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

Cocycle twisting of semidirect products and transmutation

Erik Habbestad Department of Mathematics, University of Oslo, P.O. Box 1053 Blindern, NO-0316 Oslo, Norway [email protected]  and  Sergey Neshveyev [email protected]
(Date: April 2, 2023; revised: January 11, 2024)
Abstract.

We apply Majid’s transmutation procedure to Hopf algebra maps H[T]H\to{\mathbb{C}}[T], where TT is a compact abelian group, and explain how this construction gives rise to braided Hopf algebras over quotients of TT by subgroups that are cocentral in HH. This allows us to unify and generalize a number of recent constructions of braided compact quantum groups, starting from the braided SUq(2)SU_{q}(2) quantum group, and describe their bosonizations.

Supported by the NFR project 300837 “Quantum Symmetry”.

Introduction

Assume that HH and KK are Hopf algebras together with Hopf algebra maps

p:HK,i:KH,pi=idK.p\colon H\to K,\quad i\colon K\to H,\quad p\circ i=\mathrm{id}_{K}.

This setting was considered in 1985 by Radford in his paper The structure of Hopf algebras with a projection, [Rad85]. He explained that this data is equivalent to having an object AA in the category of Yetter–Drinfeld modules over KK satisfying certain conditions reminiscent of a Hopf algebra. This object is often not a genuine Hopf algebra, because the algebra structure considered on AAA\otimes A involves the braiding on the Yetter–Drinfeld modules. Today, following Majid, it is common to call AA a braided Hopf algebra. It is an example of a Hopf algebra object in a braided monoidal category.

As for genuine Hopf algebras, there is a Tannaka–Krein type reconstruction theorem for braided Hopf algebras [Maj95, Chapter 9]. That is, under certain representability conditions, a monoidal functor 𝒞𝒟\mathcal{C}\to\mathcal{D}, where 𝒞\mathcal{C} is rigid and 𝒟\mathcal{D} is braided, gives rise to a Hopf algebra object in 𝒟\mathcal{D}. A particular example of this is when the functor is induced from a map π:HK\pi\colon H\to K, where HH is a Hopf algebra and KK is a coquasitriangular Hopf algebra: the natural monoidal functor between the categories of comodules

π:HK,\mathcal{F}_{\pi}\colon\mathcal{M}^{H}\to\mathcal{M}^{K},

gives rise to a braided Hopf algebra, which Majid calls a transmutation of HH, [Maj93].

More recently, there have been a number of constructions of braided compact quantum groups, [Ans+22], [BJR22], [Kas+16], [MR22]. For the authors of the present paper they showed up “in nature” through a connection with C-algebras arising from certain subproduct systems, [HN22]. In particular, a Cuntz–Pimsner type algebra associated to the noncommutative polynomial X1X2q¯X2X1X_{1}X_{2}-\bar{q}X_{2}X_{1} turned out to be the C-algebra of continuous functions on the braided SUq(2)SU_{q}(2) quantum group, constructed in [Kas+16]. Although the analytic/C-algebraic aspects are therefore important to us, compactness nevertheless allows one to treat an essential part of the theory purely algebraically.

In this paper our starting point is a map π:H[T]\pi\colon H\to{\mathbb{C}}[T], where HH is a Hopf *-algebra and TT is a compact abelian group. We observe that the resulting transmutation may be viewed as a braided Hopf *-algebra over the quotients of TT by subgroups T0T_{0} that are cocentral in HH. Using a theorem due to Majid we describe the corresponding Hopf algebra with projection, called bosonization, in terms of 22-cocycle twists of [T/T0]H{\mathbb{C}}[T/T_{0}]\ltimes H. This allows us to treat a number of examples in a unified and efficient way.

The paper is organized as follows. In Section 1 we collect some facts about braided Hopf *-algebras, twisting and transmutation. Section 2 contains our general results. Here we also suggest a definition of braided compact matrix quantum groups, covering in particular transmutations of compact matrix quantum groups, and discuss a connection of transmutation with a recent construction of Bochniak and Sitarz [BS19]. Section 3, which constitutes a large part of this text, consists of computing examples.


1. Generalities

We refer the reader to Majid’s book [Maj95] for more information on the notions in this section. We only consider unital algebras over the complex numbers.

1.1. Categories of comodules

Let (H,Δ,S,ε)(H,\Delta,S,\varepsilon) be a Hopf *-algebra. As is usual in the theory of Hopf algebras, we adopt Sweedler’s sumless notation: we write Δ(h)=h(1)h(2)\Delta(h)=h_{(1)}\otimes h_{(2)}, but remember that the expression represents a sum of simple tensors. We write H\mathcal{M}^{H} for the monoidal category of right HH-comodules. For an object MHM\in\mathcal{M}^{H} with corresponding map δM:MMH\delta_{M}\colon M\to M\otimes H, we write δM(m)=m(1)m(2)\delta_{M}(m)=m^{(1)}\otimes m^{(2)}.

Recall that a coquasitriangular structure on HH is a linear map R:HHR\colon H\otimes H\to{\mathbb{C}} which is convolution invertible, with convolution inverse R1R^{-1}, and satisfies

R(xy,z)=R(x,z(1))R(y,z(2)),R(x,yz)=R(x(1),z)R(x(2),y),R(xy,z)=R(x,z_{(1)})R(y,z_{(2)}),\quad R(x,yz)=R(x_{(1)},z)R(x_{(2)},y),
y(1)x(1)R(x(2),y(2))=R(x(1),y(1))x(2)y(2)y_{(1)}x_{(1)}R(x_{(2)},y_{(2)})=R(x_{(1)},y_{(1)})x_{(2)}y_{(2)}

for all x,y,zHx,y,z\in H. We say that RR is unitary if it is unitary as an element of the convolution *-algebra (HH)(H\otimes H)^{*}, or equivalently, as R=R(SS)R=R\circ(S\otimes S),

R1(x,y)=R(x,y)¯.R^{-1}(x,y)=\overline{R(x^{*},y^{*})}.

For the rest of this subsection we assume that HH is given a unitary coquasitriangular structure RR. Then H\mathcal{M}^{H} has a braiding given by

MNNM,mnR(m(2),n(2))n(1)m(1).M\otimes N\to N\otimes M,\quad m\otimes n\mapsto R(m^{(2)},n^{(2)})n^{(1)}\otimes m^{(1)}.

There is also an induced right HH-module structure on MM given by

mh=R(m(2),h)m(1),hH,mM.m\triangleleft h=R(m^{(2)},h)m^{(1)},\quad h\in H,\ m\in M. (1.1)

The right comodule and module structures are compatible in the sense that

δM(mh)=m(1)h(2)S(h(1))m(2)h(3),\delta_{M}(m\triangleleft h)=m^{(1)}\triangleleft h_{(2)}\otimes S(h_{(1)})m^{(2)}h_{(3)},

so MM becomes a Yetter–Drinfeld module over HH.

We denote by Alg(H)\mathrm{Alg}^{*}(H) the category of right HH-comodule *-algebras. In other words, an object in Alg(H)\mathrm{Alg}^{*}(H) is a *-algebra AA and a right HH-comodule such that the map δA:AAH\delta_{A}\colon A\to A\otimes H is a *-homomorphism. For A,BAlg(H)A,B\in\mathrm{Alg}^{*}(H) we denote by ARBA\otimes_{R}B the *-algebra with underlying vector space ABA\otimes B equipped with the product

(ab)(ab)=R(b(2),a(2))aa(1)b(1)b(a\otimes b)\cdot(a^{\prime}\otimes b^{\prime})=R(b^{(2)},a^{\prime(2)})aa^{\prime(1)}\otimes b^{(1)}b^{\prime}

and the *-structure

(ab)=(1b)(a1)=R(b(2),a(2))a(1)b(1).(a\otimes b)^{*}=(1\otimes b^{*})\cdot(a^{*}\otimes 1)=R(b^{(2)*},a^{(2)*})a^{(1)*}\otimes b^{(1)*}. (1.2)

The braided tensor product R\otimes_{R} turns Alg(H)\mathrm{Alg}^{*}(H) into a monoidal category Alg(H,R)\mathrm{Alg}^{*}(H,R) with equivariant *-homomorphisms as morphisms.

Considering HH as an HH-comodule *-algebra with the coaction given by the coproduct, we recover the smash product (with respect to the right action (1.1)):

H#A=HRA.H\#A=H\otimes_{R}A.

Let Hopf(H,R)\mathrm{Hopf}^{*}(H,R) denote the category of Hopf *-algebras internal to the braided monoidal category (H,R)(\mathcal{M}^{H},R). An object AHopf(A,R)A\in\mathrm{Hopf}^{*}(A,R) is thus an HH-comodule *-algebra together with HH-comodule maps

ΔA:AARA,SA:AA,εA:A,\Delta_{A}\colon A\to A\otimes_{R}A,\quad S_{A}\colon A\to A,\quad\varepsilon_{A}\colon A\to{\mathbb{C}},

which are required to fit in commutative diagrams analogous to those defining Hopf *-algebras. An object in Hopf(H,R)\mathrm{Hopf}^{*}(H,R) is called a braided Hopf *-algebra, and it is usually not a genuine Hopf *-algebra. It is, however, always closely related to one:

Definition 1.1.

The bosonization of AHopf(H,R)A\in\mathrm{Hopf}^{*}(H,R) is the Hopf *-algebra with underlying *-algebra H#AH\#A, counit ε(h#a)=εH(h)εA(a)\varepsilon(h\#a)=\varepsilon_{H}(h)\varepsilon_{A}(a), coproduct

Δ(h#a)=(h(1)#a(1)(1))(h(2)a(1)(2)#a(2)),\Delta(h\#a)=(h_{(1)}\#{a_{(1)}}^{(1)})\otimes(h_{(2)}{a_{(1)}}^{(2)}\#a_{(2)}), (1.3)

and antipode

S(h#a)=(1#SA(a(1)))(SH(ha(2))#1).S(h\#a)=(1\#S_{A}(a^{(1)}))(S_{H}(ha^{(2)})\#1).

That H#AH\#A is a Hopf algebra, is a special case of results of Radford [Rad85], who proved that a Hopf algebra object in the category of Yetter–Drinfeld-modules is equivalent to a Hopf algebra with projection.

Let AA be a braided Hopf *-algebra. We will only consider AA-comodules internal to H\mathcal{M}^{H}. Thus, the notion of an AA-comodule will be reserved for triples (M,δM,γM)(M,\delta_{M},\gamma_{M}), where δM:MMH\delta_{M}\colon M\to M\otimes H is an HH-comodule and γM:MMA\gamma_{M}\colon M\to M\otimes A is a morphism of HH-comodules that defines a comodule for the coalgebra AA in the usual sense. We record the following well-known result:

Proposition 1.2 (cf. [Maj95, Theorem 9.4.12]).

Let AHopf(H,R)A\in\mathrm{Hopf}^{*}(H,R). Then the category of AA-comodules is isomorphic to the category of (H#A)(H\#A)-comodules through the assignment

(M,δM,γM)(M,(δMι)γM).(M,\delta_{M},\gamma_{M})\mapsto(M,(\delta_{M}\otimes\iota)\gamma_{M}).

The inverse is given by

(M,δ)(M,(ι(ι#εA))δ,(ι(εH#ι))δ).(M,\delta)\mapsto(M,(\iota\otimes(\iota\#\varepsilon_{A}))\delta,(\iota\otimes(\varepsilon_{H}\#\iota))\delta).

We will say that a finite dimensional AA-comodule is unitary, if the corresponding (H#A)(H\#A)-comodule is unitary. Recall that a finite dimensional comodule (M,δ)(M,\delta) over a Hopf *-algebra HH^{\prime} is called unitary, if MM is equipped with a scalar product and

δ(m),δ(m)=(m,m)1for allm,mM,\langle\delta(m),\delta(m)\rangle=(m,m^{\prime})1\ \ \text{for all}\ \ m,m^{\prime}\in M,

where the HH^{\prime}-valued sesquilinear form ,\langle\cdot,\cdot\rangle is defined by ma,mb=(m,m)baH\langle m\otimes a,m^{\prime}\otimes b\rangle=(m,m^{\prime})b^{*}a\in H^{\prime}.

1.2. Twisting and transmutation

Assume that J:HHJ\colon H\otimes H\to{\mathbb{C}} is a Hopf 2-cocycle. This means that JJ is convolution invertible and satisfies

J(x(1),y(1))J(x(2)y(2),z)=J(y(1),z(1))J(x,y(2)z(2))J(x_{(1)},y_{(1)})J(x_{(2)}y_{(2)},z)=J(y_{(1)},z_{(1)})J(x,y_{(2)}z_{(2)})

for x,y,zHx,y,z\in H. We say that JJ is unitary if J=J1J^{*}=J^{-1} in the convolution *-algebra (HH)(H\otimes H)^{*}, that is,

J1(x,y)=J(S(x),S(y))¯.J^{-1}(x,y)=\overline{J(S(x)^{*},S(y)^{*})}.

Given JJ, we can define a convolution invertible element u:Hu\colon H\to{\mathbb{C}} by

u(x)=J(x(1),S(x(2))).u(x)=J(x_{(1)},S(x_{(2)})).

The following is the well-known twisting procedure for Hopf *-algebras, see, e.g., [Maj95, Theorem 2.3.4].

Proposition 1.3.

Let JJ be a unitary Hopf 2-cocycle on HH. There is a Hopf *-algebra HJ1J{}_{J}H_{J^{-1}} having the same coalgebra structure as HH and new product, antipode and involution defined by

xJy=J(x(1),y(1))x(2)y(2)J1(x(3),y(3)),x\star_{J}y=J(x_{(1)},y_{(1)})x_{(2)}y_{(2)}J^{-1}(x_{(3)},y_{(3)}),
SJ(x)=u(x(1))S(x(2))u1(x(3)),xJ=u1(S(x(1)))x(2)u(S(x(3)))=SJ(S1(x)).S^{J}(x)=u(x_{(1)})S(x_{(2)})u^{-1}(x_{(3)}),\quad x^{*_{J}}=u^{-1}(S(x_{(1)})^{*})x^{*}_{(2)}u(S(x_{(3)})^{*})=S^{J}(S^{-1}(x^{*})).

We remark that we are interested in unitary cocycles, while Majid in [Maj95, Section 2.3] considers so-called real ones. That the *-structure on HJ1J{}_{J}H_{J^{-1}} defined above is the correct one in the unitary case is explained in [NT13, Example  2.3.9] in the context of compact quantum groups. It is not difficult to check directly that the definition works for general Hopf *-algebras.

Next we consider a way to produce braided Hopf algebras, due to Majid [Maj93]. Recall that the adjoint comodule for HH is given by

ad:HHH,ad(x)=x(1)x(2)=x(2)S(x(1))x(3).\mathrm{ad}\colon H\to H\otimes H,\quad\mathrm{ad}(x)=x^{(1)}\otimes x^{(2)}=x_{(2)}\otimes S(x_{(1)})x_{(3)}.

Let us for the moment forget about the *-structure on HH. The process of transmutation produces a braided Hopf algebra from the HH-comodule (H,ad)(H,\mathrm{ad}).

Proposition 1.4 ([Maj93, Theorem 4.1]).

Assume HH is a Hopf algebra with a coquasitriangular structure R:HHR\colon H\otimes H\to{\mathbb{C}}. Then HH with the same coalgebra structure and new product R\cdot_{R} and antipode SRS_{R}, given by

xRy=x(2)y(3)R(x(3),S(y(1)))R(x(1),y(2)),x\cdot_{R}y=x_{(2)}y_{(3)}R(x_{(3)},S(y_{(1)}))R(x_{(1)},y_{(2)}),
SR(x)=S(x(2))R(S2(x(3))S(x(1)),x(4)),S_{R}(x)=S(x_{(2)})R(S^{2}(x_{(3)})S(x_{(1)}),x_{(4)}),

defines a braided Hopf algebra HRH_{R} over (H,R)(H,R), where the comodule structure is given by ad\mathrm{ad}.

Note that even though the coproduct is not changed, it is now considered as an algebra map HHRHH\to H\otimes_{R}H.

More generally, assume that π:HK\pi\colon H\to K is a map of Hopf algebras, where KK has a coquasitriangular structure RR. Although πR=R(ππ)\pi^{*}R=R\circ(\pi\otimes\pi) is not a coquasitriangular structure on HH in general, the same formulas as above, with RR replaced by πR\pi^{*}R, define a braided Hopf algebra HRH_{R} over (K,R)(K,R) with the restricted coaction

adπ=(ιπ)ad:HHK.\mathrm{ad}_{\pi}=(\iota\otimes\pi)\mathrm{ad}\colon H\to H\otimes K. (1.4)

The next result relates the bosonization of HRH_{R} to a Hopf algebra with the tensor product coalgebra structure.

Theorem 1.5.

