This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Twists of representations of complex reflection groups and rational Cherednik algebras

Y. Bazlov and E. Jones-Healey
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 BnB_{n} and DnD_{n} 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 AA and its representations in a category of modules over a quasitriangular Hopf algebra, or more generally a Hopf algebra equipped with a 22-cocycle, leads to a deformation — known as cocycle twist — of the associative product on AA, as well as of the representations of AA. 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 ϕ\phi between braided Cherednik algebras, constructed earlier, and cocycle twists of rational Cherednik algebras Hc(G(m,p,n))H_{c}(G(m,p,n)) of imprimitive complex reflection groups G(m,p,n)G(m,p,n) with even mm. These algebras have an action of the elementary abelian 22-group T=T(2,1,n)T=T(2,1,n) whose group algebra T\mathbb{C}T carries a non-standard, cohomologically non-trivial quasitriangular structure \mathcal{F} which is the cocycle we use. A strong property of a rational Cherednik algebra Hc(G)H_{c}(G) is its PBW-type factorisation, SGSS\otimes\mathbb{C}G\otimes S^{\prime} into three subalgebras: polynomial algebras SS, SS^{\prime} and the group algebra of GG. A consequence of this factorisation is that there is a class of representations of Hc(G(m,p,n))H_{c}(G(m,p,n)) called standard modules, which are induced from an irreducible representation of G\mathbb{C}G. 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 mp\frac{m}{p} is even, then the algebra Hc(G(m,p,n))H_{c}(G(m,p,n)) is isomorphic to its twist by \mathcal{F}. (Recall that, in the nomenclature of imprimitive complex reflection groups, mp\frac{m}{p} must be an integer.) This follows from a construction due to Kulish and Mudrov [15] which gives an isomorphism η:AA\eta\colon A_{\mathcal{F}}\to A of kk-algebras whenever the twist AA_{\mathcal{F}} is via an adjoint action of a kk-Hopf algebra on AA. The action of TT on Hc(G(m,p,n))H_{c}(G(m,p,n)) for even mp\frac{m}{p} is indeed adjoint because it factors via the embedding T=T(2,1,n)T=T(2,1,n) as a subgroup in G(m,p,n)G(m,p,n). Combining this new map η\eta with the isomorphism ϕ\phi from [2], we conclude that the braided Cherednik algebra of the mystic reflection group μ(G(m,p,n))\mu(G(m,p,n)) is isomorphic to Hc(G(m,p,n))H_{c}(G(m,p,n)). This is an example where quantum algebra methods are used to establish a result in representation theory.

The isomorphism η\eta 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 η:G(m,p,n)G(m,p,n)\eta\colon\mathbb{C}G(m,p,n)_{\mathcal{F}}\to\mathbb{C}G(m,p,n). Hence, combined with the map ϕ\phi constructed in [2], which restricts to an isomorphism μ(G(m,p,n))ϕG(m,p,n)\mathbb{C}\mu(G(m,p,n))\xrightarrow{\phi}\mathbb{C}G(m,p,n)_{\mathcal{F}}, we obtain an isomorphism

ηϕ:μ(G(m,p,n))G(m,p,n)\eta\phi\colon\mathbb{C}\mu(G(m,p,n))\to\mathbb{C}G(m,p,n)

between the group algebras of the mystic and the complex reflection group.

Despite having isomorphic group algebras over \mathbb{C}, the groups μ(G(m,p,n))\mu(G(m,p,n)) and G(m,p,n)G(m,p,n) may not be isomorphic. However, if mp\frac{m}{p} is even, these two groups are the same subgroup of the group of n×nn\times n monomial matrices, hence ηϕ\eta\phi gives us an automorphism of the group algebra of G(m,p,n)G(m,p,n). In particular, ηϕ\eta\phi induces a permutation of the set of irreducible characters of G(m,p,n)G(m,p,n) for even mp\frac{m}{p}. We study this permutation in Section 3 in the case m=2m=2, p=1p=1, that is, for the Coxeter group of type BnB_{n}. The main result is Proposition 3.2 which says that the quantum automorphism ηϕ\eta\phi of Bn\mathbb{C}B_{n} induces the permutation

χ(λ,μ)ηϕ=χ(λ,μ)\chi_{(\lambda,\mu)}\circ\eta\phi=\chi_{(\lambda,\mu^{*})}

where the irreducible characters are labelled by bipartitions (λ,μ)(\lambda,\mu) of nn, and μ\mu^{*} stands for the partition dual to μ\mu.

After formally presenting the result of the isomorphism between Hc(G(m,p,n))H_{c}(G(m,p,n)) and the corresponding negative braided Cherednik algebra as part of Theorem 4.2, we introduce a noncommutative analogue of the finite-dimensional coinvariant algebra SGS_{G} of a complex reflection group G=G(m,p,n)G=G(m,p,n). We show that the noncommutative coinvariant algebra S¯W\underline{S}_{W} of the mystic reflection group W=μ(G(m,p,n))W=\mu(G(m,p,n)) carries a regular representation of WW, in an analogy to the classical Chevalley theorem for SGS_{G}. 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 ϕ\phi intertwines the WW-action on S¯W\underline{S}_{W} with the Giaquinto-Zhang twist of the GG-action on SGS_{G}.

There is further evidence to support the view that these new algebras S¯W\underline{S}_{W} should be treated as a “quantum” analogue of SGS_{G}. Namely, the noncommutative multiplication on S¯W\underline{S}_{W} is the cocycle twist of the product on SGS_{G}, see Corollary 5.7. Recall that the coinvariant algebra SGS_{G} appears as a factor in the triangular decomposition of the finite-dimensional restricted rational Cherednik algebra H¯c(G)SGGSG\overline{H}_{c}(G)\cong S_{G}\otimes\mathbb{C}G\otimes S_{G}^{\prime}. It turns out that the twist is fully compatible with the quotient map from H0,c(G)H_{0,c}(G) onto H¯c(G)\overline{H}_{c}(G), which leads to the twisted version H¯¯c(W)\overline{\underline{H}}_{c}(W) of H¯c(G)\overline{H}_{c}(G), constructed in Section 6.

It is the construction of H¯¯c(W)\overline{\underline{H}}_{c}(W) 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 22 over the field \mathbb{Q}; in fact, the two 6464-dimensional algebras are forms of each other, i.e., they become isomorphic if the scalars are extended to \mathbb{C}. Since the twist is defined over \mathbb{Q}, 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.   HH-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 HH be a Hopf algebra over a field kk with coproduct :HHH\triangle:H\xrightarrow{}H\otimes H, counit ϵ:Hk\epsilon:H\xrightarrow{}k and antipode S:HHS:H\to H. In this section assume all structures are linear over kk, with tensor products taken over kk too. An important example of a Hopf algebra is the group algebra kTkT for a finite group TT, where (t)=tt\triangle(t)=t\otimes t, ϵ(t)=1\epsilon(t)=1 and S(t)=t1S(t)=t^{-1} for all tTt\in T, with these maps extending linearly to kTkT.

Actions of algebras and Hopf algebras will generally be denoted by \rhd, adorned if necessary. If AA is both an algebra, with product m:AAAm:A\otimes A\xrightarrow{}A, and an HH-module where HH acts by \rhd, then AA is additionally an HH-module algebra if mm is an HH-module homomorphism and h1A=ϵ(h)1AhHh\rhd 1_{A}=\epsilon(h)1_{A}\ \forall h\in H. Equivalently, HH-module algebras can be seen as algebra objects in the category of HH-modules.

In what follows, the smash product of an HH-module algebra AA with HH will be denoted by A#HA\#H. Recall this is the algebra with underlying vector space AHA\otimes H containing AA1HA\cong A\otimes 1_{H} and H1AHH\cong 1_{A}\otimes H as subalgebras, and with cross-commutation relation:

ha=(h(1)a)#h(2)hH,aA.ha=(h_{(1)}\rhd a)\#h_{(2)}\qquad\forall h\in H,a\in A. (1)

We use Sweedler notation (h)=h(1)h(2)\triangle(h)=h_{(1)}\otimes h_{(2)} 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 HH-module algebra via a 22-cocycle on HH:

Definition 2.1 ([17], Example 2.3.1 and Theorem 2.3.4).

A 22-cocycle on a Hopf algebra HH is an invertible HH\mathcal{F}\in H\otimes H such that (1)(id)()=(1)(id)()(\mathcal{F}\otimes 1)\cdot(\triangle\otimes\text{id})(\mathcal{F})=(1\otimes\mathcal{F})\cdot(\text{id}\otimes\triangle)(\mathcal{F}) and (ϵid)()=1=(idϵ)()(\epsilon\otimes\text{id})(\mathcal{F})=1=(\text{id}\otimes\epsilon)(\mathcal{F}). The Hopf algebra HH^{\mathcal{F}} is defined as having the same algebra structure, and counit, as HH, but with coproduct and antipode:

h=(h)1,S(h)=US(h)U1,hH\triangle_{\mathcal{F}}h=\mathcal{F}(\triangle h)\mathcal{F}^{-1},\ S_{\mathcal{F}}(h)=U\cdot S(h)\cdot U^{-1},\ \forall h\in H

where U:=1S(2)U:=\mathcal{F}_{1}\cdot S(\mathcal{F}_{2}) and =12\mathcal{F}=\mathcal{F}_{1}\otimes\mathcal{F}_{2} (summation suppressed).

We can now recall the notion of cocycle twist (also called Drinfeld twist) via the following result. Below let H-ModH\text{-}\mathrm{Mod} denote the category of left HH-modules.

Proposition 2.2 ([2], Proposition 4.3).

A 22-cocycle HH\mathcal{F}\in H\otimes H for a Hopf algebra HH gives rise to the functor

():Alg(H-Mod)Alg(H-Mod)(\ )_{\mathcal{F}}:\mathop{\mathrm{Alg}}(H\text{-}\mathrm{Mod})\to\mathop{\mathrm{Alg}}(H^{\mathcal{F}}\text{-}\mathrm{Mod})

which takes an object (A,m)(A,m) to (A,m=m(1))(A,m_{\mathcal{F}}=m(\mathcal{F}^{-1}\rhd-)), and an arrow (A,m)ϕ(B,m)(A,m)\xrightarrow{\phi}(B,m^{\prime}) to (A,m)ϕ(B,m)(A,m_{\mathcal{F}})\xrightarrow{\phi_{\mathcal{F}}}(B,m^{\prime}_{\mathcal{F}}), where ϕ=ϕ\phi_{\mathcal{F}}=\phi as HH-module morphisms.

2.3.   Cocycle twists of representations.

Giaquinto and Zhang [9] show that, under the restriction of working within the category of HH-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 AA is an HH-module algebra, define (H,A)-Mod(H,A)\text{-}\mathrm{Mod} to be the category whose objects VV are both HH-modules and AA-modules where the action A:AVV\rhd_{A}\colon A\otimes V\to V is a morphism of HH-modules. That is, the compatibility condition

h(aAv)=(h(1)a)A(h(2)v),hH,aA,vVh\rhd(a\rhd_{A}v)=(h_{(1)}\rhd a)\rhd_{A}(h_{(2)}\rhd v),\qquad\forall h\in H,a\in A,v\in V (2)

is satisfied. Morphisms in (H,A)-Mod(H,A)\text{-}\mathrm{Mod} are simultaneously AA-module and HH-module maps.

The following notion of twist, from Giaquinto and Zhang [9], takes objects from (H,A)-Mod(H,A)\text{-}\mathrm{Mod} to (H,A)-Mod(H^{\mathcal{F}},A_{\mathcal{F}})\text{-}\mathrm{Mod}, where AA_{\mathcal{F}} is the cocycle twist of AA, i.e. the image of AA under the functor given in Proposition 2.2. Suppose V(H,A)-ModV\in(H,A)\text{-}\mathrm{Mod}, and define V=VV_{\mathcal{F}}=V as a vector space. Note that AVA\otimes V is naturally an HHH\otimes H-module via the actions of HH on AA and VV respectively, and therefore 1\mathcal{F}^{-1}\rhd defines an endomorphism of AVA\otimes V. The map

:AVV,:=A(1)\rhd_{\mathcal{F}}:A_{\mathcal{F}}\otimes V_{\mathcal{F}}\xrightarrow{}V_{\mathcal{F}},\qquad\rhd_{\mathcal{F}}:=\rhd_{A}\circ(\mathcal{F}^{-1}\rhd) (3)

defines an AA_{\mathcal{F}}-module structure on VV_{\mathcal{F}}. This is easily checked to be a well-defined action on applying the conditions of a 22-cocycle in Definition 2.1, and using (2).

We emphasise that (3) only applies to (H,A)(H,A)-modules and not to arbitrary AA-modules. The next Proposition implies that in general not all AA-modules can be given an (H,A)(H,A)-module structure. A simple description of the (H,A)(H,A)-module category is as follows:

Proposition 2.4 ([7], Exercise 7.8.32).

The category of (H,A)(H,A)-modules is equivalent to A#H-ModA\#H\text{-}\mathrm{Mod}.

Sketch of proof.

On an (H,A)(H,A)-module VV, the formula (a#h)v=aA(hv)(a\#h)\rhd v=a\rhd_{A}(h\rhd v) gives a well-defined A#HA\#H-action: (2) is the same as compatibility of this action with the cross-commutator relation (1) of the smash product. ∎

By a result of Kulish and Mudrov, the smash product is stable under the twist:

Proposition 2.5 ([15], Proposition 2.5).

Write 1HH\mathcal{F}^{-1}\in H\otimes H as 1=ff′′\mathcal{F}^{-1}=f^{\prime}\otimes f^{\prime\prime} (summation understood). The map

A#HA#H,a#h(fa)#f′′h,A_{\mathcal{F}}\#H^{\mathcal{F}}\xrightarrow{\sim}A\#H,\qquad a\#h\mapsto(f^{\prime}\rhd a)\#f^{\prime\prime}h, (4)

is an isomorphism of algebras.

This result provides an alternative means of showing the map \rhd_{\mathcal{F}} defined by Giaquinto and Zhang in (3) is a well-defined action of AA_{\mathcal{F}} on VV: A=A#1A#HA_{\mathcal{F}}=A_{\mathcal{F}}\#1\subseteq A_{\mathcal{F}}\#H^{\mathcal{F}} embeds as a subalgebra in A#HA\#H via the morphism (4), and the map \rhd_{\mathcal{F}} can be seen to coincide with the restriction of the A#HA\#H-action on VV, arising from Proposition 2.4, onto the image of AA_{\mathcal{F}}.

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 A#HA_{\mathcal{F}}\#H^{\mathcal{F}} and A#HA\#H:

Corollary 2.6 ([9], Theorem 1.7).

The categories (H,A)-Mod(H,A)\text{-}\mathrm{Mod} and (H,A)-Mod(H^{\mathcal{F}},A_{\mathcal{F}})\text{-}\mathrm{Mod} are equivalent.

2.4.   Twists via adjoint actions.

One can say much more about the twist and the semidirect product when AA has the following special HH-module structure:

Definition 2.7 ([19], Example 7.3.3).

An action \rhd of HH on AA is called adjoint (or strongly inner) if, for some algebra homomorphism u:HAu\colon H\to A, \rhd is given by:

ha=u(h(1))au(Sh(2)).h\rhd a=u(h_{(1)})au(Sh_{(2)}).
Proposition 2.8.

Assume that AA is an HH-module algebra with an adjoint action of HH via the homomorphism u:HAu\colon H\to A, and retain the above notation. Then,

  • 1.

    The smash product A#HA\#H and the tensor product AHA\otimes H (where the two factors commute) are isomorphic as algebras, via the map a#hau(h(1))h(2)a\#h\mapsto au(h_{(1)})\otimes h_{(2)}.

  • 2.

    The map

    η:AA,a(fa)u(f′′),\eta\colon A_{\mathcal{F}}\xrightarrow{\sim}A,\qquad a\mapsto(f^{\prime}\rhd a)u(f^{\prime\prime}), (5)

    is an isomorphism of algebras.

  • 3.

    Let VV be an AA-module, and denote the action of AA as A\rhd_{A}. Define an action of HH on VV as H=A(uidV)\rhd_{H}=\rhd_{A}\circ(u\otimes\text{id}_{V}). Then VV is an object in (H,A)-Mod(H,A)\text{-}\mathrm{Mod}, and the twisted action :AVV\rhd_{\mathcal{F}}:A_{\mathcal{F}}\otimes V_{\mathcal{F}}\rightarrow V_{\mathcal{F}}, :=A(1)\rhd_{\mathcal{F}}:=\rhd_{A}\circ(\mathcal{F}^{-1}\rhd) is equal to the pullback of A\rhd_{A} along η\eta, i.e. =A(ηidV)\rhd_{\mathcal{F}}=\rhd_{A}\circ(\eta\otimes\text{id}_{V_{\mathcal{F}}}).

