Enveloping algebra is a Yetter–Drinfeld module algebra over Hopf algebra of regular functions on the automorphism group of a Lie algebra

Zoran Škoda, Martina Stojić
Abstract

We present an elementary construction of a (highly degenerate) Hopf pairing between the universal enveloping algebra U(𝔤) of a finite-dimensional Lie algebra 𝔤 over arbitrary field 𝒌 and the Hopf algebra 𝒪(Aut(𝔤)) of regular functions on the automorphism group of 𝔤. This pairing induces a Hopf action of 𝒪(Aut(𝔤)) on U(𝔤) which together with an explicitly given coaction makes U(𝔤) into a braided commutative Yetter–Drinfeld 𝒪(Aut(𝔤))-module algebra. From these data one constructs a Hopf algebroid structure on the smash product algebra 𝒪(Aut(𝔤))U(𝔤) retaining essential features from earlier constructions of a Hopf algebroid structure on infinite-dimensional versions of Heisenberg double of U(𝔤), including a noncommutative phase space of Lie algebra type, while avoiding the need of completed tensor products.

We prove a slightly more general result where algebra 𝒪(Aut(𝔤)) is replaced by 𝒪(Aut(𝔥)) and where 𝔥 is any finite-dimensional Leibniz algebra having 𝔤 as its maximal Lie algebra quotient.

Keywords: Yetter–Drinfeld module algebra, universal enveloping algebra, Lie algebra automorphism, Leibniz algebra, adjoint map

MSC 2020: 16T05, 16S40

1 Introduction

Yetter–Drinfeld modules over bialgebras (2.2 and [9, 10, 14]) are ubiquitous in quantum algebra, low dimensional topology and representation theory. More intricate structure of a braided commutative monoid in the category 𝒴H𝒟H of Yetter–Drinfeld modules is an ingredient in a construction of scalar extension bialgebroids and Hopf algebroids. Namely, given a Hopf algebra H and a braided commutative Yetter–Drinfeld H-module algebra A, the smash product algebra HA has a structure of a Hopf algebroid over A [1, 2, 11]. An important special case is the Heisenberg double AA of a finite-dimensional Hopf algebra A, where A is the dual Hopf algebra of A; the A-coaction on A is given by an explicit formula involving basis of A and the dual basis of A ([5], map β in Section 6). There are several important examples in literature, some motivated by mathematical physics, when the underlying algebra of A is the universal enveloping algebra U(𝔤) of an n-dimensional Lie algebra 𝔤. In the example of a noncommutative phase space 𝔤 of Lie algebra type [8], the Hopf algebra H is a realization via formal power series of the algebraic dual U(𝔤) with its natural topological Hopf algebra structure. Historically, 𝔤 as an algebra has been introduced by extending U(𝔤) by deformed derivatives (“momenta”) in [7] and in a rather nonrigorous treatment [13] it has been argued that 𝔤 is actually the Heisenberg double of U(𝔤); this made plausible that Hopf algebroid structure on 𝔤 could be exhibited analogously to Lu’s example of finite-dimensional double, leading to an ad hoc version of completed Hopf algebroid structure in [8]. An abstract version of U(𝔤)U(𝔤) is described in [12] as an internal Hopf algebroid in the symmetric monoidal category of filtered cofiltered vector spaces. These examples may be viewed as infinite-dimensional cases of Heisenberg double (in fact, this observation from [13] influenced the Hopf algebroid approach in [8]), but a number of results in [12] show that there are intricate conditions for which infinite-dimensional dually paired Hopf algebras H and A one can indeed form a Hopf algebroid structure on the smash product algebra HA, even in a completed sense or even when one of the Hopf algebras is the restricted dual of another. In the case of U(𝔤), formulas from [8] show that there are special elements 𝒪jiH=U(𝔤) for i,j{1,,n} such that the coaction on the generators is given by the formulas (adapted to our conventions) U(𝔤)𝔤xki𝒪kixiU(𝔤)𝔤. Moreover, elements 𝒪ji satisfy the relations which are satisfied by matrix elements of automorphisms of 𝔤. The prime motivation of this article is to demistify this phenomenon and to find a much smaller Hopf algebra H containing an abstract model for 𝒪ji and avoiding any completions in describing Yetter–Drinfeld H-module structure on U(𝔤).

We show that the Hopf algebra of regular functions H=𝒪(Aut(𝔤)) on the automorphism group of 𝔤 will do, namely that U(𝔤) is a braided commutative Yetter–Drinfeld module 𝒪(Aut(𝔤))-algebra whose structure is given by essentially the same formulas as in the case of U(𝔤) from [12]. General formulas for scalar extensions from [11] describe the Hopf algebroid structure on the smash product 𝒪(Aut(𝔤))U(𝔤). In the sequel work [15] we present several other natural examples of Hopf algebras H equipped with a Hopf algebra homomorphism 𝒪(Aut(𝔤))H and where U(𝔤) is still a Yetter–Drinfeld module algebra over H without completions. In [15] formulas for the symmetric Hopf algebroid structure are spelled out in full detail.

The main result of this article is actually proved in a slightly more general form than described above. Namely, instead of the automorphism group of a Lie algebra 𝔤 we can take the automorphism group of any finite-dimensional Leibniz algebra 𝔥 such that 𝔤 is the maximal Lie algebra quotient 𝔥Lie of 𝔥, enabling a structure of a Hopf algebroid on the smash product 𝒪(Aut(𝔥))U(𝔥Lie).