Assume that π:HK\pi\colon H\to K is a map of Hopf algebras and R:KKR\colon K\otimes K\to{\mathbb{C}} is a coquasitriangular structure on KK. Let KRHK\bowtie_{R}H denote the Hopf algebra that coincides with KHK\otimes H as a coalgebra, but has the twisted product

(kh)(kh)=R1(π(h(1)),k(1))kk(2)h(2)hR(π(h(3)),k(3)).(k\otimes h)\cdot(k^{\prime}\otimes h^{\prime})=R^{-1}(\pi(h_{(1)}),k_{(1)}^{\prime})\,kk_{(2)}^{\prime}\otimes h_{(2)}h^{\prime}\,R(\pi(h_{(3)}),k_{(3)}^{\prime}).

Then the map

KRHK#HR,khkπ(h(1))#h(2),K\bowtie_{R}H\to K\#H_{R},\quad k\otimes h\mapsto k\pi(h_{(1)})\#h_{(2)},

defines an isomorphism of Hopf algebras.

Proof.

For K=HK=H and π=id\pi=\operatorname{id}, this is [Maj95, Theorem 7.4.10]. The general case can be proved similarly or by a direct computation. ∎

Note that K1K\otimes{\mathbb{C}}1 is a Hopf subalgebra of KRHK\bowtie_{R}H. Therefore KRHK\bowtie_{R}H is a Hopf algebra with projection KRHKK\bowtie_{R}H\to K, khkπ(h)k\otimes h\mapsto k\pi(h), and then the transmuted braided Hopf algebra HRH_{R} can be viewed as a particular case of the construction of Radford [Rad85].

Remark 1.6.

The formula for the product on KRHK\bowtie_{R}H is the same as for the cocycle twisting by

J((kh),(kh))=εK(k)εH(h)R1(π(h),k),J((k\otimes h),(k^{\prime}\otimes h^{\prime}))=\varepsilon_{K}(k)\varepsilon_{H}(h^{\prime})R^{-1}(\pi(h),k^{\prime}),

but the element JJ is not a Hopf 22-cocycle on KHK\otimes H in general. It becomes a 22-cocycle, when KK is cocommutative. Note also that if KK is both commutative and cocommutative, then HH is a KK-comodule algebra with respect to the adjoint coaction and then as an algebra KRHK\bowtie_{R}H coincides with K#H=KRHK\#H=K\otimes_{R}H.


2. Transmutation over abelian groups

Let (H,Δ,S,ε)(H,\Delta,S,\varepsilon) be a Hopf *-algebra and TT be a compact abelian group. We write T^\hat{T} for the dual discrete group and T^{\mathbb{C}}\hat{T} for the group Hopf *-algebra of T^\hat{T}. We will frequently identity T^{\mathbb{C}}\hat{T} with the function algebra [T]{\mathbb{C}}[T], and will use the latter notation when it is natural to focus on the compact group TT. Throughout this section we assume that we are given a Hopf *-algebra map π:HT^\pi\colon H\to{\mathbb{C}}\hat{T}.

2.1. Braided Hopf algebras over quotients of TT

There are canonical commuting left and right coactions

δL=(πι)Δ,δR=(ιπ)Δ\delta_{L}=(\pi\otimes\iota)\Delta,\quad\delta_{R}=(\iota\otimes\pi)\Delta

by T^{\mathbb{C}}\hat{T} on HH. It follows that HH is bi-graded by T^\hat{T}. More precisely, H=a,bT^Ha,bH=\bigoplus_{a,b\in\hat{T}}H_{a,b}, where

Ha,b={xH|δL(x)=ax and δR(x)=xb}.H_{a,b}=\{x\in H\,|\,\delta_{L}(x)=a\otimes x\mbox{ and }\delta_{R}(x)=x\otimes b\}.

A consequence of coassociativity is then that for any xHx\in H we have

Δ(x)=x(1)x(2)a,b,cHa,bHb,c.\Delta(x)=x_{(1)}\otimes x_{(2)}\in\bigoplus_{a,b,c}H_{a,b}\otimes H_{b,c}. (2.1)

We use the shorthand notation x(1)a,bx(2)b,cx_{(1)}^{a,b}\otimes x_{(2)}^{b,c} to mean the part of the sum x(1)x(2)x_{(1)}\otimes x_{(2)} which is in Ha,bHb,cH_{a,b}\otimes H_{b,c}. Thus, Δ(x)=a,b,cx(1)a,bx(2)b,c\Delta(x)=\sum_{a,b,c}x_{(1)}^{a,b}\otimes x_{(2)}^{b,c}.

Consider a closed subgroup T0TT_{0}\subset T, with the corresponding restriction map q:[T][T0]q\colon{\mathbb{C}}[T]\to{\mathbb{C}}[T_{0}]. We say that T0T_{0} is HH-cocentral, or that the map qπq\pi is cocentral, if the induced adjoint coaction

adqπ=(ιqπ)ad:HH[T0]\mathrm{ad}_{q\pi}=(\iota\otimes q\pi)\mathrm{ad}\colon H\to H\otimes{\mathbb{C}}[T_{0}]

is trivial. In other words,

Ha,b=0wheneverq(a)q(b).H_{a,b}=0\quad\text{whenever}\quad q(a)\neq q(b).

This condition implies that if we view [T/T0]{\mathbb{C}}[T/T_{0}] as a Hopf *-subalgebra of [T]{\mathbb{C}}[T], then

adπ(H)H[T/T0],\mathrm{ad}_{\pi}(H)\subset H\otimes{\mathbb{C}}[T/T_{0}], (2.2)

so that (H,adπ)(H,\mathrm{ad}_{\pi}) can be viewed as a [T/T0]{\mathbb{C}}[T/T_{0}]-comodule.

Denote by Δ(T0)\Delta(T_{0}) the diagonal subgroup in T0×T0T_{0}\times T_{0}. It is ([T]H)({\mathbb{C}}[T]\otimes H)-cocentral with respect to the composition of ιπ:[T]H[T][T]\iota\otimes\pi\colon{\mathbb{C}}[T]\otimes H\to{\mathbb{C}}[T]\otimes{\mathbb{C}}[T] with the restriction map [T×T][Δ(T0)]{\mathbb{C}}[T\times T]\to{\mathbb{C}}[\Delta(T_{0})]. It follows that the spaces of right and left coinvariants coincide,

([T]H)Δ(T0)=([T]H)Δ(T0)=a,b,c:q(ab)=1aHb,c=a,b,c:q(ac)=1aHb,c,({\mathbb{C}}[T]\otimes H)^{\Delta(T_{0})}={}^{\Delta(T_{0})}({\mathbb{C}}[T]\otimes H)=\bigoplus_{\begin{subarray}{c}a,b,c:\\ q(ab)=1\end{subarray}}a\otimes H_{b,c}=\bigoplus_{\begin{subarray}{c}a,b,c:\\ q(ac)=1\end{subarray}}a\otimes H_{b,c}, (2.3)

and define a Hopf *-subalgebra of [T]H{\mathbb{C}}[T]\otimes H.

In the present setting we observe that HH equipped with adπ\mathrm{ad}_{\pi} is an object in Hopf([T],εε)\mathrm{Hopf}^{*}({\mathbb{C}}[T],\varepsilon\otimes\varepsilon) without any modifications. Moreover, by (2.2) it can also be viewed as an object in Hopf([T/T0],εε)\mathrm{Hopf}^{*}({\mathbb{C}}[T/T_{0}],\varepsilon\otimes\varepsilon). The corresponding bosonization [T/T0]H{\mathbb{C}}[T/T_{0}]\ltimes H is just the tensor product *-algebra with the coproduct defined by (1.3) (in other words, as a coalgebra, it is the smash coproduct of [T/T0]{\mathbb{C}}[T/T_{0}] and HH):

Δ[T/T0]H(ax)=d(ax(1)b,d)(ab1dx(2)d,c)\Delta_{{\mathbb{C}}[T/T_{0}]\ltimes H}(a\otimes x)=\sum_{d}(a\otimes x^{b,d}_{(1)})\otimes(ab^{-1}d\otimes x^{d,c}_{(2)})

for aT/T0^T^a\in\widehat{T/T_{0}}\subset\hat{T} and xHb,cx\in H_{b,c}.

Proposition 2.1.

Consider HH as a [T]{\mathbb{C}}[T]-comodule Hopf *-algebra under the adjoint coaction adπ\mathrm{ad}_{\pi}. Then

Θ:[T]H[T]H,Θ(ax)=aπ(x(1))x(2),\Theta\colon{\mathbb{C}}[T]\otimes H\to{\mathbb{C}}[T]\ltimes H,\quad\Theta(a\otimes x)=a\pi(x_{(1)})\otimes x_{(2)},

is an isomorphism of Hopf *-algebras. Moreover, Θ\Theta restricts to an isomorphism of Hopf *-algebras

([T]H)Δ(T0)[T/T0]H({\mathbb{C}}[T]\otimes H)^{\Delta(T_{0})}\cong{\mathbb{C}}[T/T_{0}]\ltimes H

for any HH-cocentral closed subgroup T0TT_{0}\subset T.

Proof.

That Θ\Theta is an isomorphism is easily verified, but except for the *-structure it is also a special case of Theorem 1.5. As Θ(aHb,c)=abHb,c\Theta(a\otimes H_{b,c})=ab\otimes H_{b,c} and [T/T0]{\mathbb{C}}[T/T_{0}] is spanned by aT^a\in\hat{T} such that q(a)=1q(a)=1, the second part of the proposition follows from (2.3). ∎

Remark 2.2.

Let GG be a compact group with a closed abelian subgroup TT. Assume that T0T_{0} is a closed subgroup of TZ(G)T\cap Z(G), where Z(G)Z(G) is the center of GG. As T0T_{0} acts trivially under the conjugation action (t,g)tgt1(t,g)\mapsto tgt^{-1} of TT on GG, we have an induced action on GG by the quotient group T/T0T/T_{0}. Then the map

(T/T0)G(T×G)/Δ(T0),([t],g)[(t,tg)],(T/T_{0})\ltimes G\to(T\times G)/\Delta(T_{0}),\quad([t],g)\mapsto[(t,tg)],

is a group isomorphism. The above result can be seen as a generalization of this.  \diamond

Next, fix a bicharacter β:T^×T^𝕋\beta\colon\hat{T}\times\hat{T}\to{\mathbb{T}}. This simply means that, for all a,b,ca,b,c in T^\hat{T},

β(ab,c)=β(a,c)β(b,c) and β(a,bc)=β(a,b)β(a,c).\beta(ab,c)=\beta(a,c)\beta(b,c)\quad\mbox{ and }\quad\beta(a,bc)=\beta(a,b)\beta(a,c).

We will write β\beta also for the extension of the bicharacter to a linear map T^T^{\mathbb{C}}\hat{T}\otimes{\mathbb{C}}\hat{T}\to{\mathbb{C}}. This defines a unitary coquasitriangular structure on [T]=T^{\mathbb{C}}[T]={\mathbb{C}}\hat{T}. We want to understand the corresponding transmutation HβHopf([T],β)H_{\beta}\in\mathrm{Hopf}^{*}({\mathbb{C}}[T],\beta).

By definition, the product on HβH_{\beta} is determined by

xβy=β(a1b,c1)xy,xHa,b,yHc,d,x\cdot_{\beta}y=\beta(a^{-1}b,c^{-1})xy,\quad x\in H_{a,b},\ y\in H_{c,d}, (2.4)

the coproduct and counit remain unchanged, while the antipode is determined by

Sβ(x)=β(a1b,b)S(x),xHa,b.S_{\beta}(x)=\beta(a^{-1}b,b)S(x),\quad x\in H_{a,b}. (2.5)

In the present setting we may also introduce a *-structure on HβH_{\beta}:

Lemma 2.3.

The formula

xβ=β(a1b,a1)x,xHa,b,x^{*_{\beta}}=\beta(a^{-1}b,a^{-1})x^{*},\quad x\in H_{a,b}, (2.6)

turns HβH_{\beta} into a braided Hopf *-algebra.

Proof.

Using that xHa1,b1x^{*}\in H_{a^{-1},b^{-1}} for xHa,bx\in H_{a,b}, it is straightforward to check that β*_{\beta} is involutive. Using formula (2.4), it is also easy to see that (xβy)β=yββxβ(x\cdot_{\beta}y)^{*_{\beta}}=y^{*_{\beta}}\cdot_{\beta}x^{*_{\beta}}. To show that Δ:HβHββHβ\Delta\colon H_{\beta}\to H_{\beta}\otimes_{\beta}H_{\beta} is *-preserving, notice that if xHa,bx\in H_{a,b} and yHb,cy\in H_{b,c}, then by (1.2) we have

(xy)βββ\displaystyle(x\otimes y)^{*_{\beta}\,\otimes_{\beta}\,*_{\beta}} =β(a1b,a1)β(b1c,b1)β(b1c,a1b)xy\displaystyle=\beta(a^{-1}b,a^{-1})\beta(b^{-1}c,b^{-1})\beta(b^{-1}c,a^{-1}b)\,x^{*}\otimes y^{*}
=β(a1c,a1)xy.\displaystyle=\beta(a^{-1}c,a^{-1})\,x^{*}\otimes y^{*}.

That Δ\Delta preserves the new *-structure is thus a consequence of (2.1). ∎

By (2.2) we thus get the following:

Proposition 2.4.

For any HH-cocentral closed subgroup T0TT_{0}\subset T, the transmutation HβH_{\beta} can be viewed as an object in Hopf([T/T0],iβ)\mathrm{Hopf}^{*}({\mathbb{C}}[T/T_{0}],i^{*}\beta), where i:[T/T0][T]i\colon{\mathbb{C}}[T/T_{0}]\to{\mathbb{C}}[T] is the embedding map.

The following is the main result of this section.

Theorem 2.5.

Given a bicharacter β:T^×T^𝕋\beta\colon\hat{T}\times\hat{T}\to{\mathbb{T}}, let J1,J2:(T^×T^)×(T^×T^)𝕋J_{1},J_{2}\colon(\hat{T}\times\hat{T})\times(\hat{T}\times\hat{T})\to{\mathbb{T}} be the 22-cocycles defined by

J1((a,b),(c,d))=β(b,c)¯,J2((a,b),(c,d))=β(b,cd1)¯.J_{1}((a,b),(c,d))=\overline{\beta(b,c)},\quad J_{2}((a,b),(c,d))=\overline{\beta(b,cd^{-1})}.

Then, for any HH-cocentral closed subgroup T0TT_{0}\subset T, we have Hopf *-algebra isomorphisms

(([T]H)Δ(T0))J11J1\displaystyle{}_{J_{1}}(({\mathbb{C}}[T]\otimes H)^{\Delta(T_{0})})_{J_{1}{}^{-1}} [T/T0]#Hβ,axaπ(x(1))#x(2),\displaystyle\cong{\mathbb{C}}[T/T_{0}]\#H_{\beta},\quad a\otimes x\mapsto a\pi(x_{(1)})\#x_{(2)},
([T/T0]H)J21J2\displaystyle{}_{J_{2}}({\mathbb{C}}[T/T_{0}]\ltimes H)_{J_{2}{}^{-1}} [T/T0]#Hβ,axa#x.\displaystyle\cong{\mathbb{C}}[T/T_{0}]\#H_{\beta},\quad a\otimes x\mapsto a\#x.

More pedantically, by restriction J1J_{1} defines a 22-cocycle on Δ(T0)T^×T^\Delta(T_{0})^{\perp}\subset\hat{T}\times\hat{T}, hence a Hopf 22-cocycle on [(T×T)/Δ(T0)]{\mathbb{C}}[(T\times T)/\Delta(T_{0})]. The latter gives rise to a Hopf 22-cocycle on ([T]H)Δ(T0)({\mathbb{C}}[T]\otimes H)^{\Delta(T_{0})} using the map ιπ:([T]H)Δ(T0)[(T×T)/Δ(T0)]\iota\otimes\pi\colon({\mathbb{C}}[T]\otimes H)^{\Delta(T_{0})}\to{\mathbb{C}}[(T\times T)/\Delta(T_{0})]. This is the cocycle we use to define (([T]H)Δ(T0))J11J1{}_{J_{1}}(({\mathbb{C}}[T]\otimes H)^{\Delta(T_{0})})_{J_{1}{}^{-1}}. Similarly, J2J_{2} defines a Hopf 22-cocycle on [T/T0]H{\mathbb{C}}[T/T_{0}]\ltimes H.

Proof of Theorem 2.5..

We note that ([T]H)J11J1=[T]βH{}_{J_{1}}({\mathbb{C}}[T]\otimes H)_{{J_{1}}{}^{-1}}={\mathbb{C}}[T]\bowtie_{\beta}H, where the latter Hopf algebra is defined as in Theorem 1.5. By that theorem, this immediately gives the first Hopf algebra isomorphism for trivial T0T_{0}. The isomorphism is readily verified to be *-preserving, using that by (1.2) and Definition 1.1 the involution on [T]#Hβ{\mathbb{C}}[T]\#H_{\beta} is given by

(a#x)=β(b1c,a)a1#xβ=β(b1c,ab1)a1#x,aT^,xHb,c,(a\#x)^{*}=\beta(b^{-1}c,a)\,a^{-1}\#x^{*_{\beta}}=\beta(b^{-1}c,ab^{-1})\,a^{-1}\#x^{*},\quad a\in\hat{T},\ x\in H_{b,c},

while by Proposition 1.3 the involution on ([T]H)J11J1{}_{J_{1}}({\mathbb{C}}[T]\otimes H)_{J_{1}{}^{-1}} is given by

(ax)J1=β(b1c,a)a1x,aT^,xHb,c.(a\otimes x)^{*_{J_{1}}}=\beta(b^{-1}c,a)\,a^{-1}\otimes x^{*},\quad a\in\hat{T},\ x\in H_{b,c}.

