines the actions of and on :
Lemma 5.5.
Let be the action of on , and let be the Giaquinto-Zhang twist by of the natural action of on . Then for all , as endomorphisms of the vector space .
Proof.
We bring in the Cherednik algeras. The relations in Definition 4.1 mean that the action can be written as the composite map
(19) |
where is the product map on , is the augmentation map of the group algebra , and maps a polynomial to its constant term. The domain of is the triple tensor product in (17). Thus, (19) says that the -module is the restriction to of the representation of induced from the one-dimensional trivial module of the semidirect product . Consider the diagram
where the leftmost square commutes because restricts to the identity on , and the middle square commutes because is an isomorphism of algebras.
We claim that the rightmost square commutes. This boils down to the equality of maps , where is the augmentation map of . This may not be obvious, since does not send group elements of to group elements of . Note that is multiplicative on , being the trivial character of the group . Furthermore, since carries the trivial action of the group , the Giaquinto-Zhang twist of is the same linear map , viewed as a representation of the twisted algebra . Therefore, is multiplicative on . Thus and are two algebra homomorphisms from to ; they both send grouplike generators , , and to , hence coincide on .
We conclude that the above diagram commutes. But it shows that intertwines the action of on with the -action arising from the representation of
on the space , induced from the trivial -module. The latter is the Giaquinto-Zhang twist of the trivial -module. Therefore, by Proposition 2.12, the -action on is the Giaquinto-Zhang twist of the standard -module , arising from the trivial representation of . We know that the action of on this standard module is the natural action of on , so the lemma is proved. ∎
Our next step is to replace and in Lemma 5.5 by the respective coinvariant algebras. For this, we need to show the coinvariant algebras are also defined on the same underlying vector space.
Lemma 5.6.
The ideal of and the ideal of share the same underlying subspace of . This subspace is stable under the action of the group .
Proof.
We temporarily write for -multiplication and for -multiplication on the space . Then is the -ideal generated by and as given in (16), and is the -ideal with the same generators. Hence is spanned by and where are standard monomials and . Using the facts that is -central as is even, that is a monomial, and that for all monomials , we conclude that all elements of this spanning set lie in . So . Similarly . Note that acts on the space by automorphisms, and also by algebra automorphisms on and . This action leaves invariant, and maps to , hence also maps to . ∎
Corollary 5.7.
The algebras and are defined on the same underlying vector space . As an algebra, is the cocycle twist of . The isomorphism intertwines the action of on and the Giaquinto-Zhang twist of the action of on .
5.4. The trace is unchanged under the twist.
To prove Theorem 5.4, we calculate the character of the representation of afforded by , that is, the trace of the action of on . By Corollary 5.7, this is the same as . We will now show that neither the Giaquinto-Zhang twist by nor the application of changes the trace. The former follows from a general result about Giaquinto-Zhang twists:
Proposition 5.8.
Let be a Hopf algebra with involutive antipode over a field , an -module algebra, a finite-dimensional -module where acts via , and a -cocycle. Denote the -twisted action of on the space by .
For any element of the vector space of , .
Proof.
Denote the action of on simply by ; same for the -action on and on . Recall [17, Proposition 9.3.3] that the -action on is such that the evaluation map and the coevaluation map , , are -module maps, where is the trivial -module. Equivalently, for all , , ,
(20) |
where is the antipode. Since is an -module, it follows from (2) that for ,
(21) |
By definition of trace, . Writing as (summation understood), we compute the trace of the -twisted action of :
(22) |
where we used definition (3) of , (20) and (21). We now manipulate the coevaluation tensor:
where we applied the second part of (20), the assumption that and the antipode law. Here is the counit. Since by Definition 2.1 is counital, which means is equal to the coevaluation tensor . Thus, (22) rewrites as which is . ∎
5.5. The trace is unchanged under ; proof of Theorem 5.4.
If is any finite group, consider the linear map
Of course, we recognise as the character of the regular representation of . Viewing as a linear map between the underlying vector spaces , , we show that intertwines the regular characters:
Lemma 5.9.
.
Proof.
We have , so elements of are written as where and . By [2, Proposition 5.7], there exist elements , such that for all . Moreover, the proof of [2, Proposition 5.7] shows that .