2 Preliminaries

2.1 General conventions and preliminaries on pairings

Throughout the paper we freely use Sweedler notation with or without the summation sign [6, 9] and the Kronecker symbol δji. Throughout, 𝒌 is a fixed ground field and Vec𝒌 the category of 𝒌-vector spaces. If VVec𝒌, V:=Hom𝒌(V,𝒌). If Δ is a comultiplication on a 𝒌-coalgebra C then for m1, Δm:=(idCm2Δ)(idCΔ)Δ:CCm+1. If (A,μ,η) is an associative 𝒌-algebra with multiplication μ:AAA and unit map η, consider its transpose μ:A(AA). Restricted dual AA consists of all fA such that μ(f) falls within the image of inclusion AA(AA). It follows that μ(f) belongs also to the image of AA hence the restriction μ(f)|A may be corestricted to a map ΔA:AAA making A into a coalgebra; if A is a bialgebra (resp. Hopf algebra) then A is. Pairings of vector spaces are bilinear maps into the ground field which are in this work not required to be nondegenerate. For V,WVec𝒌, a pairing ,:VW𝒌 induces a pairing between VV and WW componentwise: vv,ww:=v,w𝒌v,w. If VVec𝒌 and A is an algebra, and fA such that there exists an element fAA such that f,gh=f,gh for all g,hA, then the functional f,A and f,AA. If V=C is a coalgebra then Δ(f),AA. Thus, if Δ(c),hg=c,gh for all g,hA, then ϕ1:cc,:=ϕ1(c) is a map CA, and (ϕ1ϕ1)ΔC=ΔAϕ1. It is a coalgebra map if moreover c,1=ϵ(c). Conversely, if cc, corestricts to a coalgebra map ϕ1:CA, then the identities Δ(c),hg=c,gh and c,1=ϵ(c) hold for all cC, g,hA. Both conditions hold if and only if a,a is a map of algebras AC. A pairing between two bialgebras B and H is Hopf if ΔB(b),hk=b,hk, ϵB(b)=b,1H and the symmetric conditions bc,ΔH(h)=bc,h, ϵH(h)=1B,h. Clearly, the latter two conditions hold if and only if h,h corestricts to a coalgebra map ϕ2:HB. Alternatively, the pairing is Hopf if and only if bb, corestricts to a bialgebra map ϕ1:BH. If B and H are Hopf algebras, ϕ1 is a bialgebra map between Hopf algebras, hence it automatically respects the antipode. Thus, for every Hopf pairing between Hopf algebras, identity SB(b),h=b,SH(h) holds for all bB, hH.

Replacing Hopf pairing with a bialgebra map ϕ1:BH is useful in constructing new pairings from old. Namely, if IB is a biideal (ideal which is also a coideal in the sense that Δ(I)IB+BI and ϵ(I)=0), then there is an induced bialgebra map ϕ1:B/IH if and only if ϕ1(I)=0 that is i,h=0 for all iI,hH.

Lemma 2.1.

Let ,:BH𝐤 be a bialgebra pairing.

  1. (i)

    For the bialgebra pairing to vanish on IH where IB is a coideal, it suffices that it vanishes on IKH where KH is some set of algebra generators of H.

  2. (ii)

    If IB is a biideal, KII a set of generators of I as an ideal and KH a set of generators of H as an algebra, then the bialgebra pairing vanishes on IH if, in addition to i,h=0 for all iK and hKH, one has that Δ2(h)HSpan𝒌(KH{1})H for all hKH.

Proof.

(i) Fix any iI. Then Δ(i)=αiαbα+βbβiβ for some iαI,bβB. Then i,hh=αiα,hbα,h+βbβ,hiβ,h=0.

(ii) To show bib,h=0 one needs b,h(1)i,h(2)b,h(3)=0. The condition on Δ2(h) is sufficient for this to hold for all iI, b,bB, hKH.

One often starts by constructing an auxiliary pairing where one of the bialgebras is free [9]. If C is a coalgebra, then by the universal property of the tensor algebra T(C) there is a unique algebra map T(C)T(C)T(C) extending the composition CΔCCT(C)T(C) along inclusion CT(C); this is a comultiplication on T(C) making it into a bialgebra. Every coalgebra map CB to a bialgebra admits a unique extension to a bialgebra map T(C)B.

We need a different variant of this standard universal property. Suppose C=V𝒌u as a vector space where u is grouplike, that is, Δ(u)=uu, ϵ(u)=1. Ideal Iu in T(C) generated by u1 is a biideal because Δ(u1)=u(u1)+(u1)1 and ϵ(u1)=0. Composition T(V)T(V𝒌u)T(V𝒌u)/Iu is an isomorphism of algebras (the inverse can easily be described); we transfer the comultiplication from T(V𝒌u)/Iu to T(V) along this isomorphism. By the above universal property, any coalgebra map f:V𝒌uB, where B is a bialgebra, extends uniquely to a bialgebra map T(V𝒌u)B; if f(u)=1 then it induces a bialgebra map T(V𝒌u)/IuB.

Lemma 2.2.

Suppose VVec𝐤 and C=(V𝐤,Δ,ϵ) is a coalgebra such that 01 is grouplike. Then T(V) has a canonical bialgebra structure such that the inclusion V𝐤T(V) is a coalgebra map and for any bialgebra B, each coalgebra map CB mapping 01 to 1B admits a unique extension to a bialgebra map T(V)B.

2.2 Yetter–Drinfeld module algebras

In this subsection, fix a Hopf 𝒌-algebra H=(H,Δ,ϵ) with comultiplication Δ:hh(1)h(2) and counit ϵ:H𝒌. Recall that the category H of right H-modules is monoidal: if (M,M) and (N,N) are H-modules then their tensor product is 𝒌-module M𝒌N with H-action :(mn)h(mMh(1))(nNh(2)) and the unit object is 𝒌 with action ch=ϵ(h)c, for mM, nN, hH, c𝒌. A right H-module algebra A is a monoid in H: a right H-module (A,:AHA) with multiplication such that (ah(1))(bh(2))=(ab)h and 1h=ϵ(h)1. One can then form a smash product algebra HA with underlying 𝒌-vector space HA and associative multiplication given by (ha)(kb):=hk(1)(ak(2))b where ha is an alias for haHA. We often identify aA with a1 and hH with 1hAH (thus for a,bA, h,kH, a(hb) denotes (1a)(hb) and h(kb)=(hk)b). We extend to a right action, also denoted , of HA on A by setting a(hb):=(ah)bA.

A right-left Yetter–Drinfeld H-module (M,,λ) is a unital right H-module (M,) with a left H-coaction λ:MHM, mλ(m)=m[1]m[0], satisfying Yetter–Drinfeld compatibility condition

f(2)(mf(1))[1](mf(1))[0]=m[1]f(1)(m[0]f(2)), for all mM,fH. (1)

Morphisms of Yetter–Drinfeld modules are morphisms of underlying modules which are also morphisms of comodules and the tensor product of underlying H-modules is a Yetter–Drinfeld module via coaction mnn[1]m[1]m[0]n[0] (notice the order!). Thus we obtain a braided monoidal category 𝒴H𝒟H of (right-left) Yetter–Drinfeld H-modules with braiding σM,N:MNNM given by mn(nNm[1])m[0]. For finite-dimensional H, 𝒴H𝒟H is braided monoidally equivalent to the Drinfeld–Majid center of the monoidal category H of right H-modules.