For the same reason as in Proposition 2.1, for every HH-cocentral closed subgroup T0TT_{0}\subset T, the isomorphism ([T]H)J11J1[T]#Hβ{}_{J_{1}}({\mathbb{C}}[T]\otimes H)_{J_{1}{}^{-1}}\cong{\mathbb{C}}[T]\#H_{\beta} defines by restriction the first isomorphism in the formulation of the theorem. The second isomorphism follows from this and again Proposition 2.1. ∎

It is well-known that 22-cocycle twisting preserves the monoidal categories of comodules. We thus get the following:

Corollary 2.6.

For any HH-cocentral closed subgroup T0TT_{0}\subset T, the category [T/T0]#Hβ\mathcal{M}^{{\mathbb{C}}[T/T_{0}]\#H_{\beta}} is monoidally equivalent to [T/T0]H\mathcal{M}^{{\mathbb{C}}[T/T_{0}]\ltimes H}.

Remark 2.7.

The cocentral homomorphism qπ:H[T0]q\pi\colon H\to{\mathbb{C}}[T_{0}] defines a T^0\hat{T}_{0}-grading on the category fH\mathcal{M}^{H}_{f} of finite dimensional comodules. Then f[T]HVectfT^fH\mathcal{M}^{{\mathbb{C}}[T]\otimes H}_{f}\cong\operatorname{Vect}_{f}^{\hat{T}}\boxtimes\mathcal{M}^{H}_{f}, where VectfT^\operatorname{Vect}_{f}^{\hat{T}} is the category of finite dimensional T^\hat{T}-graded vector spaces, is bi-graded by T^×T^0\hat{T}\times\hat{T}_{0}. We can conclude that f[T/T0]#Hβ\mathcal{M}^{{\mathbb{C}}[T/T_{0}]\#H_{\beta}}_{f} is monoidally equivalent to the subcategory of VectfT^fH\operatorname{Vect}_{f}^{\hat{T}}\boxtimes\mathcal{M}^{H}_{f} generated by the homogeneous components of bi-degree (a,b)(a,b) such that q(a)b=1q(a)b=1.  \diamond

From Theorems 1.5 or 2.5 we see that we have a Hopf *-algebra inclusion

ϕ:H[T]#Hβ,ϕ(x)=π(x(1))#x(2).\phi\colon H\to{\mathbb{C}}[T]\#H_{\beta},\quad\phi(x)=\pi(x_{(1)})\#x_{(2)}. (2.7)

This map induces a monoidal functor from the category HH-comodules to the category of HβH_{\beta}-comodules. It will be convenient to have the following description of this functor.

Lemma 2.8.

Let δ:MMH\delta\colon M\to M\otimes H be an HH-comodule. Then (M,(ιπ)δ,δ)(M,(\iota\otimes\pi)\delta,\delta) is an HβH_{\beta}-comodule.

Proof.

Write δ=(ιϕ)δ\delta^{\prime}=(\iota\otimes\phi)\delta. Then

(ι(ι#ε))δ=(ι(πε)Δ)δ=(ιπ)δ,(\iota\otimes(\iota\#\varepsilon))\delta^{\prime}=(\iota\otimes(\pi\otimes\varepsilon)\Delta)\delta=(\iota\otimes\pi)\delta,

and

(ι(ε#ι))δ=(ι(επι)Δ)δ=δ.(\iota\otimes(\varepsilon\#\iota))\delta^{\prime}=(\iota\otimes(\varepsilon\pi\otimes\iota)\Delta)\delta=\delta.

Hence, by Proposition 1.2, the claim follows. ∎

2.2. Another view on HβH_{\beta}

Motivated by the recent of work of Bochniak and Sitarz [BS19], we now give another interpretation of the structure maps for HβH_{\beta}.

Using the left and right coactions of [T]{\mathbb{C}}[T] on HH, we can view HH as a [T×T]{\mathbb{C}}[T\times T]-comodule algebra. Then the new product β\cdot_{\beta} on HβH_{\beta} is obtained by cocycle twisting (see [Maj95, Section 2.3]) the original product by the 22-cocycle

(T^×T^)×(T^×T^)𝕋,((a,b),(c,d))β(a1b,c1).(\hat{T}\times\hat{T})\times(\hat{T}\times\hat{T})\to{\mathbb{T}},\quad((a,b),(c,d))\mapsto\beta(a^{-1}b,c^{-1}).

On the other hand, HH is also a [T]{\mathbb{C}}[T]-comodule coalgebra, so its coalgebra structure can be twisted by a 22-cocycle ωZ2(T^;𝕋)\omega\in Z^{2}(\hat{T};{\mathbb{T}}):

Δω(x)=cω(a1c,c1b)x(1)a,cx(2)c,b,xHa,b.\Delta_{\omega}(x)=\sum_{c}\omega(a^{-1}c,c^{-1}b)x^{a,c}_{(1)}\otimes x^{c,b}_{(2)},\quad x\in H_{a,b}.

As a consequence of the following lemma, this always gives an isomorphic comodule coalgebra.

Lemma 2.9.

Assume ωZ2(T^;𝕋)\omega\in Z^{2}(\hat{T};{\mathbb{T}}) is a normalized cocycle (so ω(1,1)=1\omega(1,1)=1) and γ:T^×T^𝕋\gamma\colon\hat{T}\times\hat{T}\to{\mathbb{T}} is a function. Then the identity

ω(a1c,c1b)=γ(a,c)γ(c,b)γ(a,b)1\omega(a^{-1}c,c^{-1}b)=\gamma(a,c)\gamma(c,b)\gamma(a,b)^{-1} (2.8)

holds for all a,b,cT^a,b,c\in\hat{T} if and only if

γ(a,b)=ω(a1,b)1f(a)g(b)\gamma(a,b)=\omega(a^{-1},b)^{-1}f(a)g(b) (2.9)

for all a,b𝕋a,b\in{\mathbb{T}}, where f,g:T^𝕋f,g\colon\hat{T}\to{\mathbb{T}} are arbitrary functions such that

f(a)g(a)=ω(a,a1),aT^.f(a)g(a)=\omega(a,a^{-1}),\quad a\in\hat{T}.
Proof.

Assume (2.8) holds. Letting c=1c=1 we get

ω(a1,b)=γ(a,1)γ(1,b)γ(a,b)1.\omega(a^{-1},b)=\gamma(a,1)\gamma(1,b)\gamma(a,b)^{-1}.

Therefore (2.9) holds with f(a)=γ(a,1)f(a)=\gamma(a,1) and g(b)=γ(1,b)g(b)=\gamma(1,b). The identity f(a)g(a)=ω(a,a1)f(a)g(a)=\omega(a,a^{-1}) is satisfied by (2.9), since by letting a=b=ca=b=c in (2.8) we see that γ(a,a)=1\gamma(a,a)=1 (recall also that ω(a,a1)=ω(a1,a)\omega(a,a^{-1})=\omega(a^{-1},a) by the cocycle identity).

Conversely, assume (2.9) holds for some functions f,g:T^𝕋f,g\colon\hat{T}\to{\mathbb{T}}. Then

γ(a,c)γ(c,b)γ(a,b)1=ω(a1,c)1ω(c1,b)1ω(a1,b)f(c)g(c).\gamma(a,c)\gamma(c,b)\gamma(a,b)^{-1}=\omega(a^{-1},c)^{-1}\omega(c^{-1},b)^{-1}\omega(a^{-1},b)f(c)g(c).

Therefore (2.8) holds if and only if

ω(a1,c)ω(a1c,c1b)ω(c1,b)=ω(a1,b)f(c)g(c).\omega(a^{-1},c)\omega(a^{-1}c,c^{-1}b)\omega(c^{-1},b)=\omega(a^{-1},b)f(c)g(c).

Using the cocycle identity twice, the left hand side equals

ω(a1,b)ω(c,c1b)ω(c1,b)=ω(a1,b)ω(c,c1)ω(1,b)=ω(a1,b)ω(c,c1),\omega(a^{-1},b)\omega(c,c^{-1}b)\omega(c^{-1},b)=\omega(a^{-1},b)\omega(c,c^{-1})\omega(1,b)=\omega(a^{-1},b)\omega(c,c^{-1}),

where the last identity holds, since ω\omega is normalized. Thus, identity (2.8) holds if and only if f(c)g(c)=ω(c,c1)f(c)g(c)=\omega(c,c^{-1}). ∎

From this we see that up to an isomorphism HβH_{\beta} can be obtained in many different ways by simultaneously twisting the product and coproduct on HH:

Proposition 2.10.

Given a bicharacter β:T^×T^𝕋\beta\colon\hat{T}\times\hat{T}\to{\mathbb{T}} and a normalized 22-cocycle ωZ2(T^;𝕋)\omega\in Z^{2}(\hat{T};{\mathbb{T}}), choose a function γ:T^×T^𝕋\gamma\colon\hat{T}\times\hat{T}\to{\mathbb{T}} satisfying (2.8) and define a 22-cocycle ΩZ2(T^×T^;𝕋)\Omega\in Z^{2}(\hat{T}\times\hat{T};{\mathbb{T}}) by

Ω((a,b),(c,d))=β(a1b,c1)γ(a,b)1γ(c,d)1γ(ac,bd).\Omega((a,b),(c,d))=\beta(a^{-1}b,c^{-1})\gamma(a,b)^{-1}\gamma(c,d)^{-1}\gamma(ac,bd).

Then we can define new product, coproduct and involution on the [T]{\mathbb{C}}[T]-comodule HH by

xΩy=Ω((a,b),(c,d))xy,xHa,b,yHc,d,x\cdot_{\Omega}y=\Omega((a,b),(c,d))xy,\quad x\in H_{a,b},\ y\in H_{c,d},
Δω(x)=cω(a1c,c1b)x(1)a,cx(2)c,b,xHa,b,\Delta_{\omega}(x)=\sum_{c}\omega(a^{-1}c,c^{-1}b)x^{a,c}_{(1)}\otimes x^{c,b}_{(2)},\quad x\in H_{a,b},
x=Ω((a,b),(a,b)1)¯x=β(a1b,a1)γ(a,b)γ(a1,b1)x,xHa,b,x^{\star}=\overline{\Omega((a,b),(a,b)^{-1})}x^{*}=\beta(a^{-1}b,a^{-1})\gamma(a,b)\gamma(a^{-1},b^{-1})x^{*},\quad x\in H_{a,b},

to get a braided Hopf *-algebra HΩ,ωHopf([T],β)H_{\Omega,\omega}\in\mathrm{Hopf}^{*}({\mathbb{C}}[T],\beta). We have an isomorphism

HβHΩ,ω,Ha,bxγ(a,b)x.H_{\beta}\cong H_{\Omega,\omega},\quad H_{a,b}\ni x\mapsto\gamma(a,b)x.
Example 2.11.

Consider T=𝕋T={\mathbb{T}} and, identifying 𝕋^\hat{\mathbb{T}} with {\mathbb{Z}}, let

β(m,n)=e2iϕmn,ω(m,n)=eiϕmn\beta(m,n)=e^{2i\phi mn},\qquad\omega(m,n)=e^{i\phi mn}

for some ϕ\phi\in{\mathbb{R}}. By taking γ(m,n)=ω(m,n)1eiϕm2=eiϕ(mnm2)\gamma(m,n)=\omega(-m,n)^{-1}e^{-i\phi m^{2}}=e^{i\phi(mn-m^{2})}, we get

Ω((k,l),(m,n))=eiϕ(knlm).\Omega((k,l),(m,n))=e^{i\phi(kn-lm)}.

Then HΩ,ωH_{\Omega,\omega} coincides with the braided Hopf algebra defined in [BS19].

2.3. Braided compact matrix quantum groups

In our examples we will mainly be interested in transmutations of compact quantum groups. A compact quantum group GG is a Hopf *-algebra [G]{\mathbb{C}}[G] that is spanned (equivalently, generated as an algebra) by matrix coefficients of finite dimensional unitary comodules. We refer the reader to [NT13] for an introduction to the subject and we will often use the terminology there. For instance, a [G]{\mathbb{C}}[G]-comodule will sometimes be called a representation of GG.

Recall that fixing a basis in the underlying vector space of an mm-dimensional HH-comodule defines a corepresentation matrix for HH. This is a matrix U=(uij)i,jMatm(H)U=(u_{ij})_{i,j}\in\operatorname{Mat}_{m}(H) such that

Δ(uij)=kuikukj,ε(uij)=δij.\Delta(u_{ij})=\sum_{k}u_{ik}\otimes u_{kj},\quad\varepsilon(u_{ij})=\delta_{ij}. (2.10)

Conversely any such matrix defines an mm-dimensional HH-comodule δU:MMH\delta_{U}\colon M\to M\otimes H, by setting

δU(ej)=ieiuij\delta_{U}(e_{j})=\sum_{i}e_{i}\otimes u_{ij} (2.11)

for a fixed vector space MM with basis (ei)i(e_{i})_{i}. If UU is unitary, then the conjugate corepresentation matrix is U¯=(uij)i,j\bar{U}=(u_{ij}^{*})_{i,j}.

Definition 2.12 ([Wor87, Wor91]).

A compact matrix quantum group GG is a Hopf *-algebra [G]{\mathbb{C}}[G] with generators uiju_{ij}, 1i,jm1\leq i,j\leq m, such that

  • (i)

    U=(uij)i,jU=(u_{ij})_{i,j} is a unitary corepresentation matrix;

  • (ii)

    U¯=(uij)i,j\bar{U}=(u_{ij}^{*})_{i,j} is equivalent to a unitary corepresentation matrix.

The coproduct Δ\Delta, counit ε\varepsilon and antipode SS are then given by

Δ(uij)=kuikukj,ε(uij)=δij,S(uij)=uji.\Delta(u_{ij})=\sum_{k}u_{ik}\otimes u_{kj},\quad\varepsilon(u_{ij})=\delta_{ij},\quad S(u_{ij})=u_{ji}^{*}.

The matrix UU is called the fundamental unitary for the compact matrix quantum group.

We remark that this is not the original definition of Woronowicz, but it is equivalent to that by a result of Dijkhuizen and Koornwinder [DK94], see also [NT13, Section 1.6].

Next, we want to introduce a braided analogue of this definition, but first we need some preparation. Suppose that AHopf(K,R)A\in\mathrm{Hopf}^{*}(K,R) for a Hopf *-algebra KK with a unitary coquasitriangular structure RR. Suppose U=(uij)i,jMatm(A)U=(u_{ij})_{i,j}\in\operatorname{Mat}_{m}(A) satisfies ΔA(uij)=kuikukj\Delta_{A}(u_{ij})=\sum_{k}u_{ik}\otimes u_{kj}, εA(uij)=δij\varepsilon_{A}(u_{ij})=\delta_{ij}. We can still define δU:MMA\delta_{U}\colon M\to M\otimes A as in (2.11) and this gives a comodule for the coalgebra AA. By definition, if Z=(zij)i,jMatm(K)Z=(z_{ij})_{i,j}\in\operatorname{Mat}_{m}(K) is a corepresentation matrix, the triple (M,δZ,δU)(M,\delta_{Z},\delta_{U}) defines an AA-comodule if and only if

(δUι)δZ=δMAδU,(\delta_{U}\otimes\iota)\circ\delta_{Z}=\delta_{M\otimes A}\circ\delta_{U},

where δMA\delta_{M\otimes A} denotes the tensor product comodule in K\mathcal{M}^{K}. We have the following characterization:

Lemma 2.13.

In the above setting, the pair (Z,U)(Z,U) defines an AA-comodule if and only if

δ(uij)=s,tustS(zis)ztj,\delta(u_{ij})=\sum_{s,t}u_{st}\otimes S(z_{is})z_{tj}, (2.12)

where δ:AAK\delta\colon A\to A\otimes K is the coaction by KK on AA. Furthermore, the AA-comodule we thus get is unitary (that is, the corresponding (K#A)(K\#A)-comodule is unitary) if and only if UU and ZZ are unitary, and then the conjugate comodule is given by the pair (Z¯,U¯Z)(\bar{Z},\bar{U}_{Z}), where

U¯Z=(u¯ijZ)i,j,u¯ijZ=s,l,tR(ztjzsl,zil)ust,\bar{U}_{Z}=(\bar{u}_{ij}^{Z})_{i,j},\quad\bar{u}_{ij}^{Z}=\sum_{s,l,t}R(z_{tj}^{*}z_{sl},z_{il}^{*})u_{st}^{*}, (2.13)

while the antipode on AA satisfies SA(uij)=ujiS_{A}(u_{ij})=u^{*}_{ji}.