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 η\eta in (5) is obtained as the composite map A=A#1A#HA#HAA_{\mathcal{F}}=A_{\mathcal{F}}\#1\subseteq A_{\mathcal{F}}\#H^{\mathcal{F}}\xrightarrow{\sim}A\#H\to A, given by (4) followed by the map a#hau(h1)ϵ(h(2))=au(h)a\#h\mapsto au(h_{1})\epsilon(h_{(2)})=au(h) which, by part 1, is a homomorphism of algebras. For Part 3, we first must check equation (2) holds. Firstly note hH(aAv)=u(h)aAvh\rhd_{H}(a\rhd_{A}v)=u(h)a\rhd_{A}v, whilst for (h)=h(1)h(2)\triangle(h)=h_{(1)}\otimes h_{(2)},

(h(1)a)A(h(2)v)\displaystyle(h_{(1)}\rhd a)\rhd_{A}(h_{(2)}\rhd v) =u(h(1)(1))au(S(h(1)(2)))A(u(h(2))Av)\displaystyle=u(h_{(1)(1)})au(S(h_{(1)(2)}))\rhd_{A}(u(h_{(2)})\rhd_{A}v)
=u(h(1))au(S(h(2)(1)h(2)(2)))Av\displaystyle=u(h_{(1)})au(S(h_{(2)(1)}h_{(2)(2)}))\rhd_{A}v
=u(ϵ(h(2))h(1))aAv=u(h)aAv\displaystyle=u(\epsilon(h_{(2)})h_{(1)})a\rhd_{A}v=u(h)a\rhd_{A}v

where in the second, third and fourth equalities we use coassociativity, and the antipode and counit axioms for Hopf algebras respectively. Now note av=(fa)A(f′′Hv)=(fa)A(u(f′′)Av)=η(a)Ava\rhd_{\mathcal{F}}v=(f^{\prime}\rhd a)\rhd_{A}(f^{\prime\prime}\rhd_{H}v)=(f^{\prime}\rhd a)\rhd_{A}(u(f^{\prime\prime})\rhd_{A}v)=\eta(a)\rhd_{A}v, as required. ∎

Remark 2.9.

An immediate corollary to Proposition 2.8(3) is that, if ρ:AEnd(V)\rho:A\xrightarrow{}\text{End}(V) is the representation corresponding to A\rhd_{A}, and ρ\rho_{\mathcal{F}} is the representation corresponding to \rhd_{\mathcal{F}}, arising as a result of the Giaquinto and Zhang twist, then we have that ρ\rho_{\mathcal{F}} is given by the pullback of ρ\rho along η\eta, i.e. ρ=ρη\rho_{\mathcal{F}}=\rho\circ\eta. Additionally, if χ\chi and χ\chi_{\mathcal{F}} are the characters of ρ\rho and ρ\rho_{\mathcal{F}} respectively, then χ=χη\chi_{\mathcal{F}}=\chi\circ\eta.

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 CC is an algebra, then an algebra factorisation C=ABC=A\cdot B is a pair of subalgebras AA, BB of an associative unital kk-algebra CC such that the restriction of the multiplication map m:CCCm\colon{C\otimes C}\to C to ABA\otimes B is an isomorphism

m|AB:ABCm|_{A\otimes B}\colon A\otimes B\xrightarrow{\sim}C (6)

of kk-vector spaces. To an algebra factorisation there is associated a linear map ΨC:BAAB\Psi_{C}\colon B\otimes A\to A\otimes B given by

ΨC:BACC𝑚C(m|AB)1AB.\Psi_{C}\colon B\otimes A\hookrightarrow C\otimes C\xrightarrow{m}C\xrightarrow{(m|_{A\otimes B})^{-1}}A\otimes B. (7)

The map ΨC\Psi_{C} obeys the “generalised braiding” equations given by Majid in [18, Proposition 21.4].

If HH is a Hopf algebra, we consider an algebra factorisation in H-ModH\text{-}\mathrm{Mod} to be an algebra factorisation C=ABC=A\cdot B in which CC is an HH-module algebra, and AA, BB are HH-module subalgebras of CC. In this case, the isomorphism (6) and the map ΨC\Psi_{C} given by (7) are HH-module homomorphisms. Furthermore, given a 22-cocycle HH\mathcal{F}\in H\otimes H, the Drinfeld twists A,BA_{\mathcal{F}},\ B_{\mathcal{F}} of A,BA,\ B are clearly HH^{\mathcal{F}}-module subalgebras of the Drinfeld twist CC_{\mathcal{F}} of CC. Additionally, the following holds,

Proposition 2.11.

If C=ABC=A\cdot B is an algebra factorisation in H-ModH\text{-}\mathrm{Mod}, then CC_{\mathcal{F}} is an algebra factorisation C=ABC_{\mathcal{F}}=A_{\mathcal{F}}\cdot B_{\mathcal{F}} in H-ModH^{\mathcal{F}}\text{-}\mathrm{Mod}, and ΨC=()ΨC(1)\Psi_{C_{\mathcal{F}}}=(\mathcal{F}\rhd)\circ\Psi_{C}\circ(\mathcal{F}^{-1}\rhd).

Proof.

Recall from Proposition 2.2 that the product mm_{\mathcal{F}} on CC_{\mathcal{F}} is given by m=m(1)m_{\mathcal{F}}=m\circ(\mathcal{F}^{-1}\rhd), hence its restriction onto ABA_{\mathcal{F}}\otimes B_{\mathcal{F}} is the composition m|AB=m|AB(1)|ABm_{\mathcal{F}}|_{A_{\mathcal{F}}\otimes B_{\mathcal{F}}}=m|_{A\otimes B}\circ(\mathcal{F}^{-1}\rhd)|_{A\otimes B} where both maps on the right-hand side are bijective. Therefore, m|AB:ABCm_{\mathcal{F}}|_{A_{\mathcal{F}}\otimes B_{\mathcal{F}}}\colon A_{\mathcal{F}}\otimes B_{\mathcal{F}}\to C_{\mathcal{F}} is bijective, as required. Substituting the formulas for mm_{\mathcal{F}} and m|ABm_{\mathcal{F}}|_{A_{\mathcal{F}}\otimes B_{\mathcal{F}}} in (7), we obtain ΨC\Psi_{C_{\mathcal{F}}} as stated. ∎

2.6.   Induced representations.

Induced representations of associative algebras were defined by Higman [11], extending the notion from group theory. Given a subalgebra BB of an associative unital algebra CC (so that CC is a CCBB bimodule via the regular action), the induction functor

IndBC:B-ModC-Mod,IndBC(V)=CBV,\mathrm{Ind}_{B}^{C}\colon B\text{-}\mathrm{Mod}\to C\text{-}\mathrm{Mod},\qquad\mathrm{Ind}_{B}^{C}(V)=C\otimes_{B}V,

is a left adjoint of the restriction functor from CC to BB, see [22]. If C=ABC=A\cdot B is an algebra factorisation and VV is a BB-module, one has

IndBC(V)AV\mathrm{Ind}_{B}^{C}(V)\cong A\otimes V

as vector spaces, and moreover, as left AA-modules. The action of CC on IndBC(V)\mathrm{Ind}_{B}^{C}(V) is given by

C:CAV(m|AB)1idAVABAVidAΨCidVAABVmBAV\rhd_{C}\colon C\otimes A\otimes V\xrightarrow{(m|_{A\otimes B})^{-1}\otimes\text{id}_{A\otimes V}}A\otimes B\otimes A\otimes V\xrightarrow{\text{id}_{A}\otimes\Psi_{C}\otimes\text{id}_{V}}A\otimes A\otimes B\otimes V\xrightarrow{m\otimes\rhd_{B}}A\otimes V (8)

where B\rhd_{B} is the action of BB on VV. As pointed out above, when ABA\cdot B is an algebra factorisation in H-ModH\text{-}\mathrm{Mod}, then (6) and (7) are HH-module homomorphisms. Therefore, if V(H,B)-ModV\in(H,B)\text{-}\mathrm{Mod}, then, in view of (8), we see that IndBC(V)(H,C)-Mod\mathrm{Ind}_{B}^{C}(V)\in(H,C)\text{-}\mathrm{Mod}.

We now show that for algebra factorisations, the induction functor commutes with the twist functor. More precisely,

Proposition 2.12.

If C=ABC=A\cdot B is an algebra factorisation in H-ModH\text{-}\mathrm{Mod} and VV is a (H,B)(H,B)-module, then

IndBC(V)(IndBC(V))\mathrm{Ind}_{B_{\mathcal{F}}}^{C_{\mathcal{F}}}(V_{\mathcal{F}})\cong(\mathrm{Ind}_{B}^{C}(V))_{\mathcal{F}} (9)

as CC_{\mathcal{F}}-modules, where BB_{\mathcal{F}} and CC_{\mathcal{F}} are Drinfeld twists, and VV_{\mathcal{F}} and (IndBC(V))(\mathrm{Ind}_{B}^{C}(V))_{\mathcal{F}} are Giaquinto and Zhang twists, as in (3). The two CC_{\mathcal{F}}-module structures on the underlying vector space AVA\otimes V are intertwined by the map 1:AVAV\mathcal{F}^{-1}\rhd\colon A\otimes V\to A\otimes V.

Proof.

Above we explained that IndBC(V)(H,C)-Mod\mathrm{Ind}_{B}^{C}(V)\in(H,C)\text{-}\mathrm{Mod}, so firstly we find that the Giaquinto and Zhang twist (IndBC(V))(\mathrm{Ind}_{B}^{C}(V))_{\mathcal{F}} is a well-defined object.

Now by Proposition 2.11, the algebra CC_{\mathcal{F}} is generated by AA_{\mathcal{F}} and BB_{\mathcal{F}}, therefore it is enough to show that the two actions of AA_{\mathcal{F}} on the space AVA\otimes V are intertwined by 1\mathcal{F}^{-1}\rhd, and to do the same for BB_{\mathcal{F}}.

In the following diagram, the top row represents the action of AA_{\mathcal{F}} on the induced module IndBC(V)\mathrm{Ind}_{B_{\mathcal{F}}}^{C_{\mathcal{F}}}(V_{\mathcal{F}}): the composite arrow is midVm_{\mathcal{F}}\otimes\text{id}_{V}, which is indeed the action of AA_{\mathcal{F}} on the free left module AVA_{\mathcal{F}}\otimes V. The bottom row is the action of AA on IndBC(V)\mathrm{Ind}_{B}^{C}(V), twisted by \mathcal{F}.

AAV\textstyle{A\otimes A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(1)idV\scriptstyle{(\mathcal{F}^{-1}\rhd)\otimes\text{id}_{V}}idA(1)\scriptstyle{\text{id}_{A}\otimes(\mathcal{F}^{-1}\rhd)}AAV\textstyle{A\otimes A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}midV\scriptstyle{m\otimes\text{id}_{V}}(id)(1)\scriptstyle{(\triangle\otimes\text{id})(\mathcal{F}^{-1})\rhd}AV\textstyle{A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{\mathcal{F}^{-1}\rhd}AAV\textstyle{A\otimes A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(id)(1)\scriptstyle{(\text{id}\otimes\triangle)(\mathcal{F}^{-1})\rhd}AAV\textstyle{A\otimes A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}midV\scriptstyle{m\otimes\text{id}_{V}}AV\textstyle{A\otimes V}

The leftmost square of the diagram commutes by the cocycle equation, and the rightmost square commutes because m:AAAm\colon A\otimes A\to A is an HH-module morphism. Hence the diagram commutes, proving that 1\mathcal{F}^{-1}\rhd intertwines the two AA_{\mathcal{F}}-actions on AVA\otimes V.

The second diagram deals with the two actions of BB_{\mathcal{F}}. The top row consists in applying ΨC\Psi_{C_{\mathcal{F}}}, which by Proposition 2.11 is ()ΨC(1)(\mathcal{F}\rhd)\circ\Psi_{C}\circ(\mathcal{F}^{-1}\rhd) to the first two legs, followed by the action B=B(1)\rhd_{B_{\mathcal{F}}}=\rhd_{B}\circ(\mathcal{F}^{-1}\rhd) on VV. By (8), this gives the action of BB_{\mathcal{F}} on IndBC(V)\mathrm{Ind}_{B_{\mathcal{F}}}^{C_{\mathcal{F}}}(V_{\mathcal{F}}). The bottom row is the \mathcal{F}-twisted action of BB on IndBC(V)\mathrm{Ind}_{B}^{C}(V). To make the diagram more compact, we write 12\mathcal{F}_{12} to denote \mathcal{F} acting on the first two legs of the tensor product, i.e., the operator (id)(\mathcal{F}\rhd\otimes\text{id}); similarly for other operators:

BAV\textstyle{B\otimes A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}121\scriptstyle{\mathcal{F}^{-1}_{12}}231\scriptstyle{\mathcal{F}_{23}^{-1}}BAV\textstyle{B\otimes A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ΨC)12\scriptstyle{(\Psi_{C})_{12}}(id)(1)\scriptstyle{(\triangle\otimes\text{id})(\mathcal{F}^{-1})\rhd}ABV\textstyle{A\otimes B\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}12\scriptstyle{\mathcal{F}_{12}}(id)(1)\scriptstyle{(\triangle\otimes\text{id})(\mathcal{F}^{-1})\rhd}ABV\textstyle{A\otimes B\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}231\scriptstyle{\mathcal{F}_{23}^{-1}}ABV\textstyle{A\otimes B\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(B)23\scriptstyle{(\rhd_{B})_{23}}(id)(1)\scriptstyle{(\text{id}\otimes\triangle)(\mathcal{F}^{-1})\rhd}AV\textstyle{A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{\mathcal{F}^{-1}\rhd}BAV\textstyle{B\otimes A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(id)(1)\scriptstyle{(\text{id}\otimes\triangle)(\mathcal{F}^{-1})\rhd}BAV\textstyle{B\otimes A\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ΨC)12\scriptstyle{(\Psi_{C})_{12}}ABV\textstyle{A\otimes B\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ABV\textstyle{A\otimes B\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ABV\textstyle{A\otimes B\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(B)23\scriptstyle{(\rhd_{B})_{23}}AV\textstyle{A\otimes V}

The first (leftmost) square commutes by the cocycle equation, the second square commutes because ΨC\Psi_{C} is an HH-module morphism, the third square is the cocycle equation again, and the rightmost square commutes because the action B\rhd_{B} of BB is a morphism in H-ModH\text{-}\mathrm{Mod}. Hence the diagram commutes, and the two actions of BB_{\mathcal{F}} are indeed intertwined by 1\mathcal{F}^{-1}\rhd, 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 BnB_{n}. Here we will extend this result, showing how twisting permutes all of the irreducible characters of BnB_{n}. Additionally, we show how twisting induces a bijective correspondence between the irreducible characters of the Coxeter group of type DnD_{n} and those of its mystic partner μ(Dn)\mu(D_{n}). In particular we describe the permutation for BnB_{n}, and the bijective correspondence between DnD_{n} and μ(Dn)\mu(D_{n}), using a partition conjugation action.

Before discussing the Coxeter groups BnB_{n} and DnD_{n} let us first recall the family of complex reflection groups G(m,p,n)G(m,p,n), and their mystic counterparts μ(G(m,p,n))\mu(G(m,p,n)). For nn\in\mathbb{N}, let 𝔾n\mathbb{G}_{n} and 𝕊n\mathbb{S}_{n} denote the groups of n×nn\times n-monomial, and permutation, matrices over \mathbb{C} respectively.

Definition 3.1.

For m,p,nm,p,n\in\mathbb{N} with p|mp|m, define the following groups:

  • G(m,p,n)G(m,p,n) is the subgroup of 𝔾n\mathbb{G}_{n} formed by matrices where all the non-zero entries are mm-th roots of unity, and the product of all non-zero entries is an mp\frac{m}{p}-th root of unity.

  • μ(G(m,p,n))\mu(G(m,p,n)) is the subgroup of 𝔾n\mathbb{G}_{n} with non-zero entries being mm-th roots of unity, and such that the determinant of the matrix is an mp\frac{m}{p}-th root of unity.

We write (×)n(\mathbb{C}^{\times})^{n} to denote the subgroup of all diagonal matrices in 𝔾n\mathbb{G}_{n} and put T(m,p,n)=(×)nG(m,p,n)T(m,p,n)=(\mathbb{C}^{\times})^{n}\cap G(m,p,n). The diagonal matrix whose (i,i)(i,i)-th entry is ϵ\epsilon and the rest of diagonal entries are 11 will be denoted by ti(ϵ)t_{i}^{(\epsilon)}. The elements ti(1)t_{i}^{(-1)} play a special role in the paper, and we abbreviate ti(1)t_{i}^{(-1)} to tit_{i}.