If A is a right H-module algebra with a left H-coaction λ and if we identify the underlying vector spaces of HA and HA, then the Yetter–Drinfeld compatibility may be rewritten in terms of the multiplication in HA, as

f(2)λ(af(1))=λ(a)f, for all aA,fH. (2)

Monoids in 𝒴H𝒟H are called (right-left) Yetter–Drinfeld H-module algebras. They are Yetter–Drinfeld modules with multiplication such that they become H-module algebras and Hop-comodule algebras. Notice that an H-comodule is the same thing as an Hop-comodule, but saying that it is a comodule algebra is different. If (A,,λ) is a Yetter–Drinfeld module and μ:AAA a 𝒌-linear map, then μ is braided commutative if σA,Aμ=μ, that is, (ab[1])b[0]=ba for all a,bA. An H-module algebra (A,) is braided commutative if its multiplication is braided commutative.

Lemma 2.3.

Consider an H-module algebra (A,) with multiplication μ and a coaction λ so that (A,,λ) is a Yetter–Drinfeld H-module.

  1. (i)

    Multiplication μ is braided commutative in 𝒴H𝒟H if and only if for the extended action of the smash product HA relation aλ(b)=ba holds.

  2. (ii)

    Multiplication μ is braided commutative if and only if all elements of the form 1a, aA, commute with all elements of the form λ(b), bA, viewed inside algebra HA.

  3. (iii)

    Suppose μ is braided commutative. Then A is an Hop-comodule algebra (hence also a Yetter–Drinfeld module algebra) if and only if λ considered as a map with values in smash product algebra HA is antimultiplicative.

Proof.

For (i) indeed, the left hand side is a(b[1]b[0])=(ab[1])b[0]. Parts (ii) and (iii) are left to the reader. They are implicit in [2].

2.3 Leibniz algebras

Left and right Leibniz algebras are nonassociative algebras slightly generalizing Lie algebras by dropping the condition of antisymmetry.

A 𝒌-vector space 𝔥 equipped with a linear map [,]:𝔥𝒌𝔥𝔥 is a left Leibniz algebra [4] if for every x𝔥 the map adx:y[x,y] is a derivation on 𝔥, that is, if left Leibniz identity [x,[y,z]]=[[x,y],z]+[y,[x,z]] holds for all x,y,z𝔥. Let 𝔥l be a copy of vector space 𝔥, with elements denoted lx, x𝔥, with operations transported via xlx. Denote by 𝔥Lie the Lie algebra obtained as a quotient of 𝔥 by two-sided ideal I[x,x],x𝔥 generated by all commutators [x,x], x𝔥. It is a Lie algebra and it is maximal in the sense that if char𝒌2, every map 𝔥𝔤 to a Lie algebra 𝔤 factors through 𝔥Lie (if char𝒌=2, relation [x,x]=0 is stronger than the antisymmetry).

Lemma 2.4.

Let 𝔥 be a left Leibniz algebra. Universal enveloping algebra U(𝔥Lie) of Lie algebra 𝔥Lie𝔥/I[x,x],x𝔥 is isomorphic to T(𝔥l)/Il, where Il is the ideal in T(𝔥l) generated by l[x,y]lxly+lylx, x,y𝔥.

Proof.

Denote by x¯ the image of x𝔥 in 𝔥Lie. Denote by l¯x the image of lxT(𝔥l) in T(𝔥l)/Il. Denote by x~ the image of x𝔥 in U(𝔥Lie). The dashed map in the following diagram satisfies that [x,x]0 for every x𝔥 since by the composition of maps 𝔥T(𝔥l) in the top row [x,x]l[x,x]=l[x,x]lxlx+lxlxIlT(hl).

𝔥𝔥lT(𝔥l)T(𝔥l)/Ilxlx

Therefore, there exists the map 𝔥LieT(𝔥l)/Il, x¯l¯x, and consequently the map T(𝔥Lie)T(𝔥l)/Il. This map satisfies [x,y]¯x¯y¯+y¯x¯l¯[x,y]l¯xl¯y+l¯yl¯x=0T(𝔥l)/Il. Therefore, we have induced the map U(𝔥Lie)T(𝔥l)/Il.

𝔥𝔥lT(𝔥l)𝔥LieT(𝔥l)/IlT(𝔥Lie)U(𝔥Lie)

On the other hand, in the following diagram

𝔥𝔥lT(𝔥l)𝔥LieT(𝔥l)/IlT(𝔥Lie)U(𝔥Lie)

the map 𝔥U(𝔥Lie) induces T(𝔥l)U(𝔥Lie). By this map l[x,y]lxly+lylx[x,y]~x~y~+y~x~=[x~,y~]x~y~+y~x~=0U(𝔥Lie) for all x,y𝔥. Therefore, we obtain a well defined map T(𝔥l)/IlU(𝔥Lie).

These two maps are inverse to each other.

We say that 𝒌-vector space 𝔥 together with a linear map [,]:𝔥𝒌𝔥𝔥 is a right Leibniz algebra if for every x𝔥 the map y[y,x] is a derivation on 𝔥, that is, if the right Leibniz identity [[x,y],z]=[[x,z],y]+[x,[y,z]] holds for all x,y,z𝔥. By quotienting 𝔥 by the ideal generated by [x,x],x𝔥, we get a maximal quotient Lie algebra, 𝔥𝔥Lie.

Lemma 2.5.

Let 𝔥 be a right Leibniz algebra. Universal enveloping algebra U(𝔥Lie) of Lie algebra 𝔥Lie𝔥/I[x,x],x𝔥 is isomorphic to T(𝔥r)/Ir, where Ir is the ideal in T(𝔥r) generated by r[x,y]rxry+ryrx, x,y𝔥.

Proof.

Analogous to proof of Lemma 2.4.

3 U(𝔥Lie) as a Yetter–Drinfeld 𝒪(Aut(𝔥))-module algebra

In this section, we prove the central result of this article: for any finite-dimensional Leibniz algebra 𝔥 over any field 𝒌, the universal enveloping algebra U(𝔥Lie) of its maximal quotient Lie algebra 𝔥Lie is a braided commutative Yetter–Drinfeld module algebra over the Hopf algebra 𝒪(Aut(𝔥)) of regular functions on the algebraic group of automorphisms of 𝔥. This result immediately implies that the smash product algebra 𝒪(Aut(𝔥))U(𝔥Lie) has a structure of a Hopf algebroid over base algebras U(𝔥Lie)op,U(𝔥Lie) [11]. Its structure is, for the case when 𝔥 is a Lie algebra, spelled out in detail in [15]. Moreover, this structure is related to other, more geometric, examples in [15].