Here by the conjugate AA-comodule we mean the comodule obtained by taking the conjugate (K#A)K\#A)-comodule.

Proof.

The condition (δUι)δZ=δMAδU(\delta_{U}\otimes\iota)\circ\delta_{Z}=\delta_{M\otimes A}\circ\delta_{U} is satisfied if and only if

tustztj=kukj(1)zskukj(2)\sum_{t}u_{st}\otimes z_{tj}=\sum_{k}u_{kj}^{(1)}\otimes z_{sk}u_{kj}^{(2)}

for 1s,jm1\leq s,j\leq m, where we write δ(a)=a(1)a(2)\delta(a)=a^{(1)}\otimes a^{(2)}, aAa\in A. Multiplying both sides by S(zis)S(z_{is}) and summing over ss yields

s,tustS(zis)ztj=k,sukj(1)S(zis)zskukj(2)=uij(1)uij(2).\sum_{s,t}u_{st}\otimes S(z_{is})z_{tj}=\sum_{k,s}u_{kj}^{(1)}\otimes S(z_{is})z_{sk}u_{kj}^{(2)}=u_{ij}^{(1)}\otimes u_{ij}^{(2)}.

This implies the first statement.

For the second one, note that the AA-comodule defined by (Z,U)(Z,U) corresponds to the (K#A)(K\#A)-comodule given by the matrix

W=(wij)i,j=Z#U,sowij=kzik#ukj.W=(w_{ij})_{i,j}=Z\#U,\quad\text{so}\quad w_{ij}=\sum_{k}z_{ik}\#u_{kj}.

Using that ι#εA:K#AK\iota\#\varepsilon_{A}\colon K\#A\to K is a *-homomorphism, it is easy to see that WW is unitary if and only ZZ and UU are unitary.

When UU is unitary, the equality SA(uij)=ujiS_{A}(u_{ij})=u_{ji}^{*} holds by the antipode identity. The claim about the conjugate comodule follows from the *-structure on K#AK\#A and the fact that U¯Z=(εK#ι)(W¯)\bar{U}_{Z}=(\varepsilon_{K}\#\iota)(\bar{W}). Alternatively, we can use the formula for the antipode in Definition 1.1 to get

u¯ijZ=(εK#ι)S(wji)=rR(SA(ujr)(2),SK(zri))SA(ujr)(1).\bar{u}^{Z}_{ij}=(\varepsilon_{K}\#\iota)S(w_{ji})=\sum_{r}R(S_{A}(u_{jr})^{(2)},S_{K}(z_{ri}))S_{A}(u_{jr})^{(1)}.

As SA(ujr)=urjS_{A}(u_{jr})=u_{rj}^{*}, we recover (2.13). ∎

We remark that it is important to keep track of both RR and ZZ in the definition of U¯Z\bar{U}_{Z} above. However, we stick to the notation U¯Z\bar{U}_{Z} for the rest of the paper, as RR will always be given by a fixed bicharacter β\beta.

Definition 2.14.

Let TT be a compact abelian group with a fixed unitary corepresentation matrix ZMatm([T])Z\in\operatorname{Mat}_{m}({\mathbb{C}}[T]) and a bicharacter β\beta on T^\hat{T}. A braided compact matrix quantum group over the triple (T,Z,β)(T,Z,\beta) is an object AHopf([T],β)A\in\mathrm{Hopf}^{*}({\mathbb{C}}[T],\beta) generated as a *-algebra by elements uiju_{ij}, 1i,jm1\leq i,j\leq m, such that, for U=(uij)i,jU=(u_{ij})_{i,j},

  • (i)

    (Z,U)(Z,U) defines a unitary AA-comodule;

  • (ii)

    (Z¯,U¯Z)(\bar{Z},\bar{U}_{Z}) defines a unitarizable AA-comodule.

We say that UU is the fundamental unitary for AA, while the pair (Z,U)(Z,U) is the fundamental unitary representation.

More explicitly, by Lemma 2.13, conditions (i) and (ii) mean that UMatm(A)U\in\operatorname{Mat}_{m}(A) is unitary, there is FGLm()F\in\operatorname{GL}_{m}({\mathbb{C}}) such that both FZ¯F1F\bar{Z}F^{-1} and FU¯ZF1F\bar{U}_{Z}F^{-1} are unitary, and the structure maps for AA satisfy the following properties: the coaction of [T]{\mathbb{C}}[T] on AA is given by (2.12), and

ΔA(uij)=kuikukj,εA(uij)=δij,SA(uij)=uji.\Delta_{A}(u_{ij})=\sum_{k}u_{ik}\otimes u_{kj},\qquad\varepsilon_{A}(u_{ij})=\delta_{ij},\qquad S_{A}(u_{ij})=u_{ji}^{*}.

We remark that in view of Lemma 2.13 we can similarly define a braided compact matrix quantum group over (K,Z,R)(K,Z,R) for any Hopf *-algebra KK with a unitary coquasitriangular structure RR and a unitary corepresentation matrix Z=(zij)i,jMatm(K)Z=(z_{ij})_{i,j}\in\operatorname{Mat}_{m}(K), but the adjective “compact” in this generality might be somewhat misleading.

Proposition 2.15.

Given a compact abelian group TT, a unitary corepresentation matrix ZMatm([T])Z\in\operatorname{Mat}_{m}({\mathbb{C}}[T]) and a bicharacter β\beta on T^\hat{T}, the bosonization of any braided compact matrix quantum group AA over (T,Z,β)(T,Z,\beta) is a compact quantum group.

Proof.

By working in an orthonormal basis where ZZ is diagonal, we see that the *-algebra [T]#A{\mathbb{C}}[T]\#A is generated by the matrix coefficients of the fundamental unitary representation and the characters of TT. ∎

Proposition 2.16.

Let GG be a compact matrix quantum group with fundamental unitary U=(uij)i,j=1mU=(u_{ij})^{m}_{i,j=1}. Assume that TT is a compact abelian group with a Hopf *-algebra map π:[G][T]\pi\colon{\mathbb{C}}[G]\to{\mathbb{C}}[T], and let β\beta be a bicharacter on T^\hat{T}. Then the transmutation [G]β{\mathbb{C}}[G]_{\beta} is a braided compact matrix quantum group over (T,π(U),β)(T,\pi(U),\beta) with fundamental unitary UU.

Proof.

Put Z=π(U)Z=\pi(U). Recall that by (2.7) we have a Hopf *-algebra map

ϕ:[G][T]#[G]β,ϕ(x)=π(x(1))#x(2).\phi\colon{\mathbb{C}}[G]\to{\mathbb{C}}[T]\#{\mathbb{C}}[G]_{\beta},\quad\phi(x)=\pi(x_{(1)})\#x_{(2)}.

By Lemma 2.8 this implies that the pair (Z,U)(Z,U) defines a unitary [G]β{\mathbb{C}}[G]_{\beta}–comodule, with the corresponding ([T]#[G]β)({\mathbb{C}}[T]\#{\mathbb{C}}[G]_{\beta})-comodule given by the unitary ϕ(U)=Z#U\phi(U)=Z\#U. The conjugate ([T]#[G]β)({\mathbb{C}}[T]\#{\mathbb{C}}[G]_{\beta})-comodule is given by ϕ(U¯)\phi(\bar{U}). As U¯Matm([G])\bar{U}\in\operatorname{Mat}_{m}({\mathbb{C}}[G]) is unitarizable, by Lemma 2.13 we see that both conditions (i) and (ii) in Definition 2.14 are satisfied.

It remains to check that [G]β{\mathbb{C}}[G]_{\beta} is generated by the matrix coefficients of UU as a *-algebra. This becomes clear if we work in an orthonormal basis where ZZ is diagonal, as then the products of the elements uiju_{ij} and their adjoints in [G]{\mathbb{C}}[G] and [G]β{\mathbb{C}}[G]_{\beta} coincide up to phase factors. ∎

Remark 2.17.

Even though [G]β{\mathbb{C}}[G]_{\beta} has fundamental representation (π(U),U)(\pi(U),U), condition (2.12) can be satisfied for another pair (Z,U)(Z^{\prime},U), which is then also a fundamental representation. A particularly interesting situation is when ZMatm([T/T0])Z^{\prime}\in\operatorname{Mat}_{m}({\mathbb{C}}[T/T_{0}]) for a [G]{\mathbb{C}}[G]-cocentral subgroup T0TT_{0}\subset T. In this case we can view [G]β{\mathbb{C}}[G]_{\beta} as a braided compact matrix quantum group over (T/T0,Z,iβ)(T/T_{0},Z^{\prime},i^{*}\beta).   \diamond

We record a useful lemma related to the above remark.

Lemma 2.18.

Assume AHopf([T],β)A\in\mathrm{Hopf}^{*}({\mathbb{C}}[T],\beta) and take wT^w\in\hat{T}. Then (Z,U)(Z,U) defines an AA-comodule if and only if (wZ,U)(wZ,U) defines an AA-comodule. If in addition ZZ and UU are unitary, then we have the relation

U¯wZ=DU¯ZD1,D=(β(zij,w))i,j.\bar{U}_{wZ}=D\bar{U}_{Z}D^{-1},\quad D=(\beta(z_{ij}^{*},w))_{i,j}.
Proof.

The first claim is obvious from condition (2.12). The second claim is easy to check in an orthonormal basis where ZZ is diagonal, in which case it follows immediately from (2.13). More conceptually, one can check that for the corepresentation matrix W=(kzik#ukj)i,jW=(\sum_{k}z_{ik}\#u_{kj})_{i,j} for [T]#A{\mathbb{C}}[T]\#A we have

(w#1)W=XW(w#1)X1,(w\#1)W=XW(w\#1)X^{-1},

where X=(β(zij,w))i,jX=(\beta(z_{ij},w))_{i,j}, which is a matrix commuting with ZZ. This implies that U¯wZ=X¯U¯ZX¯1\bar{U}_{wZ}=\bar{X}\bar{U}_{Z}\bar{X}^{-1}. It remains to observe that β(x,w)¯=β(x,w)\overline{\beta(x,w)}=\beta(x^{*},w) for all x[T]x\in{\mathbb{C}}[T] to see that X¯=D\bar{X}=D. ∎


3. Examples: transmuting matrix quantum groups

Before we embark on the examples, we remark that in a number of recent papers (see, e.g., [Kas+16, MR22, BJR22, Ans+22]) braided quantum groups are constructed in a C-algebraic setting. However, the corresponding bosonizations are C-algebraic compact quantum groups, and these always have dense *-subalgebras of matrix coefficients, which leads to purely algebraic results. Conversely, in our examples the bosonizations will be compact quantum groups by Theorem 2.5 (as unitary cocycle twisting preserves compactness) or Proposition 2.15, and hence they can be completed to C-algebraic compact quantum groups. We can therefore go back and forth between the *-algebraic and C-algebraic settings. Below we will not dwell on the specific details of this but rather stick to the algebraic picture.

3.1. Braided SUq(2)SU_{q}(2)

Fix q>0q>0 and recall that H:=[SUq(2)]H:={\mathbb{C}}[SU_{q}(2)] is the universal unital *-algebra with generators α\alpha and γ\gamma subject to the relations

αγ=qγα,αγ=qγα,γγ=γγ,\alpha\gamma=q\gamma\alpha,\quad\alpha\gamma^{*}=q\gamma^{*}\alpha,\quad\gamma^{*}\gamma=\gamma\gamma^{*},
αα+γγ=1,αα+q2γγ=1.\quad\alpha^{*}\alpha+\gamma^{*}\gamma=1,\quad\alpha\alpha^{*}+q^{2}\gamma\gamma^{*}=1.

It is a Hopf *-algebra with coproduct

Δ(α)=ααqγγ,Δ(γ)=γα+αγ.\Delta(\alpha)=\alpha\otimes\alpha-q\gamma^{*}\otimes\gamma,\quad\Delta(\gamma)=\gamma\otimes\alpha+\alpha^{*}\otimes\gamma.

Consider the map

π:H[𝕋]=[z,z1],π(α)=z,π(γ)=0.\pi\colon H\to{\mathbb{C}}[{\mathbb{T}}]={\mathbb{C}}[z,z^{-1}],\quad\pi(\alpha)=z,\quad\pi(\gamma)=0.

Under the identification =𝕋^{\mathbb{Z}}=\hat{\mathbb{T}}, we have

αH1,1,γH1,1,\alpha\in H_{1,1},\qquad\gamma\in H_{{-1},1},

and the restricted right adjoint coaction adπ\mathrm{ad}_{\pi} is determined by

adπ(α)=α1,adπ(γ)=γz2.\mathrm{ad}_{\pi}(\alpha)=\alpha\otimes 1,\quad\mathrm{ad}_{\pi}(\gamma)=\gamma\otimes z^{2}.

For λ𝕋\lambda\in{\mathbb{T}}, define a bicharacter on =𝕋^{\mathbb{Z}}=\hat{\mathbb{T}} by βλ(m,n)=λmn\beta_{\lambda}(m,n)=\lambda^{-mn}. To find relations in the transmutation Hλ=HβλH_{\lambda}=H_{\beta_{\lambda}} we write ab=aβλba\cdot b=a\cdot_{\beta_{\lambda}}b and aλ=aβλa^{*_{\lambda}}=a^{*_{\beta_{\lambda}}}. Then, by (2.4) and (2.6),

αλ=βλ(0,1)α=α,γλ=βλ(2,1)γ=λ2γ,\alpha^{*_{\lambda}}=\beta_{\lambda}(0,{-1})\alpha^{*}=\alpha^{*},\qquad\gamma^{*_{\lambda}}=\beta_{\lambda}(2,1)\gamma^{*}=\lambda^{-2}\gamma^{*},

and

αγ=βλ(0,1)αγ=αγ=qγα=qβλ(2,1)1γα=qλ2γα,\alpha\cdot\gamma=\beta_{\lambda}(0,1)\alpha\gamma=\alpha\gamma=q\gamma\alpha=q\beta_{\lambda}(2,{-1})^{-1}\gamma\cdot\alpha=q\lambda^{-2}\gamma\cdot\alpha,
αγλ=βλ(0,1)αγλ=αγλ=qγλα=qβλ(2,1)1γλα=qλ2γλα,\alpha\cdot\gamma^{*_{\lambda}}=\beta_{\lambda}(0,{-1})\alpha\gamma^{*_{\lambda}}=\alpha\gamma^{*_{\lambda}}=q\gamma^{*_{\lambda}}\alpha=q\beta_{\lambda}({-2},{-1})^{-1}\gamma^{*_{\lambda}}\cdot\alpha=q\lambda^{2}\gamma^{*_{\lambda}}\cdot\alpha,
γλγ=βλ(2,1)γλγ=γγ=γγ=βλ(2,1)γγλ=γγλ,\gamma^{*_{\lambda}}\cdot\gamma=\beta_{\lambda}({-2},1)\gamma^{*_{\lambda}}\gamma=\gamma^{*}\gamma=\gamma\gamma^{*}=\beta_{\lambda}(2,{-1})\gamma\gamma^{*_{\lambda}}=\gamma\cdot\gamma^{*_{\lambda}},
ααλ=αα,αα=ααλ.\alpha\cdot\alpha^{*_{\lambda}}=\alpha\alpha^{*},\qquad\alpha^{*}\alpha=\alpha\cdot\alpha^{*_{\lambda}}.

Defining q=qλ2q^{\prime}=q\lambda^{2} we get the following relations in HλH_{\lambda}:

αγ=q¯γα,αγλ=qγλα,γλγ=γγλ,\alpha\cdot\gamma=\bar{q}\,^{\prime}\gamma\cdot\alpha,\quad\alpha\cdot\gamma^{*_{\lambda}}=q^{\prime}\gamma^{*_{\lambda}}\cdot\alpha,\quad\gamma^{*_{\lambda}}\cdot\gamma=\gamma\cdot\gamma^{*_{\lambda}},
αλα+γλγ=1,ααλ+|q|2γγλ=1.\quad\alpha^{*_{\lambda}}\cdot\alpha+\gamma^{*_{\lambda}}\cdot\gamma=1,\quad\alpha\cdot\alpha^{*_{\lambda}}+|q^{\prime}|^{2}\gamma\cdot\gamma^{*_{\lambda}}=1.

It is not difficult to see that these relations completely describe the transmuted algebra; in the next subsection we will prove a more general result. The coproduct remains unchanged, so we have

Δ(α)=ααqγλγ,Δ(γ)=γα+αλγ.\Delta(\alpha)=\alpha\otimes\alpha-q^{\prime}\gamma^{*_{\lambda}}\otimes\gamma,\qquad\Delta(\gamma)=\gamma\otimes\alpha+\alpha^{*_{\lambda}}\otimes\gamma.