So if , , and , then so . On the other hand, because guarantees that is a linear combination of group elements of , none of which is .
In the case , , we have , hence and are both if and are both if . This shows that and agree on all elements of , proving the Lemma. ∎
We are ready to finish the proof of Theorem 5.4.
Proof of Theorem 5.4.
We need to show that the trace, , of the action of on , is equal to . By Corollary 5.7, the trace equals , which by Proposition 5.8 coincides with . By Chevalley’s Theorem 5.1, the latter is which by Lemma 5.9 is , as claimed.
Our application of Proposition 5.8 is valid because the Hopf algebra has involutive antipode. Indeed, not only the antipode of any commutative or cocommutative Hopf algebra is involutive [21, Corollary 7.1.11], but the antipode of is, in fact, the identity map because every element of the group is self-inverse. ∎
6 Twists of restricted rational Cherednik algebras
6.1. Restricted rational Cherednik algebras.
Let be a field, a finite-dimensional vector space over , and a finite non-modular reflection group. The classical coinvariant algebra of appears as a PBW-type factor in the algebra factorisation of the restricted rational Cherednik algebra . These finite-dimensional quotients of rational Cherednik algebras at originate from Gordon’s 2003 paper [10] and have attracted considerable interest since.
Of interest to us is the case , and in this section we take to be a subfield of over which the reflection representation of is defined; the smallest such is therefore the th cyclotomic extension of . The -rational Cherednik algebra , given by Definition 4.1, is defined over as long as the parameter is -valued. Thiel [23] defines the restricted rational Cherednik algebra as
where is the ideal generated by homogeneous -invariants of positive degree in , is the similar ideal in , and is the ideal generated by and in . One has the following algebra factorisation,
where is the coinvariant algebra of from the previous section, and is defined in the same way from the action of on the dual space .
6.2. Twists of restricted rational Cherednik algebras.
Let now be the mystic reflection group which is the mystic counterpart of . We define a finite-dimensional quotient of the negative braided Cherednik algebra by
where , are ideals generated by homogeneous invariants of positive degree. This can be called the “restricted negative braided Cherednik algebra”. The results of Lemma 5.6, Corollary 5.7 and [2, Theorem 5.2] immediately lead us to the following,
Theorem 6.1.
The cocycle twist of by the cocycle given in (15) is isomorphic to . This algebra has triangular decomposition
into the group algebra of and two noncommutative coinvariant algebras.
We further observe the following,
Proposition 6.2.
If is even, then and are isomorphic as algebras.
Proof.
Indeed, Proposition 2.8 applies because the action of the Hopf algebra on is adjoint, as it factors via the embedding of as the subgroup in . ∎
6.3. A case where the restricted rational Cherednik algebra is not isomorphic to its twist.
We stress that Proposition 2.8 establishes an isomorphism between and over an arbitrary field , as long as the -module structure on and the cocycle are defined over , and the -action on is adjoint. We will now give an explicit example where the action is not adjoint, and the restricted rational Cherednik algebra is not isomorphic to its twist by .
Of course, we need to be odd, and we consider the case . Note is the Klein four-group generated by the commuting involutions and . On the other hand, is the cyclic group of order generated by the mystic reflection . We see the minimal field over which the reflection representation of and the mystic reflection representation of are defined is the rationals, .
The proof of the following result is partly based on an explicit computer calculation (over the rational field).
Proposition 6.3.
If , then the -algebras and are not isomorphic.
Proof.
Note that is the non-reduced Coxeter group of type , so
where is the reflection group of order which operates on the -dimensional space. Thus, is a -dimensional algebra which is a tensor product of two (commuting) -dimensional subalgebras. Explicitly, is presented as
It follows that
By results of [12], the centre is two-dimensional. Consider . We claim that is central in . Indeed, , and
so commutes with . Similarly, commutes with . We conclude that . Observe that
Since in our example, . In particular, the algebra does not contain elements of multiplicative order , hence such elements do not exist in . On the other hand, we claim that
is a central element of . Moreover, is not a scalar whereas . This is a result of a computer calculation. Hence, if , the centre of contains an element of multiplicative order , which shows that the two algebras are not isomorphic over . ∎
We remark that the above example does not work over , or indeed over an extension of containing . If scalars are extended to , the two algebras become isomorphic. We currently have no explicit example where we could disprove isomorphism over .