3.1 Hopf algebra 𝒪(Aut(L))

Let (L,L) be any nonassociative algebra of finite dimension n over a field 𝒌. The general linear group of the underlying vector space, GL(L) is an affine algebraic group with algebra of regular functions 𝒪(GL(L)) that is therefore a Hopf algebra via Δ(f)(M,N)=f(MN) and ϵ(f)=f(1) for any M,NGL(L) [3]. For a chosen ordered basis 𝐛=(x1,,xn) of L, interpreting matrices as operators amounts to an isomorphism ι𝐛:GL(n,𝒌)GL(L). Structure constants Cijk=C𝐛ijk are defined by

xixj=k=1nC𝐛ijkxk,i,j{1,,n}, (3)

and we introduce as algebra generators of 𝒪(GL(n,𝒌)) regular functions Uji:MMji,U¯ji:M(M1)ji, where Mji is the (i,j)-th entry of matrix MGL(n,𝒌). As an abstract algebra, 𝒪(GL(n,𝒌)) is the free algebra on n2 generators Uji,U¯ji modulo the n2 relations kUkiU¯jk=δji. The comultiplication is then given by Δ(Uji)=kUkiUjk and Δ(U¯ji)=kU¯jkU¯ki with counit ϵ(Uji)=ϵ(U¯ji)=δji. By definition, an element ψGL(L) is an automorphism if ψ(a)Lψ(b)=ψ(aLb) for all a,bL. These relations cut out the subgroup Aut(L)GL(L). To see that it is a Zariski closed subgroup, write a=kakxk, b=kbkxk and observe that this condition amounts to a system of n3 polynomial equations in GL(n,𝒌),

rCijrψrk=l,mψilψjmClmk.

In other words, ι𝐛 induces an identification ι𝐛:𝒪(Aut(L))𝒪(Aut(L))𝐛 with the quotient 𝒪(Aut(L))𝐛 of 𝒪(GL(n,𝒌)) by the ideal IAut(L)𝐛 generated by relations

l,mClmkUilUjm=rUrkCijr. (4)

Regarding that the inclusion of subvarieties Aut(L)GL(L) is also an inclusion of groups, this ideal is Hopf and 𝒪(Aut(L)) is the quotient Hopf algebra of functions on the subgroup. One can also directly check that the ideal IAut(L)𝐛 is a Hopf ideal.

Denote by 𝒢ji=𝒢𝐛ji=Uji+IAut(L)𝐛 and 𝒢¯ji=𝒢¯𝐛ji=U¯ji+IAut(L)𝐛 the generators of 𝒪(Aut(L))𝐛. If T=(Tji)i,j=1n is a transition matrix to a basis 𝐛=(x1,,xn), xj=iTjixi, then ι𝐛1ι𝐛:GL(n,𝒌)GL(n,𝒌), ATAT1. Then 𝒢𝐛jil,mTli𝒢𝐛mlT1jm extends to a Hopf algebra isomorphism θ𝐛𝐛:𝒪(Aut(L))𝐛𝒪(Aut(L))𝐛 and θ𝐛𝐛=θ𝐛𝐛1. This implies ι𝐛1(𝒢𝐛ji)=l,mTliι𝐛1(𝒢𝐛ml)T1jm within 𝒪(Aut(𝔥)). When it is clear which basis 𝐛 is fixed, ι𝐛1(𝒢𝐛ji)𝒪(Aut(L)) will also be denoted by 𝒢𝐛ji or simply 𝒢ji. Assuming the identification ι𝐛1, we write

𝒢𝐛ji=l,mTli𝒢𝐛mlT1jm. (5)

Assuming the identification ι𝐛1, if ψ is an automorphism of L, i𝒢𝐛ji(ψ)xi=ψ(xj). The standard reasoning above is summarized in the following proposition.

Proposition 3.1.

Let L be a nonassociative algebra of finite dimension n with a 𝐤-basis 𝐛 and structure constants Cijk=C𝐛ijk (3). Hopf algebra 𝒪(Aut(L)) of regular functions on the affine algebraic group of automorphisms of L is as an algebra isomorphic to a commutative algebra 𝒪(Aut(L))𝐛 with 2n2-generators 𝒢ji, 𝒢¯ji, i,j{1,,n} and defining relations

l,mClmk𝒢il𝒢jm=r𝒢rkCijr,k𝒢ki𝒢¯jk=δji=k𝒢¯ki𝒢jk,i,j,k{1,,n}. (6)

As a direct consequence the following identities hold for all i,j,k{1,,n}:

m,p𝒢piCmjp𝒢¯km=mCkmi𝒢jm, (7)
l,mClmk𝒢¯il𝒢¯jm=r𝒢¯rkCijr. (8)

Isomorphism 𝒪(Aut(L))𝐛𝒪(Aut(L)) is a Hopf algebra isomorphism if 𝒪(Aut(L))𝐛 is given the unique comultiplication Δ and counit ϵ which are algebra maps satisfying

Δ(𝒢ji)=k𝒢ki𝒢jk,Δ(𝒢¯ji)=k𝒢¯jk𝒢¯ki,ϵ(𝒢ji)=ϵ(𝒢¯ji)=δji, (9)

and the antipode S satisfying S(𝒢¯ji)=𝒢ji, S(𝒢ji)=𝒢¯ji for all i,j{1,,n}.

3.2 Hopf pairing

If (B,Δ,ϵ) is a 𝒌-bialgebra, then a differentiation of B is any 𝒌-linear map D:B𝒌 such that Leibniz rule D(bc)=D(b)ϵ(c)+ϵ(b)D(c) holds. In other words, it is a 𝒌ϵ-valued derivation of B, where 𝒌ϵ is 𝒌 with the trivial B-(bi)module structure coming from the counit. In Hopf algebraic language, a differentiation is a primitive element in the restricted dual bialgebra B. The following lemma is standard and elementary.

Lemma 3.2.

Let B be any bialgebra such that its underlying algebra is the free unital commutative algebra with a set of free generators FB. There is a canonical isomorphism between the vector space 𝐤FB of set maps FB𝐤 and the space of differentiations of B which extend these maps.

Assume VVec𝒌 and C=V𝒌 is a coalgebra such that Δ(1)=11 and Δ(v)=1v+v1 for all vV. Suppose V is paired with B such that map vv, corestricts to a coalgebra map ϕ1:CB for which ϕ1(1)=1B=ϵB. Then ϕ1(v) is a differentiation of B. Conversely, by Lemma 3.2 each such ϕ1(v) is determined by ϕ1(v)|FB, where the values for the latter can be chosen independently.

Proposition 3.3.