These formulas are the same as for the braided quantum group SUq(2)SU_{q^{\prime}}(2) constructed in [Kas+16], modulo a small but important nuance. By Proposition 2.4, HλH_{\lambda} can be viewed as a braided quantum group over different tori. Namely, we see that HλH_{\lambda} can be viewed as a braided compact matrix quantum group over both triples

(𝕋,(z00z1),βλ) and (𝕋/T0,(z2001),iβλ),({\mathbb{T}},\,\begin{pmatrix}z&0\\ 0&z^{-1}\end{pmatrix},\,\beta_{\lambda})\quad\mbox{ and }\quad({\mathbb{T}}/T_{0},\,\begin{pmatrix}z^{2}&0\\ 0&1\end{pmatrix},\,i^{*}\beta_{\lambda}),

where T0={1,1}𝕋T_{0}=\{-1,1\}\subset{\mathbb{T}}, see Remark 2.17. As [𝕋/T0]{\mathbb{C}}[{\mathbb{T}}/T_{0}] is generated by z2z^{2} we have the isomorphism

f:[𝕋/T0][w,w1],f(z2)=w.f\colon{\mathbb{C}}[\mathbb{T}/T_{0}]\to{\mathbb{C}}[w,w^{-1}],\quad f(z^{2})=w.

Moreover, (if1)βλ=βζ(i\circ f^{-1})^{*}\beta_{\lambda}=\beta_{\zeta}, where ζ=λ4=q/q¯\zeta=\lambda^{4}=q^{\prime}/\bar{q}\,^{\prime}. Therefore we can consider HλH_{\lambda} as a braided compact matrix quantum group over the triple

(𝕋,(w001),βζ).({\mathbb{T}},\,\begin{pmatrix}w&0\\ 0&1\end{pmatrix},\,\beta_{\zeta}).

This is the braided quantum group [SUq(2)]Hopf([𝕋],βζ){\mathbb{C}}[SU_{q^{\prime}}(2)]\in\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}],\beta_{\zeta}) considered in [Kas+16].

Finally, let us consider the bosonizations. By Theorem 2.5 we have

[z,z1]#Hλ([z,z1][SUq(2)])J1J,{\mathbb{C}}[z,z^{-1}]\#H_{\lambda}\cong{}_{J}({\mathbb{C}}[z,z^{-1}]\otimes{\mathbb{C}}[SU_{q}(2)])_{J^{-1}},

where J((m,n),(m,n))=λnmJ((m,n),({m^{\prime}},{n^{\prime}}))=\lambda^{nm^{\prime}}. In other words, the bosonization is a cocycle twist of the compact quantum group 𝕋×SUq(2){\mathbb{T}}\times SU_{q}(2).

On the other hand, by the same theorem, the bosonization [w,w1]#Hλ{\mathbb{C}}[w,w^{-1}]\,\#\,H_{\lambda} is a cocycle twist of (𝕋×SUq(2))/Δ(T0)({\mathbb{T}}\times SU_{q}(2))/\Delta(T_{0}). It is easy to see that the latter quantum group is isomorphic to Uq(2)U_{q}(2), similarly to the classical isomorphism

(𝕋×SU(2))/Δ(T0)U(2),[(z,U)](z00z)U,({\mathbb{T}}\times SU(2))/{\Delta(T_{0})}\cong U(2),\quad[(z,U)]\mapsto\begin{pmatrix}z&0\\ 0&z\end{pmatrix}U,

and therefore its cocycle twist must be one of the quantum deformations of U(2)U(2) studied in [ZZ05], cf. [Kas+16].

3.2. Braided free orthogonal quantum groups

Let m2m\geq 2 be a natural number, FGLm()F\in\mathrm{GL}_{m}({\mathbb{C}}) and assume that FF¯=±1F\bar{F}=\pm 1. Let [OF+]{\mathbb{C}}[O_{F}^{+}] be the universal unital *-algebra generated by elements uiju_{ij}, 1i,jm1\leq i,j\leq m, subject to the relations

U=(uij)i,jis unitary andU=FU¯F1.U=(u_{ij})_{i,j}\quad\mbox{is unitary and}\quad U=F\bar{U}F^{-1}.

The Hopf *-algebra structure on [OF+]{\mathbb{C}}[O_{F}^{+}] is defined as in Definition 2.12, and OF+O_{F}^{+} is called a free orthogonal quantum group.

Let TT be a compact abelian group and ZMatm([T])Z\in\operatorname{Mat}_{m}({\mathbb{C}}[T]) be a unitary corepresentation matrix satisfying FZ¯F1=ZF\bar{Z}F^{-1}=Z. By the universality of [OF+]{\mathbb{C}}[O_{F}^{+}] there is a Hopf *-algebra map π:[OF+][T]\pi\colon{\mathbb{C}}[O_{F}^{+}]\to{\mathbb{C}}[T] such that π(uij)=zij\pi(u_{ij})=z_{ij}. Fix a bicharacter β:T^×T^𝕋\beta\colon\hat{T}\times\hat{T}\to{\mathbb{T}} and consider the transmutation [OF+]β{\mathbb{C}}[O_{F}^{+}]_{\beta}. It is natural to call it a braided free orthogonal quantum group.

Proposition 3.1.

The braided Hopf *-algebra [OF+]β{\mathbb{C}}[O_{F}^{+}]_{\beta} is a braided compact matrix quantum group over (T,Z,β)(T,Z,\beta) with fundamental unitary U=(uij)i,jU=(u_{ij})_{i,j}. As a *-algebra, it is a universal unital *-algebra with generators uiju_{ij} satisfying the relations

U=(uij)i,jis unitary andU=FU¯ZF1,U=(u_{ij})_{i,j}\quad\mbox{is unitary and}\quad U=F\bar{U}_{Z}F^{-1}, (3.1)

where U¯Z=(u¯ijZ)i,j\bar{U}_{Z}=(\bar{u}_{ij}^{Z})_{i,j} and u¯ijZ=s,l,tβ(ztjzsl,zil)ust\bar{u}_{ij}^{Z}=\sum_{s,l,t}\beta(z_{tj}^{*}z_{sl},z_{il}^{*})u_{st}^{*}.

Proof.

For the purpose of this proof let us denote the fundamental unitary of OF+O^{+}_{F} by V=(vij)i,jV=(v_{ij})_{i,j} and write AA for [OF+]β{\mathbb{C}}[O_{F}^{+}]_{\beta}. The first claim follows from Proposition 2.16. Relations (3.1) are obtained by considering, as in the proof of that proposition, the Hopf *-algebra map ϕ:[OF+][T]#A\phi\colon{\mathbb{C}}[O^{+}_{F}]\to{\mathbb{C}}[T]\#A, ϕ(x)=π(x(1))#x(2)\phi(x)=\pi(x_{(1)})\#x_{(2)}, and using that ϕ(V)=Z#U\phi(V)=Z\#U, ϕ(V¯)=Z¯#U¯Z\phi(\bar{V})=\bar{Z}\#\bar{U}_{Z}.

It remains to show that as a *-algebra AA is completely described by relations (3.1). Consider a universal unital *-algebra A~\tilde{A} with generators u~ij\tilde{u}_{ij} satisfying these relations, and let ρ:A~A\rho\colon\tilde{A}\to A be the *-homomorphism such that ρ(u~ij)=uij\rho(\tilde{u}_{ij})=u_{ij}.

Working in a basis where ZZ is diagonal, it is not difficult to check that A~\tilde{A} is a [T]{\mathbb{C}}[T]-comodule *-algebra, with the coaction of [T]{\mathbb{C}}[T] given by

δ~(u~ij)=s,tu~stzsiztj,or(ιδ~)(U)=Z13U~12Z13.\tilde{\delta}(\tilde{u}_{ij})=\sum_{s,t}\tilde{u}_{st}\otimes z_{si}^{*}z_{tj},\quad\text{or}\quad(\iota\otimes\tilde{\delta})(U)=Z^{*}_{13}\tilde{U}_{12}Z_{13}.

Consider the smash product [T]#A~=[T]βA~{\mathbb{C}}[T]\#\tilde{A}={\mathbb{C}}[T]\otimes_{\beta}\tilde{A}. It is again not difficult to check that we have a *-homomorphism

ϕ~:[OF+][T]#A~,ϕ~(vij)=kzik#u~kj.\tilde{\phi}\colon{\mathbb{C}}[O^{+}_{F}]\to{\mathbb{C}}[T]\#\tilde{A},\quad\tilde{\phi}(v_{ij})=\sum_{k}z_{ik}\#\tilde{u}_{kj}.

Define linear maps

ψ~:[T][OF+][T]#A~,ψ~(xa)=xϕ~(a),\displaystyle\tilde{\psi}\colon{\mathbb{C}}[T]\otimes{\mathbb{C}}[O^{+}_{F}]\to{\mathbb{C}}[T]\#\tilde{A},\qquad\tilde{\psi}(x\otimes a)=x\tilde{\phi}(a),
ψ:[T][OF+][T]#A,ψ(xa)=xπ(a(1))#a(2).\displaystyle\psi\colon{\mathbb{C}}[T]\otimes{\mathbb{C}}[O^{+}_{F}]\to{\mathbb{C}}[T]\#A,\qquad\psi(x\otimes a)=x\pi(a_{(1)})\#a_{(2)}.

Then ψ=(ι#ρ)ψ~\psi=(\iota\#\rho)\tilde{\psi}. The map ψ\psi is a linear isomorphism, e.g., by Theorem 2.5. On the other hand, the map ψ~\tilde{\psi} is surjective, which becomes particularly clear if we work in a basis where ZZ is diagonal and therefore ϕ~(vij)=zii#u~ij\tilde{\phi}(v_{ij})=z_{ii}\#\tilde{u}_{ij}. (Alternatively, we can observe that ψ~\tilde{\psi} defines a homomorphism [T]#[OF+][T]#A~{\mathbb{C}}[T]\#{\mathbb{C}}[O^{+}_{F}]\to{\mathbb{C}}[T]\#\tilde{A}, cf. Remark 1.6, and its image contains the elements 1#u~ij1\#\tilde{u}_{ij}.) It follows that ψ~\tilde{\psi} is a linear isomorphism and hence ρ\rho is an isomorphism as well. ∎

Next, we want to change the perspective on the braided free orthogonal quantum groups and show how they can be associated with a larger class of matrices than FF as above.

Proposition 3.2.

Let AGLm()A\in\mathrm{GL}_{m}({\mathbb{C}}) (m2m\geq 2) be a matrix such that AA¯A\bar{A} is unitary, and choose a sign τ=±1\tau=\pm 1, with τ=1\tau=1 if mm is odd. Then there are a compact abelian group TT, a unitary corepresentation matrix X=(xij)i,jMatm([T])X=(x_{ij})_{i,j}\in\operatorname{Mat}_{m}({\mathbb{C}}[T]), a character wT^w\in\hat{T} and a bicharacter β:T^×T^𝕋\beta\colon\hat{T}\times\hat{T}\to{\mathbb{T}} such that

A(w2X¯)A1=XinMatm([T])A(w^{2}\bar{X})A^{-1}=X\quad\mbox{in}\quad\operatorname{Mat}_{m}({\mathbb{C}}[T]) (3.2)

and ACAC¯=τ1AC\overline{AC}=\tau 1, where C=(β(xij,w))i,jC=(\beta(x_{ij}^{*},w))_{i,j}. For every such quadruple (T,X,w,β)(T,X,w,\beta), consider a universal unital *-algebra [OAX,β]{\mathbb{C}}[O_{A}^{X,\beta}] with generators uiju_{ij} satisfying the relations

U=(uij)i,jis unitary andU=AU¯XA1,U=(u_{ij})_{i,j}\quad\mbox{is unitary and}\quad U=A\bar{U}_{X}A^{-1},

where U¯X=(u¯ijX)i,j\bar{U}_{X}=(\bar{u}_{ij}^{X})_{i,j} and u¯ijX=s,l,tβ(xtjxsl,xil)ust\bar{u}_{ij}^{X}=\sum_{s,l,t}\beta(x_{tj}^{*}x_{sl},x_{il}^{*})u_{st}^{*}. Then [OAX,β]{\mathbb{C}}[O_{A}^{X,\beta}], equipped with the coaction

δ(uij)=s,tustxsixtj,\delta(u_{ij})=\sum_{s,t}u_{st}\otimes x_{si}^{*}x_{tj},

is a braided compact matrix quantum group over (T,X,β)(T,X,\beta) with fundamental unitary UU.

Proof.

Assume first that a quadruple (T,X,w,β)(T,X,w,\beta) as in the formulation indeed exists. Define F=ACF=AC. By our assumptions this matrix satisfies FF¯=τ1F\bar{F}=\tau 1 and, as CC commutes with X¯\bar{X}, we have

F(wX¯)F1=w1X.F(w\bar{X})F^{-1}=w^{-1}X.

By universality there is a Hopf *-homomorphism π:[OF+][T]\pi\colon{\mathbb{C}}[O_{F}^{+}]\to{\mathbb{C}}[T] sending the fundamental representation to Z=w1XZ=w^{-1}X. We claim that the corresponding transmutation [OF+]β{\mathbb{C}}[O_{F}^{+}]_{\beta} satisfies all the required properties of [OAX,β]{\mathbb{C}}[O_{A}^{X,\beta}].

Indeed, by Lemma 2.18, in [OF+]β{\mathbb{C}}[O_{F}^{+}]_{\beta} we have

U¯X=U¯wZ=DU¯ZD1,\bar{U}_{X}=\bar{U}_{wZ}=D\bar{U}_{Z}D^{-1},

where D=(β(zij,w))i,j=(β(wxij,w))i,j=β(w,w)CD=(\beta(z^{*}_{ij},w))_{i,j}=(\beta(wx^{*}_{ij},w))_{i,j}=\beta(w,w)C. Then A=β(w,w)FD1A=\beta(w,w)FD^{-1}, and the claim follows from Proposition 3.1.

Next we explain the existence of (T,X,w,β)(T,X,w,\beta). By [HN21, Proposition 1.5], we can find a unitary vv such that vAvtvAv^{t} has the form

(0ama10),aia¯mi+1=λi𝕋.\begin{pmatrix}0&&a_{m}\\ &\iddots&\\ a_{1}&&0\end{pmatrix},\quad a_{i}\bar{a}_{m-i+1}=\lambda_{i}\in{\mathbb{T}}. (3.3)

If we can find a quadruple (T,X,w,β)(T,X,w,\beta) for this matrix, then (T,vX()v,w,β)(T,v^{*}X(\cdot)v,w,\beta) is a quadruple for AA. Thus, we may assume that AA has the above form.

We will construct TT and XX such that XX is diagonal, so X(t)=diag(x1(t),,xm(t))X(t)=\operatorname{diag}(x_{1}(t),\dots,x_{m}(t)) for some characters xix_{i}. The conditions (3.2) and ACAC¯=τ1AC\overline{AC}=\tau 1 for C=diag(β(x11,w),,β(xm1,w))C=\operatorname{diag}(\beta(x_{1}^{-1},w),\dots,\beta(x_{m}^{-1},w)) mean then that

xixmi+1=w2andβ(xi1xmi+1,w)=τλi.x_{i}x_{m-i+1}=w^{2}\qquad\text{and}\qquad\beta(x_{i}^{-1}x_{m-i+1},w)=\tau\lambda_{i}. (3.4)

If m=2km=2k, these conditions can be easily satisfied for the dual TT of a free abelian group with independent generators x1,,xk,wx_{1},\dots,x_{k},w by letting xmi+1=w2xi1x_{m-i+1}=w^{2}x_{i}^{-1} for 1ik1\leq i\leq k. If m=2k+1m=2k+1, then τ=1\tau=1, λk+1=1\lambda_{k+1}=1 and the conditions can be satisfied for the dual TT of a free abelian group with independent generators x1,,xk+1x_{1},\dots,x_{k+1} by letting w=xk+1w=x_{k+1} and xmi+1=w2xi1x_{m-i+1}=w^{2}x_{i}^{-1} for 1ik1\leq i\leq k. ∎

As is clear from the proof of this proposition, the braided quantum groups OAX,βO^{X,\beta}_{A} lie within the class of braided free orthogonal quantum groups that we defined by transmutation. Namely, we have the following:

Corollary 3.3.

The braided Hopf *-algebra [OAX,β]Hopf([T],β){\mathbb{C}}[O_{A}^{X,\beta}]\in\mathrm{Hopf}^{*}({\mathbb{C}}[T],\beta) is isomorphic to the transmutation [OF+]β{\mathbb{C}}[O_{F}^{+}]_{\beta} with respect to the map [OF+][T]{\mathbb{C}}[O_{F}^{+}]\to{\mathbb{C}}[T], Uw1XU\mapsto w^{-1}X, where F=ACF=AC.

Remark 3.4.

A moment’s reflection shows that in the proof of Proposition 3.2 we could take a slightly smaller group TT and arrange XX to be faithful. Namely, if m=2km=2k, instead of taking ww as a separate independent generator, we could let w=x1jw=x_{1}^{j} for any j0,1j\neq 0,1. Similarly, for m=2k+1m=2k+1 we could take xk+1=w=x1jx_{k+1}=w=x_{1}^{j} for any j0,1j\neq 0,1. In both cases we cannot choose groups of a smaller rank in general, since the numbers τλi\tau\lambda_{i} generate a group of rank up to k=[m/2]k=[m/2].