We finish with the following conjecture.
Conjecture 6.4.
Let be even and be odd, and let . The -algebras and are not isomorphic.
References
- [1] Y. Bazlov and A. Berenstein. Mystic reflection groups. Symmetry, Integrability and Geometry: Methods and Applications, 10, April 2014.
- [2] Y. Bazlov, A. Berenstein, E. Jones-Healey, and A. Mcgaw. Twists of rational Cherednik algebras. The Quarterly Journal of Mathematics, 74(2):511–528, 2022.
- [3] A. Berenstein and K. Schmidt. Factorizable module algebras. International Mathematics Research Notices, 2019(21):6711–6764, 2019.
- [4] I. N. Bernstein, I. M. Gelfand, and S. I. Gelfand. Structure of representations that are generated by vectors of highest weight. Funckcional. Anal. i Priložen., 5(1):1–9, 1971.
- [5] A. Broer, V. Reiner, L. Smith, and P. Webb. Extending the coinvariant theorems of Chevalley, Shephard-Todd, Mitchell, and Springer. Proceedings of the London Mathematical Society, 103(5):747–785, 2011.
- [6] C. Chevalley. Invariants of finite groups generated by reflections. American Journal of Mathematics, 77(4):778–782, 1955.
- [7] P. Etingof, S. Gelaki, D. Nikshych, and V. Ostrik. Tensor Categories. Mathematical Surveys and Monographs. American Mathematical Society, 2016.
- [8] M. Geck and G. Pfeiffer. Characters of finite Coxeter groups and Iwahori-Hecke algebras. London Mathematical Society monographs ; 21. Clarendon Press, Oxford, 2000.
- [9] A. Giaquinto and J. J. Zhang. Bialgebra actions, twists, and universal deformation formulas. Journal of Pure and Applied Algebra, 128(2):133–151, 1998.
- [10] I Gordon. Baby Verma modules for rational cherednik algebras. Bulletin of the London Mathematical Society, 35(3):321–336, 2003.
- [11] D. G. Higman. Induced and produced modules. Canadian J. Math., 7:490–508, 1955.
- [12] N. Hird. An explicit presentation of the centre of the restricted rational cherednik algebra. arXiv preprint arXiv:2301.03316, 2023.
- [13] I. M. Isaacs. Character theory of finite groups. Pure and applied mathematics, a series of monographs and textbooks ; v. 69. Academic Press, New York, 1976.
- [14] E. Kirkman, J. Kuzmanovich, and J. J. Zhang. Shephard-Todd-Chevalley theorem for skew polynomial rings. Algebras and Representation Theory, 13:127–158, 2008.
- [15] P. Kulish and A. Mudrov. Twisting adjoint module algebras. Lett. Math. Phys., 95(3):233–247, 2011.
- [16] G. Lehrer and D. E. Taylor. Unitary reflection groups. Australian Mathematical Society lecture series ; 20. Cambridge University Press, Cambridge, 2009.
- [17] S. Majid. Foundations of Quantum Group Theory. Cambridge University Press, 1995.
- [18] S. Majid. A Quantum Groups Primer, volume 292 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2002.
- [19] S. Montgomery. Hopf algebras and their actions on rings, volume 82 of CBMS Regional Conference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1993.
- [20] M. D. Neusel and L. Smith. Invariant theory of finite groups, volume 94 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2002.
- [21] D. E. Radford. Hopf Algebras. World Scientific, 2011.
- [22] M. A. Rieffel. Induced representations of rings. Canadian Journal of Mathematics, 27(2):261–270, 1975.
- [23] U. Thiel. Restricted rational Cherednik algebras. In Representation theory—current trends and perspectives, EMS Ser. Congr. Rep., pages 681–745. Eur. Math. Soc., Zürich, 2017.
- [24] D-N. Verma. Structure of certain induced representations of complex semisimple Lie algebras. Bulletin of the American Mathematical Society, 74(1):160 – 166, 1968.