Let 𝔥 be a left Leibniz 𝐤-algebra with a vector space basis 𝐛=(x1,,xn) and structure constants Cjki determined from [xj,xk]=iCjkixi, j,k{1,,n}. In the notation of Subsection 3.1, 𝒢ji,𝒢¯ji, i,j{1,,n} are the generators of the algebra 𝒪(Aut(𝔥)). Denote also by x~ the image of x𝔥 in 𝔥Lie.

Then there is a well defined and unique Hopf pairing

,:U(𝔥Lie)𝒪(Aut(𝔥))𝒌

such that x~k,𝒢ji=Ckji for all i,j,k{1,,n}. This Hopf pairing does not depend on the choice of basis.

Proof.

Notice that Ckji=(adxk)ji, the (i,j)-th matrix element of adxk, hence x~,𝒢ji=(adx)ji for all x~𝔥.

Uniqueness. If such a pairing exists, then, x𝔥, 0=ϵU(𝔥Lie)(x~)=x~,1, hence

0 =x~,δji1=x~,m𝒢mi𝒢¯jm=ΔU(𝔥Lie)(x~),m𝒢mi𝒢¯jm
=mx~,𝒢mi1,𝒢¯jm+1,𝒢mix~,𝒢¯jm=m(adx~)miϵ(𝒢¯jm)+ϵ(𝒢mi)x~,𝒢¯mi.

Thus, x~,𝒢¯ji=m(adx~)miϵ(𝒢¯jm)=(adx~)ji and, in particular, x~k,𝒢¯ji=Ckji.

Denote 𝒪1+:=Span𝒌{𝒢ji}i,j=1n𝒪(Aut(𝔥)). By (9), Δ(𝒪1+)𝒪1+𝒪1+, hence x~k1x~km,𝒢ji=x~k1x~km,Δm1(𝒢ji) is a polynomial in expressions of the form x~k,𝒢sr. Similarly, x~k1x~km,𝒢¯ji are determined by x~k,𝒢¯sr.

For any vU(𝔥Lie), v,𝒢~j1i1𝒢~jmim=Δm1(v),𝒢~j1i1𝒢~jmim, where each 𝒢~jpip stands for either 𝒢jpip or 𝒢¯jpip. After expanding Δm1(v), the right-hand side is written in terms of expressions of the form x~k1x~km,𝒢~ji. Therefore, if such a pairing exists, it is unique.

Existence. We first consider the free commutative algebra Bn on the 2n2 generators, still denoted 𝒢ji,𝒢¯ji, and with the same rule (9) for a bialgebra structure (this is the bialgebra of regular functions on the variety of pairs of arbitrary n×n matrices). A unique pairing of 𝔥l with B is by Lemma 3.2 extending lxk,𝒢ji=Ckji, lxk,𝒢¯ji=Ckji by Leibniz rule, requiring that lxk, is a differentiation of Bn. We now want to show that there is an induced pairing between 𝔥l and the quotient Hopf algebra Bn/I=𝒪(Aut(𝔥)); functionals lxk, remain differentiations on the quotient. We need to show that the pairing restricted to 𝔥lI vanishes. The biideal of relations I has a generating set KI of all elements of the form l,mClmk𝒢il𝒢jm𝒢rkCijr, k𝒢ki𝒢¯jkδji or k𝒢¯ki𝒢jkδji. Observe that ϵ(s)=0 for all sKI. Thus for differentiation D=lxk, we obtain D(bs)=D(b)ϵ(s)+ϵ(b)D(s)=0 for all bB and sKI. Therefore if the pairing vanishes on KI then it vanishes on the ideal generated by KI.

Thus we need to check lxp,r,mCrmk𝒢ir𝒢jm=lxp,n𝒢nkCijn and lxp,j𝒢ji𝒢¯kj=lxp,δki=lxp,j𝒢¯ji𝒢kj for all i,j,k,p{1,,n}. The first equation is

rCrjkCpir+mCimkCpjm=nCpnkCijn,

which is left Leibniz identity [[xp,xi],xj]+[xi,[xp,xj]]=[xp,[xi,xj]] in terms of the structure constants. By using the differentiation rule, the second equation is simply lxp,𝒢ki+lxp,𝒢¯ki=0, which holds for generators by definition. Therefore, there is a well defined pairing (𝔥l𝒌)𝒪(Aut(𝔥))𝒌 such that 1,f=ϵ(f) for all f𝒪(Aut(𝔥)) and lx, is a differentiation of 𝒪(Aut(𝔥)) for all x𝔥, that is,

lx,fg=lx,fϵ(g)+ϵ(f)lx,g,x𝔥,f,g𝒪(Aut(𝔥)). (10)

This means that 𝔥l𝒌 is equipped with a comultiplication Δ such that Δ(lx)=1lx+lx1 and ϵ(lx)=0 and the pairing respects Δ: in the notation of Subsection 2.1, cc, restricts to a coalgebra map ϕ1:𝔥l𝒌𝒪(Aut(𝔥)) sending 1 to 1𝒪(Aut(𝔥))=ϵ𝒪(Aut(𝔥)). By Lemma 2.2 and the equivalence between Hopf pairings and bialgebra maps T(𝔥l)𝒪(Aut(𝔥)), we extend this pairing to a unique Hopf pairing T(𝔥l)𝒪(Aut(𝔥))𝒌; it is determined by the formula

lxi1lxim,f=lxi1lxim,Δm1(f).

Denote by Il the ideal in T(𝔥l) generated by ex,y:=l[x,y]lxly+lylx, x,y𝔥. By Lemma 2.4, U(𝔥Lie)T(𝔥l)/Il. Moreover, ΔT(𝔥l)(ex,y)=1ex,y+ex,y1 and ϵ(ex,y)=0, hence Il is a biideal. Clearly, U(𝔥Lie)T(𝔥l)/Il as Hopf algebras as well.

We now check that the ideal generators ex,y of Il are paired with every element of 𝒪(Aut(𝔥)) as 0. Relation l[xk,xn]+lxnlxk,𝒢ji=lxklxn,𝒢ji is equivalent to mCknmCmji+pCnpiCkjp=pCkpiCnjp, which restates the left Leibniz identity [[xk,xn],xj]+[xn,[xk,xj]]=[xk,[xn,xj]]. Similarly, l[xk,xn]+lxnlxk,𝒢¯ji=lxklxn,𝒢¯ji computes to the same identity. Since I is a biideal and 𝒪1:=Span𝒌{𝒢ji,𝒢¯ji}i,j=1n satisfies Δ𝒪(Aut(𝔥))(𝒪1)𝒪1𝒪1, we can apply Lemma 2.1, part (ii), for KH=𝒪1 to conclude that the pairing vanishes on the entire I𝒪(Aut(𝔥)). Therefore, there is a well defined Hopf pairing U(𝔥Lie)𝒪(Aut(𝔥))𝒌 satisfying x~k,𝒢ji=Ckji.