Note that if mp\frac{m}{p} is even, then μ(G(m,p,n))=G(m,p,n)\mu(G(m,p,n))=G(m,p,n). Bazlov and Berenstein [1] defined a family of algebra isomorphisms Jc:𝔾n𝔾nJ_{c}:\mathbb{C}\mathbb{G}_{n}\xrightarrow{\sim}\mathbb{C}\mathbb{G}_{n}, which for certain choices of c×c\in\mathbb{C}^{\times} and m,pm,p\in\mathbb{N}, restrict to an isomorphism from μ(G(m,p,n))\mathbb{C}\mu(G(m,p,n)) to G(m,p,n)\mathbb{C}G(m,p,n). For c×,t(×)n,w𝕊nc\in\mathbb{C}^{\times},\ t\in(\mathbb{C}^{\times})^{n},\ w\in\mathbb{S}_{n}, the map JcJ_{c} is given by

Jc(wt):=wt{i<j|w(i)>w(j)}14((c+c1)(1titj)+(cc1+2)ti+(c1c+2)tj)J_{c}(wt):=wt\prod_{\{i<j|w(i)>w(j)\}}\frac{1}{4}\big{(}(c+c^{-1})(1-t_{i}t_{j})+(c-c^{-1}+2)t_{i}+(c^{-1}-c+2)t_{j}\big{)} (10)

When w=1w=1, the set {i<j|w(i)>w(j)}{\{i<j|w(i)>w(j)\}} is empty, and so Jc(t)=tt(×)nJ_{c}(t)=t\ \forall t\in(\mathbb{C}^{\times})^{n}. Therefore the interesting behaviour of JcJ_{c} comes from how it evaluates on the standard generators si, 1i<ns_{i},\ 1\leq i<n, of 𝕊n\mathbb{S}_{n}. We will be particularly interested in the cases where c=1c=1 and c=𝐢=1c=-\mathbf{i}=\sqrt{-1}. These evaluate on sis_{i} as follows:

J1(si)\displaystyle J_{1}(s_{i}) =12si(1+ti+ti+1titi+1)\displaystyle=\frac{1}{2}s_{i}(1+t_{i}+t_{i+1}-t_{i}t_{i+1}) (11)
J𝐢(si)\displaystyle J_{-\mathbf{i}}(s_{i}) =12si((1𝐢)ti+(1+𝐢)ti+1)\displaystyle=\frac{1}{2}s_{i}((1-\mathbf{i})t_{i}+(1+\mathbf{i})t_{i+1}) (12)

Note that when mp\frac{m}{p} is even, J1J_{1} restricts to an isomorphism μ(G(m,p,n))G(m,p,n)\mathbb{C}\mu(G(m,p,n))\xrightarrow{\sim}\mathbb{C}G(m,p,n). Later in Section 4.1 we will see that, when mp\frac{m}{p} is even, J1J_{1} in fact coincides precisely with the embedding map ηϕ\eta\phi of Theorem 4.2, after restricting ηϕ\eta\phi to the subalgebra μ(G(m,p,n))\mathbb{C}\mu(G(m,p,n)) of the negative braided Cherednik algebra H¯c¯(μ(G(m,p,n)))\underline{H}_{\underline{c}}(\mu(G(m,p,n))). Here ϕ\phi is an isomorphism, constructed in [2], between the H¯c¯(μ(G(m,p,n)))\underline{H}_{\underline{c}}(\mu(G(m,p,n))) and the twist Hc(G(m,p,n))H_{c}(G(m,p,n))_{\mathcal{F}} of the rational Cherednik algebra, and η\eta is a new map which arises from the results given in Section 2.

Now, since J1=ηϕJ_{1}=\eta\phi is an isomorphism, the pullback of a representation ρ:G(m,p,n)End(V)\rho:\mathbb{C}G(m,p,n)\xrightarrow{}\text{End}(V) along J1J_{1} gives rise to a representation ρJ1\rho\circ J_{1} of the group μ(G(m,p,n))\mu(G(m,p,n)). Based on Remark 2.9, pulling back ρ\rho along η\eta corresponds to twisting the representation of G(m,p,n)\mathbb{C}G(m,p,n) (in the sense of Giaquinto and Zhang) into a representation G(m,p,n)\mathbb{C}G(m,p,n)_{\mathcal{F}}. Pulling back ρη\rho\circ\eta along the map ϕ\phi (from [2, Theorem 5.2]) allows us to interpret the twisted representation ρη\rho\circ\eta as another group representation, in particular of μ(G(m,p,n))\mu(G(m,p,n)). Therefore we interpret the action of pulling back a representation of G(m,p,n)G(m,p,n) along J1J_{1} as the action on representations induced by twisting.

In Section 3.1 we describe this action explicity for the groups G(2,1,n)G(2,1,n), which are the Coxeter groups of type BnB_{n}. We then turn our attention to the map J𝐢J_{-\mathbf{i}} in (12). Although this map doesn’t directly arise from any twisting constructions, it turns out that pulling back the irreducible characters of BnB_{n} via J𝐢J_{-\mathbf{i}}, instead of J1J_{1}, induces the same permutation of characters. The advantage of J𝐢J_{-\mathbf{i}} however is that it restricts to an isomorphism μ(G(m,p,n))G(m,p,n)\mathbb{C}\mu(G(m,p,n))\xrightarrow{\sim}\mathbb{C}G(m,p,n) when mp\frac{m}{p} is odd and even. We can therefore use J𝐢J_{-\mathbf{i}} to map the irreducible characters of Coxeter group of type DnD_{n}, given by G(2,2,n)G(2,2,n), to those of μ(G(2,2,n))\mu(G(2,2,n)). We describe this mapping explicity in Section 3.2.

3.1.   Twisting irreducible characters of BnB_{n}.

Recall the group G(2,1,n)G(2,1,n) is isomorphic to the Coxeter group of type BnB_{n}. This group is well-known [8, Section 1.6.3] to have irreducible characters labelled by bipartitions of nn, i.e. ordered pairs (λ,μ)(\lambda,\mu) where λ\lambda and μ\mu are partitions of a,ba,b\in\mathbb{N} respectively, where a+b=na+b=n. Since mp\frac{m}{p} is even for this group, the map ηϕ\eta\phi of Theorem 4.2 restricts to an isomorphism μ(G(2,1,n))G(2,1,n)\mathbb{C}\mu(G(2,1,n))\xrightarrow{\sim}\mathbb{C}G(2,1,n). It is clear from Definition 3.1 that μ(G(2,1,n))=G(2,1,n)\mu(G(2,1,n))=G(2,1,n), and therefore we see ηϕ\eta\phi gives rise to an automorphism of G(2,1,n)\mathbb{C}G(2,1,n). So the pullback of an irreducible character of BnB_{n} along ηϕ\eta\phi will be another irreducible character of BnB_{n}. The following result describes this permutation of the characters explicity.

Proposition 3.2.

χ(λ,μ)ηϕ=χ(λ,μ)\chi_{(\lambda,\mu)}\circ\eta\phi=\chi_{(\lambda,\mu^{*})} where λa\lambda\vdash a and μb\mu\vdash b with a+b=na+b=n.

Proof.

By [8, Section 5.5.4], the irreducible characters of BnB_{n} are given as

χ(λ,μ)=IndBa×BbBn(χ~λ(ϵbχ~μ))\chi_{(\lambda,\mu)}=\text{Ind}^{B_{n}}_{B_{a}\times B_{b}}(\tilde{\chi}_{\lambda}\boxtimes(\epsilon^{\prime}_{b}\otimes\tilde{\chi}_{\mu}))

where χ~μIrr(Ba)\tilde{\chi}_{\mu}\in\text{Irr}(B_{a}) is the pullback of the irreducible character χλSa\chi_{\lambda}\in S_{a} (see [8, Definition 5.4.4]) along the projection map BaSaB_{a}\rightarrow S_{a}, \boxtimes denotes outer tensor product, and ϵb\epsilon^{\prime}_{b} is the restriction to BbB_{b} of the linear character ϵ\epsilon^{\prime} on BnB_{n} which sends ti1t_{i}\mapsto-1 and si+1s_{i}\mapsto+1. Precomposing with ηϕ\eta\phi we find:

χ(λ,μ)ηϕ\displaystyle\chi_{(\lambda,\mu)}\circ\eta\phi =IndBa×BbBn(χ~λ(ϵbχ~μ))ηϕ\displaystyle=\text{Ind}^{B_{n}}_{B_{a}\times B_{b}}(\tilde{\chi}_{\lambda}\boxtimes(\epsilon^{\prime}_{b}\otimes\tilde{\chi}_{\mu}))\circ\eta\phi
=IndBa×BbBn(χ(λ,)(ϵbχ(μ,)))ηϕ\displaystyle=\text{Ind}^{B_{n}}_{B_{a}\times B_{b}}(\chi_{(\lambda,\varnothing)}\boxtimes(\epsilon^{\prime}_{b}\otimes\chi_{(\mu,\varnothing)}))\circ\eta\phi
=IndBa×BbBn(χ(λ,)χ(,μ)))ηϕ.\displaystyle=\text{Ind}^{B_{n}}_{B_{a}\times B_{b}}(\chi_{(\lambda,\varnothing)}\boxtimes\chi_{(\varnothing,\mu)}))\circ\eta\phi.

The first equality just uses the definition of χ(λ,μ)\chi_{(\lambda,\mu)}, whilst the second rewrites χ~λ\tilde{\chi}_{\lambda} as χ(λ,)\chi_{(\lambda,\varnothing)}, i.e. via the unique bipartition of aa which characterises it. We do similarly for χ~μ\tilde{\chi}_{\mu}. In the 3rd equality we apply [8, Theorem 5.5.6(c)] which says ϵχ(λ,μ)=χ(μ,λ)\epsilon^{\prime}\otimes\chi_{(\lambda,\mu)}=\chi_{(\mu,\lambda)}.

Now note that the functor which sends a character ψ\psi of Ba×BbB_{a}\times B_{b} to the character IndBa×BbBn(ψ)ηϕ\text{Ind}^{B_{n}}_{B_{a}\times B_{b}}(\psi)\circ\eta\phi is a composition of two functors, the induction functor IndBa×BbBn:Rep(Ba×Bb)Rep(Bn)\text{Ind}^{B_{n}}_{B_{a}\times B_{b}}\colon\text{Rep}(B_{a}\times B_{b})\to\text{Rep}(B_{n}) followed by the autofunctor Precompηϕ\text{Precomp}_{\eta\phi} of Rep(Bn)\text{Rep}(B_{n}) given by precomposing a representation of Bn\mathbb{C}B_{n} with ηϕ\eta\phi. Note that Precompηϕ\text{Precomp}_{\eta\phi} is a permutation of irreducible characters of BnB_{n} which is involutive, because (ηϕ)2=(J1)2=id(\eta\phi)^{2}=(J_{1})^{2}=\text{id}. Hence Precompηϕ\text{Precomp}_{\eta\phi} is a self-adjoint functor. It follows, by Frobenius reciprocity, that PrecompηϕIndBa×BbBn\text{Precomp}_{\eta\phi}\circ\text{Ind}^{B_{n}}_{B_{a}\times B_{b}} is adjoint to the functor ResBa×BbBnPrecompηϕ:Rep(Bn)Rep(Ba×Bb)\text{Res}^{B_{n}}_{B_{a}\times B_{b}}\circ\text{Precomp}_{\eta\phi}\colon\text{Rep}(B_{n})\to\text{Rep}(B_{a}\times B_{b}). But since ηϕ\eta\phi restricted to BaBn\mathbb{C}B_{a}\subseteq\mathbb{C}B_{n} is an automorphism of the group algebra Ba\mathbb{C}B_{a}, same for BbB_{b}, it is obvious that this functor is the same as (Precompηϕ|BaPrecompηϕ|Bb)ResBa×BbBn(\text{Precomp}_{\eta\phi|_{B_{a}}}\boxtimes\text{Precomp}_{\eta\phi|_{B_{b}}})\circ\text{Res}^{B_{n}}_{B_{a}\times B_{b}}. Taking adjoints again, we conclude that

χ(λ,μ)ηϕ=IndBa×BbBn((χ(λ,)ηϕ|Ba)(χ(,μ)ηϕ|Bb)).\chi_{(\lambda,\mu)}\circ\eta\phi=\text{Ind}^{B_{n}}_{B_{a}\times B_{b}}((\chi_{(\lambda,\varnothing)}\circ\eta\phi|_{B_{a}})\boxtimes(\chi_{(\varnothing,\mu)}\circ\eta\phi|_{B_{b}})).

At this point we claim and prove the following,

χ(λ,)ηϕ|Ba=χ(λ,)χ(,μ)ηϕ|Bb=χ(,μ)\chi_{(\lambda,\varnothing)}\circ\eta\phi|_{B_{a}}=\chi_{(\lambda,\varnothing)}\hskip 28.45274pt\chi_{(\varnothing,\mu)}\circ\eta\phi|_{B_{b}}=\chi_{(\varnothing,\mu^{*})} (13)

The left hand identity follows since χ(λ,)=χλπ\chi_{(\lambda,\varnothing)}=\chi_{\lambda}\circ\pi for π:BnSn\pi:B_{n}\rightarrow S_{n} the standard projection map, and on noting that πηϕ=π\pi\circ\eta\phi=\pi, we deduce χ(λ,)ηϕ=χλπηϕ=χλπ=χ(λ,)\chi_{(\lambda,\varnothing)}\circ\eta\phi=\chi_{\lambda}\circ\pi\circ\eta\phi=\chi_{\lambda}\circ\pi=\chi_{(\lambda,\varnothing)}.

It remains to check the right hand identity. Firstly, (ϵbηϕ)(ti)=ϵb(ti)=1=ϵb(ti),i=a+1,,n(\epsilon^{\prime}_{b}\circ\eta\phi)(t_{i})=\epsilon^{\prime}_{b}(t_{i})=-1=\epsilon_{b}(t_{i}),\forall i=a+1,\dots,n. Also, (ϵbηϕ)(si)=ϵb(si12(1+ti+ti+1titi+1))=112(1111)=1=ϵb(si),i=a+1,,n1(\epsilon^{\prime}_{b}\circ\eta\phi)(s_{i})=\epsilon^{\prime}_{b}(s_{i}\cdot\frac{1}{2}(1+t_{i}+t_{i+1}-t_{i}t_{i+1}))=1\cdot\frac{1}{2}(1-1-1-1)=-1=\epsilon_{b}(s_{i}),\forall i=a+1,\dots,n-1. Since ϵbηϕ\epsilon^{\prime}_{b}\circ\eta\phi and ϵb\epsilon_{b} are algebra homomorphisms Bb\mathbb{C}B_{b}\to\mathbb{C} which agree on generators, one has ϵbηϕ=ϵb\epsilon^{\prime}_{b}\circ\eta\phi=\epsilon_{b}. So χ(,μ)ηϕ|Bb=(ϵbχ(μ,))ηϕ|Bb=(ϵbηϕ|Bb)(χ(μ,)ηϕ|Bb)=ϵbχ(μ,)=χ(,μ)\chi_{(\varnothing,\mu)}\circ\eta\phi|_{B_{b}}=(\epsilon^{\prime}_{b}\otimes\chi_{(\mu,\varnothing)})\circ\eta\phi|_{B_{b}}=(\epsilon^{\prime}_{b}\circ\eta\phi|_{B_{b}})\otimes(\chi_{(\mu,\varnothing)}\circ\eta\phi|_{B_{b}})=\epsilon_{b}\otimes\chi_{(\mu,\varnothing)}=\chi_{(\varnothing,\mu^{*})}, where the final equality again applies [8, Theorem 5.5.6(c)] which says ϵχ(λ,μ)=χ(μ,λ)\epsilon\otimes\chi_{(\lambda,\mu)}=\chi_{(\mu^{*},\lambda^{*})}.

With these two identities we find,

χ(λ,μ)ηϕ\displaystyle\chi_{(\lambda,\mu)}\circ\eta\phi =IndBa×BbBn((χ(λ,)ηϕ|Ba)(χ(,μ)ηϕ|Bb))\displaystyle=\text{Ind}^{B_{n}}_{B_{a}\times B_{b}}((\chi_{(\lambda,\varnothing)}\circ\eta\phi|_{B_{a}})\boxtimes(\chi_{(\varnothing,\mu)}\circ\eta\phi|_{B_{b}}))
=IndBa×BbBn(χ(λ,)χ(,μ))\displaystyle=\text{Ind}^{B_{n}}_{B_{a}\times B_{b}}(\chi_{(\lambda,\varnothing)}\boxtimes\chi_{(\varnothing,\mu^{*})})
=χ(λ,μ)\displaystyle=\chi_{(\lambda,\mu^{*})}

as required. ∎

The above result shows how ηϕ\eta\phi gives rise to permutation of the irreducible characters of the Coxeter group of type BnB_{n}. In the next section we consider the Coxeter groups of type DnD_{n}, which are isomorphic to the groups G(2,2,n)G(2,2,n). Since mp\frac{m}{p} is odd in this case, we cannot pull back the irreducible characters of G(2,2,n)G(2,2,n) along ηϕ\eta\phi. However, it was shown in [1, proof of Theorem 2.8] that the map JcJ_{c} in (10) with c=𝐢c=\mathbf{i} restricts to an isomorphism G(m,p,n)μ(G(m,p,n))\mathbb{C}G(m,p,n)\xrightarrow{\sim}\mathbb{C}\mu(G(m,p,n)) when mp\frac{m}{p} is both even and odd. Since (Jc)1=Jc1(J_{c})^{-1}=J_{c^{-1}}, we find J𝐢J_{-\mathbf{i}} restricts to an isomorphism μ(G(m,p,n))G(m,p,n)\mathbb{C}\mu(G(m,p,n))\xrightarrow{\sim}\mathbb{C}G(m,p,n). We can therefore ask how precomposition with J𝐢J_{-\mathbf{i}} acts on the irreducible characters of both BnB_{n} and DnD_{n}. It follows from the next result that J𝐢J_{-\mathbf{i}} permutes the characters of the BnB_{n} in exactly the same way as J1J_{1} (or ηϕ\eta\phi).