Remark 3.5.

Once (3.2) is satisfied, condition ACAC¯=τ1AC\overline{AC}=\tau 1 can be formulated as follows. Let twTt_{w}\in T be the element such that x(tw)=β(x,w)x(t_{w})=\beta(x,w) for all xT^x\in\hat{T}, so that C=X(tw)¯C=\overline{X(t_{w})}. Then the requirement is

AA¯=τβ(w,w)2X(tw)2.A\bar{A}=\tau\beta(w,w)^{2}X(t_{w})^{-2}. (3.5)

As a prerequisite for constructing OAX,βO^{X,\beta}_{A} this condition can be written as

AA¯=cX(tw)2for somec𝕋.A\bar{A}=c\,X(t_{w})^{-2}\quad\text{for some}\quad c\in{\mathbb{T}}. (3.6)

Indeed, assume (3.2) and (3.6) are satisfied. Then applying complex conjugation and conjugation by AA to the last identity we get

AA¯=c¯β(w,w)4X(tw)2.A\bar{A}=\bar{c}\,\beta(w,w)^{4}X(t_{w})^{-2}.

Hence c=c¯β(w,w)4c=\bar{c}\,\beta(w,w)^{4} and therefore τ:=cβ(w,w)2=±1\tau:=c\,\beta(w,w)^{-2}=\pm 1, so that (3.5) is satisfied for this τ\tau. Note that the sign must be +1+1 for odd mm, which becomes obvious if we choose a unitary vv such that vAX(tw)¯vtvA\,\overline{X(t_{w})}v^{t} is of the form (3.3). Thus, the braided compact matrix quantum groups OAX,βO^{X,\beta}_{A} are defined under assumptions (3.2) and (3.6).  \diamond

In the setting of Proposition 3.1, assume now that T0TT_{0}\subset T is a closed [OF+]{\mathbb{C}}[O_{F}^{+}]-cocentral subgroup. Since the elements uij[OF+]u_{ij}\in{\mathbb{C}}[O^{+}_{F}] are linearly independent, this means that the matrices Z(t)Z(t), tT0t\in T_{0}, are scalar. Then the condition Z=FZ¯F1Z=F\bar{Z}F^{-1} implies that Z(t)=±1Z(t)=\pm 1, so we get a character χ:T0{±1}\chi\colon T_{0}\to\{\pm 1\}. When it is nontrivial, it defines the standard (/2)({\mathbb{Z}}/2{\mathbb{Z}})-grading on RepOF+\operatorname{Rep}O^{+}_{F}.

It is known that the quantum group OF+O^{+}_{F} is monoidally equivalent to SUq(2)SU_{q}(2) for an appropriate qq (see [NT13, Theorem 2.5.11]), and this equivalence respects the (/2)({\mathbb{Z}}/2{\mathbb{Z}})-gradings. Combining this with Remark 2.7, one can conclude that the bosonization of the braided Hopf *-algebra [OF+]βHopf([T/T0],iβ){\mathbb{C}}[O^{+}_{F}]_{\beta}\in\mathrm{Hopf}^{*}({\mathbb{C}}[T/T_{0}],i^{*}\beta) is monoidally equivalent to

(T×SUq(2))/(id×χ)Δ(T0).(T\times SU_{q}(2))/{(\operatorname{id}\times\chi)\Delta(T_{0})}.

Together with Corollary 3.3 this leads to the following conclusion.

Proposition 3.6.

In the setting of Proposition 3.2, let q[1,1]{0}q\in[-1,1]\setminus\{0\} be such that sgnq=τ\operatorname{sgn}q=-\tau and |q+q1|=Tr(AA)|q+q^{-1}|=\operatorname{Tr}(A^{*}A). Assume T0TT_{0}\subset T is a closed subgroup satisfying X(t)=±w(t)1X(t)=\pm w(t)1 for all tT0t\in T_{0}, and let χ:T0{±1}\chi\colon T_{0}\to\{\pm 1\} be the character such that X=χw1X=\chi w1 on T0T_{0}. Then the bosonization of [OAX,β]Hopf([T/T0],iβ){\mathbb{C}}[O^{X,\beta}_{A}]\in\mathrm{Hopf}^{*}({\mathbb{C}}[T/T_{0}],i^{*}\beta) is a compact quantum group monoidally equivalent to

(T×SUq(2))/(id×χ)Δ(T0).(T\times SU_{q}(2))/{(\operatorname{id}\times\chi)\Delta(T_{0})}. (3.7)

We remark that the sign of qq is not uniquely determined by a monoidal equivalence between (3.7) and the bosonization of OAX,βO^{X,\beta}_{A} in general, since Uq(2)U_{q}(2) and Uq(2)U_{-q}(2) are cocycle twists of each other.

Next, it is possible to obtain a classification of braided free orthogonal quantum groups up to isomorphism similar to the known classification of ordinary free orthogonal quantum groups, see again [NT13, Theorem 2.5.11].

Proposition 3.7.

Consider a compact abelian group TT and a bicharacter β:T^×T^𝕋\beta\colon\hat{T}\times\hat{T}\to{\mathbb{T}}. Assume AGLm()A\in\mathrm{GL}_{m}({\mathbb{C}}) (m2m\geq 2) is a matrix such that AA¯A\bar{A} is unitary and X=(xij)i,jMatm([T])X=(x_{ij})_{i,j}\in\operatorname{Mat}_{m}({\mathbb{C}}[T]) is a unitary corepresentation matrix such that conditions (3.2) and (3.6) are satisfied for some wT^w\in\hat{T}. Assume AGLm()A^{\prime}\in\mathrm{GL}_{m^{\prime}}({\mathbb{C}}) (m2m^{\prime}\geq 2) and X=(xij)i,jMatm([T])X^{\prime}=(x^{\prime}_{ij})_{i,j}\in\operatorname{Mat}_{m^{\prime}}({\mathbb{C}}[T]) is another such pair, with (3.2) and (3.6) satisfied for some wT^w^{\prime}\in\hat{T}. Then the braided Hopf *-algebras [OAX,β]{\mathbb{C}}[O^{X,\beta}_{A}] and [OAX,β]{\mathbb{C}}[O^{X^{\prime},\beta}_{A^{\prime}}] over ([T],β)({\mathbb{C}}[T],\beta) are isomorphic if and only if m=mm=m^{\prime} and there exist a unitary matrix vU(m)v\in\operatorname{U}(m) and a character χT^\chi\in\hat{T} such that

vXv=χXandvADvt=A,vXv^{*}=\chi X^{\prime}\quad\text{and}\quad vADv^{t}=A^{\prime}, (3.8)

where D=(β(xij,χ))i,j=X(tχ)¯D=(\beta(x^{*}_{ij},\chi))_{i,j}=\overline{X(t_{\chi})}.

Proof.

Denote by UU and UU^{\prime} the fundamental unitaries of [OAX,β]{\mathbb{C}}[O^{X,\beta}_{A}] and [OAX,β]{\mathbb{C}}[O^{X^{\prime},\beta}_{A^{\prime}}], resp. If conditions (3.8) are satisfied, then U¯χ1X=D1U¯XD\bar{U}_{\chi^{-1}X}=D^{-1}\bar{U}_{X}D by Lemma 2.18, hence U=ADU¯χ1X(AD)1U=AD\bar{U}_{\chi^{-1}X}(AD)^{-1}, and it is easy to check that we get an isomorphism [OAX,β][OAX,β]{\mathbb{C}}[O^{X,\beta}_{A}]\cong{\mathbb{C}}[O^{X^{\prime},\beta}_{A^{\prime}}] of braided Hopf *-algebras such that UvUvU\mapsto v^{*}U^{\prime}v.

Conversely, assume we have an isomorphism ρ:[OAX,β][OAX,β]\rho\colon{\mathbb{C}}[O^{X,\beta}_{A}]\to{\mathbb{C}}[O^{X^{\prime},\beta}_{A^{\prime}}]. We then get an isomorphism ι#ρ:[T]#[OAX,β][T]#[OAX,β]\iota\#\rho\colon{\mathbb{C}}[T]\#{\mathbb{C}}[O^{X,\beta}_{A}]\to{\mathbb{C}}[T]\#{\mathbb{C}}[O^{X^{\prime},\beta}_{A^{\prime}}] of the bosonizations, which in turn defines a monoidal equivalence of the corresponding C-tensor categories of finite dimensional unitary comodules. For [T]#[OAX,β]{\mathbb{C}}[T]\#{\mathbb{C}}[O^{X,\beta}_{A}], as we discussed before Proposition 3.6, this C-tensor category is RepTRepSUq(2)\operatorname{Rep}T\boxtimes\operatorname{Rep}SU_{q}(2) for suitable q[1,1]{0}q\in[-1,1]\setminus\{0\}. The simple noninvertible objects of the smallest intrinsic dimension (equal to |q+q1||q+q^{-1}|) in this category are the tensor products of characters χT^\chi\in\hat{T} with the fundamental representation of SUq(2)SU_{q}(2). At the level of [T]#[OAX,β]{\mathbb{C}}[T]\#{\mathbb{C}}[O^{X,\beta}_{A}], these objects are defined by the unitary corepresentation matrices χX#U\chi X\#U. For the same reason, the simple noninvertible objects of the smallest intrinsic dimension in the category of finite dimensional unitary comodules of [T]#[OAX,β]{\mathbb{C}}[T]\#{\mathbb{C}}[O^{X^{\prime},\beta}_{A^{\prime}}] are defined by the unitary corepresentation matrices χX#U\chi X^{\prime}\#U^{\prime}. It follows that there exist χT^\chi\in\hat{T} and a unitary v:mmv\colon{\mathbb{C}}^{m}\to{\mathbb{C}}^{m^{\prime}} such that

v(ι#ρ)(X#U)v=χX#U.v(\iota\#\rho)(X\#U)v^{*}=\chi X^{\prime}\#U^{\prime}.

In particular, we must have m=mm=m^{\prime}. As (ι#ρ)(X#U)=X#ρ(U)(\iota\#\rho)(X\#U)=X\#\rho(U), by applying ι#ε[OAX,β]\iota\#\varepsilon_{{\mathbb{C}}[O^{X^{\prime},\beta}_{A^{\prime}}]} we first conclude that vXv=χXvXv^{*}=\chi X^{\prime} and then that vρ(U)v=Uv\rho(U)v^{*}=U^{\prime}.

As vχ1Xv=Xv\chi^{-1}Xv^{*}=X^{\prime} and vρ(U)v=Uv\rho(U)v^{*}=U^{\prime}, we have v¯ρ(U¯χ1X)v¯=U¯X\bar{v}\rho(\bar{U}_{\chi^{-1}X})\bar{v}^{*}=\bar{U}^{\prime}_{X^{\prime}}. Recall also that U¯χ1X=D1U¯XD\bar{U}_{\chi^{-1}X}=D^{-1}\bar{U}_{X}D for D=(β(xij,χ))i,jD=(\beta(x^{*}_{ij},\chi))_{i,j}. By the relations in [OAX,β]{\mathbb{C}}[O^{X,\beta}_{A}] we then get

U=vρ(U)v=vρ(ADU¯χ1X(AD)1)v=vADvtU¯Xv¯(AD)1v.U^{\prime}=v\rho(U)v^{*}=v\rho(AD\bar{U}_{\chi^{-1}X}(AD)^{-1})v^{*}=vADv^{t}\bar{U}^{\prime}_{X^{\prime}}\bar{v}(AD)^{-1}v^{*}.

On the other hand, U=AU¯XA1U^{\prime}=A^{\prime}\bar{U}^{\prime}_{X^{\prime}}{A^{\prime}}^{-1} by the relations in [OAX,β]{\mathbb{C}}[O^{X^{\prime},\beta}_{A^{\prime}}]. Since the matrix coefficients of UU^{\prime} are linearly independent, it follows that

vADvt=λAfor someλ×.vADv^{t}=\lambda A^{\prime}\quad\text{for some}\quad\lambda\in{\mathbb{C}}^{\times}.

As both AA¯A^{\prime}\bar{A^{\prime}} and ADAD¯AD\overline{AD} are unitary (recall that D=X(tχ)¯D=\overline{X(t_{\chi})} is unitary and ADAD coincides with D¯A\bar{D}A up to a phase factor), we must have λ𝕋\lambda\in{\mathbb{T}}. Hence, by multiplying vv by a phase factor we can achieve that vADvt=AvADv^{t}=A^{\prime}, while the equality vXv=χXvXv^{*}=\chi X^{\prime} is still satisfied. ∎

Note that, for fixed (A,X)(A,X) and (A,X)(A^{\prime},X^{\prime}), there can only be finitely many χ\chi such vXv=χXvXv^{*}=\chi X^{\prime} for some vv. Furthermore, if T^\hat{T} is torsion-free (equivalently, TT is connected), then the only possible candidate for such χ\chi is ww1w{w^{\prime}}^{-1}, since a nontrivial translation of a finite symmetric subset of T^\hat{T} is never symmetric. Once χ\chi is fixed, the question whether there is vv satisfying both conditions in (3.8) can be solved by writing (AD,χ1X)(AD,\chi^{-1}X) and (A,X)(A^{\prime},X^{\prime}) in a standard form in the following sense.

Lemma 3.8.

Assume AGLm()A\in\mathrm{GL}_{m}({\mathbb{C}}) (m2m\geq 2) is a matrix such that AA¯A\bar{A} is unitary and X=(xij)i,jMatm([T])X=(x_{ij})_{i,j}\in\operatorname{Mat}_{m}({\mathbb{C}}[T]) is a unitary corepresentation matrix such that A(w0X¯)A1=XA(w_{0}\bar{X})A^{-1}=X for some w0T^w_{0}\in\hat{T}. Then there is vU(m)v\in\operatorname{U}(m) such that vAvtvAv^{t} and vXvvXv^{*} are block-diagonal matrices diag(A1,,Ak)\operatorname{diag}(A_{1},\dots,A_{k}) and diag(X1,,Xk)\operatorname{diag}(X_{1},\dots,X_{k}), where the blocks are

  • either 2×22\times 2 matrices

    Ai=(0θiλi1λi0),Xi=(χi00w0χ¯i),A_{i}=\begin{pmatrix}0&\theta_{i}\lambda_{i}^{-1}\\ \lambda_{i}&0\end{pmatrix},\qquad X_{i}=\begin{pmatrix}\chi_{i}&0\\ 0&w_{0}\bar{\chi}_{i}\end{pmatrix},

    with 0<λi10<\lambda_{i}\leq 1, θi𝕋\theta_{i}\in{\mathbb{T}} and χiT^\chi_{i}\in\hat{T} such that if λi=1\lambda_{i}=1, then either 0<argθiπ0<\arg\theta_{i}\leq\pi or (θi=1\theta_{i}=1 and χi2w0\chi^{2}_{i}\neq w_{0});

  • or 1×11\times 1 matrices Ai=1A_{i}=1, Xi=χiX_{i}=\chi_{i}, with χiT^\chi_{i}\in\hat{T} such that χi2=w0\chi^{2}_{i}=w_{0}.

Proof.

The lemma can be viewed as a refinement of [HN21, Proposition 1.5], but the proof is almost the same, so we will be brief.

First one observes that classification of the pairs (A,X)(A,X) up to the transformations AvAvtA\mapsto vAv^{t} and XvXvX\mapsto vXv^{*} is the same as classification of the pairs (AJ,X)(AJ,X) up to unitary conjugacy, where J:mmJ\colon{\mathbb{C}}^{m}\to{\mathbb{C}}^{m} is the complex conjugation. Consider the polar decomposition AJ=u|AJ|AJ=u|AJ|, so |AJ||AJ| is positive and uu is anti-unitary. Then the joint spectrum, considered together with multiplicities, of the commuting operators |AJ||AJ|, u2u^{2} and X(t)X(t) (tTt\in T) is a complete invariant of the unitary conjugacy class, and a standard form of (A,X)(A,X) as in the formulation of the lemma is obtained from an appropriate orthonormal basis diagonalizing these operators, as follows.

For λ>0\lambda>0, θ𝕋\theta\in{\mathbb{T}} and χT^\chi\in\hat{T}, consider the space

Hλ,θ,χ={ξm:|AJ|ξ=λξ,u2ξ=θξ,X(t)ξ=χ(t)ξfor alltT}.H_{\lambda,\theta,\chi}=\{\xi\in{\mathbb{C}}^{m}:|AJ|\xi=\lambda\xi,\ u^{2}\xi=\theta\xi,\ X(t)\xi=\chi(t)\xi\ \text{for all}\ t\in T\}.