To show that the pairing does not depend on the choice of basis 𝐛, note that we started from a pairing ,𝐛 defined on 𝔥lSpan𝒌{𝒢𝐛ji,𝒢¯𝐛ji}i,j=1n𝔥lBn by lx,𝒢𝐛ji𝐛=ad(x)𝐛ji and lx,𝒢¯𝐛ji𝐛=ad(x)𝐛ji. For a base change by a numerical matrix T, lx,𝒢𝐛ji𝐛=(ad(x)𝐛)ji=m,lTmi(ad(x)𝐛)lmT1jl=m,lTmilx,𝒢𝐛lm𝐛T1jl=lx,(T𝒢𝐛T1)ji𝐛=lx,θ𝐛𝐛(𝒢𝐛ji)𝐛. Likewise for 𝒢¯. Induced pairing U(𝔥Lie)𝒪(Aut(𝔥))𝒌 is uniquely defined by the pairing on 𝔥lSpan𝒌{𝒢𝐛ji,𝒢¯𝐛ji}i,j=1n and by abstract properties of the extension. Thus, it respects bialgebra isomorphism θ𝐛𝐛 in the second argument. Once we quotient from Bn down to 𝒪(Aut(𝔥)), θ𝐛𝐛 becomes an identification ι𝐛1ι𝐛 (extending (5)), yielding the invariance.

Proposition 3.4.

Let 𝔥 be a right Leibniz 𝐤-algebra and 𝐛=(y1,,yn) a 𝐤-basis of 𝔥. Denote by Cjki structure constants determined from [yj,yk]=Cjkiyi, for j,k{1,,n}. Let 𝒢ji,𝒢¯ji,i,j{1,,n} be the generators of the algebra 𝒪(Aut(𝔥)) from Subsection 3.1. Denote by y~ the image of y𝔥 in 𝔥Lie.

Then there is a well defined and unique Hopf pairing

,:U(𝔥Lie)𝒪(Aut(𝔥))𝒌

such that y~k,𝒢ji=Cjki for all i,j,k{1,,n}. This Hopf pairing does not depend on the choice of basis.

Proof.

Notice that y~k,𝒢ji=(adryk)ji where adry:z[z,y] is the right adjoint action; thus the main difference from Proposition 3.3 is change of side.

The entire proof is analogous to the proof of Proposition 3.3, hence we skip it. One first observes that, if the pairing exists, y~k,𝒢¯ji𝐛=C𝐛jki must hold. We are presenting U(𝔥Lie) as T(𝔥r)/Ir from Lemma 2.5. Key calculations with elements lx,x𝔥, which in Proposition 3.3 boil down to the left Leibniz identity are now replaced by calculations with elements ry,y𝔥, (from Lemma 2.5) and boil down to the right Leibniz identity. For example, ryp,s,mCsmk𝒢is𝒢jm=ryp,n𝒢nkCijn is

sCsjkCips+mCimkCjpm=nCnpkCijn,

which is right Leibniz identity [[yi,yp],yj]+[yi,[yj,yp]]=[[yi,yj],yp].

Remark 3.5.

(Geometric origin of the pairing.) If 𝒌 is or and 𝔥 is a Lie algebra 𝔤 over 𝒌, then Aut(𝔤) is a linear Lie group and its Lie algebra is Der(𝔤). Differential dfid of function f𝒪(Aut(𝔤)) at the unit id of Aut(𝔤) is a linear functional on Tid(Aut(𝔤))Der(𝔤), and therefore dfidDer(𝔤). Let adX:𝔤𝔤, adX:Z[X,Z]. Then adXDer(𝔤).

We prove that the pairing U(𝔤)𝒪(Aut(𝔤))𝒌 from Proposition 3.3, in the case when 𝒌 is or and 𝔥 is a Lie algebra 𝔤, agrees on subset 𝔤𝒪(Aut(𝔤)) of its domain with the pairing ,:𝔤𝒪(Aut(𝔤))𝒌 defined by

X,f=dfid(adX), for X𝔤 and f𝒪(Aut(𝔤)).

First we check that indeed d(𝒢ji)id(adXk)=Ckji. The exponential map exp maps a neighborhood of 0 in Der(𝔤) to a neighborhood of id in Aut(𝔤). We have that

d(𝒢ji)id(adXk) =(adXk)(𝒢ji)(id)=limt0ddt𝒢ji(exp(tadXk))
=limt0ddt𝒢ji(r=0(tadXk)rr!)=limt0ddt(r=0(tadXk)rr!)ji
=limt0(r=1tr1(adXk)r(r1)!)ji=(adXk)ji=Ckji.

Similarly, one checks that d(𝒢¯ji)id(adXk)=Ckji, by using that exp(tadXk)1=exp(tad(Xk)). By linearity, we conclude that the pairings agree for all X𝔤 and generators 𝒢ji,𝒢¯ji,i,j{1,,n}. Since the pairing also has the property

X,fg=X1+1X,fg, for X𝔤 and f,g𝒪(Aut(𝔤)),

we conclude that they agree for all X𝔤 and f𝒪(Aut(𝔤)).

3.3 Main theorem

Theorem 3.6.

Let 𝔥 be a left Leibniz 𝐤-algebra with vector space basis 𝐛=(x1,,xn) and structure constants Cijk determined from [xi,xj]=kCijkxk, i,j{1,,n}. Let 𝒢ji,𝒢¯ji,i,j{1,,n} be the generators of the algebra 𝒪(Aut(𝔥)) from Subsection 3.1. Denote by x~ the image of x𝔥 in 𝔥Lie. Then the following holds.

  1. (i)

    Hopf pairing U(𝔥Lie)𝒪(Aut(𝔥))𝒌 from Proposition 3.3 induces a right Hopf action :U(𝔥Lie)𝒪(Aut(𝔥))U(𝔥Lie) by formula

    x~f:=x~(1),fx~(2), for x~U(𝔥Lie) and f𝒪(Aut(𝔥)), (11)

    which further induces the structure of a smash product algebra 𝒪(Aut(𝔥))U(𝔥Lie). This action and the smash product do not depend on the choice of basis 𝐛.

  2. (ii)

    There is a unique 𝒌-linear unital antimultiplicative map

    λ:U(𝔥Lie)𝒪(Aut(𝔥))U(𝔥Lie)

    such that

    λ(x~j)=i𝒢¯jix~i, for j{1,,n}. (12)

    Map λ does not depend on the choice of basis 𝐛.