Proposition 3.3.

χ(λ,μ)J1=χ(λ,μ)J𝐢\chi_{(\lambda,\mu)}\circ J_{1}=\chi_{(\lambda,\mu)}\circ J_{-\mathbf{i}}.

Proof.

This follows by almost exactly the same argument as used in the proof of Proposition 3.2, with ηϕ\eta\phi replaced with J𝐢J_{-\mathbf{i}}. The only part of the proof which does not immediately follow also for J𝐢J_{-\mathbf{i}} is the version of equations (13) with J𝐢J_{-\mathbf{i}} in place of ηϕ\eta\phi. But we see these hold too since, for the first equation, (πJ𝐢)(t)=π(t)(\pi\circ J_{-\mathbf{i}})(t)=\pi(t) and (πJ𝐢)(si)=π(12si((1𝐢)ti+(1+𝐢)ti+1))=12((1𝐢)si+(1+𝐢)si)=si=π(si)(\pi\circ J_{-\mathbf{i}})(s_{i})=\pi(\frac{1}{2}s_{i}((1-\mathbf{i})t_{i}+(1+\mathbf{i})t_{i+1}))=\frac{1}{2}((1-\mathbf{i})s_{i}+(1+\mathbf{i})s_{i})=s_{i}=\pi(s_{i}), so πJ𝐢=π\pi\circ J_{-\mathbf{i}}=\pi as required. For the second equation: (ϵbJ𝐢)(ti)=ϵb(ti)=1=ϵb(ti)i=a+1,,n(\epsilon^{\prime}_{b}\circ J_{-\mathbf{i}})(t_{i})=\epsilon^{\prime}_{b}(t_{i})=-1=\epsilon_{b}(t_{i})\ \forall i=a+1,\dots,n and also, (ϵbJ𝐢)(si)=ϵb(12si((1𝐢)ti+(1+𝐢)ti+1))=1=ϵb(si)i=a+1,,n1(\epsilon^{\prime}_{b}\circ J_{-\mathbf{i}})(s_{i})=\epsilon^{\prime}_{b}(\frac{1}{2}s_{i}((1-\mathbf{i})t_{i}+(1+\mathbf{i})t_{i+1}))=-1=\epsilon_{b}(s_{i})\ \forall i=a+1,\dots,n-1. ∎

What is the reason why J𝐢J_{-\mathbf{i}} acts the same way on irreducible characters of BnB_{n} as J1=ηϕJ_{1}=\eta\phi? Using the fact (J1)2=id(J_{1})^{2}=\text{id}, we can reframe Proposition 3.3 as saying

χ(λ,μ)(J𝐢J1)=χ(λ,μ)\chi_{(\lambda,\mu)}\circ(J_{-\mathbf{i}}\circ J_{1})=\chi_{(\lambda,\mu)} (14)

So the characters of BnB_{n} are invariant under J𝐢J1J_{-\mathbf{i}}\circ J_{1}. This means that J𝐢J1J_{-\mathbf{i}}\circ J_{1} is an inner automorphism of G(2,1,n)\mathbb{C}G(2,1,n):

Lemma 3.4.

Suppose that GG is a finite group and J:GGJ\colon\mathbb{C}G\to\mathbb{C}G is an automorphism of its group algebra such that χJ=χ\chi\circ J=\chi for all irreducible characters χ\chi of GG. Then there exists invertible XGX\in\mathbb{C}G such that J(u)=XuX1J(u)=XuX^{-1} for all uGu\in\mathbb{C}G.

Proof.

Note that the condition χJ=χ\chi\circ J=\chi means that for all finite-dimensional G\mathbb{C}G-modules VV and for all uGu\in\mathbb{C}G, the trace TrV(J(u))\text{Tr}_{V}(J(u)) is equal to the trace TrV(u)\text{Tr}_{V}(u).

The centre Z(G)Z(\mathbb{C}G) of G\mathbb{C}G is spanned by primitive idempotents eχe_{\chi} labelled by irreducible characters χ\chi of GG. As an automorphism, JJ must permute the set {eχ}\{e_{\chi}\}. Suppose that J(eχ)=eψJ(e_{\chi})=e_{\psi}, and let VχV_{\chi} be a G\mathbb{C}G-module which affords χ\chi. Then TrVχ(eψ)=TrVχ(eχ)=dimVχ\text{Tr}_{V_{\chi}}(e_{\psi})=\text{Tr}_{V_{\chi}}(e_{\chi})=\dim V_{\chi}. However, if ψχ\psi\neq\chi then eψe_{\psi} acts on VχV_{\chi} by zero. Hence ψ=χ\psi=\chi, meaning that J|Z(G)J|_{Z(\mathbb{C}G)} is the identity map.

But then JJ fixes each component in the decomposition G=χeχG\mathbb{C}G=\prod_{\chi}e_{\chi}\mathbb{C}G of G\mathbb{C}G into the direct product of central simple algebras. By the Skolem-Noether theorem, J|eχGJ|_{e_{\chi}\mathbb{C}G} is given by uXχuXχ1u\mapsto X_{\chi}uX_{\chi}^{-1} for all χ\chi. Put X=χXχX=\prod_{\chi}X_{\chi}. ∎

We can conjecture from this that J1J_{1} and J𝐢J_{-\mathbf{i}} twist the characters of all the groups G(m,p,n)G(m,p,n), for mp\frac{m}{p} even, in the same way.

Conjecture 3.5.

If mp\frac{m}{p} is even and χ\chi is a character of G(m,p,n)G(m,p,n), then χJ1=χJ𝐢\chi\circ J_{1}=\chi\circ J_{-\mathbf{i}}.

3.2.   Twisting irreducible characters of DnD_{n}.

Having addressed the Coxeter group of type BnB_{n}, we turn to two of its index 22 normal subgroups, the Coxeter group of type DnD_{n} and its mystic counterpart μ(Dn)\mu(D_{n}), given by G(2,2,n)G(2,2,n) and μ(G(2,2,n))\mu(G(2,2,n)) respectively. We will use the map J𝐢J_{-\mathbf{i}} to twist the characters of DnD_{n} into those of μ(Dn)\mu(D_{n}). Let us recall how the irreducible characters of these groups relate to those of BnB_{n}.

The irreducibles of BnB_{n} are indexed by bipartitions (λ,μ)(\lambda,\mu), and, by [13, Corollary 6.19], the restriction to DnD_{n} of an irreducible character of BnB_{n} is either irreducible, or the sum of two distinct irreducibles. More specifically, for an irreducible V(λ,μ)V_{(\lambda,\mu)} of BnB_{n}, if λμ\lambda\neq\mu, then the restriction to DnD_{n} remains irreducible. Whereas when λ=μ\lambda=\mu, the restriction is a direct sum of two non-isomorphic irreducibles of DnD_{n}.

Now μ(Dn)\mu(D_{n}) is also an index 22 subgroup of BnB_{n}, so again by [13, Corollary 6.19] the restriction to μ(Dn)\mu(D_{n}) of an irreducible character of BnB_{n} will either be an irreducible, or the sum of two distinct irreducibles. In particular, μ(Dn)\mu(D_{n}) is the kernel of the linear character ϵ:Bn×,t1,s1\epsilon:B_{n}\rightarrow\mathbb{C}^{\times},\ t\mapsto-1,s\mapsto-1, and ϵV(λ,μ)=V(μ,λ)\epsilon\otimes V_{(\lambda,\mu)}=V_{(\mu^{*},\lambda^{*})}. We therefore find that, for λμ\lambda\neq\mu^{*}, the irreducible V(λ,μ)V_{(\lambda,\mu)} of BnB_{n} restricts to an irreducible of μ(Dn)\mu(D_{n}). Whilst for λ=μ\lambda=\mu^{*}, the restriction decomposes as a direct sum of two non-isomorphic irreducibles of μ(Dn)\mu(D_{n}).

In the following, let χ(λ,μ)\chi_{(\lambda,\mu)} denote the character of the irreducible representation V(λ,μ)V_{(\lambda,\mu)} of BnB_{n}. When λμ\lambda\neq\mu, let χ(λ,μ)Dn\chi^{D_{n}}_{(\lambda,\mu)} denote the irreducible character arising from the restriction of χ(λ,μ)\chi_{(\lambda,\mu)} to DnD_{n}, i.e. ResDn(χ(λ,μ))\mathrm{Res}^{D_{n}}(\chi_{(\lambda,\mu)}). Likewise, if λμ\lambda\neq\mu^{*}, let χ(λ,μ)μ(Dn):=Resμ(Dn)(χ(λ,μ))\chi^{\mu(D_{n})}_{(\lambda,\mu)}:=\mathrm{Res}^{\mu(D_{n})}(\chi_{(\lambda,\mu)}).

Proposition 3.6.

J𝐢J_{-\mathbf{i}} induces a bijection between the irreducible characters of DnD_{n} and the irreducible characters of μ(Dn)\mu(D_{n}). In particular, for λμ\lambda\neq\mu, χ(λ,μ)Dnχ(λ,μ)μ(Dn)\chi^{D_{n}}_{(\lambda,\mu)}\mapsto\chi^{\mu(D_{n})}_{(\lambda,\mu^{*})}.

Proof.

Precomposition with J𝐢J_{-\mathbf{i}} indeed gives a bijection of characters since J𝐢J_{-\mathbf{i}} restricts to an isomorphism μ(Dn)Dn\mathbb{C}\mu(D_{n})\xrightarrow{}\mathbb{C}D_{n}. For brevity let us identify J𝐢J_{-\mathbf{i}} with this restriction. We then have,

χ(λ,μ)DJ𝐢=ResDn(χ(λ,μ))J𝐢=Resμ(Dn)(χ(λ,μ)J𝐢)=Resμ(Dn)(χ(λ,μ))\chi^{D}_{(\lambda,\mu)}\circ J_{-\mathbf{i}}=\mathrm{Res}^{D_{n}}(\chi_{(\lambda,\mu)})\circ J_{-\mathbf{i}}=\mathrm{Res}^{\mu(D_{n})}(\chi_{(\lambda,\mu)}\circ J_{-\mathbf{i}})=\mathrm{Res}^{\mu(D_{n})}(\chi_{(\lambda,\mu^{*})})

where the third equality applies Proposition 3.2 and Proposition 3.3. Since λμ\lambda\neq\mu, we have λ(μ)\lambda\neq(\mu^{*})^{*}, and therefore restricting χ(λ,μ)\chi_{(\lambda,\mu^{*})} onto μ(Dn)\mu(D_{n}) gives an irreducible character, in particular χ(λ,μ)μ(Dn)\chi^{\mu(D_{n})}_{(\lambda,\mu^{*})}. ∎

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 nn\in\mathbb{N}, let VV be an nn-dimensional \mathbb{C}-vector space with basis x1,,xnx_{1},\dots,x_{n}, and also let y1,,yny_{1},\dots,y_{n} be the dual basis for VV^{*}. For 1i,j,kn1\leq i,j,k\leq n and ϵ×\epsilon\in\mathbb{C}^{\times} define the following maps in End(V)\text{End}(V):

sij(ϵ)(xk):={xk,ki,j,ϵ1xj,k=i,ϵxi,k=j.ti(ϵ)(xk):=ϵδikxk,σij(ϵ)(xk):={xkki,j,ϵ1xjk=i,ϵxik=j.s_{ij}^{(\epsilon)}(x_{k}):=\begin{cases}x_{k},&k\neq i,j,\\ \epsilon^{-1}x_{j},&k=i,\\ \epsilon x_{i},&k=j.\end{cases}\qquad t_{i}^{(\epsilon)}(x_{k}):=\epsilon^{\delta_{ik}}x_{k},\qquad\sigma_{ij}^{(\epsilon)}(x_{k}):=\begin{cases}x_{k}&k\neq i,j,\\ \epsilon^{-1}x_{j}&k=i,\\ -\epsilon x_{i}&k=j.\end{cases}

Fix parameters tt\in\mathbb{C} and c={c1,cζ|ζCmp\{1}}c=\{c_{1},\ c_{\zeta}\in\mathbb{C}\ |\ \zeta\in C_{\frac{m}{p}}\backslash\{1\}\} and c={c1,cζ|ζCmp\{1}}c^{\prime}=\{c^{\prime}_{1},\ c^{\prime}_{\zeta}\in\mathbb{C}\ |\ \zeta\in C_{\frac{m}{p}}\backslash\{1\}\}.

Definition 4.1.
  • The rational Cherednik algebra Ht,c(G(m,p,n))H_{t,c}(G(m,p,n)) is the \mathbb{C}-algebra generated by x1,,xnVx_{1},\dots,x_{n}\in V, y1,,ynVy_{1},\dots,y_{n}\in V^{*}, and gG(m,p,n)g\in G(m,p,n), subject to the relations:

    xixj\displaystyle x_{i}x_{j} =xjxi\displaystyle=x_{j}x_{i} gxi\displaystyle gx_{i} =g(xi)g\displaystyle=g(x_{i})g yixjxjyi\displaystyle y_{i}x_{j}-x_{j}y_{i} =c1ϵCmϵsij(ϵ)\displaystyle=c_{1}\sum_{\epsilon\in C_{m}}\epsilon s^{(\epsilon)}_{ij}
    yiyj\displaystyle y_{i}y_{j} =yjyi\displaystyle=y_{j}y_{i} gyi\displaystyle gy_{i} =g(yi)g\displaystyle=g(y_{i})g yixixiyi\displaystyle y_{i}x_{i}-x_{i}y_{i} =tc1jiϵCmsij(ϵ)ζCmp\{1}cζti(ζ)\displaystyle=t-c_{1}\sum_{j\neq i}\sum_{\epsilon\in C_{m}}s^{(\epsilon)}_{ij}-\sum_{\zeta\in C_{\frac{m}{p}}\backslash\{1\}}c_{\zeta}t_{i}^{(\zeta)}

    for 1i,jn1\leq i,j\leq n with iji\neq j.

  • The negative braided Cherednik algebra H¯t,c(μ(G(m,p,n)))\underline{H}_{t,c^{\prime}}(\mu(G(m,p,n))) is the \mathbb{C}-algebra generated by x1,,xnVx_{1},\dots,x_{n}\in V, y1,,ynVy_{1},\dots,y_{n}\in V^{*}, and gμ(G(m,p,n))g\in\mu(G(m,p,n)), subject to the relations:

    xixj\displaystyle x_{i}x_{j} =xjxi\displaystyle=-x_{j}x_{i} gxi\displaystyle gx_{i} =g(xi)g\displaystyle=g(x_{i})g yixj+xjyi\displaystyle y_{i}x_{j}+x_{j}y_{i} =c1ϵCmϵσij(ϵ)\displaystyle=c^{\prime}_{1}\sum_{\epsilon\in C_{m}}\epsilon\sigma^{(\epsilon)}_{ij}
    yiyj\displaystyle y_{i}y_{j} =yjyi\displaystyle=-y_{j}y_{i} gyi\displaystyle gy_{i} =g(yi)g\displaystyle=g(y_{i})g yixixiyi\displaystyle y_{i}x_{i}-x_{i}y_{i} =t+c1jiϵCmσij(ϵ)+ζCmp\{1}cζti(ζ)\displaystyle=t+c^{\prime}_{1}\sum_{j\neq i}\sum_{\epsilon\in C_{m}}\sigma^{(\epsilon)}_{ij}+\sum_{\zeta\in C_{\frac{m}{p}}\backslash\{1\}}c^{\prime}_{\zeta}t_{i}^{(\zeta)}

    for 1i,jn1\leq i,j\leq n with iji\neq j.

Let T:=t1,,tn|titj=tjti,ti2=1T:=\langle t_{1},\dots,t_{n}\ |\ t_{i}t_{j}=t_{j}t_{i},\ t_{i}^{2}=1\rangle, isomorphic to the nn-fold direct product of the cyclic group of order 22. By [2, Proposition 5.1], for mm even, Hc(G(m,p,n))H_{c}(G(m,p,n)) is a T\mathbb{C}T-module algebra under the following action:

tig=tigti,tixj=(1)δijxj,tiyj=(1)δijyjt_{i}\rhd g=t_{i}gt_{i},\hskip 10.00002ptt_{i}\rhd x_{j}=(-1)^{\delta_{ij}}x_{j},\hskip 10.00002ptt_{i}\rhd y_{j}=(-1)^{\delta_{ij}}y_{j}

for 1i,jn1\leq i,j\leq n and gG(m,p,n)g\in G(m,p,n). Also recall from [2, Lemma 4.5] that the following is a counital 22-cocycle of T\mathbb{C}T,

=1=1j<infij, where fij=12(11+γi1+1γjγiγj)\mathcal{F}=\mathcal{F}^{-1}=\prod_{1\leq j<i\leq n}f_{ij}\text{, where }f_{ij}=\frac{1}{2}(1\otimes 1+\gamma_{i}\otimes 1+1\otimes\gamma_{j}-\gamma_{i}\otimes\gamma_{j}) (15)

and by [2, Theorem 5.2], we have H¯c¯(μ(G(m,p,n)))Hc(G(m,p,n))\underline{H}_{\underline{c}}(\mu(G(m,p,n)))\xrightarrow{\sim}H_{c^{\prime}}(G(m,p,n))_{\mathcal{F}} where c1=c¯1c^{\prime}_{1}=-\underline{c}_{1} and cζ=c¯ζζCmp\{1}c^{\prime}_{\zeta}=-\underline{c}_{\zeta}\ \forall\zeta\in C_{\frac{m}{p}}\backslash\{1\}. Below in Theorem 4.2 we will denote the Drinfeld twist Hc(G(m,p,n))H_{c^{\prime}}(G(m,p,n))_{\mathcal{F}} by (Hc(G(m,p,n)),)(H_{c^{\prime}}(G(m,p,n)),\star), where \star is the twisted product of Hc(G(m,p,n))H_{c^{\prime}}(G(m,p,n)).