As u|AJ|=|AJ|1uu|AJ|=|AJ|^{-1}u and w0uX=uw¯0X=Xuw_{0}uX=u\bar{w}_{0}X=Xu, we have uHλ,θ,χ=Hλ1,θ¯,w0χ¯uH_{\lambda,\theta,\chi}=H_{\lambda^{-1},\bar{\theta},w_{0}\bar{\chi}}. In particular, the spaces Hλ,θ,χ+Hλ1,θ¯,w0χ¯H_{\lambda,\theta,\chi}+H_{\lambda^{-1},\bar{\theta},w_{0}\bar{\chi}} are invariant under AJAJ and XX. For every pair of triples (λ,θ,χ)(\lambda,\theta,\chi) and (λ1,θ¯,w0χ¯)(\lambda^{-1},\bar{\theta},w_{0}\bar{\chi}) such that Hλ,θ,χ0H_{\lambda,\theta,\chi}\neq 0, pick a representative and consider three cases.

1) Assume (λ,θ,χ)(λ1,θ¯,w0χ¯)(\lambda,\theta,\chi)\neq(\lambda^{-1},\bar{\theta},w_{0}\bar{\chi}). Then, by changing the representative if necessary, we may also assume that either λ<1\lambda<1 or (λ=1\lambda=1 and 0argθπ0\leq\arg\theta\leq\pi). Choose an orthonormal basis (ξj)j(\xi_{j})_{j} in Hλ,θ,χH_{\lambda,\theta,\chi}. Then (uξj)j(u\xi_{j})_{j} is an orthonormal basis in Hλ1,θ¯,w0χ¯H_{\lambda^{-1},\bar{\theta},w_{0}\bar{\chi}}, and the restrictions of AJAJ and XX to every 22-dimensional space with basis {ξj,uξj}\{\xi_{j},u\xi_{j}\} have the required form.

2) Assume λ=1\lambda=1, θ=1\theta=-1, χ2=w0\chi^{2}=w_{0}. In this case uξξu\xi\perp\xi for every ξH1,1,χ\xi\in H_{1,-1,\chi}. Hence we can find an orthonormal system (ξj)j(\xi_{j})_{j} in H1,1,χH_{1,-1,\chi} such that (ξj)j(uξj)j(\xi_{j})_{j}\cup(u\xi_{j})_{j} is an orthonormal basis in H1,1,χH_{1,-1,\chi}. The restrictions of AJAJ and XX to every 22-dimensional space with basis {ξj,uξj}\{\xi_{j},u\xi_{j}\} have the required form.

3) Finally, assume λ=1\lambda=1, θ=1\theta=1, χ2=w0\chi^{2}=w_{0}. Then {ξH1,1,χuξ=ξ}\{\xi\in H_{1,1,\chi}\mid u\xi=\xi\} is a real form of H1,1,χH_{1,1,\chi}. By choosing an orthonormal basis in this Euclidean space we get a decomposition of H1,1,χH_{1,1,\chi} into a direct sum of one-dimensional spaces that are invariant under AJAJ and XX. ∎

From the proof one can see that the only source of nonuniqueness of the standard form of (A,X)(A,X), apart from the order of the blocks, is the choice of a representative from each pair of the triples (1,±1,χ)(1,\pm 1,\chi) and (1,±1,w0χ¯)(1,\pm 1,w_{0}\bar{\chi}) with χ2w0\chi^{2}\neq w_{0}.

Remark 3.9.

If T^\hat{T} has zero 22-torsion, then there is at most one χ\chi such that χ2=w0\chi^{2}=w_{0}. Then, as in [HN21], we can choose an orthonormal basis (ξj)j=1l(\xi_{j})^{l}_{j=1} in H1,1,χH_{1,1,\chi} such that uξj=ξlj+1u\xi_{j}=\xi_{l-j+1}. It follows that instead of making vAvtvAv^{t} block-diagonal, we can make it to be of the form (3.3), with 0<ai10<a_{i}\leq 1 for i[m+12]i\leq[\frac{m+1}{2}] and 0argaiπ0\leq\arg a_{i}\leq\pi for i>[m+12]i>[\frac{m+1}{2}] whenever |ai|=1|a_{i}|=1.  \diamond

We can now parameterize the isomorphism classes of braided free orthogonal quantum groups OAX,βO^{X,\beta}_{A}, at least when T^\hat{T} is torsion-free. We already know that it suffices to consider only the transmutations of free orthogonal quantum groups, that is, we may assume that AA¯=±1A\bar{A}=\pm 1 and w=1w=1 (the trivial character), and that then the only choice for χ\chi in (3.8) is χ=1\chi=1. Hence we get the following result.

Corollary 3.10.

Consider a compact connected abelian group TT and a bicharacter β:T^×T^𝕋\beta\colon\hat{T}\times\hat{T}\to{\mathbb{T}}. Then representatives [OAX,β]Hopf([T],β){\mathbb{C}}[O^{X,\beta}_{A}]\in\mathrm{Hopf}^{*}({\mathbb{C}}[T],\beta) of all isomorphism classes of braided free orthogonal quantum groups are obtained by considering the pairs (A,X)(A,X) such that

  • the matrix AA has the form (3.3), with 0<ai10<a_{i}\leq 1 for i[m+12]i\leq[\frac{m+1}{2}] and aiami+1=τa_{i}a_{m-i+1}=\tau for all ii, where τ=±1\tau=\pm 1 (hence am+12=1a_{\frac{m+1}{2}}=1 and τ=1\tau=1 if mm is odd);

  • the corepresentation matrix XX is diagonal, X=diag(χ1,,χm)X=\operatorname{diag}(\chi_{1},\dots,\chi_{m}), with χiχmi+1=1\chi_{i}\chi_{m-i+1}=1 for all ii (hence χm+12=1\chi_{\frac{m+1}{2}}=1 if mm is odd).

Two such pairs (A,X)(A,X) and (A,X)(A^{\prime},X^{\prime}) define isomorphic braided Hopf *-algebras over ([T],β)({\mathbb{C}}[T],\beta) if and only if (A,X)(A^{\prime},X^{\prime}) can be obtained from (A,X)(A,X) by

  • permuting the pairs (ai,ami+1)(a_{i},a_{m-i+1}) and (χi,χmi+1)(\chi_{i},\chi_{m-i+1}) (i[m2]i\leq[\frac{m}{2}]);

  • replacing χi\chi_{i} by χmi+1\chi_{m-i+1} and χmi+1\chi_{m-i+1} by χi\chi_{i} for some i[m2]i\leq[\frac{m}{2}] such that ai=1a_{i}=1.

Now, observe that the quantum group (3.7) is isomorphic to Uq(2)U_{q}(2) when χ\chi is nontrivial and T/kerχ𝕋T/\ker\chi\cong{\mathbb{T}}. All compact quantum groups monoidally equivalent to Uq(2)U_{q}(2) are known by the work of Mrozinski [Mro14]. Our next goal is to describe the subclass of these quantum groups that can be obtained as the bosonizations of [OAX,β]Hopf([T/T0],iβ){\mathbb{C}}[O^{X,\beta}_{A}]\in\mathrm{Hopf}^{*}({\mathbb{C}}[T/T_{0}],i^{*}\beta) for suitable T0T_{0}.

Assume that BGLm()B\in\mathrm{GL}_{m}({\mathbb{C}}) is such that BB¯B\bar{B} is unitary. Following the conventions of [HN22], consider the quantum group O~B+\tilde{O}_{B}^{+} of unitary transformations leaving the noncommutative polynomial P=i,jBjiXiXjP=\sum_{i,j}B_{ji}X_{i}X_{j} invariant up to a phase factor: [O~B+]{\mathbb{C}}[\tilde{O}_{B}^{+}] is the universal unital *-algebra generated by a unitary dd and elements wijw_{ij}, 1i,jm1\leq i,j\leq m, such that

W=(wij)i,j is unitary and W=BW¯B1d,W=(w_{ij})_{i,j}\mbox{ is unitary and }W=B\bar{W}B^{-1}d, (3.9)

and the coproduct is given by Δ(wij)=kwikwkj\Delta(w_{ij})=\sum_{k}w_{ik}\otimes w_{kj}, Δ(d)=dd\Delta(d)=d\otimes d. We remark that in [Mro14] the Hopf *-algebra [O~B+]{\mathbb{C}}[\tilde{O}_{B}^{+}] is denoted by Ao~(B¯t)A_{\tilde{o}}(\bar{B}^{t}).

Proposition 3.11.

Assume that BGLm()B\in\mathrm{GL}_{m}({\mathbb{C}}) (m2m\geq 2) is such that BB¯B\bar{B} is unitary. Then the following conditions are equivalent:

  • 1)

    [O~B+]{\mathbb{C}}[\tilde{O}_{B}^{+}] is the bosonization of [OBX,β]Hopf([𝕋/{±1}],iβ){\mathbb{C}}[O^{X,\beta}_{B}]\in\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}/\{\pm 1\}],i^{*}\beta) for some unitary corepresentation matrix XMatm([𝕋])X\in\operatorname{Mat}_{m}({\mathbb{C}}[{\mathbb{T}}]) and a bicharacter β\beta on =𝕋^{\mathbb{Z}}=\hat{\mathbb{T}} as in Proposition 3.2 with X(1)=1X(-1)=1 and w=z1w=z^{-1};

  • 2)

    [O~B+]{\mathbb{C}}[\tilde{O}_{B}^{+}] is the bosonization of a braided Hopf *-algebra over ([𝕋],β)({\mathbb{C}}[{\mathbb{T}}],\beta) for some bicharacter β\beta on =𝕋^{\mathbb{Z}}=\hat{\mathbb{T}};

  • 3)

    mm is even and the spectrum of BB¯B\bar{B} consists of odd powers of a single number λ𝕋\lambda\in{\mathbb{T}}.

Proof.

1) \Rightarrow 2) is obvious.

2) \Rightarrow 3): As O~B+\tilde{O}_{B}^{+} has fusion rules of U(2)U(2), the group-like elements of [O~B+]{\mathbb{C}}[\tilde{O}_{B}^{+}] are powers of dd. It follows that if we can identify [O~B+]{\mathbb{C}}[\tilde{O}_{B}^{+}] with [𝕋]#A{\mathbb{C}}[{\mathbb{T}}]\#A for some AHopf([𝕋],β)A\in\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}],\beta), then z#1=dz\#1=d or z#1=d1z\#1=d^{-1}. By replacing BB by B¯1\bar{B}^{-1} if necessary, we may assume that z#1=dz\#1=d, since there is an isomorphism O~B+O~B¯1+\tilde{O}_{B}^{+}\cong\tilde{O}_{\bar{B}^{-1}}^{+} mapping dd into d1d^{-1}, see [Mro14, Proposition 3.4].

Consider the Hopf *-algebra map ι#εA:[𝕋]#A[𝕋]\iota\#\varepsilon_{A}\colon{\mathbb{C}}[{\mathbb{T}}]\#A\to{\mathbb{C}}[{\mathbb{T}}]. Put X=(ι#εA)(W)X=(\iota\#\varepsilon_{A})(W). Then from the defining relations for O~B+\tilde{O}_{B}^{+} we get that XX is a unitary corepresentation matrix and X=BX¯B1zX=B\bar{X}B^{-1}z. By Remark 3.9 we may assume that

B=(0bmb10),bib¯mi+1=λi𝕋,B=\begin{pmatrix}0&&b_{m}\\ &\iddots&\\ b_{1}&&0\end{pmatrix},\quad b_{i}\bar{b}_{m-i+1}=\lambda_{i}\in{\mathbb{T}}, (3.10)

and X=diag(zk1,zk2,,zkm)X=\mathrm{diag}(z^{k_{1}},z^{k_{2}},...,z^{k_{m}}) for some kik_{i}\in{\mathbb{Z}}, 1im1\leq i\leq m. The identity X=BX¯B1zX=B\bar{X}B^{-1}z means then that ki+kmi+1=1k_{i}+k_{m-i+1}=1 for all ii. This implies that mm must be even.

Next, consider U=(ε[𝕋]#ι)(W)Matm(A)U=(\varepsilon_{{\mathbb{C}}[{\mathbb{T}}]}\#\iota)(W)\in\operatorname{Mat}_{m}(A). On the one hand, by (2.12), we have

δA(uij)=uijzkjki.\delta_{A}(u_{ij})=u_{ij}\otimes z^{k_{j}-k_{i}}.

As W=X#U=(zki#uij)i,jW=X\#U=(z^{k_{i}}\#u_{ij})_{i,j} and z#1=dz\#1=d, by the definition of the smash product [𝕋]#A{\mathbb{C}}[{\mathbb{T}}]\#A it follows that

dWd=(β(kikj,1)wij)i,j.dWd^{*}=(\beta(k_{i}-k_{j},1)w_{ij})_{i,j}.

On the other hand, the relation W=BW¯B1dW=B\bar{W}B^{-1}d implies W¯=dB¯WB¯1\bar{W}=d^{*}\bar{B}W\bar{B}^{-1}, hence

dWd=BB¯W(BB¯)1=(λiλ¯jwij)i,j.dWd^{*}=B\bar{B}W(B\bar{B})^{-1}=(\lambda_{i}\bar{\lambda}_{j}w_{ij})_{i,j}.

We therefore have that β(kikj,1)=λiλ¯j\beta(k_{i}-k_{j},1)=\lambda_{i}\bar{\lambda}_{j} for all i,ji,j, whence λi=λζki\lambda_{i}=\lambda\zeta^{k_{i}} for some λ𝕋\lambda\in{\mathbb{T}}, where ζ=β(1,1)\zeta=\beta(1,1). As λ¯i=λmi+1\bar{\lambda}_{i}=\lambda_{m-i+1} and ki+kmi+1=1k_{i}+k_{m-i+1}=1, we must have λ2=ζ\lambda^{-2}=\zeta and thus λi=λ2ki+1\lambda_{i}=\lambda^{-2k_{i}+1}.

3) \Rightarrow 1): We may assume that BB is of the form (3.10). By the assumption on the spectrum there exist lil_{i}\in{\mathbb{Z}} for i=1,,m/2i=1,\dots,m/2 such that λi=λ2li1\lambda_{i}=\lambda^{-2l_{i}-1}. Let li=1lmi+1l_{i}=-1-l_{m-i+1} for i=m/2+1,,mi=m/2+1,\dots,m. Then λi=λ2li1\lambda_{i}=\lambda^{-2l_{i}-1} for all ii.

Put X=diag(zl1,zl2,,zlm)X=\mathrm{diag}(z^{l_{1}},z^{l_{2}},...,z^{l_{m}}). Define a bicharacter β\beta on 𝕋^\hat{\mathbb{T}} by β(k,l)=λ2kl\beta(k,l)=\lambda^{-2kl}. We then have X=BX¯B1z1X=B\bar{X}B^{-1}z^{-1} and BB¯=λ1X(λ2)B\bar{B}=\lambda^{-1}X(\lambda^{-2}). Consider the double cover p:𝕋𝕋p\colon{\mathbb{T}}\to{\mathbb{T}}, tt2t\mapsto t^{2}. Denote by β1/4\beta^{1/4} the bicharacter on 𝕋^\hat{\mathbb{T}} defined in the same way as β\beta, but with λ2\lambda^{2} replaced by one of its fourth roots. Therefore if i:[𝕋][𝕋]i\colon{\mathbb{C}}[{\mathbb{T}}]\to{\mathbb{C}}[{\mathbb{T}}] is defined by i(z)=z2i(z)=z^{2}, we get iβ1/4=βi^{*}\beta^{1/4}=\beta. By Remark 3.5 the braided quantum group OBXp,β1/4O_{B}^{Xp,\beta^{1/4}} is well-defined. As (Xp)(1)=1(Xp)(-1)=1, we can view [OBXp,β1/4]{\mathbb{C}}[O_{B}^{Xp,\beta^{1/4}}] as an object AA in Hopf([𝕋/{±1}],iβ1/4)=Hopf([𝕋],β)\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}/\{\pm 1\}],i^{*}\beta^{1/4})=\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}],\beta).

The defining relations in AA say that BB is an intertwiner of the AA-comodules defined by (z1X¯,U¯X)(z^{-1}\bar{X},\bar{U}_{X}) and (X,U)(X,U). In other words, for W=X#U=(zli#uij)i,jW^{\prime}=X\#U=(z^{l_{i}}\#u_{ij})_{i,j} we have

B(z1#1)W¯=WB.B(z^{-1}\#1)\overline{W^{\prime}}=W^{\prime}B.

The relation W=BW¯B1dW=B\bar{W}B^{-1}d in [O~B+]{\mathbb{C}}[\tilde{O}^{+}_{B}] can be written as

dW=BddW¯B1.d^{*}W=Bd^{*}\overline{d^{*}W}B^{-1}.

It follows that we have a well-defined Hopf *-algebra map [O~B+][𝕋]#A{\mathbb{C}}[\tilde{O}^{+}_{B}]\to{\mathbb{C}}[{\mathbb{T}}]\#A such that dz#1d\mapsto z\#1 and W=(wij)i,jzX#U=(zli+1#uij)i,jW=(w_{ij})_{i,j}\mapsto zX\#U=(z^{l_{i}+1}\#u_{ij})_{i,j}. Using the relations in both algebras it is also easy to construct the inverse map. ∎

Remark 3.12.