  3. (iii)

    Elements of Imλ commute with elements of 1U(𝔥Lie) in 𝒪(Aut(𝔥))U(𝔥Lie).

  4. (iv)

    Map λ is a left 𝒪(Aut(𝔥))-coaction on U(𝔥Lie).

  5. (v)

    (U(𝔥Lie),,λ) is a braided commutative right-left Yetter–Drinfeld module algebra over 𝒪(Aut(𝔥)).

Proof.

(i) Every Hopf pairing induces a right Hopf action in this way. By Proposition 3.3, the pairing, hence also the action, does not depend on the choice of basis.

(ii) We prove that such λ exists. We first define auxiliary map λ~ as a linear map 𝔥l𝒪(Aut(𝔥))U(𝔥Lie) such that λ~(lxj)=i𝒢¯jix~i for j{1,,n}, then expand it to λ~:T(𝔥l)𝒪(Aut(𝔥))U(𝔥Lie) by antimultiplicativity and then we check that λ~(Il)={0}, where Il is the ideal in T(𝔥l) generated by l[x,y]lxly+lylx, x,y𝔥. We compute

λ~(lxilxj) =λ~(lxj)λ~(lxi)=(k𝒢¯jkx~k)(m𝒢¯imx~m)=k,m,p𝒢¯jk𝒢¯ip(x~k𝒢¯pm)x~m
=k,m,p𝒢¯jk𝒢¯ip(δpmx~kCkpm)x~m=k,m𝒢¯jk𝒢¯imx~kx~mk,m,p𝒢¯jk𝒢¯ipCkpmx~m
=(8)k,m𝒢¯jk𝒢¯imx~kx~mk,m𝒢¯mkCjimx~k.

Analogously, λ~(lxjlxi)=k,m𝒢¯im𝒢¯jkx~mx~kk,m𝒢¯kmCijkx~m. After subtracting,

λ~(lxilxjlxjlxi) =k,m𝒢¯im𝒢¯jk[x~k,x~m]k,m𝒢¯mkCjimx~k+k,m𝒢¯kmCijkx~m
=k,m𝒢¯im𝒢¯jk[xk,xm]~k,m𝒢¯mkCjimx~k+k,m𝒢¯kmCijkx~m
=k,m𝒢¯im𝒢¯jkCkmpx~pk,m𝒢¯mkCjimx~k+k,m𝒢¯kmCijkx~m
=mCjim𝒢¯mpx~pk,m𝒢¯mkCjimx~k+k,m𝒢¯kmCijkx~m
=k,m𝒢¯kmCijkx~m.

On the other hand, λ~(l[xi,xj])=λ~(pCijplxp)=p,mCijp𝒢¯pmx~m. Equality λ~(lxilxjlxjlxi)=λ~(l[xi,xj]) is now proven. Therefore, by quotienting the domain of λ~ by ideal Il, we induce a well defined map λ:U(𝔥Lie)𝒪(Aut(𝔥))U(𝔥Lie). Additionally, we note that clearly

λ(vz)=λ(z)λ(v), for all v,zU(𝔥Lie). (13)

To see that λ=λ𝐛 defined by λ𝐛(x~j)=i𝒢¯𝐛jix~i (12) does not depend on the basis 𝐛, we compute λ𝐛(x~j)=s𝒢¯𝐛jsx~s=s,i𝒢¯𝐛jsTsix~i=i,j,r,sTliT1sl𝒢¯𝐛rsTjrx~i=(5)i,l,m,sTli𝒢¯𝐛jlx~i=sTjiλ𝐛(x~i)=λ𝐛(sTjix~i)=λ𝐛(x~j).

(iii) First we check that λ(x~j) and x~k commute for all j,k{1,,n}.

x~kλ(x~j) =x~ki𝒢¯jix~i=i,m𝒢¯jm(x~k𝒢¯mi)x~i
=i,m𝒢¯jm(δmix~k+Cmki)x~i=m𝒢¯jm(x~kx~m+[x~m,x~k])=
=m𝒢¯jmx~mx~k=λ(x~j)x~k.

By using (13), it is easy to prove the claim inductively for all elements of U(𝔥Lie),

zλ(v)=λ(v)z, for all v,zU(𝔥Lie). (14)

(iv) Coaction axiom (Δid)λ=(idλ)λ on generators x~j, j{1,,n}, is apparent from definitions (9) and (12). Both sides of it evaluate to k,i𝒢¯jk𝒢¯kix~i. It is now sufficient to show that, if the coaction axiom is true for v,zU(𝔥Lie), then it is true for the product vzU(𝔥Lie). We compute

λ(zv) =(13)λ(v)λ(z)=v[1]v[0]z[1]z[0]=(14)v[1]z[1]z[0]v[0], (15)

from which it follows that, because v and z are assumed to satisfy the coaction axiom identity,

((idλ)λ)(vz) =v[1]z[1]λ(z[0]v[0])
=v[1]z[1]v[0][1]z[0][1]z[0][0]v[0][0]
=v[1](1)z[1](1)v[1](2)z[1](2)z[0]v[0]
=(v[1]z[1])(1)(v[1]z[1])(2)z[0]v[0]

and, on the other hand,

((Δid)λ)(zv) =(Δid)(v[1]z[1]z[0]v[0])
=(v[1]z[1])(1)(v[1]z[1])(2)z[0]v[0].

Counitality of λ is checked first for generators, ((ϵid)λ)(x~j)=iϵ(𝒢¯ji)x~i=x~j for every j{1,,n}, and then easily proven inductively by using formula (15).

(v) First, we prove the Yetter–Drinfeld property:

f(2)λ(vf(1))=λ(v)f, for all vU(𝔥Lie) and f𝒪(Aut(𝔥)).

It is 𝒌-linear both in v and in f, hence it is sufficient to show it for v and f being words in generators, by induction on the length of a word. For v=x~k and f=𝒢ji,

m𝒢jmλ(x~k𝒢mi) =m𝒢jmλ(δmix~k+Ckmi)=𝒢jiλ(x~k)+mCkmi𝒢jm,
λ(x~k)𝒢ji =m𝒢¯kmx~m𝒢ji=m,p𝒢¯km𝒢pi(x~m𝒢jp)=m,p𝒢¯km𝒢pi(δjpx~m+Cmjp)
=m𝒢¯km𝒢jix~m+m,p𝒢¯km𝒢piCmjp
=𝒢jiλ(x~k)+m,p𝒢piCmjp𝒢¯km=(7)𝒢jiλ(x~k)+mCkmi𝒢jm.