Note that when mp\frac{m}{p} is even, we have a natural embedding u:THc(G(m,p,n))u:\mathbb{C}T\rightarrow H_{c}(G(m,p,n)), and the action of T\mathbb{C}T on Hc(G(m,p,n))H_{c}(G(m,p,n)) is seen to be adjoint with respect to this embedding. Therefore, by Proposition 2.8(2), Hc(G(m,p,n))Hc(G(m,p,n))H_{c^{\prime}}(G(m,p,n))\cong H_{c^{\prime}}(G(m,p,n))_{\mathcal{F}}, and the module categories of these algebras must therefore be equivalent. In Theorem 4.2 we compute the isomorphism Hc(G(m,p,n))Hc(G(m,p,n))H_{c^{\prime}}(G(m,p,n))\cong H_{c^{\prime}}(G(m,p,n))_{\mathcal{F}} 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 H¯c¯(μ(G(m,p,n)))\underline{H}_{\underline{c}}(\mu(G(m,p,n))) and the rational Cherednik algebra Hc(G(m,p,n))H_{c}(G(m,p,n)). Note this isomorphism becomes an embedding when mp\frac{m}{p} is odd.

4.1.   Application to Cherednik algebras.

Theorem 4.2.

Consider the negative braided Cherednik algebra H¯c¯(μ(G(m,p,n))\underline{H}_{\underline{c}}(\mu(G(m,p,n)) where mm is even and n2n\geq 2. Define pp^{\prime}, c1c_{1} and cζc_{\zeta} as follows,

p={pif mp is even,p2if mp is odd,\displaystyle p^{\prime}=\begin{cases}p&\text{if }\frac{m}{p}\text{ is even,}\\ \frac{p}{2}&\text{if }\frac{m}{p}\text{ is odd,}\end{cases} c1=c¯1,cζ={c¯ζif ζCmp{1}0if ζCmpCmp.\displaystyle\hskip 20.00003ptc_{1}=-\underline{c}_{1},\hskip 20.00003ptc_{\zeta}=\begin{cases}-\underline{c}_{\zeta}&\text{if }\zeta\in C_{\frac{m}{p}}\setminus\{1\}\\ 0&\text{if }\zeta\in C_{\frac{m}{p^{\prime}}}\setminus C_{\frac{m}{p}}.\end{cases}

Then H¯c¯(μ(G(m,p,n))\underline{H}_{\underline{c}}(\mu(G(m,p,n)) embeds inside the rational Cherednik algebra Hc(G(m,p,n))H_{c}(G(m,p^{\prime},n)) via the following mapping of generators:

x¯i\displaystyle\underline{x}_{i} xiti1ti2t1,\displaystyle\mapsto x_{i}t_{i-1}t_{i-2}\dots t_{1},
y¯i\displaystyle\underline{y}_{i} yiti1ti2t1,\displaystyle\mapsto y_{i}t_{i-1}t_{i-2}\dots t_{1},
σi\displaystyle\sigma_{i} 12(si+s¯i+σiσi1)\displaystyle\mapsto\frac{1}{2}(s_{i}+\bar{s}_{i}+\sigma_{i}-\sigma^{-1}_{i})
t\displaystyle t ttT(m,p,n)\displaystyle\mapsto t\ \forall t\in T(m,p,n)
Proof.

Using the isomorphism ϕ\phi from [2, Theorem 5.2] we have an isomorphism H¯c¯(μ(G(m,p,n)))(Hc(G(m,p,n)),)\underline{H}_{\underline{c}}(\mu(G(m,p,n)))\xrightarrow{\sim}(H_{c^{\prime}}(G(m,p,n)),\star) where c1=c¯1c^{\prime}_{1}=-\underline{c}_{1} and cζ=c¯ζζCmp\{1}c^{\prime}_{\zeta}=-\underline{c}_{\zeta}\ \forall\zeta\in C_{\frac{m}{p}}\backslash\{1\}. Note that the rational Cherednik algebra Hc(G(m,p,n))H_{c^{\prime}}(G(m,p,n)) is a subalgebra of Hc(G(m,p2,n))H_{c}(G(m,\frac{p}{2},n)), with c1=c1c_{1}=c^{\prime}_{1} and

cζ:={cζif ζCmp0if ζC2mp\Cmpc_{\zeta}:=\begin{cases}c^{\prime}_{\zeta}&\text{if }\zeta\in C_{\frac{m}{p}}\\ 0&\text{if }\zeta\in C_{\frac{2m}{p}}\backslash C_{\frac{m}{p}}\end{cases}

Additionally for u:THc(G(m,p2,n))u:\mathbb{C}T\xrightarrow{}H_{c}(G(m,\frac{p}{2},n)) taken to be the natural embedding, T\mathbb{C}T acts adjointly (in the sense of Definition 2.7) on Hc(G(m,p2,n))H_{c}(G(m,\frac{p}{2},n)). On restricting to the subalgebra Hc(G(m,p,n))H_{c^{\prime}}(G(m,p,n)) this action coincides with the action given in [2, Proposition 5.1] which is used for twisting Hc(G(m,p,n))H_{c^{\prime}}(G(m,p,n)). Therefore the twisted algebra (Hc(G(m,p,n),)(H_{c^{\prime}}(G(m,p,n),\star) is a subalgebra of twisted algebra (Hc(G(m,p2,n)),)(H_{c}(G(m,\frac{p}{2},n)),\star).

Using the fact that the action of T\mathbb{C}T on Hc(G(m,p2,n))H_{c}(G(m,\frac{p}{2},n)) is adjoint, we can apply Proposition 2.8(1) to find (Hc(G(m,p2,n)),)Hc(G(m,p2,n))(H_{c}(G(m,\frac{p}{2},n)),\star)\cong H_{c}(G(m,\frac{p}{2},n)). We deduce that H¯c¯(μ(G(m,p,n))\underline{H}_{\underline{c}}(\mu(G(m,p,n)) is isomorphic to a subalgebra of Hc(G(m,p2,n))H_{c}(G(m,\frac{p}{2},n)) as required.

Next we inspect how the generators of H¯c¯(μ(G(m,p,n))\underline{H}_{\underline{c}}(\mu(G(m,p,n)) are mapped into Hc(G(m,p2,n))H_{c}(G(m,\frac{p}{2},n)). Recall that under ϕ\phi, σis¯i,tt,x¯ixi,y¯iyi\sigma_{i}\mapsto\bar{s}_{i},\ t\mapsto t,\ \underline{x}_{i}\mapsto x_{i},\ \underline{y}_{i}\mapsto y_{i}. It remains to show how the map η\eta defined in Proposition 2.8 maps s¯i,t,xi\bar{s}_{i},t,x_{i} and yiy_{i}. Let us use the cocycle given in (15). Let ηfij(a):=k,l=01(1)kl(γika)u(γjl)\eta_{f_{ij}}(a):=\sum_{k,l=0}^{1}(-1)^{kl}(\gamma_{i}^{k}\rhd a)u(\gamma_{j}^{l}). Then, using the commutativity of T\mathbb{C}T, we see η(a)=ηf2,1ηf3,1ηfn,n1\eta(a)=\eta_{f_{2,1}}\circ\eta_{f_{3,1}}\circ\dots\circ\eta_{f_{n,n-1}}, i.e. the composition (in any order) of ηfij\eta_{f_{ij}} for 1j<in1\leq j<i\leq n. Now u(γi)=tiu(\gamma_{i})=t_{i}, and it is easy to verify that

ηfij(xk)={xkif ikxktjif i=k\eta_{f_{ij}}(x_{k})=\begin{cases}x_{k}&\text{if }i\neq k\\ x_{k}t_{j}&\text{if }i=k\end{cases}

Therefore,

η(xk)\displaystyle\eta(x_{k}) =ηfk,1ηfk,2ηfk,k1(xk)\displaystyle=\eta_{f_{k,1}}\circ\eta_{f_{k,2}}\circ\dots\eta_{f_{k,k-1}}(x_{k})
=ηfk,1ηfk,2ηfk,k2(xktk1)\displaystyle=\eta_{f_{k,1}}\circ\eta_{f_{k,2}}\circ\dots\eta_{f_{k,k-2}}(x_{k}t_{k-1})
\displaystyle\dots
=xktk1t1\displaystyle=x_{k}t_{k-1}\dots t_{1}

One similarly shows that η(yk)=yktk1t1\eta(y_{k})=y_{k}t_{k-1}\dots t_{1}. Also, since T(m,p2,n)\mathbb{C}T(m,\frac{p}{2},n) is commutative, the adjoint action of the subalgebra T\mathbb{C}T on T(m,p2,n)\mathbb{C}T(m,\frac{p}{2},n) is trivial, so ηfij(t)=12(t1+t1+ttjttj)=t\eta_{f_{ij}}(t)=\frac{1}{2}(t\cdot 1+t\cdot 1+t\cdot t_{j}-t\cdot t_{j})=t and η(t)=t\eta(t)=t. Finally we check η(s¯i)\eta(\bar{s}_{i}). Using pj+qj=1p_{j}+q_{j}=1 we find

ηfij(s¯k)=s¯kpj+(γis¯k)qj={s¯kif ik,k+1s¯kpj+skqjif i=k,k+1\eta_{f_{ij}}(\bar{s}_{k})=\bar{s}_{k}p_{j}+(\gamma_{i}\rhd\bar{s}_{k})q_{j}=\begin{cases}\bar{s}_{k}&\text{if }i\neq k,k+1\\ \bar{s}_{k}p_{j}+s_{k}q_{j}&\text{if }i=k,k+1\end{cases}

Additionally,

ηfij(sk)={skif ik,k+1skpj+s¯kqjif i=k,k+1\eta_{f_{ij}}(s_{k})=\begin{cases}s_{k}&\text{if }i\neq k,k+1\\ s_{k}p_{j}+\bar{s}_{k}q_{j}&\text{if }i=k,k+1\end{cases}

So (ηfk+1,jηfk,j)(s¯k)=s¯k(\eta_{f_{k+1,j}}\circ\eta_{f_{k,j}})(\bar{s}_{k})=\bar{s}_{k} (using the fact pjp_{j} and qjq_{j} are orthogonal idempotents). Therefore η(s¯k)=ηfk+1,k(s¯k)=s¯kpk+skqk=12(sk+s¯k+σkσk1)\eta(\bar{s}_{k})=\eta_{f_{k+1,k}}(\bar{s}_{k})=\bar{s}_{k}p_{k}+s_{k}q_{k}=\frac{1}{2}(s_{k}+\bar{s}_{k}+\sigma_{k}-\sigma^{-1}_{k}). ∎

Since the embedding map of Theorem 4.2 is an isomorphism when mp\frac{m}{p} is even, we deduce:

Corollary 4.3.

For mp\frac{m}{p} even, the category of modules over the negative braided Cherednik algebra H¯c¯(μ(G(m,p,n))\underline{H}_{\underline{c}}(\mu(G(m,p,n)) is equivalent to the category of modules over the rational Cherednik algebra Hc(G(m,p,n))H_{c}(G(m,p,n)).

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 WW an irreducible complex reflection group with reflection representation VV, let =Hc(W)\mathcal{H}=H_{c}(W) be the associated rational Cherednik algebra (with t=1t=1).

Definition 4.4.

Let τ\tau be a simple W\mathbb{C}W-module, and extend to a S(V)WS(V)\rtimes W-module in which VV acts by zero, i.e. for pS(V),vτp\in S(V),v\in\tau, pv:=p(0)vp\rhd v:=p(0)v. Then the standard module associated to τ\tau is the Hc(W)H_{c}(W)-module Mc(τ):=Hc(W)S(V)WτM_{c}(\tau):=H_{c}(W)\otimes_{S(V)\rtimes W}\tau.

In the following we suppose mm is even, and let W=G(m,p,n)W=G(m,p,n) and μ(W)=μ(G(m,p,n))\mu(W)=\mu(G(m,p,n)). Let ¯=H¯c¯(μ(W))\underline{\mathcal{H}}=\underline{H}_{\underline{c}}(\mu(W)) be the negative braided Cherednik algebra isomorphic to the twist of \mathcal{H} under the map ϕ\phi from [2, Theorem 5.2]. Note that ¯\underline{\mathcal{H}} contains the subalgebra S1(V)μ(W)S_{-1}(V)\rtimes\mu(W), and ϕ:¯\phi:\underline{\mathcal{H}}\xrightarrow{\sim}\mathcal{H}_{\mathcal{F}} restricts to an isomorphism S1(V)μ(W)(S(V)W)S_{-1}(V)\rtimes\mu(W)\xrightarrow{\sim}(S(V)\rtimes W)_{\mathcal{F}}. We have also have the following natural analogue of Definition 4.4: for τ\tau^{\prime} a simple μ(W)\mathbb{C}\mu(W)-module, extend to a S1(V)μ(W)S_{-1}(V)\rtimes\mu(W)-module by letting VV act on τ\tau^{\prime} by 0. Then define the standard ¯\underline{\mathcal{H}}-module as M¯c¯(τ):=¯S1(V)μ(W)τ\underline{M}_{\underline{c}}(\tau^{\prime}):=\underline{\mathcal{H}}\otimes_{S_{-1}(V)\rtimes\mu(W)}\tau^{\prime}.

In the following we will prove several cases where the standard ¯\underline{\mathcal{H}}-module is (up to an isomorphism of categories) isomorphic to the Giaquinto and Zhang twist of the standard \mathcal{H}-module, Mc(τ)M_{c}(\tau)_{\mathcal{F}}. To prove this we first require the following elementary result.

Suppose ϕ:BB\phi:B\xrightarrow{\sim}B^{\prime} is an algebra isomorphism, AA is a subalgebra of BB, and A=ϕ(A)A^{\prime}=\phi(A). ϕ\phi defines an isomorphism of categories B-ModB-ModB^{\prime}\text{-}\mathrm{Mod}\xrightarrow{\sim}B\text{-}\mathrm{Mod}, whereby a BB^{\prime}-module (V,:BVV)(V,\ \rhd:B^{\prime}\otimes V\xrightarrow{}V) is sent to the BB-module (Vϕ:=V,ϕ:=(ϕidV))(V^{\phi}:=V,\ \rhd^{\phi}:=\rhd\circ(\phi\otimes\text{id}_{V})). Note that the map ϕ|A:AA\phi|_{A}:A\xrightarrow{\sim}A^{\prime} also defines an isomorphism of categories A-ModA-ModA^{\prime}\text{-}\mathrm{Mod}\xrightarrow{\sim}A\text{-}\mathrm{Mod}. The image of an AA^{\prime}-module (V,:AVV)(V,\ \rhd:A^{\prime}\otimes V\xrightarrow{}V) under this functor will be similarly denoted (Vϕ:=V,ϕ=(ϕ|AidV))(V^{\phi}:=V,\ \rhd^{\phi}=\rhd\circ(\phi|_{A}\otimes\text{id}_{V})).

Lemma 4.5.

If VV is an AA^{\prime}-module, then (IndAB(V))ϕIndAB(Vϕ)(\mathrm{Ind}_{A^{\prime}}^{B^{\prime}}(V))^{\phi}\cong\mathrm{Ind}_{A}^{B}(V^{\phi}) as BB-modules.

Proof.