The proof of the proposition implies that the procedure described in the proof of the implication 3) \Rightarrow 1) is the only way of getting decompositions [O~B+]=[𝕋]#A{\mathbb{C}}[\tilde{O}^{+}_{B}]={\mathbb{C}}[{\mathbb{T}}]\#A such that AHopf([𝕋],β)A\in\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}],\beta) and d=z#1d=z\#1.

Remark 3.13.

As in the proof of the proposition, it is not difficult to see that given BGLm()B\in\mathrm{GL}_{m}({\mathbb{C}}) such that BB¯B\bar{B} is unitary, a unitary corepresentation matrix XMatm([𝕋])X\in\mathrm{Mat}_{m}({\mathbb{C}}[{\mathbb{T}}]) such that X=BX¯B1zX=B\bar{X}B^{-1}z exists if and only if mm is even. Therefore, for even mm, we always get a Hopf *-algebra map p:[O~B+][𝕋]p:{\mathbb{C}}[\tilde{O}_{B}^{+}]\to{\mathbb{C}}[{\mathbb{T}}] such that p(d)=zp(d)=z with a right inverse zdz\mapsto d. By Radford’s theorem [Rad85] we then get a Hopf *-algebra object in the braided category 𝒴𝒟(𝕋)\mathcal{YD}({\mathbb{T}}) of 𝕋{\mathbb{T}}-Yetter–Drinfeld modules. From this perspective the above proposition characterizes when this object lies in the subcategory ([𝕋],β)𝒴𝒟(𝕋)(\mathcal{M}^{{\mathbb{C}}[{\mathbb{T}}]},\beta)\subset\mathcal{YD}({\mathbb{T}}) for some β\beta.  \diamond

Finally, let us compare the transmutations of [OF+]{\mathbb{C}}[O_{F}^{+}] to the braided quantum groups constructed in [MR22]. Fix numbers d1,d2,,dm,dd_{1},d_{2},...,d_{m},d\in{\mathbb{Z}} and consider the representation

X:𝕋Matm(),X(t)=(td100tdm).X\colon{\mathbb{T}}\to\operatorname{Mat}_{m}({\mathbb{C}}),\quad X(t)=\begin{pmatrix}t^{d_{1}}&&0\\ &\ddots&\\ 0&&t^{d_{m}}\end{pmatrix}.

Take Ω=(ωij)i,jGLm()\Omega=(\omega_{ij})_{i,j}\in\mathrm{GL}_{m}({\mathbb{C}}) and assume

ωij0di+dj=d,or equivalently,Ωt(zdX¯)(Ωt)1=X,\omega_{ij}\neq 0\,\Rightarrow\,d_{i}+d_{j}=d,\quad\text{or equivalently},\quad\Omega^{t}(z^{d}\bar{X})(\Omega^{t})^{-1}=X,

where z[𝕋]z\in{\mathbb{C}}[{\mathbb{T}}] is the generator. Assume further that there is ζ𝕋\zeta\in{\mathbb{T}} such that

Ω¯Ω=cX(ζd)for somec𝕋.\bar{\Omega}\Omega=c\,X(\zeta^{d})\quad\text{for some}\quad c\in{\mathbb{T}}.

Define a bicharacter on 𝕋^=\hat{{\mathbb{T}}}={\mathbb{Z}} by β(m,n)=ζmn\beta(m,n)=\zeta^{mn}.

In [MR22], Meyer and Roy construct an object Ao(Ω,X)Hopf([𝕋],β)A_{o}(\Omega,X)\in\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}],\beta) from this data. By [MR22, Theorem 2.6] and [BJR22, Remark 2.20(2)], in our terminology, Ao(Ω,X)A_{o}(\Omega,X) is defined as a universal braided compact matrix quantum group over (𝕋,X,β)({\mathbb{T}},X,\beta) with fundamental unitary U=(uij)i,jU=(u_{ij})_{i,j} subject to the relations

U=AU¯XA1,U=A\bar{U}_{X}A^{-1},

where A=(ωjiζddj)i,j=ΩtX(ζd)=ζd2X(ζd)ΩtA=(\omega_{ji}\zeta^{dd_{j}})_{i,j}=\Omega^{t}X(\zeta^{d})=\zeta^{d^{2}}X(\zeta^{-d})\Omega^{t}. Note that we have U¯X=(ζdi(djdi)uij)i,j\bar{U}_{X}=(\zeta^{d_{i}(d_{j}-d_{i})}u_{ij}^{*})_{i,j} and AA¯=cζd2X(ζd)A\bar{A}=c\,\zeta^{d^{2}}X(\zeta^{-d}).

Assume that d=2nd=2n. Put w(t)=tnw(t)=t^{n}. Then we see from the description above and Remark 3.5 that Ao(Ω,X)[OAX,β]A_{o}(\Omega,X)\cong{\mathbb{C}}[O_{A}^{X,\beta}] in Hopf([𝕋],β)\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}],\beta).

Next, assume that dd is odd. Then, as in the proof of Proposition 3.11, mm is even. Similarly to that proof, consider the double cover p:𝕋𝕋p\colon{\mathbb{T}}\to{\mathbb{T}}, tt2t\mapsto t^{2}, to write the character zdz^{d} as a square. Write β1/4\beta^{1/4} for the bicharacter on 𝕋^\hat{\mathbb{T}} defined in the same way as β\beta, but with ζ\zeta replaced by a fourth root ζ1/4\zeta^{1/4}. Define a character ww on the double cover by w(t)=tdw(t)=t^{d}. Then

Ao(Ω,X)[OAXp,β1/4],A_{o}(\Omega,X)\cong{\mathbb{C}}[O_{A}^{Xp,\beta^{1/4}}],

where we now view the latter braided Hopf *-algebra as an object in Hopf([𝕋/{±1}],iβ1/4)=Hopf([𝕋],β)\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}/\{\pm 1\}],i^{*}\beta^{1/4})=\mathrm{Hopf}^{*}({\mathbb{C}}[{\mathbb{T}}],\beta).

3.3. Braided free unitary quantum groups

Let m2m\geq 2 be a natural number. We recall the definition of the free unitary quantum group UF+U_{F}^{+}. Let F=(fij)i,jGLm()F=(f_{ij})_{i,j}\in\mathrm{GL}_{m}({\mathbb{C}}) be such that Tr(FF)=Tr((FF)1)\operatorname{Tr}(F^{*}F)=\operatorname{Tr}((F^{*}F)^{-1}). Then [UF+]{\mathbb{C}}[U_{F}^{+}] denotes the universal *-algebra with generators uiju_{ij}, 1i,jm1\leq i,j\leq m, and relations determined by

U=(uij)i,j and FU¯F1 are unitaries in Matm([UF+]).U=(u_{ij})_{i,j}\mbox{ and }F\bar{U}F^{-1}\mbox{ are unitaries in }\operatorname{Mat}_{m}({\mathbb{C}}[U_{F}^{+}]).

The Hopf *-algebra structure on [UF+]{\mathbb{C}}[U_{F}^{+}] is defined so that UF+U_{F}^{+} is a compact matrix quantum group as in Definition 2.12.

Similarly to the previous example, we fix a compact abelian group TT together with a unitary corepresentation matrix ZMatm([T])Z\in\operatorname{Mat}_{m}({\mathbb{C}}[T]) such that FZ¯F1F\bar{Z}F^{-1} is unitary, equivalently, Z¯\bar{Z} commutes with |F||F|. Then, by the universality of [UF+]{\mathbb{C}}[U_{F}^{+}], there is a Hopf *-algebra map π:[UF+][T]\pi\colon{\mathbb{C}}[U_{F}^{+}]\to{\mathbb{C}}[T] mapping UU to ZZ. Let β:T^×T^𝕋\beta\colon\hat{T}\times\hat{T}\to{\mathbb{T}} be a bicharacter, and consider the transmutation [UF+]β{\mathbb{C}}[U^{+}_{F}]_{\beta}. Then, similarly to Proposition 3.1, we get the following result.

Proposition 3.14.

The braided Hopf *-algebra [UF+]β{\mathbb{C}}[U_{F}^{+}]_{\beta} is a braided compact matrix quantum group over (T,Z,β)(T,Z,\beta) with fundamental unitary U=(uij)i,jU=(u_{ij})_{i,j}. As a *-algebra, it is a universal unital *-algebra with generators uiju_{ij} satisfying the relations

UandFU¯ZF1are unitaries,U\quad\mbox{and}\quad F\bar{U}_{Z}F^{-1}\quad\text{are unitaries},

where U¯Z=(u¯ijZ)i,j\bar{U}_{Z}=(\bar{u}_{ij}^{Z})_{i,j} and u¯ijZ=s,l,tβ(ztjzsl,zil)ust\bar{u}_{ij}^{Z}=\sum_{s,l,t}\beta(z_{tj}^{*}z_{sl},z_{il}^{*})u_{st}^{*}.

By Theorem 2.5, the bosonization of [UF+]β{\mathbb{C}}[U_{F}^{+}]_{\beta} is a compact quantum group that is a cocycle twist of T×UF+T\times U^{+}_{F}.

When T=𝕋T={\mathbb{T}} and ZZ is such that both ZZ and FZ¯F1F\bar{Z}F^{-1} are diagonal matrices, we recover the braided free unitary quantum groups defined in [BJR22]. We remark that we can of course always assume that one of the matrices ZZ or FZ¯F1F\bar{Z}F^{-1} is diagonal by choosing an appropriate orthonormal basis, but it is usually impossible to make both of them diagonal simultaneously.

3.4. Anyonic quantum permutation groups

Let N2N\geq 2 be a natural number. The quantum symmetric group SN+S_{N}^{+} is the universal compact matrix quantum group with fundamental unitary representation U=(uij)i,j=0N1U=(u_{ij})_{i,j=0}^{N-1} subject to the relations

iuij=1=juij and uij=uij2=uij.\sum_{i}u_{ij}=1=\sum_{j}u_{ij}\quad\mbox{ and }\quad u_{ij}^{*}=u_{ij}^{2}=u_{ij}.

We can view the cyclic group /N{\mathbb{Z}}/N{\mathbb{Z}} as a subgroup of SNSN+S_{N}\subset S_{N}^{+}, so we get a Hopf *-algebra map

π:[SN+][/N],π(uij)=δji,\pi\colon{\mathbb{C}}[S_{N}^{+}]\to{\mathbb{C}}[{\mathbb{Z}}/N{\mathbb{Z}}],\quad\pi(u_{ij})=\delta_{j-i},

where δk[/N]\delta_{k}\in{\mathbb{C}}[{\mathbb{Z}}/N{\mathbb{Z}}] are the usual delta-functions.

To describe the transmutation we want to express the relations in [SN+]{\mathbb{C}}[S_{N}^{+}] in terms of homogeneous elements with respect to the bi-grading by /N{\mathbb{Z}}/N{\mathbb{Z}}. Fix a primitive NN-th root of unity ω\omega. By Lemma 4.9 in [Ans+22] such generators can be obtained by considering the elements aija_{ij} defined by

(aij)i,j=Ω1UΩ,Ω=(1Nωij)i,j=0N1.(a_{ij})_{i,j}=\Omega^{-1}U\Omega,\quad\Omega=\Big{(}\dfrac{1}{N}\omega^{-ij}\Big{)}_{i,j=0}^{N-1}.

The elements aija_{ij} are then bi-graded by

(πι)Δ(aij)=ziaij,(ιπ)Δ(uij)=aijzj,(\pi\otimes\iota)\Delta(a_{ij})=z^{i}\otimes a_{ij},\quad(\iota\otimes\pi)\Delta(u_{ij})=a_{ij}\otimes z^{j},

where z[/N]z\in{\mathbb{C}}[{\mathbb{Z}}/N{\mathbb{Z}}] is the function z(k)=ωkz(k)=\omega^{k}. In terms of the new generators the relations in [SN+]{\mathbb{C}}[S_{N}^{+}] become

a0i=ai0=δi,0,aij=ai,j,a_{0i}=a_{i0}=\delta_{i,0},\quad a_{ij}^{*}=a_{-i,-j},
ak,i+j=lakl,ialj,ai+j,k=lajlai,kl.a_{k,i+j}=\sum_{l}a_{k-l,i}a_{lj},\quad a_{i+j,k}=\sum_{l}a_{jl}a_{i,k-l}.

Define a bicharacter by β(zi,zj)=ωij\beta(z^{i},z^{j})=\omega^{-ij}. It is then readily verified, by using formulas (2.4) and (2.6), that the transmutation [SN+]β{\mathbb{C}}[S_{N}^{+}]_{\beta} is described by the relations

a0i=ai0=δi,0,aij=ωi(ji)ai,j,a_{0i}=a_{i0}=\delta_{i,0},\quad a_{ij}^{*}=\omega^{i(j-i)}a_{-i,-j},
ak,i+j=lωl(ik+l)akl,ialj,ai+j,k=lωi(lj)ajlai,kl.a_{k,i+j}=\sum_{l}\omega^{-l(i-k+l)}a_{k-l,i}a_{lj},\quad a_{i+j,k}=\sum_{l}\omega^{-i(l-j)}a_{jl}a_{i,k-l}.

These are exactly the relations in [Ans+22, Definition 2.7]. Finally, by Theorem 2.5, the bosonization of [SN+]β{\mathbb{C}}[S_{N}^{+}]_{\beta} is a cocycle twist of the quantum group (/N)×SN+({\mathbb{Z}}/N{\mathbb{Z}})\times S_{N}^{+}.

References

  • [Ans+22] Anshu, Suvrajit Bhattacharjee, Atibur Rahaman and Sutanu Roy “Anyonic quantum symmetries of finite spaces” arXiv, 2022 DOI: 10.48550/ARXIV.2207.08153
  • [BJR22] Suvrajit Bhattacharjee, Soumalya Joardar and Sutanu Roy “Braided quantum symmetries of graph C\mathrm{C}^{*}-algebras” arXiv, 2022 DOI: 10.48550/ARXIV.2201.09885
  • [BS19] Arkadiusz Bochniak and Andrzej Sitarz “Braided Hopf algebras from twisting” In J. Algebra Appl. 18.9, 2019, pp. 1950178\bibrangessep18 DOI: 10.1142/S0219498819501780
  • [DK94] Mathijs S. Dijkhuizen and Tom H. Koornwinder “CQG algebras: a direct algebraic approach to compact quantum groups” In Lett. Math. Phys. 32.4, 1994, pp. 315–330 DOI: 10.1007/BF00761142
  • [HN21] Erik Habbestad and Sergey Neshveyev “Subproduct systems with quantum group symmetry” Preprint, 2021 DOI: 10.48550/ARXIV.2111.10911
  • [HN22] Erik Habbestad and Sergey Neshveyev “Subproduct systems with quantum group symmetry. II” Preprint, 2022 URL: https://arxiv.org/abs/2212.08512
  • [Kas+16] Paweł Kasprzak, Ralf Meyer, Sutanu Roy and Stanisław Lech Woronowicz “Braided quantum SU(2)\rm SU(2) groups” In J. Noncommut. Geom. 10.4, 2016, pp. 1611–1625 DOI: 10.4171/JNCG/268
  • [Maj93] Shahn Majid “Braided groups” In J. Pure Appl. Algebra 86.2, 1993, pp. 187–221 DOI: 10.1016/0022-4049(93)90103-Z
  • [Maj95] Shahn Majid “Foundations of quantum group theory” Cambridge University Press, Cambridge, 1995, pp. x+607 DOI: 10.1017/CBO9780511613104
  • [MR22] Ralf Meyer and Sutanu Roy “Braided free orthogonal quantum groups” In Int. Math. Res. Not. IMRN, 2022, pp. 8890–8915 DOI: 10.1093/imrn/rnaa379
  • [Mro14] Colin Mrozinski “Quantum groups of GL(2)\rm GL(2) representation type” In J. Noncommut. Geom. 8.1, 2014, pp. 107–140 DOI: 10.4171/JNCG/150
  • [NT13] Sergey Neshveyev and Lars Tuset “Compact quantum groups and their representation categories” 20, Cours Spécialisés [Specialized Courses] Société Mathématique de France, Paris, 2013
  • [Rad85] David E. Radford “The structure of Hopf algebras with a projection” In J. Algebra 92.2, 1985, pp. 322–347 DOI: 10.1016/0021-8693(85)90124-3
  • [Wor87] S.. Woronowicz “Compact matrix pseudogroups” In Comm. Math. Phys. 111.4, 1987, pp. 613–665 URL: http://projecteuclid.org/euclid.cmp/1104159726
  • [Wor91] S.. Woronowicz “A remark on compact matrix quantum groups” In Lett. Math. Phys. 21.1, 1991, pp. 35–39 DOI: 10.1007/BF00414633
  • [ZZ05] Xiaoxia Zhang and Ervin Yunwei Zhao “The compact quantum group Uq(2)U_{q}(2). I” In Linear Algebra Appl. 408, 2005, pp. 244–258 DOI: 10.1016/j.laa.2005.06.004