The Yetter–Drinfeld property for generators v=x~k and f=𝒢¯ji is proven analogously.

If the identity is true for some v,zU(𝔥Lie) and any f𝒪(Aut(𝔥)) of the form 𝒢ji, then it also holds for the product vz and all generators of 𝒪(Aut(𝔥)), because Δ(𝒢ji)=m𝒢mi𝒢jm and Δ(G¯ji)=m𝒢¯jm𝒢¯mi. Indeed,

f(2)λ((vz)f(1)) =f(3)λ((vf(1))(zf(2)))=(13)f(3)λ(zf(2))λ(vf(1))
=λ(z)f(2)λ(vf(1))=λ(z)λ(v)f=(13)λ(vz)f.

Therefore, by induction, the identity is true for all vU(𝔥Lie) and f being 𝒢ji or 𝒢¯ji.

If the identity holds for some f and g in 𝒪(Aut(𝔥)) and all vU(𝔥Lie), then it also holds for the product fg𝒪(Aut(𝔥)) and all vU(𝔥Lie), by

(fg)(2)λ(v(fg)(1)) =f(2)g(2)λ(v(f(1)g(1)))
=f(2)g(2)λ((vf(1))g(1))
=f(2)λ(vf(1))g=λ(v)fg.

We conclude inductively that the Yetter–Drinfeld property holds.

Next, the comodule algebra property is actually proven in (15), by using (14).

Finally, let us prove the braided commutativity property:

zλ(v)=vz, for all v,zU(𝔥Lie).

First we check this on generators. For any two j,k{1,,n} we have

x~ki𝒢¯jix~i=i(δjix~kx~iCkjix~i)=x~kx~j[x~k,x~j]=x~jx~k.

Next, we use induction on the length of the word acted on by λ(x~j) on the right, for every x~j,j{1,,n}. The step of induction is

(vz)i𝒢¯jix~i=i,m(v𝒢¯jm)(z𝒢¯mi)x~i=i,m(v𝒢¯jm)x~mz=x~jvz,v,zU(𝔥Lie).

At last, the step of induction on the length of the word on the right is

wλ(zv)=(wλ(v))λ(z)=(vw)λ(z)=(zv)w,w,z,vU(𝔥Lie).

Theorem 3.7.

Let 𝔥 be a right Leibniz 𝐤-algebra with 𝐤-basis 𝐛=(y1,,yn) and structure constants Cijk determined from [yi,yj]=kCijkyk, i,j{1,,n}. Let 𝒢ji,𝒢¯ji,i,j{1,,n} be the generators of the algebra 𝒪(Aut(𝔥)) from Subsection 3.1. Denote by y~ the image of y𝔥 in 𝔥Lie.

Then the Hopf pairing U(𝔥Lie)𝒪(Aut(𝔥))𝐤 defined in Proposition 3.4 induces a right Hopf action :U(𝔥Lie)𝒪(Aut(𝔥))U(𝔥Lie) by formula

y~f:=y~(1),fy~(2), for y~U(𝔥Lie) and f𝒪(Aut(𝔥)),

which further induces the structure of a smash product algebra 𝒪(Aut(𝔥))U(𝔥Lie).

Then there also exists a unique 𝐤-linear unital antimultiplicative map λ:U(𝔥Lie)𝒪(Aut(𝔥))U(𝔥Lie) such that

λ(y~j)=i𝒢jiy~i, for j{1,,n}.

This unique map λ is a left coaction.

Furthermore, (U(𝔥Lie),,λ) is a braided commutative right-left Yetter–Drinfeld 𝒪(Aut(𝔥))-module algebra. Maps and λ do not depend on the choice of basis 𝐛.

Proof.

Analogous to the proof of Theorem 3.6.

References

  • [1] G. Böhm, Hopf algebroids, in Handbook of Algebra, Vol. 6, edited by M. Hazewinkel, Elsevier 2009, 173–236, arXiv:0805.3806.
  • [2] T. Brzeziński, G. Militaru, Bialgebroids, ×A-bialgebras and duality, J. Alg. 251 (2002) 279–294 math.QA/0012164
  • [3] P. Cartier, A primer of Hopf algebras, Frontiers in number theory, physics, and geometry II, 537–615, Springer 2007.; preprint IHÉS M-06-40, 2006.
  • [4] J.-L. Loday, T. Pirashvili, Universal enveloping algebras of Leibniz algebras and (co)homology, Math. Ann. 296 (1993) 139–158.
  • [5] J-H. Lu, Hopf algebroids and quantum groupoids, Int. J. Math. 7 (1996) 47–70, q-alg/9505024
  • [6] S. Majid, Foundations of quantum group theory, Cambridge University Press 1995.
  • [7] S. Meljanac, Z. Škoda, Leibniz rules for enveloping algebras, arXiv:0711.0149
  • [8] S. Meljanac, Z. Škoda, M. Stojić, Lie algebra type noncommutative phase spaces are Hopf algebroids, Lett. Math. Phys. 107:3, (2017) 475–503 arXiv:1409.8188
  • [9] D. E. Radford, Hopf algebras, World Scientific 2012.
  • [10] D. E. Radford, J. Towber, Yetter–Drinfeld categories associated to an arbitrary bialgebra, J. Pure Appl. Algebra 87 (1993) 259–279
  • [11] M. Stojić, Scalar extension Hopf algebroids, Journal of Algebra and its Applications (2023) arXiv:2208.11696
  • [12] M. Stojić, Completed Hopf algebroids, doctoral dissertation in Croatian language (of the title Upotpunjeni Hopfovi algebroidi), University of Zagreb (2017)
  • [13] Z. Škoda, Heisenberg double versus deformed derivatives, Int. J. Mod. Phys. A 26, Nos. 27 & 28 (2011) 4845–4854. arXiv:0806.0978
  • [14] Z. Škoda, M. Stojić, Hopf algebroids with balancing subalgebra, J. Alg. 598 (2022) 445–469 arXiv:1610.03837
  • [15] M. Stojić, Z. Škoda, Examples of scalar extension Hopf algebroids over a universal enveloping algebra, preprint.

Authors’ Addresses:

Zoran Škoda, Department of Teachers’ Education, University of Zadar, Franje Tudjmana 24, 23000 Zadar, Croatia zskoda@unizd.hr

Martina Stojić, Department of Mathematics, University of Zagreb, Bijenička cesta 30, 10000 Zagreb, Croatia stojic@math.hr