We already proved a special case of this in the proof of Proposition 3.2. Note the functor Precompϕ\text{Precomp}_{\phi} has an adjoint given by Precompϕ1\text{Precomp}_{\phi^{-1}}. Also IndAB\mathrm{Ind}^{B}_{A} and IndAB\mathrm{Ind}^{B^{\prime}}_{A^{\prime}} are adjoint to ResAB\mathrm{Res}^{B}_{A} and ResAB\mathrm{Res}^{B^{\prime}}_{A^{\prime}} respectively. Therefore PrecompϕIndAB\text{Precomp}_{\phi}\circ\mathrm{Ind}^{B^{\prime}}_{A^{\prime}} has adjoint ResABPrecompϕ1\mathrm{Res}^{B^{\prime}}_{A^{\prime}}\circ\text{Precomp}_{\phi^{-1}}, which is easily seen to be equal to Precompϕ1|AResAB\text{Precomp}_{\phi^{-1}|_{A^{\prime}}}\circ\mathrm{Res}^{B}_{A}. The adjoint of this is then IndABPrecompϕ|A\mathrm{Ind}^{B}_{A}\circ\text{Precomp}_{\phi|_{A}}, as required. ∎

Recall the map J𝐢J_{-\mathbf{i}} of (12) restricts to an isomorphism μ(W)W\mathbb{C}\mu(W)\xrightarrow{\sim}\mathbb{C}W (regardless of the parity of mp\frac{m}{p}). This map therefore induces an isomorphism of categories W-Modμ(W)-Mod,ττ¯\mathbb{C}W\text{-}\mathrm{Mod}\xrightarrow{\sim}\mathbb{C}\mu(W)\text{-}\mathrm{Mod},\tau\mapsto\underline{\tau}, where τ¯=τ\underline{\tau}=\tau as a \mathbb{C}-vector space, and if :Wττ\rhd:\mathbb{C}W\otimes\tau\rightarrow\tau is the action of W\mathbb{C}W on τ\tau, then the action of μ(W)\mathbb{C}\mu(W) on τ¯\underline{\tau} is given by

(J𝐢idτ¯):μ(W)τ¯τ¯\rhd\circ(J_{-\mathbf{i}}\otimes\text{id}_{\underline{\tau}}):\mathbb{C}\mu(W)\otimes\underline{\tau}\rightarrow\underline{\tau}

In the above notation, τ¯=τJ𝐢\underline{\tau}=\tau^{J_{-\mathbf{i}}}. Similarly, the map ϕ:¯\phi:\underline{\mathcal{H}}\xrightarrow{\sim}\mathcal{H}_{\mathcal{F}} defines an isomorphism of categories -Mod¯-Mod\mathcal{H}_{\mathcal{F}}\text{-}\mathrm{Mod}\xrightarrow{\sim}\underline{\mathcal{H}}\text{-}\mathrm{Mod}, and below we take Mc(τ)ϕM_{c}(\tau)_{\mathcal{F}}^{\phi} to denote the image of Mc(τ)M_{c}(\tau)_{\mathcal{F}} under this functor.

Proposition 4.6.

Suppose that τ\tau is either a simple G(2,1,n)G(2,1,n)-module, or a simple G(2,2,n)G(2,2,n)-module corresponding to a bipartition (λ,μ)(\lambda,\mu) of nn where λμ\lambda\neq\mu (see Section 3.2). Then Mc(τ)ϕM¯c¯(τ¯)M_{c}(\tau)_{\mathcal{F}}^{\phi}\cong\underline{M}_{\underline{c}}(\underline{\tau}) as ¯\underline{\mathcal{H}}-modules.

Proof.

In following we will denote G(2,1,n)G(2,1,n) by BB, G(2,2,n)G(2,2,n) by DD, and μ(G(2,2,n))\mu(G(2,2,n)) by μ(D)\mu(D). Also let WW denote an arbitrary group G(m,p,n)G(m,p,n), where mm is even. Note Mc(τ)=IndS(V)W(τ)M_{c}(\tau)=\text{Ind}_{S(V)\rtimes W}^{\mathcal{H}}(\tau) and M¯c¯(τ¯)=IndS1(V)μ(W)¯(τ¯)\underline{M}_{\underline{c}}(\underline{\tau})=\mathrm{Ind}_{S_{-1}(V)\rtimes\mu(W)}^{\underline{\mathcal{H}}}(\underline{\tau}).

We wish to apply Proposition 2.12 to the Giaquinto-Zhang twist Mc(τ)=(IndS(V)W(τ))M_{c}(\tau)_{\mathcal{F}}=(\text{Ind}_{S(V)\rtimes W}^{\mathcal{H}}(\tau))_{\mathcal{F}}. To apply the proposition we first must note that =S(V)(S(V)W)\mathcal{H}=S(V^{*})\cdot(S(V)\rtimes W) is indeed an algebra factorisation in T-Mod\mathbb{C}T\text{-}\mathrm{Mod}, where recall TT is the group isomorphic to (C2)n(C_{2})^{n}. We also require τ(T,S(V)W)-Mod\tau\in(\mathbb{C}T,S(V)\rtimes W)\text{-}\mathrm{Mod}, which we check next. Note this category is well-defined since S(V)WS(V)\rtimes W is a T\mathbb{C}T-module algebra under the adjoint action of T\mathbb{C}T (in particular it is a T\mathbb{C}T-submodule algebra of \mathcal{H}). When mp\frac{m}{p} is even, then T\mathbb{C}T is a subalgebra of W\mathbb{C}W, and therefore τ\tau is naturally a T\mathbb{C}T-module by restricting the S(V)WS(V)\rtimes W action to T\mathbb{C}T. This action can equivalently be seen as pulling back the action of S(V)WS(V)\rtimes W along the embedding map TS(V)W\mathbb{C}T\hookrightarrow S(V)\rtimes W, and therefore we can apply Proposition 2.8(3) to deduce τ(T,S(V)W)-Mod\tau\in(\mathbb{C}T,S(V)\rtimes W)\text{-}\mathrm{Mod}.

However TT does not embed in D=G(2,2,n)D=G(2,2,n). By assumption though, τ\tau is equal to the simple D\mathbb{C}D-module τ(λ,μ)D\tau^{D}_{(\lambda,\mu)}, where λμ\lambda\neq\mu. On recalling Section 3.2, this module is the restriction to DD of either of the simple B\mathbb{C}B-modules τ(λ,μ)B\tau^{B}_{(\lambda,\mu)} or τ(μ,λ)B\tau^{B}_{(\mu,\lambda)}. Since T\mathbb{C}T is a subalgebra of B\mathbb{C}B, each of these B\mathbb{C}B-modules defines an action of T\mathbb{C}T on τ\tau. Note that on extending τ\tau to one these B\mathbb{C}B-modules, we can apply the mp\frac{m}{p} even case above to deduce that τ(T,S(V)B)-Mod\tau\in(\mathbb{C}T,S(V)\rtimes B)\text{-}\mathrm{Mod}. But then the compatibility condition (2) will still hold if we restrict the action of S(V)BS(V)\rtimes B on τ\tau to S(V)DS(V)\rtimes D, and therefore τ(T,S(V)D)-Mod\tau\in(\mathbb{C}T,S(V)\rtimes D)\text{-}\mathrm{Mod} as required. Later we show that this proposition holds independently of which action is chosen.

We can now apply Proposition 2.12 to deduce:

Mc(τ)Ind(S(V)W)(τ)M_{c}(\tau)_{\mathcal{F}}\cong\text{Ind}_{(S(V)\rtimes W)_{\mathcal{F}}}^{\mathcal{H}_{\mathcal{F}}}(\tau_{\mathcal{F}})

Here τ\tau_{\mathcal{F}} is the Giaquinto and Zhang twist of τ\tau (see (3)), so the action of (S(V)W)(S(V)\rtimes W)_{\mathcal{F}} on τ\tau_{\mathcal{F}} is given by :=(1)\rhd_{\mathcal{F}}:=\rhd\circ(\mathcal{F}^{-1}\blacktriangleright\ -\ ), where \rhd denotes the action of S(V)WS(V)\rtimes W on τ\tau, and \blacktriangleright denotes the action TT\mathbb{C}T\otimes\mathbb{C}T on (S(V)W)τ(S(V)\rtimes W)\otimes\tau. We now apply Lemma 4.5 to find,

Mc(τ)ϕ(Ind(S(V)W)(τ))ϕIndS1(V)μ(W)¯(τϕ)M_{c}(\tau)^{\phi}_{\mathcal{F}}\cong(\text{Ind}_{(S(V)\rtimes W)_{\mathcal{F}}}^{\mathcal{H}_{\mathcal{F}}}(\tau_{\mathcal{F}}))^{\phi}\cong\mathrm{Ind}^{\underline{\mathcal{H}}}_{S_{-1}(V)\rtimes\mu(W)}(\tau_{\mathcal{F}}^{\phi})

where τϕ=τ\tau^{\phi}_{\mathcal{F}}=\tau as a \mathbb{C}-vector space, and has action of S1(V)μ(W)S_{-1}(V)\rtimes\mu(W) given by ϕ=(ϕidτ)\rhd^{\phi}_{\mathcal{F}}=\rhd_{\mathcal{F}}\circ(\phi\otimes\text{id}_{\tau}).

To complete the proof it remains to check τϕ=τ¯\tau^{\phi}_{\mathcal{F}}=\underline{\tau} as S1(V)μ(W)S_{-1}(V)\rtimes\mu(W)-modules. We consider the cases of mp\frac{m}{p} even, and G(2,2,n)G(2,2,n), separately.

Let mp\frac{m}{p} be even. The action of T\mathbb{C}T on τ\tau is just a restriction of the action of S(V)WS(V)\rtimes W on τ\tau, and the action of T\mathbb{C}T on S(V)WS(V)\rtimes W is adjoint, so we can apply Proposition 2.8(3). We deduce that the action \rhd_{\mathcal{F}} of τ\tau_{\mathcal{F}} is given by the pulling back \rhd along the map η\eta, i.e. =(ηidτ)\rhd_{\mathcal{F}}=\rhd\circ(\eta\otimes\text{id}_{\tau}), where η:(S(V)W)S(V)W\eta:(S(V)\rtimes W)_{\mathcal{F}}\xrightarrow{\sim}S(V)\rtimes W was defined in the proof of Theorem 4.2. Then ϕ=(ηϕidτ)\rhd^{\phi}_{\mathcal{F}}=\rhd\circ(\eta\phi\otimes\text{id}_{\tau}), where ηϕ\eta\phi is regarded here as a map S1(V)μ(W)S(V)WS_{-1}(V)\rtimes\mu(W)\xrightarrow{\sim}S(V)\rtimes W.

Note that the map ηϕ\eta\phi is degree-preserving, and, by definition, \rhd is such that elements of degree 1\geq 1 in S(V)WS(V)\rtimes W act on τ\tau by 0. Therefore ϕ\rhd^{\phi}_{\mathcal{F}} is determined by how it behaves on the subalgebra μ(W)\mathbb{C}\mu(W) of S1(V)μ(W)S_{-1}(V)\rtimes\mu(W). Recall that the restriction of the map ηϕ\eta\phi to the subalgebra μ(W)\mathbb{C}\mu(W) coincides precisely with the map J1J_{1}, given in (11). Also, in Proposition 3.3 it was proven that χJ1=χJ𝐢\chi\circ J_{1}=\chi\circ J_{-\mathbf{i}}, for χ\chi an irreducible character of BnB_{n}. Therefore, as actions of μ(W)\mathbb{C}\mu(W) on τ\tau, we have (J1idτ)=(J𝐢idτ)\rhd\circ(J_{1}\otimes\text{id}_{\tau})=\rhd\circ(J_{-\mathbf{i}}\otimes\text{id}_{\tau}). Hence, as μ(W)\mathbb{C}\mu(W)-modules, τϕ=τ¯\tau^{\phi}_{\mathcal{F}}=\underline{\tau}. This implies they are also equal as S1(V)μ(W)S_{-1}(V)\rtimes\mu(W)-modules, since VV acts by 0 for both τϕ\tau^{\phi}_{\mathcal{F}} and τ¯\underline{\tau}. This proves the result for the group B=G(2,1,n)B=G(2,1,n).

Let us now check τϕ=τ¯\tau^{\phi}_{\mathcal{F}}=\underline{\tau}, as S1(V)μ(D)S_{-1}(V)\rtimes\mu(D)-modules, when τ=τ(λ,μ)D\tau=\tau^{D}_{(\lambda,\mu)} for λμ\lambda\neq\mu. As mentioned before, τ\tau can be extended to either of the simple B\mathbb{C}B-modules τ(λ,μ)B\tau^{B}_{(\lambda,\mu)} or τ(μ,λ)B\tau^{B}_{(\mu,\lambda)}. Suppose, for now, that we extend τ\tau to τ(λ,μ)B\tau^{B}_{(\lambda,\mu)}. Let us consider (τ(λ,μ)B)ϕ(\tau^{B}_{(\lambda,\mu)})^{\phi}_{\mathcal{F}}. From the mp\frac{m}{p} even case above we know that the action of B\mathbb{C}B on (τ(λ,μ)B)ϕ(\tau^{B}_{(\lambda,\mu)})^{\phi}_{\mathcal{F}} is given by (ηϕ|Bidτ)=(J1idτ)\rhd\circ(\eta\phi|_{\mathbb{C}B}\otimes\text{id}_{\tau})=\rhd\circ(J_{1}\otimes\text{id}_{\tau}). Therefore by Proposition 3.2, (τ(λ,μ)B)ϕ=τ(λ,μ)B(\tau^{B}_{(\lambda,\mu)})^{\phi}_{\mathcal{F}}=\tau^{B}_{(\lambda,\mu^{*})}. Since λ(μ)\lambda\neq(\mu^{*})^{*}, restricting the B\mathbb{C}B-action on τ(λ,μ)B\tau^{B}_{(\lambda,\mu^{*})} to μ(D)\mathbb{C}\mu(D) gives the irreducible representation τ(λ,μ)μ(D)\tau^{\mu(D)}_{(\lambda,\mu^{*})}. Let us now inspect τ¯\underline{\tau}, which is defined to be (τ(λ,μ)D)J𝐢(\tau^{D}_{(\lambda,\mu)})^{J_{-\mathbf{i}}}. By Proposition 3.6, (τ(λ,μ)D)J𝐢=τ(λ,μ)μ(D)(\tau^{D}_{(\lambda,\mu)})^{J_{-\mathbf{i}}}=\tau^{\mu(D)}_{(\lambda,\mu^{*})}. Therefore τϕ=τ¯\tau^{\phi}_{\mathcal{F}}=\underline{\tau} as μ(D)\mathbb{C}\mu(D)-modules, and so also as S1(V)μ(D)S_{-1}(V)\rtimes\mu(D)-modules, since each of these μ(D)\mathbb{C}\mu(D)-modules is extended by taking VV to act by 0.

Finally suppose we had instead extended τ(λ,μ)D\tau^{D}_{(\lambda,\mu)} to τ(μ,λ)B\tau^{B}_{(\mu,\lambda)}. Then on applying Proposition 3.2 we would have found (τ(μ,λ)B)ϕ=τ(μ,λ)B(\tau^{B}_{(\mu,\lambda)})^{\phi}_{\mathcal{F}}=\tau^{B}_{(\mu,\lambda^{*})}. Upon restricting this to μ(D)\mu(D) we arrive at the irreducible representation τ(μ,λ)μ(D)\tau^{\mu(D)}_{(\mu,\lambda^{*})}. But we see that τ(μ,λ)μ(D)=τ(λ,μ)μ(D)\tau^{\mu(D)}_{(\mu,\lambda^{*})}=\tau^{\mu(D)}_{(\lambda,\mu^{*})} since by [8, Theorem 5.5.6(c)], ϵτ(λ,μ)B=τ(μ,λ)B\epsilon\otimes\tau^{B}_{(\lambda,\mu)}=\tau^{B}_{(\mu^{*},\lambda^{*})}, where ϵ\epsilon is the linear character which satisfies μ(D)=ker(ϵ)\mu(D)=\ker(\epsilon).

5   Twisted coinvariant algebras

5.1.   Classical (commutative) coinvariant algebras.

Let VV be a finite-dimensional vector space over a field kk. Each finite subgroup GG of GL(V)\text{GL}(V) acts on the symmetric kk-algebra S(V)S(V) of VV. Classically, invariant theory of finite groups studies the graded algebra S(V)GS(V)^{G} of GG-invariants. Of particular interest is the case when GG is a finite reflection group, that is, when {sG:rank(idVs)=1}\{s\in G:\text{rank}(\text{id}_{V}-s)=1\} generates GG. In the non-modular case (when the characteristic of kk does not divide |G||G|), the work of Chevalley, Sheppard and Todd, and Serre in the 1950s established that GG is a reflection group, if and only if S(V)GS(V)^{G} is a polynomial algebra. This is known as the Chevalley-Shephard-Todd theorem, see for example [20, Theorem 7.1.4].

The coinvariant algebra SGS_{G} is defined as the quotient algebra S(V)/IGS(V)/I_{G} where IGI_{G} is the ideal of S(V)S(V) generated by homogeneous GG-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 kk is an arbitrary field and GG is a finite reflection group generated by reflections whose orders are prime to the characteristic of kk. Then the coinvariant algebra, viewed as a GG-module, affords the regular representation of GG.

5.2.   The noncommutative coinvariant algebra of a mystic reflection group.

Let kk be a field of characteristic 0. An extension of the Chevalley-Shephard-Todd theorem to skew-polynomial kk-algebras S𝐪(V)S_{\mathbf{q}}(V) is given by Kirkman, Kuzmanovich and Zhang in [14]. It characterises mystic reflection groups as finite subgroups WGL(V)W\subset\text{GL}(V) which act by algebra automorphisms on S𝐪(V)S_{\mathbf{q}}(V) such that the subalgebra S𝐪(V)WS_{\mathbf{q}}(V)^{W} is again skew-polynomial. Set k=k=\mathbb{C}; it turns out that the “building blocks” of mystic reflection groups are

  • the usual complex reflection groups, which act on the commutative polynomial algebra S(V)S(V),

  • the family μ(G(m,p,n))\mu(G(m,p,n)) of Definition 3.1 where mm is even, which act on the skew-polynomial algebra S1(V)S_{-1}(V).

Recall that the group μ(G(m,p,n))\mu(G(m,p,n)) is defined via its faithful action on a space VV with a fixed basis x1,,xnx_{1},\dots,x_{n}. We regard the graded vector space

S=(k1,,kn)0nx1k1xnkn,S=\bigoplus_{(k_{1},\dots,k_{n})\in\mathbb{Z}_{\geq 0}^{n}}\mathbb{C}x_{1}^{k_{1}}\dots x_{n}^{k_{n}},

spanned by the standard monomials, as the underlying space for both S(V)S(V) and S1(V)S_{-1}(V). The space SS contains

pk(m)=i=1nxikm,k=1,,n1;r(mp)=x1mpx2mpxnmp,p_{k}^{(m)}=\sum_{i=1}^{n}x_{i}^{km},\quad k=1,\dots,n-1;\qquad r^{(\frac{m}{p})}=x_{1}^{\frac{m}{p}}x_{2}^{\frac{m}{p}}\dots x_{n}^{\frac{m}{p}}, (16)

which are homogeneous algebraically independent generators, in S(V)S(V), of the subalgebra of G(m,p,n)G(m,p,n)-invariants, see [16, Chapter 2, Section 8]. The subalgebra of μ(G(m,p,n))\mu(G(m,p,n))-invariants in S1(V)S_{-1}(V) happens to form a commutative algebra isomorphic to S(V)G(m,p,n)S(V)^{G(m,p,n)}, as noted in [14].

Theorem 5.2 ([1], Theorem 2.6).

Let W=μ(G(m,p,n))W=\mu(G(m,p,n)). Then S1(V)WS_{-1}(V)^{W} is a commutative polynomial algebra, freely generated by p1(m),,pn1(m)p_{1}^{(m)},\dots,p_{n-1}^{(m)} and r(mp)r^{(\frac{m}{p})}.

Given that the mystic reflection groups μ(G(m,p,n))\mu(G(m,p,n)) have polynomial invariants in S1(V)S_{-1}(V), we define their coinvariant algebras and obtain the noncommutative analogue of Chevalley’s classical result.

In the following, take G=G(m,p,n)G=G(m,p,n) and W=μ(G(m,p,n))W=\mu(G(m,p,n)).

Definition 5.3.

Let IWI_{W} be the two-sided ideal of the algebra S1(V)S_{-1}(V) generated by homogeneous WW-invariants of positive degree. The noncommutative coinvariant algebra of WW is the quotient algebra S¯W=S1(V)/IW\underline{S}_{W}=S_{-1}(V)/I_{W}.

It is clear from the definition that the ideal IWI_{W} is a WW-submodule of S1(V)S_{-1}(V), and so WW acts on S¯W\underline{S}_{W} by algebra automorphisms — the action descends from S1(V)S_{-1}(V).

Theorem 5.4.

The noncommutative coinvariant algebra S¯W\underline{S}_{W} of the mystic reflection group W=μ(G(m,p,n))W=\mu(G(m,p,n)) affords the regular representation of WW.

We prove Theorem 5.4 over the next few sections, using the techniques we developed above for working with cocycle twists.

5.3.   WW-actions and Giaquinto-Zhang twists of GG-actions.

We are going to use the rational Cherednik algebra H:=H0,0(G)H:=H_{0,0}(G) of G=G(m,p,n)G=G(m,p,n) as well as the negative braided Cherednik algebra H¯:=H¯0,0(W)\underline{H}:=\underline{H}_{0,0}(W). We will make use only of the semidirect product relations in these algebras, so for simplicity we set the parameters t,ct,c to 0. To state the next Lemma, we need to recall from [2, Remark 5.10] that the group algebra W\mathbb{C}W is the middle term in the PBW-type algebra factorisation

H¯=S1(V)WS1(V).\underline{H}=S_{-1}(V)\otimes\mathbb{C}W\otimes S_{-1}(V^{*}). (17)

The result [2, Theorem 5.2], establishes an algebra isomorphism ϕ:H¯H\phi\colon\underline{H}\to H_{\mathcal{F}} between H¯\underline{H} and the cocycle twist of HH by \mathcal{F}. The map ϕ\phi is defined on generators of H¯\underline{H} by

ϕ(σi)=s¯i,ϕ(t)=t,ϕ(xj)=xj,ϕ(yj)=yj\phi(\sigma_{i})=\bar{s}_{i},\quad\phi(t)=t,\quad\phi(x_{j})=x_{j},\quad\phi(y_{j})=y_{j}

for i[n1]i\in[n-1], tT(m,p,n)t\in T(m,p,n) and j[n]j\in[n]. In particular, ϕ\phi restricts to an isomorphism W(G)\mathbb{C}W\to(\mathbb{C}G)_{\mathcal{F}}. Furthermore,

ϕ|S1(V)=idS:S1(V)S(V),\phi|_{S_{-1}(V)}=\text{id}_{S}\colon S_{-1}(V)\to S(V)_{\mathcal{F}}, (18)

i.e., ϕ\phi is the identity on the underlying vector space SS because it maps every standard monomial in x1,,xnx_{1},\dots,x_{n} to itself [2, section 5.9]. We show that ϕ\phi intertwines the actions of W\mathbb{C}W and (G)(\mathbb{C}G)_{\mathcal{F}} on SS:

Lemma 5.5.

Let ¯\underline{\rhd} be the action of W\mathbb{C}W on S1(V)S_{-1}(V), and let :(G)S(V)S(V)\rhd_{\mathcal{F}}\colon(\mathbb{C}G)_{\mathcal{F}}\otimes S(V)_{\mathcal{F}}\to S(V)_{\mathcal{F}} be the Giaquinto-Zhang twist by \mathcal{F} of the natural action of G\mathbb{C}G on S(V)S(V). Then for all uWu\in\mathbb{C}W, u¯=ϕ(u)u\mathop{\underline{\rhd}}=\phi(u)\rhd_{\mathcal{F}} as endomorphisms of the vector space SS.

Proof.

We bring in the Cherednik algeras. The relations in Definition 4.1 mean that the action ¯:WS1(V)S1(V)\underline{\rhd}\colon\mathbb{C}W\otimes S_{-1}(V)\to S_{-1}(V) can be written as the composite map

WS1(V)H¯H¯m¯H¯idS1(V)ϵWϵVS,\mathbb{C}W\otimes S_{-1}(V)\hookrightarrow\underline{H}\otimes\underline{H}\xrightarrow{\underline{m}}\underline{H}\xrightarrow{\text{id}_{S_{-1}(V)}\otimes\epsilon_{W}\otimes\epsilon_{V^{*}}}S, (19)

where m¯\underline{m} is the product map on H¯\underline{H}, ϵW:W\epsilon_{W}\colon\mathbb{C}W\to\mathbb{C} is the augmentation map of the group algebra W\mathbb{C}W, and ϵV:S1(V)\epsilon_{V^{*}}\colon S_{-1}(V^{*})\to\mathbb{C} maps a polynomial to its constant term. The domain of idS1(V)ϵWϵV\text{id}_{S_{-1}(V)}\otimes\epsilon_{W}\otimes\epsilon_{V^{*}} is the triple tensor product in (17). Thus, (19) says that the W\mathbb{C}W-module S1(V)S_{-1}(V) is the restriction to W\mathbb{C}W of the representation of H¯=S1(V)(S1(V)W)\underline{H}=S_{-1}(V)\cdot(S_{-1}(V^{*})\rtimes W) induced from the one-dimensional trivial module ϵWϵV\epsilon_{W}\otimes\epsilon_{V^{*}} of the semidirect product S1(V)WS_{-1}(V^{*})\rtimes W. Consider the diagram

WS1(V)\textstyle{\mathbb{C}W\otimes S_{-1}(V)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\subset}ϕid\scriptstyle{\phi\otimes\text{id}}H¯H¯\textstyle{\underline{H}\otimes\underline{H}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}m¯\scriptstyle{\underline{m}}ϕϕ\scriptstyle{\phi\otimes\phi}H¯\textstyle{\underline{H}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idϵWϵV\scriptstyle{\text{id}\otimes\epsilon_{W}\otimes\epsilon_{V^{*}}}ϕ\scriptstyle{\phi}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id(G)S(V)\textstyle{(\mathbb{C}G)_{\mathcal{F}}\otimes S(V)_{\mathcal{F}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\subset}HH\textstyle{H_{\mathcal{F}}\otimes H_{\mathcal{F}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}m\scriptstyle{m_{\mathcal{F}}}H\textstyle{H_{\mathcal{F}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idϵGϵV\scriptstyle{\text{id}\otimes\epsilon_{G}\otimes\epsilon_{V^{*}}}S\textstyle{S}

where the leftmost square commutes because ϕ\phi restricts to the identity on S1(V)S_{-1}(V), and the middle square commutes because ϕ\phi is an isomorphism of algebras.

We claim that the rightmost square commutes. This boils down to the equality eGϕ=eWe_{G}\phi=e_{W} of maps W\mathbb{C}W\to\mathbb{C}, where eG:(G)e_{G}\colon(\mathbb{C}G)_{\mathcal{F}}\to\mathbb{C} is the augmentation map of G\mathbb{C}G. This may not be obvious, since ϕ\phi does not send group elements of WW to group elements of GG. Note that eG:Ge_{G}\colon\mathbb{C}G\to\mathbb{C} is multiplicative on G\mathbb{C}G, being the trivial character of the group GG. Furthermore, since \mathbb{C} carries the trivial action of the group TT, the Giaquinto-Zhang twist of eGe_{G} is the same linear map eGe_{G}, viewed as a representation eG:(G)e_{G}\colon(\mathbb{C}G)_{\mathcal{F}}\to\mathbb{C} of the twisted algebra (G)(\mathbb{C}G)_{\mathcal{F}}. Therefore, eGe_{G} is multiplicative on (G)(\mathbb{C}G)_{\mathcal{F}}. Thus eGϕe_{G}\phi and eWe_{W} are two algebra homomorphisms from W\mathbb{C}W to \mathbb{C}; they both send grouplike generators σi\sigma_{i}, i[n]i\in[n], and tT(m,p,n)t\in T(m,p,n) to 11, hence coincide on W\mathbb{C}W.

We conclude that the above diagram commutes. But it shows that ϕ\phi intertwines the action ¯\underline{\rhd} of W\mathbb{C}W on S1(V)S_{-1}(V) with the (G)(\mathbb{C}G)_{\mathcal{F}}-action arising from the representation of

H=S(V)(G)S(V),H_{\mathcal{F}}=S(V)_{\mathcal{F}}\otimes(\mathbb{C}G)_{\mathcal{F}}\otimes S(V^{*})_{\mathcal{F}},

on the space S(V)S(V)_{\mathcal{F}}, induced from the trivial (G)S(V)(\mathbb{C}G)_{\mathcal{F}}S(V^{*})_{\mathcal{F}}-module. The latter is the Giaquinto-Zhang twist of the trivial (G)S(V)(\mathbb{C}G)S(V^{*})-module. Therefore, by Proposition 2.12, the HH_{\mathcal{F}}-action on S(V)S(V)_{\mathcal{F}} is the Giaquinto-Zhang twist of the standard HH-module S(V)S(V), arising from the trivial representation of G\mathbb{C}G. We know that the action of G\mathbb{C}G on this standard module is the natural action of G\mathbb{C}G on S(V)S(V), so the lemma is proved. ∎

Our next step is to replace S(V)S(V) and S1(V)S_{-1}(V) 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 IGI_{G} of S(V)S(V) and the ideal IWI_{W} of S1(V)S_{-1}(V) share the same underlying subspace II of SS. This subspace is stable under the action of the group TT.

Proof.

We temporarily write \cdot for S(V)S(V)-multiplication and \star for S1(V)S_{-1}(V)-multiplication on the space SS. Then IGI_{G} is the ()(\cdot)-ideal generated by p1(m),,pn1(m)p_{1}^{(m)},\dots,p_{n-1}^{(m)} and r(mp)r^{(\frac{m}{p})} as given in (16), and IWI_{W} is the ()(\star)-ideal with the same generators. Hence IWI_{W} is spanned by Mpi(m)NM\star p_{i}^{(m)}\star N and Mr(mp)NM\star r^{(\frac{m}{p})}\star N where M,NM,N are standard monomials and i[n]i\in[n]. Using the facts that pi(m)p_{i}^{(m)} is \star-central as mm is even, that r(mp)r^{(\frac{m}{p})} is a monomial, and that MN=±MNM\star N=\pm M\cdot N for all monomials M,NM,N, we conclude that all elements of this spanning set lie in IGI_{G}. So IWIGI_{W}\subseteq I_{G}. Similarly IGIWI_{G}\subseteq I_{W}. Note that T=γ1,,γnT=\langle\gamma_{1},\dots,\gamma_{n}\rangle acts on the space SS by automorphisms, and also by algebra automorphisms on S(V)S(V) and S1(V)S_{-1}(V). This action leaves pi(m)p_{i}^{(m)} invariant, and maps r(mp)r^{(\frac{m}{p})} to ±r(mp)\pm r^{(\frac{m}{p})}, hence also maps IGI_{G} to IGI_{G}. ∎

Corollary 5.7.

The algebras S¯W\underline{S}_{W} and SGS_{G} are defined on the same underlying vector space S/IS/I. As an algebra, S¯W\underline{S}_{W} is the cocycle twist (SG)(S_{G})_{\mathcal{F}} of SGS_{G}. The isomorphism ϕ:W(G)\phi\colon\mathbb{C}W\to(\mathbb{C}G)_{\mathcal{F}} intertwines the action ¯\underline{\rhd} of W\mathbb{C}W on S¯W\underline{S}_{W} and the Giaquinto-Zhang twist \rhd_{\mathcal{F}} of the action \rhd of G\mathbb{C}G on SGS_{G}.

Proof.

This follows from the fact that S1(V)=S(V)S_{-1}(V)=S(V)_{\mathcal{F}} (18) and Lemma 5.5, by passing to the quotient modulo I=IG=IWI=I_{G}=I_{W} (Lemma 5.6), which is an ideal in both algebras and is stable under all actions considered here. ∎

5.4.   The trace is unchanged under the twist.

To prove Theorem 5.4, we calculate the character of the representation of WW afforded by S¯W\underline{S}_{W}, that is, the trace TrS¯W(g¯)\text{Tr}_{\underline{S}_{W}}(g\underline{\rhd}) of the action ¯\underline{\rhd} of gWg\in W on S¯W\underline{S}_{W}. By Corollary 5.7, this is the same as Tr(SG)(ϕ(g))\text{Tr}_{(S_{G})_{\mathcal{F}}}(\phi(g)\rhd_{\mathcal{F}}). We will now show that neither the Giaquinto-Zhang twist by \mathcal{F} nor the application of ϕ\phi changes the trace. The former follows from a general result about Giaquinto-Zhang twists:

Proposition 5.8.

Let HH be a Hopf algebra with involutive antipode over a field kk, AA an HH-module algebra, VV a finite-dimensional (H,A)(H,A)-module where AA acts via \rhd, and HH\mathcal{F}\in H\otimes H a 22-cocycle. Denote the \mathcal{F}-twisted action of AA_{\mathcal{F}} on the space VV by \rhd_{\mathcal{F}}.

For any element aa of the vector space of AA, TrV(a)=TrV(a)\text{Tr}_{V}(a\rhd)=\text{Tr}_{V}(a\rhd_{\mathcal{F}}).

Proof.

Denote the action of hHh\in H on aAa\in A simply by h(a)h(a); same for the HH-action on VV and on VV^{*}. Recall [17, Proposition 9.3.3] that the HH-action on VV^{*} is such that the evaluation map ,:VVk\langle\cdot,\cdot\rangle\colon V^{*}\otimes V\to k and the coevaluation map kVVk\to V\otimes V^{*}, 1ixiyi1\mapsto\sum_{i}x_{i}\otimes y_{i}, are HH-module maps, where kk is the trivial HH-module. Equivalently, for all hHh\in H, xVx\in V, yVy\in V^{*},

y,h(x)=S1h(y),x,ixih(yi)=iS1h(xi)yi,\textstyle\langle y,h(x)\rangle=\langle S^{-1}h(y),x\rangle,\qquad\sum_{i}x_{i}\otimes h(y_{i})=\sum_{i}S^{-1}h(x_{i})\otimes y_{i}, (20)

where S:HHS\colon H\to H is the antipode. Since VV is an (H,A)(H,A)-module, it follows from (2) that for aAa\in A,

h(a)x=h(1)(aSh(2)(x)).h(a)\rhd x=h_{(1)}(a\rhd Sh_{(2)}(x)). (21)

By definition of trace, TrV(a)=iyi,axi\text{Tr}_{V}(a\rhd)=\sum_{i}\langle y_{i},a\rhd x_{i}\rangle. Writing 1HH\mathcal{F}^{-1}\in H\otimes H as ff′′f^{\prime}\otimes f^{\prime\prime} (summation understood), we compute the trace of the \mathcal{F}-twisted action of aa:

TrV(a)=iyi,axi\displaystyle\text{Tr}_{V}(a\rhd_{\mathcal{F}})=\sum_{i}\langle y_{i},a\rhd_{\mathcal{F}}x_{i}\rangle =iyi,f(a)f′′(xi)\displaystyle=\sum_{i}\langle y_{i},f^{\prime}(a)\rhd f^{\prime\prime}(x_{i})\rangle
=iyi,f(1)(aSf(2)f′′(xi))\displaystyle=\sum_{i}\langle y_{i},f^{\prime}_{(1)}(a\rhd Sf^{\prime}_{(2)}f^{\prime\prime}(x_{i}))\rangle
=iS1f(1)(yi),aSf(2)f′′(xi)\displaystyle=\sum_{i}\langle S^{-1}f^{\prime}_{(1)}(y_{i}),a\rhd Sf^{\prime}_{(2)}f^{\prime\prime}(x_{i})\rangle (22)

where we used definition (3) of \rhd_{\mathcal{F}}, (20) and (21). We now manipulate the coevaluation tensor:

Sf(2)f′′(xi)S1f(1)(yi)\displaystyle Sf^{\prime}_{(2)}f^{\prime\prime}(x_{i})\otimes S^{-1}f^{\prime}_{(1)}(y_{i}) =S2f(1)Sf(2)f′′(xi)yi\displaystyle=S^{-2}f^{\prime}_{(1)}Sf^{\prime}_{(2)}f^{\prime\prime}(x_{i})\otimes y_{i}
=f(1)Sf(2)f′′(xi)yi\displaystyle=f^{\prime}_{(1)}Sf^{\prime}_{(2)}f^{\prime\prime}(x_{i})\otimes y_{i}
=ϵ(f)f′′(xi)yi\displaystyle=\epsilon(f^{\prime})f^{\prime\prime}(x_{i})\otimes y_{i}

where we applied the second part of (20), the assumption that S2=idHS^{2}=\text{id}_{H} and the antipode law. Here ϵ:Hk\epsilon\colon H\to k is the counit. Since by Definition 2.1 \mathcal{F} is counital, (ϵidH)1=1(\epsilon\otimes\text{id}_{H})\mathcal{F}^{-1}=1 which means ϵ(f)f′′(xi)yi\epsilon(f^{\prime})f^{\prime\prime}(x_{i})\otimes y_{i} is equal to the coevaluation tensor xiyix_{i}\otimes y_{i}. Thus, (22) rewrites as iyi,axi\sum_{i}\langle y_{i},a\rhd x_{i}\rangle which is TrV(a)\text{Tr}_{V}(a\rhd). ∎

5.5.   The trace is unchanged under ϕ\phi; proof of Theorem 5.4.

If Γ\Gamma is any finite group, consider the linear map

χΓ:Γ, defined on the basis {ggΓ} of Γ by χΓ(g)={|Γ|g=1Γ,0g1Γ.\chi_{\mathbb{C}\Gamma}\colon\mathbb{C}\Gamma\to\mathbb{C},\text{ defined on the basis $\{g\mid g\in\Gamma\}$ of $\mathbb{C}\Gamma$ by }\chi_{\mathbb{C}\Gamma}(g)=\begin{cases}|\Gamma|&g=1_{\Gamma},\\ 0&g\neq 1_{\Gamma}.\end{cases}

Of course, we recognise χΓ\chi_{\mathbb{C}\Gamma} as the character of the regular representation of Γ\Gamma. Viewing ϕ:W(G)\phi\colon\mathbb{C}W\to(\mathbb{C}G)_{\mathcal{F}} as a linear map between the underlying vector spaces W\mathbb{C}W, G\mathbb{C}G, we show that ϕ\phi intertwines the regular characters:

Lemma 5.9.

χGϕ=χW\chi_{\mathbb{C}G}\circ\phi=\chi_{\mathbb{C}W}.

Proof.

We have G=G(m,p,n)=𝕊nT(m,p,n)G=G(m,p,n)=\mathbb{S}_{n}\ltimes T(m,p,n), so elements of GG are written as wtwt where w𝕊nw\in\mathbb{S}_{n} and tT(m,p,n)t\in T(m,p,n). By [2, Proposition 5.7], there exist elements θwT(2,1,n)\theta_{w}\in\mathbb{C}T(2,1,n), such that ϕ1(wt)=wθwt\phi^{-1}(wt)=w\theta_{w}t for all w,tw,t. Moreover, the proof of [2, Proposition 5.7] shows that θ1𝕊n=1\theta_{1_{\mathbb{S}_{n}}}=1.

So if w𝕊nw\in\mathbb{S}_{n}, w1𝕊nw\neq 1_{\mathbb{S}_{n}}, and tT(m,p,n)t\in T(m,p,n), then wt1Gwt\neq 1_{G} so χG(wt)=0\chi_{\mathbb{C}G}(wt)=0. On the other hand, χW(ϕ1(wt))=χW(wθwt)=0\chi_{\mathbb{C}W}(\phi^{-1}(wt))=\chi_{\mathbb{C}W}(w\theta_{w}t)=0 because w1𝕊nw\neq 1_{\mathbb{S}_{n}} guarantees that wθwtw\theta_{w}t is a linear combination of group elements of WW, none of which is 1W1_{W}.

In the case w=1𝕊nw=1_{\mathbb{S}_{n}}, tT(m,p,n)t\in T(m,p,n), we have ϕ1(t)=t\phi^{-1}(t)=t, hence χG(t)\chi_{\mathbb{C}G}(t) and χW(ϕ1(t))\chi_{\mathbb{C}W}(\phi^{-1}(t)) are both 0 if t1Gt\neq 1_{G} and are both |W|=|G||W|=|G| if t=1Gt=1_{G}. This shows that χWϕ1\chi_{\mathbb{C}W}\circ\phi^{-1} and χG\chi_{\mathbb{C}G} agree on all elements of GG, 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, TrS¯W(g¯)\text{Tr}_{\underline{S}_{W}}(g\underline{\rhd}), of the action of gWg\in W on S¯W\underline{S}_{W}, is equal to χW(g)\chi_{\mathbb{C}W}(g). By Corollary 5.7, the trace equals Tr(SG)(ϕ(g))\text{Tr}_{(S_{G})_{\mathcal{F}}}(\phi(g)\rhd_{\mathcal{F}}), which by Proposition 5.8 coincides with TrSG(ϕ(g))\text{Tr}_{S_{G}}(\phi(g)\rhd). By Chevalley’s Theorem 5.1, the latter is χG(ϕ(g))\chi_{\mathbb{C}G}(\phi(g)) which by Lemma 5.9 is χW(g)\chi_{\mathbb{C}W}(g), as claimed.

Our application of Proposition 5.8 is valid because the Hopf algebra T\mathbb{C}T 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 T\mathbb{C}T is, in fact, the identity map because every element of the group TT is self-inverse. ∎

6   Twists of restricted rational Cherednik algebras

6.1.   Restricted rational Cherednik algebras.

Let kk be a field, VV a finite-dimensional vector space over kk, and GGL(V)G\subset GL(V) a finite non-modular reflection group. The classical coinvariant algebra of GG appears as a PBW-type factor in the algebra factorisation of the restricted rational Cherednik algebra H¯c(G)\overline{H}_{c}(G). These finite-dimensional quotients of rational Cherednik algebras H0,c(G)H_{0,c}(G) at t=0t=0 originate from Gordon’s 2003 paper [10] and have attracted considerable interest since.

Of interest to us is the case G=G(m,p,n)G=G(m,p,n), and in this section we take kk to be a subfield of \mathbb{C} over which the reflection representation of G(m,p,n)G(m,p,n) is defined; the smallest such kk is therefore the mmth cyclotomic extension of \mathbb{Q}. The kk-rational Cherednik algebra H0,c(G)H_{0,c}(G), given by Definition 4.1, is defined over kk as long as the parameter cc is kk-valued. Thiel [23] defines the restricted rational Cherednik algebra as

H¯c(G)=H0,c(G)/(I,I)\overline{H}_{c}(G)=H_{0,c}(G)/(I,I^{\prime})

where II is the ideal generated by homogeneous GG-invariants of positive degree in S(V)S(V), II^{\prime} is the similar ideal in S(V)S(V^{*}), and (I,I)(I,I^{\prime}) is the ideal generated by II and II^{\prime} in H0,c(G)H_{0,c}(G). One has the following algebra factorisation,

H¯c(G)S(V)/IkGS(V)/I=SGkGSG,\overline{H}_{c}(G)\cong S(V)/I\otimes kG\otimes S(V^{*})/I^{\prime}=S_{G}\otimes kG\otimes S_{G}^{\prime},

where SGS_{G} is the coinvariant algebra of GG from the previous section, and SGS_{G}^{\prime} is defined in the same way from the action of GG on the dual space VV^{*}.

6.2.   Twists of restricted rational Cherednik algebras.

Let now W=μ(G(m,p,n))W=\mu(G(m,p,n)) be the mystic reflection group which is the mystic counterpart of GG. We define a finite-dimensional quotient of the negative braided Cherednik algebra H¯0,c¯(W)\underline{H}_{0,\underline{c}}(W) by

H¯¯c¯(W)=H¯0,c¯(W)/(IW,IW)\overline{\underline{H}}_{\underline{c}}(W)=\underline{H}_{0,\underline{c}}(W)/(I_{W},I_{W}^{\prime})

where IWS1(V)I_{W}\subseteq S_{-1}(V), IWS1(V)I_{W}^{\prime}\subseteq S_{-1}(V^{*}) 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 H¯c(G)\overline{H}_{c}(G) by the cocycle \mathcal{F} given in (15) is isomorphic to H¯¯c¯(W)\underline{\overline{H}}_{\underline{c}}(W). This algebra has triangular decomposition

H¯¯c¯(W)S1(V)/IWkGS1(V)/IW=S¯WkGS¯W\underline{\overline{H}}_{\underline{c}}(W)\cong S_{-1}(V)/I_{W}\otimes kG\otimes S_{-1}(V^{*})/I_{W}^{\prime}=\underline{S}_{W}\otimes kG\otimes\underline{S}_{W}^{\prime}

into the group algebra of WW and two noncommutative coinvariant algebras.

We further observe the following,

Proposition 6.2.

If mp\frac{m}{p} is even, then H¯¯c¯(W)\underline{\overline{H}}_{\underline{c}}(W) and H¯c(G)\overline{H}_{c}(G) are isomorphic as algebras.

Proof.

Indeed, Proposition 2.8 applies because the action of the Hopf algebra kTkT on H¯c(G(m,p,n))\overline{H}_{c}(G(m,p,n)) is adjoint, as it factors via the embedding of TT as the subgroup T(2,1,n)T(2,1,n) in G(m,p,n)G(m,p,n). ∎

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 AA_{\mathcal{F}} and AA over an arbitrary field kk, as long as the HH-module structure on AA and the cocycle \mathcal{F} are defined over kk, and the HH-action on AA 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 \mathcal{F}.

Of course, we need mp\frac{m}{p} to be odd, and we consider the case m=p=n=2m=p=n=2. Note G(2,2,2)G(2,2,2) is the Klein four-group generated by the commuting involutions s12s_{12} and s¯12\bar{s}_{12}. On the other hand, μ(G(2,2,2))\mu(G(2,2,2)) is the cyclic group of order 44 generated by the mystic reflection σ12=(0110)\sigma_{12}=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}. We see kk the minimal field over which the reflection representation of G(2,2,2)G(2,2,2) and the mystic reflection representation of μ(G(2,2,2))\mu(G(2,2,2)) are defined is the rationals, k=k=\mathbb{Q}.

The proof of the following result is partly based on an explicit computer calculation (over the rational field).

Proposition 6.3.

If c=1c=1, then the \mathbb{Q}-algebras H¯c(G(2,2,2))\overline{H}_{c}(G(2,2,2)) and H¯¯c(μ(G(2,2,2)))\overline{\underline{H}}_{-c}(\mu(G(2,2,2))) are not isomorphic.

Proof.

Note that G(2,2,2)G(2,2,2) is the non-reduced Coxeter group of type A1×A1A_{1}\times A_{1}, so

H¯c(G(2,2,2))H¯c(G(2,1,1))H¯c(G(2,1,1))\overline{H}_{c}(G(2,2,2))\cong\overline{H}_{c}(G(2,1,1))\otimes\overline{H}_{c}(G(2,1,1))

where G(2,1,1)G(2,1,1) is the reflection group {±1}\{\pm 1\} of order 22 which operates on the 11-dimensional space. Thus, H¯c(G(2,2,2))\overline{H}_{c}(G(2,2,2)) is a 6464-dimensional algebra which is a tensor product of two (commuting) 88-dimensional subalgebras. Explicitly, H¯c(G(2,1,1))\overline{H}_{c}(G(2,1,1)) is presented as

y,s,xs2=1,sx=xs,ys=sy,yx=xy2cs,x2=y2=0.\langle y,s,x\mid s^{2}=1,\ sx=-xs,\ ys=-sy,\ yx=xy-2cs,\ x^{2}=y^{2}=0\rangle.

It follows that

Z(H¯c(G(2,2,2)))Z(H¯c(G(2,1,1)))Z(H¯c(G(2,1,1))).Z(\overline{H}_{c}(G(2,2,2)))\cong Z(\overline{H}_{c}(G(2,1,1)))\otimes Z(\overline{H}_{c}(G(2,1,1))).

By results of [12], the centre Z(H¯c(G(2,1,1)))Z(\overline{H}_{c}(G(2,1,1))) is two-dimensional. Consider z=xycsz=xy-cs. We claim that zz is central in H¯c(G(2,1,1)))\overline{H}_{c}(G(2,1,1))). Indeed, sz=zssz=zs, and

zx=xyxcsx=x(xy2cs)+cxs=0cxs=cxs,xz=cxs=cxszx=xyx-csx=x(xy-2cs)+cxs=0-cxs=-cxs,\quad xz=-cxs=-cxs

so zz commutes with xx. Similarly, zz commutes with yy. We conclude that Z(H¯c(G(2,1,1)))=1+zZ(\overline{H}_{c}(G(2,1,1)))=\mathbb{Q}1+\mathbb{Q}z. Observe that

z2=zxyczs=(cxs)y+cxsy+c2=c2.z^{2}=zxy-czs=(-cxs)y+cxsy+c^{2}=c^{2}.

Since c=1c=1 in our example, z2=1z^{2}=1. In particular, the algebra Z(H¯c(G(2,1,1)))Z(\overline{H}_{c}(G(2,1,1))) does not contain elements of multiplicative order 44, hence such elements do not exist in Z(H¯c(G(2,2,2)))Z(\overline{H}_{c}(G(2,2,2))). On the other hand, we claim that

γ=\displaystyle\gamma= 14[2c22cσ+2c2σ2+2cσ3\displaystyle\frac{1}{4}\big{[}2c^{2}-2c\sigma+2c^{2}\sigma^{2}+2c\sigma^{3}
+cx1σy1cx1σy22x1σ2y2+cx2σy1\displaystyle+cx_{1}\sigma y_{1}-cx_{1}\sigma y_{2}-2x_{1}\sigma^{2}y_{2}+cx_{2}\sigma y_{1}
+cx2σy2+2x2σ2y1+2x12y12+cx1σ3y1\displaystyle+cx_{2}\sigma y_{2}+2x_{2}\sigma^{2}y_{1}+2x_{1}^{2}y_{1}^{2}+cx_{1}\sigma^{3}y_{1}
+cx1σ3y2cx2σ3y1+cx2σ3y2]\displaystyle+cx_{1}\sigma^{3}y_{2}-cx_{2}\sigma^{3}y_{1}+cx_{2}\sigma^{3}y_{2}\big{]}

is a central element of H¯¯c(μ(G(2,2,2)))\overline{\underline{H}}_{c}(\mu(G(2,2,2))). Moreover, γ2\gamma^{2} is not a scalar whereas γ4=c4\gamma^{4}=c^{4}. This is a result of a computer calculation. Hence, if c=±1c=\pm 1, the centre of H¯¯c(μ(G(2,2,2)))\overline{\underline{H}}_{c}(\mu(G(2,2,2))) contains an element of multiplicative order 44, which shows that the two algebras are not isomorphic over \mathbb{Q}. ∎

We remark that the above example does not work over \mathbb{C}, or indeed over an extension of \mathbb{Q} containing 1\sqrt{-1}. If scalars are extended to (1)\mathbb{Q}(\sqrt{-1}), the two algebras become isomorphic. We currently have no explicit example where we could disprove isomorphism over \mathbb{C}.

We finish with the following conjecture.

Conjecture 6.4.

Let mm be even and mp\frac{m}{p} be odd, and let k=(1m)k=\mathbb{Q}(\sqrt[m]{1}). The kk-algebras H¯¯c¯(μ(G(m,p,n)))\underline{\overline{H}}_{\underline{c}}(\mu(G(m,p,n))) and H¯c(G(m,p,n))\overline{H}_{c}(G(m,p,n)) 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.